1: \documentclass[12pt,reqno]{amsart}
2: \usepackage{amssymb}
3:
4: %\input iso.def
5: \input epsf.tex
6: \newdimen\xsize
7: \newdimen\oldbaselineskip
8: \newdimen\oldlineskiplimit
9: \xsize=.7\hsize
10: \def\nolineskip{\oldbaselineskip=\baselineskip\baselineskip=0pt%
11: \oldlineskiplimit=\lineskiplimit\lineskiplimit=0pt}
12: \def\restorelineskip{\baselineskip=\oldbaselineskip%
13: \lineskiplimit=\oldlineskiplimit}
14: %
15: \def\putm[#1][#2]#3{% #1 X-axis #2 Y-axis(downwords) #3 $"mathobject"$
16: \hbox{\vbox to 0pt{\parindent=0pt%
17: \vskip#2\xsize\hbox to0pt{\hskip#1\xsize $#3$\hss}\vss}}}%
18: %\enddef\putm
19: %
20: %\def\putm[#1][#2]#3{% #1 X-axis #2 Y-axis(downwords) #3 $"mathobject"$
21: %\vbox to 0pt{\noindent\hskip#1\xsize\lower#2\xsize\hbox to0pt{$#3$\hss}\vss}}%
22: %\enddef\putm
23: %
24: \def\putt[#1][#2]#3{% #1 X-axis #2 Y-axis(downwords) #3 "text"
25: \vbox to 0pt{\noindent\hskip#1\xsize\lower#2\xsize%
26: \vtop{\restorelineskip#3}\vss}}
27: %\enddef\putt
28:
29:
30: %
31: \font\Bigbf=cmbx12 scaled 1200
32: \font\bigbf=cmbx12
33: %
34: \thinmuskip = 2mu
35: \medmuskip = 2.5mu plus 1.5mu minus 2.1mu %minus 2.5mu
36: \thickmuskip = 4mu plus 6mu
37: %
38: %\font\teneusm=eusm10
39: %\font\seveneusm=eusm7
40: %\font\fiveeusm=eusm5
41: %\newfam\eusmfam
42: %\textfont\eusmfam=\teneusm
43: %\scriptfont\eusmfam=\seveneusm
44: %\scriptscriptfont\eusmfam=\fiveeusm
45: %\def\cal#1{{\fam\eusmfam\relax#1}}
46: %
47: %\def\cal#1{{\mathcal{#1}}}
48:
49:
50:
51: %\font\tenscr=rsfs10 % scaled \magstep1
52: %\font\sevenscr=rsfs7 % scaled \magstep1
53: %\font\fivescr=rsfs5 % scaled \magstep1
54: %\skewchar\tenscr='177 \skewchar\sevenscr='177 \skewchar\fivescr='177
55: %\newfam\scrfam
56: %\textfont\scrfam=\tenscr
57: %\scriptfont\scrfam=\sevenscr
58: %\scriptscriptfont\scrfam=\fivescr
59: %\def\scr#1{{\fam\scrfam \tenscr{#1}}}
60: \DeclareFontFamily{U}{rsf}{\skewchar\font'177}%
61: \DeclareFontShape{U}{rsf}{m}{n}{<-6>rsfs5<6-8>rsfs7<8->rsfs10}{}%
62: \DeclareFontShape{U}{rsf}{b}{n}{<-6>rsfs5<6-8>rsfs7<8->rsfs10}{}%
63: \DeclareMathAlphabet\RSFS{U}{rsf}{m}{n}
64: \SetMathAlphabet\RSFS{bold}{U}{rsf}{b}{n}
65: \DeclareOption{psamsfonts}{%
66: \DeclareFontFamily{U}{rsf}{}%
67: \DeclareFontShape{U}{rsf}{m}{n}{<-6>rsfs5<6-8>rsfs7<8->rsfs10}{}%
68: \DeclareFontShape{U}{rsf}{b}{n}{<-6>rsfs5<6-8>rsfs7<8->rsfs10}{}%
69: }
70: \ProcessOptions
71:
72: \def\rfs#1{{\RSFS{#1}}}
73: \let\scr=\rfs
74: %\let\scr=\mathcal
75: %\def\scr#1{{\mathcal{#1}}}
76:
77:
78:
79: %
80: %\font\tenmib=cmmib10
81: %\font\sevenmib=cmmib7
82: %\font\fivemib=cmmib5
83: %\newfam\mibfam
84: %\textfont\mibfam=\tenmib
85: %\scriptfont\mibfam=\sevenmib
86: %\scriptscriptfont\mibfam=\fivemib
87: %\def\mib{\fam\mibfam}
88: \def\mib#1{\boldsymbol{#1}}
89:
90: %\font\tensl=cmsl10
91: %\font\sevensl=cmsl7 %cmsl10 scaled 694
92: %\font\fivesl=cmsl5 %cmsl10 scaled 482
93: %
94: %\font\sevensl=cmsl10 scaled 694
95: %\font\fivesl=cmsl10 scaled 482
96: %\newfam\slfam
97: %\textfont\slfam=\tensl
98: %\scriptfont\slfam=\sevensl
99: %\scriptscriptfont\slfam=\fivesl
100: %\def\sl{\fam\slfam\tensl}
101:
102:
103: %
104: %\font\tensf=cmss10
105: %\font\sevensf=cmss8 scaled 833
106: %\font\fivesf=cmss8 scaled 578
107: %\newfam\sffam
108: %\textfont\sffam=\tensf
109: %\scriptfont\sffam=\sevensf
110: %\scriptscriptfont\sffam=\fivesf
111: \def\sf#1{{\mathsf{#1}}}%\def\sf{\fam\sffam}
112:
113: %
114: %\font\tenfr=eufm10 scaled 1200
115: %\font\sevenfr=eufm7 scaled 1200
116: %\font\fivefr=eufm5 scaled 1200
117: %\newfam\frfam
118: %\textfont\frfam=\tenfr
119: %\scriptfont\frfam=\sevenfr
120: %\scriptscriptfont\frfam=\fivefr
121: %\def\fr#1{{\mathfrak{#1}}
122:
123:
124:
125: %
126: %\font\mathnine=cmmi9
127: %\font\rmnine=cmr9
128: %\font\cmsynine=cmsy9
129: %\font\cmexnine=cmex9 %cmex10 scaled 913
130: %\font\ninesl=cmsl9
131: %\font\ninesf=cmss9
132: %\font\ninemib=cmmib9
133: \def\msmall#1{\mathchoice{\hbox{\small$\displaystyle {#1}$}}{#1}{#1}{#1}}
134: %\def\msmall#1{\hbox{$\displaystyle
135: %\textfont0=\rmnine \textfont1=\mathnine
136: %\textfont2=\cmsynine \textfont3=\cmexnine
137: %\textfont\slfam=\ninesl \font\tensf=cmss9
138: %\font\tenmib=cmmib9
139: %{#1}$}}
140:
141:
142: %\font\mathnine=cmmi9
143: %\font\rmnine=cmr9
144: %\font\cmsynine=cmsy9
145: %\font\cmexnine=cmex10 scaled 905
146: %\def\msmall#1{\hbox{$\displaystyle \textfont0=\rmnine
147: %\textfont1=\mathnine \textfont2=\cmsynine \textfont3=\cmexnine
148: %{#1}$}}
149: %
150: \hyphenation{Lip-schit-zian Lip-schitz com-pact-ness}
151: %
152: \let\3=\ss
153: \def\cc{{\mathbb C}}
154: \def\jj{{\mathbb J}}
155: \def\rr{{\mathbb R}}
156: \def\nn{{\mathbb N}}
157: \def\pp{{\mathbb P}}
158: \def\cp{\cc\pp}
159: \def\zz{{\mathbb Z}}
160: \def\sss{{\mathbb S}}
161: \def\ttt{{\mathbb T}}
162: \def\bfg{{\mathbf G}}
163: \def\cinf{C^\infty}
164: %
165: \def\arg{{\sf{Arg}}}
166: \def\aut{{\sf{Aut}}}
167: \def\loc{{\sf{loc}}}
168: \def\st{_{\mathsf{st}}}
169: %\def\st{\mathchoice{_\sf{st}}{_\sf{st}}%
170: %{_{\hbox{\fivesf st}}}{_{\hbox{\fivesf st}}}}
171: \def\area{\sf{area}}
172: \def\sfb{\sf{b}}
173: \def\br{\sf{Br}}
174: \def\cosh{\sf{cosh}}
175: \def\cos{\sf{cos}}
176: \def\coker{\sf{Coker}\,}
177: \def\codim{\sf{codim}\,}
178: \def\deg{\sf{deg}\,}
179: \def\diam{\sf{diam}\,}
180: \def\dim{\sf{dim}\,}
181: \def\dimc{\dim_\cc}
182: \def\dimr{\dim_\rr}
183: \def\dfrm{\sf{dfrm}}
184: \def\dist{\sf{dist}\,}
185: \def\endo{\sf{End}}
186: \def\endr{\sf{End}\vph_\rr}
187: \def\endrn{\endr(\cc^n)}
188: \def\endc{\sf{End}\vph_\cc}
189: \def\ev{\sf{ev}}
190: \def\gh{\sf{gh}}
191: \def\ind{\sf{ind}}
192: \def\sym{\sf{Sym}}
193: \def\sind{\sf{S\hbox{-}ind}\,}
194: \def\exp{\sf{exp}}
195: \def\gcd{\sf{gcd}}
196: \def\sfh{\sf{H}}
197: \def\hom{\sf{Hom}\,}
198: \def\homr{\sf{Hom}\vph_\rr}
199: \def\id{\sf{Id}}
200: \def\im{\sf{Im}\,}
201: \def\bint{{-}\mathchoice{\mkern-18mu}{\mkern-15mu}{\mkern-15mu}{\mkern-15mu}\int}
202: \def\sfl{\sf{L}}
203: \def\re{\sf{Re}\,}
204: \def\ker{\sf{Ker}\,}
205: \def\lim{\mathop{\sf{lim}}}
206: \def\limsup{\mathop{\sf{lim\,sup}}}
207: \def\log{\sf{log}\,}
208: \def\max{\sf{max}}
209: \def\min{\sf{min}}
210: \def\nod{{\sf{nod}}}
211: \def\ord{\sf{ord}}
212: \def\osc{\sf{osc}}
213: \def\op{\sf{op}}
214: \def\pr{\sf{pr}}
215: \def\pt{\sf{pt}}
216: \def\rank{\sf{rank}}
217: \def\res{\sf{Res}}
218: \def\sinh{\sf{sinh}}
219: \def\sin{\sf{sin}}
220: %\def\spec{\sf{Sp}}
221: \def\supp{\sf{supp}\,}
222: \def\sup{\sf{sup}\,}
223: \def\frs{{\frak S}}
224: \def\vol{\sf{Vol}}
225: \def\sfz{\sf{Z}}
226:
227:
228: \def\bfone{\boldsymbol{1}}
229: \def\mbfe{{\mib{e}}}
230: \def\mbfk{{\mib{k}}}
231: \def\mbfl{{\mib{l}}}
232: \def\mbfp{{\mib{p}}}
233: \def\mbfs{{\mib{s}}}
234: \def\mbfu{{\mib{u}}}
235: \def\mbfv{{\mib{v}}}
236: \def\mbfw{{\mib{w}}}
237: \def\mbfx{{\mib{x}}}
238: \def\mbfz{{\mib{z}}}
239:
240: \def\vpp{\,\,\vec{\vphantom{p}}\!\!p'{}}
241: \def\vdp{\,\,\vec{\vphantom{d}}\!\!d'{}}
242:
243: \def\eps{\varepsilon}
244: \def\epsi{\varepsilon}
245: \let\vkappa=\varkappa
246: \def\<{\langle}\let\la=\<
247: \def\>{\rangle}\let\ra=\>
248: \def\bss{\backslash} \let\bs=\bss
249: \def\comp{\Subset}
250: \def\d{\partial}
251: \def\dbar{{\barr\partial}}
252: \def\dbarj{\dbar_J}
253: \def\ddef{\mathrel{{=}\raise0.3pt\hbox{:}}}
254: \def\deff{\mathrel{\raise0.3pt\hbox{\rm:}{=}}}
255: \def\inv{^{-1}}
256: \def\hook{\hookrightarrow}
257: \def\fraction#1/#2{\mathchoice{{\msmall{ #1\over#2}}}%
258: {{ #1\over #2 }}{{#1/#2}}{{#1/#2}}}
259: \def\norm#1{\Vert #1 \Vert}
260: \def\half{{\fraction1/2}}
261: \def\le{\leqslant}
262: \def\vph{^{\mathstrut}}
263: \def\lrar{\longrightarrow}
264: \def\emptyset{\varnothing}
265: \def\scirc{\mathop{\mathchoice{\hbox{\small$\circ$}}{\hbox{\small$\circ$}}%
266: {{\scriptscriptstyle\circ}}{{\scriptscriptstyle\circ}}}}
267: %
268: \def\longpoints{\leaders\hbox to 0.5em{\hss.\hss}\hfill \hskip0pt}
269: \def\stateskip{\smallskip}
270: \def\state#1. {\stateskip\noindent{\bf#1. }} %\medskip
271: \def\statep#1. {\stateskip\noindent{\bf#1 }} %\medskip
272: \def\step#1{{\sl Step #1)}}
273: \def\proof{\state Proof. \.}
274: \def\Chi{\raise 2pt\hbox{$\chi$}}
275: \def\eg{\hskip1pt plus1pt{\sl{e.g.\/\ \hskip1pt plus1pt}}}
276: \def\ie{\hskip1pt plus1pt{\sl i.e.\/\ \hskip1pt plus1pt}}
277: \def\iff{if and only if }
278: %{\hskip1pt plus1pt{\sl iff\/\hskip2pt plus1pt }}
279: \def\wrt{with respect to }
280: \def\isl{{\mathrm i}}
281: \def\sli{{\sl i)} } \def\slip{{\sl i)}}
282: \def\slii{{\sl i$\!$i)} } \def\sliip{{\sl i$\!$i)}}
283: \def\sliii{{\sl i$\!$i$\!$i)} }\def\sliiip{{\sl i$\!$i$\!$i)}}
284: \def\sliv{{\sl i$\!$v)} } \def\slivp{{\sl i$\!$v)}}
285: \def\slv{{\sl v)} } \def\slvp{{\sl v)}}
286: \def\slvi{{\sl v$\!$i)} }
287: \def\reg{^\sf{reg}}
288: \def\sing{^\sf{sing}}
289: \def\mat{\sf{Mat}}
290: \def\matn{\mat(n,\cc)}
291: \def\diff{{\scr D}\mskip -2mu i\mskip -5mu f\mskip-6.3mu f}
292: \def\star{\mathop{\msmall{*}}}
293: \def\barr#1{\mskip1mu\overline{\mskip-1mu{#1}\mskip-1mu}\mskip1mu}
294: %\def\barr#1{\overline{#1}\vphantom{#1}}
295: \def\usig{U\!\bfsig}
296:
297: %
298: \def\Chi{\raise 2pt\hbox{$\chi$}}
299: \def\yps{\Upsilon}
300: \let\phI=\phi\let\phi=\varphi\let\varphi=\phI\def\vphi{\varphi}
301: %
302: \def\igam{{\mathchar"0100}} % Italic \Gamma
303: \def\idel{{\mathchar"0101}} % Italic \Delta
304: \def\ithe{{\mathchar"0102}} % Italic \Theta
305: \def\ilam{{\mathchar"0103}} % Italic \Lambda
306: \def\ixi{{\mathchar"0104}} % Italic \Xi
307: \def\ipi{{\mathchar"0105}} % Italic \Pi
308: \def\isig{{\mathchar"0106}} % Italic \Sigma
309: \def\iyp{{\mathchar"0107}} % Italic \Upsilon
310: \def\iphi{{\mathchar"0108}} % Italic \Phi
311: \def\ipsi{{\mathchar"0109}} % Italic \Psi
312: \def\iom{{\mathchar"010A}} % Italic \Omega
313:
314: %
315: \def\bfalpha{{\boldsymbol\alpha}}%
316: \def\bfbeta{{\boldsymbol\beta}}%
317: \def\bfgamma{{\boldsymbol\gamma}}%
318: \def\bfdelta{{\boldsymbol\delta}}%
319: \def\bfepslon{{\boldsymbol\varepsilon}}%
320: \def\bfvarepsilon{{\boldsymbol\epslon}}%
321: \def\bfzeta{{\boldsymbol\zeta}}%
322: \def\bfeta{{\boldsymbol\eta}}%
323: \def\bftheta{{\boldsymbol\theta}}%
324: \def\bfvartheta{{\boldsymbol\vartheta}}%
325: \def\bfiota{{\boldsymbol\iota}}%
326: \def\bfkappa{{\boldsymbol\varkappa}}%
327: \def\bfvarkappa{{\boldsymbol\kappa}}%
328: \def\bflambda{{\boldsymbol\lambda}} \let\bflamb=\bflambda%
329: \def\bfmu{{\boldsymbol\mu}}%
330: \def\bfnu{{\boldsymbol\nu}}%
331: \def\bfxi{{\boldsymbol\xi}}%
332: \def\bfpi{{\boldsymbol\pi}}%
333: \def\bfvarpi{{\boldsymbol\varpi}}%
334: \def\bfrho{{\boldsymbol\rho}}%
335: \def\bfvarrho{{\boldsymbol\varrho}}%
336: \def\bfsigma{{\boldsymbol{\sigma}}}
337: \def\bfsig{\bfsigma}%
338: \def\bfvarsigma{{\boldsymbol\varsigma}}%
339: \def\bftau{{\boldsymbol\tau}}%
340: \def\bfupsilon{{\boldsymbol\upsilon}}%
341: \def\bfphi{{\boldsymbol\varphi}}%
342: \def\bfvarphi{{\boldsymbol\phi}}%
343: \def\bfchi{{\boldsymbol\chi}}%
344: \def\bfpsi{{\boldsymbol\psi}}%
345: \def\bfomega{{\boldsymbol\omega}}%
346: \def\bfell{{\boldsymbol\ell}}%
347: %
348: \def\scra{\scr{A}}
349: \def\scrb{\scr{B}}
350: \def\scrc{\scr{C}}
351: \def\scrd{\scr{D}}
352: \def\scrh{\scr{H}}
353: \def\scre{\scr{E}}
354: \def\scrf{\scr{F}}
355: \def\scri{\scr{I}}
356: \def\scrj{\scr{J}}
357: \def\scrk{\scr{K}}
358: \def\scrl{\scr{L}}
359: \def\scrm{\scr{M}}
360: \def\whcalm{\wh{\scr M}}
361: \def\barm{{\mskip5mu\barr{\mskip-5mu\scr M}}{}}
362: \def\frm{{\frak{M}}}
363: \def\scrn{{\scr N}}
364: \def\scro{{\scr O}}
365: \def\scrp{{\scr P}}
366: \def\scrpp{{\scr{P\!\!P}}}
367: % \def\whcalp{\wh{\scr P}}
368: \def\scrq{{\scr Q}}
369: \def\scrs{{\scr S}}
370: \def\scrt{{\scr T}}
371: \def\scru{{\scr U}}
372: \def\scrv{{\scr V}}
373: \def\scrw{{\scr W}}
374: \def\scrx{{\scr X}}
375: \def\scry{{\scr Y}}
376: \def\scrz{{\scr Z}}
377:
378: \def\trans{\pitchfork}
379: \def\eps{\varepsilon}
380: \def\epsi{\varepsilon}
381: \def\bs{\backslash}
382: \def\ogran{{\hskip0.7pt\vrule height6pt depth2.5pt\hskip0.7pt}}
383: \def\comp{\Subset}
384: \def\d{\partial}
385: \def\dbar{{\barr\partial}}
386: \def\1{{1\mkern-5mu{\rom l}}}
387: \def\dbarj{\dbar_J}
388: \def\ge{\geqslant}
389: \def\inv{^{-1}}
390: \let\wh=\widehat
391: \let\wt=\widetilde
392: \def\fraction#1/#2{\mathchoice{{\msmall{ #1\over#2}}}%
393: {{ #1\over #2 }}{{#1/#2}}{{#1/#2}}}
394: \def\half{{\fraction1/2}}
395: \def\le{\leqslant}
396: \def\vph{^{\mathstrut}}
397: \def\emptyset{\varnothing}
398: \def\.{\thinspace}
399: \def\parnin{\par\noindent}
400:
401: \def\ti#1{{\tilde{#1}}}
402: \def\term#1#2{\underline{#2}_{[{#1}]}}
403:
404: \def\qed{\ \ \hfill\hbox to .1pt{}\hfill\hbox to .1pt{}\hfill $\square$\par}
405:
406: \def\comment#1\endcomment{}
407:
408: \abovedisplayskip=5pt plus3pt minus2pt
409: \belowdisplayskip=\abovedisplayskip
410:
411: %\show\abovedisplaylimit
412: %\def\o#1{\buildrel\circ \over{#1}}$\o u$
413:
414:
415: %
416: %%%
417:
418: \def\lineeqqno(#1){\hfill\llap{\vbox to 10pt%
419: {\vss\begin{align} \eqqno(#1)\end{align}\vss}}\vskip1pt}
420:
421: \textwidth= 6.3 truein
422: \textheight=9.4 truein
423: \vsize=9.5 truein
424: \voffset= -.5truein
425: \oddsidemargin 0.1in
426: \evensidemargin 0in
427: \advance\headheight 1.2pt
428:
429: \numberwithin{equation}{subsection}
430: %\renewcommand{\theequation}{\thesection\thesubsection\arabic{equation}}
431:
432:
433: \def\newsection[#1]#2{\section{#2}\label{sec:#1}}
434: \def\refsection#1{{\sl Section \ref{sec:#1}$\,$}}
435: \def\newsubsection[#1]#2{\subsection{#2}\label{sec:#1}\showlabel{\tt #1}}
436: \def\refsubsection#1{{\sl Paragraph \ref{sec:#1}$\,$}}
437:
438:
439:
440:
441: \def\inDex#1{\index{#1}}
442: \def\Zpage#1{ page\ #1}
443:
444: \newtheorem{thm}{Theorem}[subsection]
445: \def\newthm#1{\begin{thm} \showlabel{\tt #1}\label{#1}%
446: \index{{\sl Theorem \ref{#1}}, \ label {\tt #1}|Zpage}}
447: \def\refthm#1{{\sl Theorem \ref{#1}$\,$}}
448: \newtheorem{lem}[thm]{Lemma}
449: \def\newlemma#1{\begin{lem} \showlabel{\tt #1}\label{#1}%
450: \inDex{{\sl Lemma \ref{#1}}, \ label {\tt #1}|Zpage}}
451: \def\lemma#1{{\sl Lemma \ref{#1}$\,$}}
452: \newtheorem{prop}[thm]{Proposition}
453: \def\newprop#1{\begin{prop} \showlabel{\tt #1}\label{#1}%
454: \inDex{{\sl Proposition \ref{#1}}, \ label {\tt #1}|Zpage}}
455: \def\propo#1{{\sl Proposition \ref{#1}$\,$}}
456: \newtheorem{corol}[thm]{Corollary}
457: \def\newcorol#1{\begin{corol} \showlabel{\tt #1}\label{#1}%
458: \inDex{{\sl Corollary \ref{#1}}, \ label {\tt #1}|Zpage}}
459: \def\refcorol#1{{\sl Corollary \ref{#1}$\,$}}
460: \newtheorem{defi}{Definition}[subsection]
461: \def\newdefi#1{\begin{defi} \showlabel{\tt #1}\label{#1}\rm %
462: \inDex{{\sl Definition \ref{#1}}, \ label {\tt #1}|Zpage}}
463: \def\refdefi#1{{\sl Definition \ref{#1}$\,$}}
464:
465: \def\eqqno(#1){\label{eq#1}%
466: \index{{\sl Eqtn \ref{eq#1}}, \ label {\tt(#1)}|Zpage}}
467: \def\eqqref(#1){\eqref{eq#1}}
468: %\def\eeq{\end{equation}}
469:
470:
471: %\def\showlabel#1{}
472: \def\showlabel#1{}
473: % Uncomment this line if you want to see labels
474: %\def\showlabel#1{{\rm[}{\tt #1}{\rm]}}
475:
476:
477: \makeindex{}
478:
479: \begin{document}
480: \baselineskip=14.0pt plus 2pt
481:
482:
483:
484: \title[Pseudoholomorphic curves and the symplectic isotopy problem]%
485: %[Symplectic isotopy and pseudoholomorphic curves]%
486: {Pseudoholomorphic curves
487: \\[5pt]
488: and
489: the symplectic isotopy problem%
490: }
491: \author[V.~Shevchishin]{Vsevolod V.~Shevchishin}
492: \address{Fakult\"at f\"ur Mathematik\\
493: Ruhr-Universit\"at Bochum\\
494: Universit\"atsstrasse 150\\
495: 44780 Bochum\\
496: Germany}
497: \email{sewa@@cplx.ruhr-uni-bochum.de}
498: %\curraddr{}
499: \dedicatory{}
500: \subjclass{}
501: %\thanks{The author was supported in part by ...}
502: %\date{This version: October 2000.
503: %3rd version: March 2000.
504: %2nd version: October 1999.
505: %1st version: June 1999.}
506: \keywords{}
507: \begin{abstract}
508: The deformation problem for pseudoholomorphic curves and related
509: geometrical properties of the total moduli space of pseudoholomorphic curves
510: are studied. A sufficient condition for the saddle point property
511: of the total moduli space is established. The local symplectic isotopy problem
512: is formulated and solved for the case of imbedded pseudoholomorphic curves.
513: It is shown that any two symplectically imbedded surfaces $\Sigma_0, \Sigma_1
514: \subset \cp^2$ of the same degree $d\le 6$ are symplectically isotopic.
515: \end{abstract}
516: \maketitle
517: %\pagenumbering{roman}
518: \setcounter{tocdepth}{2}
519: %\setcounter{tocdepth}{1}
520: %\tableofcontents
521:
522:
523:
524: \setcounter{section}{-1}
525: %\pagebreak[1]
526:
527: \newsection[intro]{Introduction}
528:
529: \baselineskip =14.0pt plus .5pt
530: The symplectic isotopy problem can be formulated as follows:
531:
532: {\it For a given symplectic 4-dimensional manifold $(X,\omega)$ and symplectically
533: imbedded compact connected oriented surfaces $\Sigma_0, \Sigma_1 \subset X$
534: in the same homology class $[A] \in \sfh_2(X,\zz)$, does there exist an isotopy
535: $\{\Sigma_t\}_{t\in [0,1]}$ connecting $\Sigma_0$ with $\Sigma_1$ such that all\/
536: $\Sigma_t$ are also symplectically imbedded?}
537:
538: In this case $\Sigma_0$ and $\Sigma_1$ are called {\sl symplectically isotopic}.
539: Note that by the {\sl genus formula} for pseudoholomorphic curves (see
540: \refsection{1}) $\Sigma_0$ and $\Sigma_1$ have the same genus.
541: The example of Fintushel and Stern \cite{Fi-St} (see \refsection{6} for details)
542: shows that in general the answer is negative. On the other hand, Sikorav
543: \cite{Sk-3} gives an affirmative answer in the case of surfaces of positive degree
544: $d\le 3$ in $\cp^2$. So it is natural to ask under which conditions on $(X,\omega)$
545: and $[A] \in \sfh_2(X,\zz)$ the symplectic isotopy does exist.
546:
547: In this paper techniques involving pseudoholomorphic curves are developed in
548: directions needed for solving the symplectic isotopy problem. The main result
549: is the following
550:
551: \state Theorem 1. {\it Two symplectically imbedded compact connected oriented
552: surfaces $\Sigma_0, \Sigma_1 \subset \cp^2$ of the same positive degree $d\le 6$
553: are symplectically isotopic}.
554:
555: \smallskip
556: The proof is based on the solution of two problems which are closely related to
557: the symplectic isotopy problem and concern the geometry of the total moduli space
558: of pseudoholomorphic curves. The first result is a sufficient condition for the
559: {\it saddle point property} of the total moduli space of pseudoholomorphic
560: curves. This removes one of the obstacles to the existence of a symplectic isotopy
561: and is applied to solve the {\it local symplectic isotopy problem} for imbedded
562: pseudoholomorphic curves. The latter appears as a necessary part of the global
563: problem.
564:
565:
566: \smallskip
567: The obtained progress gives hope that the symplectic isotopy problem has
568: an affirmative solution in the case $c_1(X,\omega)[A] >0$. Note that compact
569: symplectic 4-manifolds with this property are classified: Up to the case when
570: $[A]$ is represented by an exceptional sphere, these are symplectic blow-ups of
571: a rational or ruled symplectic manifold, see \cite{McD-Sa-3}, {\sl Corollary 1.5},
572: and also \refsection{6}.
573:
574: \newsubsection[0.1]{Overview of main results} There is essentially only one known
575: method of constructing symplectic isotopies. Having its origin in Gromov's
576: celebrated article \cite{Gro}, it utilizes moduli spaces of pseudoholomorphic
577: curves. One fixes a homotopy $h(t) \deff J_t$, $t\in [0,1]$, of
578: $\omega$-tame almost complex structures and considers the relative moduli space,
579: $$
580: \scrm_h \deff \bigr\{ (C,t) :
581: C \text{ is an imbedded $J_t$-holomorphic
582: curve in the homology class }[A]
583: \bigl\},
584: $$
585: equipped with a natural topology and with the projection $\pi_h: (C,t) \in \scrm_h
586: \mapsto t \in [0,1]$. It follows essentially from \cite{Gro} that for a generic
587: homotopy $h(t) = J_t$ the space $\scrm_h$ is a smooth manifold of expected
588: dimension $\dimr\scrm_h = 2c_1(X,\omega)[A] + 2g-1$ and the projection $\pi_h$ is
589: also smooth. Moreover, if this dimension is positive, then there exists a generic
590: path $h(t)=J_t$ such that the original surface $\Sigma_0$ (resp.\ $\Sigma_1$) is
591: $J_0$-holomorphic (resp.\ $J_1$-holomorphic) so that $(\Sigma_i, i) \in \scrm_h$
592: for $i=0,1$. Using a trivial but crucial observation that for every $(C,t) \in
593: \scrm_h$ the curve $C$ is an $\omega$-symplectic real surface, one tries to find
594: the desired isotopy $\Sigma_t$ by constructing a continuous section $\sigma: t
595: \in [0,1] \mapsto (\Sigma_t, t) \in \scrm_h$ connecting $(\Sigma_0,0)$ with
596: $(\Sigma_1,1)$.
597:
598: One can easily see possible obstacles to the existence of such a section $\sigma:
599: [0,1]\to \scrm_h$. The first one is that the projection $\pi_h: \scrm_h \to [0,1]$,
600: considered as a function, can have local maxima and minima. Indeed, if some
601: $(C^*, t^*)\in \scrm_h$ appears as a local maximum of $\pi_h$, then for all $t>
602: t^*$ there exists no $J_t$-holomorphic curves sufficiently close to $C^*$. Observe
603: that the mere fact of existence of a $J_t$-holomorphic curve $C_t$
604: for $t>t^*$ does not help much, because this does not imply that such a curve $C_t$
605: is symplectically isotopic to $\Sigma_0$ or $\Sigma_1$. Note that exactly the
606: existence of $J$-holomorphic curves is the main technical tool in the Gromov's
607: article \cite{Gro}.
608:
609: On the other hand, this obstacle does not appear in the case when the projection
610: $\pi_h$ has the following {\sl saddle point property}: the Hesse matrix
611: of $\pi_h$ at any critical point has at least one positive and at least one negative
612: eigenvalues. In \refsection{3} we prove
613:
614: \state Theorem 2. {\it Assume that $c_1(X,\omega)[A]>0$. Then for a generic
615: homotopy $h(t)$ all critical points of the projection $\pi_h: \scrm_h \to
616: [0,1]$ are saddle}.
617:
618: \smallskip
619: On the other hand, we also show that in the case $c_1(X,\omega)[A]\le 0$ local
620: maxima and minima of the projection $\pi_h: \scrm_h \to [0,1]$ do exist in the case
621: of an apropriate generic choice of $h$. Note that in the example of Fintushel
622: and Stern \cite{Fi-St} one has $c_1(X,\omega)[A]= 0$.
623:
624: \medskip
625: The next obstacle to existence of the desired section $\sigma: [0,1]\to \scrm_h$
626: comes from the fact that in general $\scrm_h$ is not compact and the projection
627: $\pi_h$ is not proper. This means that while attempting to construct the section
628: $\sigma: [0,1]\to \scrm_h$ we go to ``infinity'' in the space $\scrm_h$. The
629: Gromov compactness theorem provides control on the limiting behavior of curves
630: $C_t$ in this case. It says that some subsequence, say $C_{t_n}$, converges in
631: a certain weak sense to a pseudoholomorphic curve $C^*$ in the same homology class
632: $[A]$. The curve $C^*$ can have several irreducible components, some of them
633: multiple, and also several singular points. Thus we come to a problem of
634: describing symplectic isotopy classes of imbedded pseudoholomorphic
635: curves close to a given singular curve $C^*$.
636:
637: Here we have essentially two different difficulties. The first one comes from
638: multiple components. At the moment, we have no remedy for this problem. So we
639: appempt to avoid the appearence of multiple components. This is done
640: by imposing the following additional constraint. We consider the curves $C_t$
641: which pass through fixed points $\mbfx=(x_1,\ldots,x_k)$ on $X$. This means,
642: however, that now we must consider a new moduli space $\scrm_{h,\mbfx}$ with a
643: new projection $\pi_{h,\mbfx}: \scrm_{h,\mbfx} \to [0,1]$. In \refsection{3}
644: we also show that if the number $k$ of the fixed points is strictly less than
645: $c_1(X,\omega)[A]$, then the saddle point property for $\pi_{h,\mbfx}$ remains
646: valid. An easy calculation shows that fixing $k=3d-1$ generic points on $\cp^2$,
647: we can avoid the appearance of multiple components in pseudoholomorphic curves
648: of degree $d\le6$. This explains the restriction in the main theorem.
649:
650: \smallskip
651: In the case when $C^*$ has no multiple components we still have to describe the
652: possible symplectic isotopy classes of imbedded pseudoholomorphic curves close to
653: $C^*$. Recall that by the result of Micallef and White (see \refsubsection{1.2})
654: every singular point of $C^*$ is isolated and topologically equivalent to a
655: singular point of a usual holomorphic curve. In this way we come to the {\sl local
656: symplectic isotopy problem} which asks about possible symplectic isotopy types of
657: imbedded pseudoholomorphic curves in a neighborhood of a given isolated
658: singularity, see \refsubsection{6.2} for details.
659:
660: The solution of the local symplectic isotopy problem is based on the simple
661: observation that for {\sl holomorphic curves} this problem has a trivial solution.
662: Namely, a generic holomorphic deformation of a given (local) holomorphic curve
663: $C^*$ gives a non-singular curve $C$, and the set of such curves is open and
664: connected. Using this fact, we expoit essentially the same method as in the case
665: of the global problem and prove that it is possible to deform isotopically an
666: imbedded pseudoholomorphic curve $C$ sufficiently close to a given singular curve
667: $C^*$ into a genuine holomorphic curve. In this way we prove
668:
669: \state Theorem 3. {\it There exists a unique symplectic isotopy class of
670: non-singular pseudoholomorphic curves which are close to a given pseudoholomorphic
671: curve $C^*$ without multiple components}.
672:
673: \smallskip
674: It should be noted that in the proof the saddle point property from {\sl Theorem
675: 2\/} is used in an essential way. {\sl Theorem 1\/} follows now by
676: the procedure of avoiding multiple components as it is explained above.
677:
678: \medskip
679: Further technical results of the paper are as follows. \refsection{4} is devoted
680: to the deformation problem of pseudoholomorphic maps with prescribed singularities.
681: It is shown that the subspace of such maps is an immerced submanifold of expected
682: codimension in the total moduli space of pseudoholomorphic maps. In \refsection{3}
683: the second variation of the $\dbar$-equation is computed. The result establishes
684: the relationship between the geometry of a pseudoholomorphic curve $C^*$
685: corresponding to a critical point of the projection $\pi_h: \scrm_h \to [0,1]$
686: and the eigenvalues of the Hesse matix $d^2\pi_h$ at this point. Combined with
687: transversality results, this yields the proof of {\sl Theorem 2}. Finally,
688: in \refsection{5} the problem of smoothing of nodal points on pseudoholomorphic
689: curves is studied.
690:
691:
692: \bigskip
693: {\it Acknowledgements.}
694: The author would like to express his gratitude
695: to A.~Huckle\-berry, S.~Ivash\-ko\-vich,
696: St.~Ne\-mi\-rov\-ski, St.~Orev\-kov, B.~Sie\-bert and J.-C.~Siko\-rav for
697: numerous useful conversations, suggestions and remarks.
698:
699: %\newpage
700: %\tableofcontents
701: %\newpage
702:
703: %\bigskip
704: \newsection[1]{Deformation and the normal sheaf of pseudoholomorphic curves}
705:
706: In this section we give a brief description of pseudoholomorphic %(ps.-hol.)
707: curves and the related deformation theory.
708:
709: \newsubsection[1.1]{Pseudoholomorphic curves}
710: First we collect some facts from the Gromov's theory.
711: %It was introduced as a tool for studying symplectic manifolds in \cite{Gro},
712: %and appears to be very effective.
713: Since there are several books devoted to or
714: treating this theme (see \eg \cite{McD-Sa-1} or \cite{McD-Sa-2})
715: we only mention the basic definitions and results we shall use later.
716:
717: \newdefi{def1.1.0}
718: An {\sl almost complex structure} on a manifold $X$ is an
719: endomorphism $J \in TM$ of the tangent bundle such that $J^2 = - \id$.
720: The pair $(X,J)$ is called an {\sl almost complex manifold}.
721: One of the most important classes of such manifolds appears in symplectic
722: geometry. An almost complex structure on a symplectic manifold $(X,\omega)$ is
723: called {\sl$\omega$-tame} if $\omega(v,Jv)>0$ for any non-zero tangent vector
724: $v$. It is well-known that the set $\scrj_\omega$ of $\omega$-tame almost complex
725: structures is non-empty and contractible, (see \eg \cite{Gro},
726: \cite{McD-Sa-1}, or \cite{McD-Sa-2}). In
727: particular, any two $\omega$-tame almost complex structures $J_0$ and $J_1$ can be
728: connected by a homotopy (path) $J_t, t\in [0,1],$ inside $\scrj_\omega$.
729: \end{defi}
730:
731: \newdefi{def1.1.1} A {\sl parameterized $J$-holomorphic curve} in an almost
732: complex manifold $(X,J)$ is given by a (connected) Riemann surface $S$ with a
733: complex structure $J_S$ on $S$ and a non-constant $C^1$-map $u: S \to X$
734: satisfying the Cauchy-Riemann equation
735: \begin{equation}
736: du + J \scirc du \scirc J_S =0.
737: \eqqno(1.1.1)
738: \end{equation}
739: In this case we call $u$ a $(J_S, J)$-holomorphic map, or simply
740: $J$-holomorphic map. Here we use the fact that if $u$ is not constant, then
741: such a structure $J_S$ is unique. We shall also use the notion {\sl $J$-curve}
742: which, depending on the context, will mean a map $u:S \to X$, \ie a
743: parameterized curve in $X$, or an image $u(S)$ of $J$-holomorphic map, taken
744: with appropriate multiplicity, \ie a {\sl non-parameterized $J$-curve}.
745: \end{defi}
746:
747:
748: The equation \eqqref(1.1.1) is elliptic with the Cauchy-Riemann symbol. This
749: provides regularity properties for $u$. In particular, $u$ is H\"older
750: $C^{l+1,\alpha}$-smooth, $u\in C^{l+1,\alpha}(S,X)$, provided $J\in C^{l,\alpha}$
751: with integer $l\ge 1$ and $0< \alpha <1$. To simplify the notations we set $\ell
752: \deff l+ \alpha$ and write $C^\ell$ to indicate $C^{l,\alpha}$-smoothness.
753: In what follows we shall assume that almost complex structures $J$ on $X$ are
754: $C^\ell$-smooth for some fixed sufficiently big non-integer $\ell$.
755:
756:
757:
758: An easy consequence of the tameness condition is that any $J$-holomorphic
759: {\sl imbedding} $u: S \to X$ with $J \in \scrj_\omega$ is {\sl symplectic}
760: \ie the pull-back $u^*\omega$ is non-degenerate on $S$. The converse is also
761: true: Any $C^{\ell+1}$-smooth symplectic imbedding $u:S \to X$ with $\ell >1$
762: is $J$-holomorphic for some $C^\ell$-smooth $\omega$-tame structure
763: $J$. For {\sl immersions} the situation is more complicated. We state a result
764: in a setting which will be relevant later on
765: (see, \eg \cite{Gro} or \cite{McD-Sa-2} for details).
766: %We state a related result only for the special case we are interested in.
767:
768: \newlemma{lem1.1.1}
769: Let $(X,\omega)$ be a symplectic manifold with $\dim_\rr X =4$, and $u: S \to X$
770: an $\omega$-symplectic $C^1$-map such that $u(S)$ has only simple transversal
771: {\sl positive} self-intersection points.
772:
773: Then there exist an $\omega$-tame almost complex structure $J$ on $X$ and a complex
774: structure $J_S$ on $S$ and making $u$ a $J$-holomorphic map.
775: \end{lem}
776:
777: %The claim is standard, see \eg \cite{Gro} or \cite{McD-Sa-2} for details.
778: It is worth to make the following remark. If $x \in X$ is a self-intersection
779: point of $u$, $x= u(z_1)= u(z_2)$ with $z_1 \not= z_2$, such that the tangent
780: planes $du(T_{z_i}) \subset T_xX$ are transversal and {\sl complex} \wrt some
781: structure $J_x$ in $T_xX$, then the intersection index of planes $du(T_{z_i})$
782: in $x$ is {\sl positive}. However, it is possible that two {\sl symplectic} planes
783: $L_i$ in $(\rr^4, \omega)$ have {\sl negative} intersection index.
784:
785:
786: \smallskip
787: More detailed considerations lead to
788: %Making similar considerations in more details one obtains
789: the {\sl genus formula} (also called {\sl adjunction formula}) for immersed
790: symplectic surfaces in symplectic four-folds. For this let $(X,\omega)$ be
791: a symplectic
792: manifold of dimension 4, $S\deff \bigsqcup_{j=1}^dS_j$ a compact oriented surface
793: and $u: S \to X$ an immersion with only transversal self-intersection points.
794: Denote by $g_j$ the genus of $S_j$, by $[C]$ the homology class of the image
795: $C \deff u(S)$, $[C]^2$ the homological self-intersection number of $[C]$, and
796: by $c_1(X)$ the first Chern class of $(X, \omega)$. Define the {\sl geometric
797: self-intersection number} $\delta$ of $M=u(S)$ as the algebraic number of pairs
798: $z' \not= z'' \in S$ with $u(z') = u(z'')$, taken with the sign corresponding
799: to the intersection index.
800:
801:
802: \newlemma{lem1.1.2}
803: Suppose that $u: S \to X$ is a symplectic immersion which is
804: compatible with the orientation on each component $S_j$ of $S$. Then
805: \begin{equation}
806: \sum_{j=1}^d g_j = {[C]^2 - c_1(X)[C]\over2} + d - \delta .
807: \eqqno(1.1.3)
808: \end{equation}
809: \end{lem}
810:
811: \smallskip
812: An elementary proof uses the fact that for a {\sl symplectic} immersion
813: $u:S \to (X,\omega)$ one has $c_1(X)[C]= \chi(S) + \chi(N)$, where
814: $N$ is the normal bundle and $\chi$ denotes the Euler characteristic. Finally,
815: one observes that $\chi(N)= [C]^2 -2\delta$. For details, see \cite{Iv-Sh-1}.
816:
817: \newsubsection[1.2]{Local structure of pseudoholomorphic curves}
818: For the most results of this paragraph we refer to \cite{Mi-Wh} where a very
819: precise description of the local structure of pseudoholomorphic curves is
820: given. As a rough summary, one can say that the
821: %The main pathos is that a
822: local behavior of (non-parameterized) pseudoholomorphic curves is essentially
823: the same as for usual holomorphic curves.
824:
825: \newlemma{lem1.2.1} {\rm (\cite{Mi-Wh})} Let $(X,J)$ be an almost complex
826: manifold of $\dimc X
827: \allowbreak
828: =n$, $u: S \to X$ a~$J$-holomorphic map, and $x\in X$ a
829: point. Suppose that $J\in C^2$ and that for any $x'\in X$ sufficiently close to
830: $x$ the pre-image $u\inv(x)$ is finite. Then there exist neighborhoods $U
831: \subset X$ of $x$, $U' \subset \cc^n$ of\/ $0\in \cc^n$ and a
832: $C^1$-diffeomorphism $\phi: U \to U'$ such that $C' \deff \phi\bigl( u(S)
833: \cap U\bigr)$ is a proper analytic curve in $U'$ and such that $\phi_*(J_x)=J\st$,
834: where $J\st$ is the standard complex structure in $\cc^n$.
835: \end{lem}
836:
837: In particular, the notion of a (local) irreducible component of a
838: $J$-holomorphic curve $C=u(S)$ is well-defined. Further, in the special case
839: when $(X,J)$ is an {\sl almost complex surface} one can correctly define
840:
841: \noindent
842: \sli {\sl the intersection index $\delta_{ij}(x)\in \nn$} of two local
843: components $C_i$ and $C_j$ at $x\in X$, and
844:
845: \noindent
846: \slii {\sl the nodal number $\delta_i(x) \in \nn$} of a local component
847: $C_i$ at $x\in X$ (see \cite{Mil}, \S\.10 and \refdefi{def6.2.1}).
848:
849: The main properties of these local invariants are summarized in
850:
851: \newlemma{lem1.2.2}
852: \sli If $x\in C_i \cap C_j$, then $\delta_{ij}(x)
853: \ge 1$. The equality holds \iff $C_i$ and $C_j$ are smooth and
854: intersect transversally in $x$;
855:
856: \slii The set $\{ z \in S\;:\; \delta_i(u(z)) >0 \}$ is discrete in $S$;
857:
858: \sliii Suppose additionally that $S= \sqcup_{j=1}^d S_j$ is a closed surface
859: and $u: S \to X$ is an imbedding almost everywhere on $S$. Set $C\deff u(S)$.
860: Denote by $\delta$ the sum of all local intersection indices $\delta_{ij}(x)$ and
861: all local nodal numbers $\delta_i(x)$, the homology class of $C$ by $[C]$, and
862: the genera of particular components $S_j$ by $g_j$. Then
863: \begin{equation}
864: {\textstyle \sum_{j=1}^d} g_j = {[C]^2 - c_1(X)[C]\over2} + d -\delta.
865: \eqqno(1.2.1)
866: \end{equation}
867: \end{lem}
868:
869: The formula \eqqref(1.2.1) is the {\sl genus formula for pseudoholomorphic
870: curves}. We shall also apply a local version of this result. Here we say that a
871: pseudoholomorphic curve $C$ in a symplectic manifold $X$ is {\sl parameterized by
872: a real surface $S$} if there exists a map $u: S \to X$ which is an imbedding
873: outside a discrete subset in $S$. Such a surface $S$, possibly not connected, can
874: be constructed as the normalization of $C \subset X$.
875:
876: \newlemma{lem1.2.2a} Let $B \subset \rr^4$ be the unit ball equipped with the
877: standard symplectic structure $\omega\st$, and $C_1$, $C_2$ pseudoholomorphic
878: curves in $B$.
879: Assume that the boundaries of the curves $\d C_i$ are imbedded in the boundary of
880: the ball $\d B$, are sufficiently close to each other, and that every $C_i$ meets
881: transversally $\d B$. Denote by $\Chi_i$ the Euler characteristic of the surface
882: $S_i$ parameterizing $C_i$ and by $\delta_i$ the sum of the nodal number of
883: singular points of $C_i$. Then
884: \begin{equation}\eqqno(1.2.1a)
885: \Chi_1 - 2\delta_1 = \Chi_2 - 2\delta_2
886: \end{equation}
887: \end{lem}
888:
889:
890:
891: \proof It is shown in \cite{Iv-Sh-1} that every $C_i$ can be perturbed
892: to a nearby pseudoholomorphic curve $C'_i$ in such a way that every singular point
893: $x \in C_i$
894: with the nodal number $\delta_x(C_i)\ge 2$ ``splits'' into $\delta_x(C_i)$ {\sl
895: nodal points} on the perturbed curve $C'_i$, \ie the points where exactly two
896: branches of $C'_i$ meet transversally. By this procedure the topology of each $S_i$
897: and the whole nodal number of every $C_i$ remain unchanged. After this, one can
898: replace a sufficiently small neighborhood of every nodal point $x \in C'_i$ by a
899: symplectically imbedded handle. This ``symplectic surgery of $C'_i$'' produce {\sl
900: imbedded} pseudoholomorphic curves $C''_i$ with $\Chi(C''_i)= \Chi_i - 2\delta_i$.
901:
902:
903: Moreover, all this can be carried out with the boundaries $\d C_i$ unchanged.
904: Further, the hypothesis of the lemma implies that the boundaries $\d C_i$ are
905: transversal to the standard symplectic structure on $\d B=S^3$ and are isotopic
906: as {\sl transversal links}, see \cite{Iv-Sh-1} and \cite{Eli}. Now one applies
907: the theorem of Bennequin \cite{Bn}, see also \cite{Eli}, which claims that, up
908: to sign convention, $\Chi(C''_i)$ is the {\sl Bennequin index of $\d C_i$} and
909: depends only on the {\sl transversal isotopy} class of $\d C_i$. The lemma follows.
910: \qed
911:
912: \medskip
913: The result of Micallef and White, \lemma{lem1.2.1}, is not sufficient for
914: our purpose, because it does not allow us to control local structure of
915: pseudoholomorphic curves under deformation. A necessary tool is provided
916: by the following statement proven in \cite{Iv-Sh-1}. Here and thereafter
917: $\Delta$ denotes the unit disc in $\cc$ equipped with the standard complex
918: structure.
919:
920: \newlemma{lem1.2.3}
921: Suppose that a $f\in L^{1,2}_\loc(\Delta, \cc^n)$ is not identically $0$ and
922: satisfies {\sl a.e.}
923: the~inequality
924: \begin{equation}
925: | \dbar f(z)| \le h(z)\cdot |z|^k \cdot |f(z)|
926: \eqqno(1.2.2)
927: \end{equation}
928: for some $k\in \nn$ and nonnegative $h\in L^p_\loc(\Delta)$ with $2<p<\infty$.
929: Then
930: \begin{equation}
931: f(z)=z^\mu\bigl(P^{(k)}(z) + z^k g(z)\bigr),
932: \eqqno(1.2.3)
933: \end{equation}
934: where $\mu\in\nn$, $P^{(k)}$ is a polynomial in $z$ of degree $\le k$ with
935: $P^{(k)}(0) \not=0$, and $g\in L^{1, p}_\loc(\Delta,\cc^n)\hookrightarrow C^{0,
936: \alpha}$, $\alpha=1-{2\over p}$, with $g(0)=0$.
937: \end{lem}
938:
939:
940: \medskip
941: Using this result one can obtain the following description of the local
942: behavior of a pseudoholomorphic map. Note that on a given almost complex
943: manifold $(X,J)$ in a neighborhood of a given point $x_0\in X$ there exist
944: an (integrable) complex structure $J^*$ with $J^*(x_0)= J(x_0)$ and
945: $J^*$-holomorphic coordinates $w_1, \ldots, w_n$, $n=\dimc X$.
946:
947: \newlemma{lem1.2.4} \sli Assume that $J$ is $C^1$-smooth and $u: \Delta \to
948: X$ is a non-constant $J$-holomorphic map with $u(0)=x_0$. Then in coordinates
949: $w_1, \ldots, w_n$ chosen as above in a neighborhood of $x_0\in X$ the map $u$
950: has the form
951: \begin{equation}
952: u(z) = z^\mu\cdot P^{(\mu-1)}(z) + z^{2\mu-1} \cdot v(z),
953: \eqqno(1.2.4)
954: \end{equation}
955: where $\mu\in \nn$, $P^{(\mu-1)}(z)$ is a complex $\cc^n$-valued polynomial
956: of degree $\le \mu-1$ with $P^{(\mu-1)}(0)\not=0$, and $v(z) \in
957: L^{1,p}(\Delta, \cc^n)$ with $v(0)=0$.
958:
959: \slii Assume that $J$ is $C^1$-smooth and let $u_1, u_2: \Delta \to X$
960: be $J$-holomorphic maps such that $u_1$ is an immersion and $u_2(0) \in
961: u_1(\Delta)$. Then there exists $r\in ]0,1[$ such that either $u_2(\Delta(r))
962: \subset u_1(\Delta)$ or $u_2(\Delta(r)) \cap u_1(\Delta)= u_2(0)$.
963:
964: \sliii Let $J$ be a $C^\ell$-smooth almost complex structure
965: on the ball $B \subset \cc^n$ with $J(0) = J\st(0)$, and let $u_1, u_2 :
966: \Delta \to B$ be $J$-holomorphic maps with $u_1(0) =u_2(0) =0 \in B$,
967: such that $u_1 \not = u_2$.
968:
969: Then there exists a uniquely defined $\nu \in \nn$ and $w(z) \in C^1(
970: \Delta, \cc^n)$ such that
971: \begin{equation}
972: u_1(z) - u_2(z) = z^\nu w(z).
973: \eqqno(1.2.5)
974: \end{equation}
975: \end{lem}
976:
977: \proof The first and second parts of the lemma are proven in \cite{Iv-Sh-1}.
978: For the third see {\sl Remark 1.6} and {\sl Section 6} of \cite{Mi-Wh}.
979:
980:
981:
982: \smallskip
983: \newdefi{def1.2.1} If for an appropriate local complex coordinate $z$ on $S$
984: a $J$-holomorphic map $u:S \to X$ has the form \eqqref(1.2.4), then we call
985: the (uniquely defined) $\mu$ the {\sl multiplicity of $u$ at the point $z=0$}.
986: \end{defi}
987:
988: \newdefi{def1.2.2} A $J$-holomorphic map $u: S \to X$ is {\sl multiple}
989: if there exists a non-empty $U \subset S$ such that the restriction
990: $u\ogran_U$ can be represented as a composition $u\ogran_U = u' \scirc \phi$
991: where $u': \Delta \to X$ is a $J$-holomorphic map and $\phi: U \to \Delta$ is
992: a (branched) covering of degree $m\ge 2$. In other words, $u$ is multiple
993: if some part of the image $u(S)$ is multiply covered by $u$.
994: \end{defi}
995:
996:
997:
998: \newsubsection[1.3]{Deformation of pseudoholomorphic maps and the Gromov operator
999: $D_{u, J}$}
1000: Roughly speaking, the main idea of the Gromov's theory is to construct and
1001: study $J$-holomorphic curves in a symplectic manifold $(X, \omega)$ for some
1002: special (\eg {\sl integrable}) $J$. Often, one can show the existence of a
1003: $J_0$-holomorphic curve with some other almost complex structure $J_0$, see \eg
1004: {\sl Lemma \ref{lem1.1.1}}. If both $J$ and $J_0$ are $\omega$-tame, then
1005: there exists a homotopy $\{J_t\}_{t\in [0,1]}$ from $J_0$ to $J_1=J$. Hence
1006: one could try to deform the constructed $J_0$-holomorphic map $u_0:S \to X$
1007: into a $J_1$-holomorphic one using the continuity principle. The first step in
1008: this direction is to study the linearization of (\ie the first variation) the
1009: equation \eqqref(1.1.1). This means that we are interested in the first
1010: differential of the section $\sigma_\dbar$.
1011:
1012: \smallskip
1013: Fix a compact surface $S$ of genus $g$. Denote by $\scrj_S$ the Banach
1014: manifold of $C^{1,\alpha}$-smooth complex structures on $S$ with some fixed
1015: $\alpha \in\; ]0,1[$. Thus
1016: \begin{equation}
1017: \scrj_S = \{ J_S\in C^{1,\alpha}(S, \endo(TS)):
1018: J_S^2=-\id \}
1019: \end{equation}
1020: and the tangent space to $\scrj_S$ at $J_S$ is
1021: \begin{equation}
1022: T_{J_S}\scrj_S = \{ I\in C^{1,\alpha}(S, \endo(TS)) :
1023: J_S I + I J_S = 0\} \equiv C^{1,\alpha}(S, \Lambda^{0,1}S\otimes TS),
1024: \end{equation}
1025: where $\Lambda^{0,1}S$ denotes the line bundle of $(0,1)$-form on $S$.
1026:
1027: Let $\scrj$ be an open {\sl connected} subset in the Banach manifold of all
1028: $C^\ell$-smooth almost complex structures on $X$ for some fixed non-integer $\ell>
1029: 2$. In our context the most interesting example is the set $\scrj_\omega$ of
1030: $C^\ell$-smooth $\omega$-tame almost complex structures on $X$. The tangent space
1031: to $\scrj$ at $J$ consists of $C^\ell$-smooth $J$-antilinear endomorphisms of
1032: $TX$,
1033: \begin{equation}
1034: T_J\scrj = \{ I\in C^\ell(X, \endo(TX)) : JI + IJ = 0\}
1035: \equiv C^\ell(X, \Lambda^{0,1}X \otimes TX),
1036: \end{equation}
1037: where $\Lambda^{0,1}X$ denotes the complex bundle of $(0,1)$-forms on $X$.
1038:
1039: Fix $p$ with $2<p<\infty$. Then the set $L^{1,p}(S, X)$ of all Sobolev
1040: $L^{1,p}$-smooth maps from $S$ to $X$ is a Banach manifold. For $u\in L^{1,p}(
1041: S, X)$ we denote
1042: \begin{equation}
1043: E_u \deff u^*TX.
1044: \eqqno(1.3.1)
1045: \end{equation}
1046: In this notation, the tangent space at $u \in L^{1,p}(S, X)$ is $T_uL^{1,p}(S,
1047: X) = L^{1,p}(S, E_u)$, the space of $L^{1,p}$-smooth sections
1048: of the pulled-back tangent bundle of $X$.
1049:
1050: Fix a homology class $[C] \in \sfh_2(X, \zz)$ and consider the set
1051: \begin{equation}
1052: \scrs = \{ u\in L^{1,p}(S, X):u(S)\in [C]\}
1053: \end{equation}
1054: of maps $u$ representing the class $[C]$. Then $\scrs$ is open
1055: in $L^{1,p}(S, X)$ and has the same tangent space, $T_u\scrs = L^{1,p}(S,
1056: E_u)$. Since $\scrj$ is connected, the first Chern class $c_1(X,J)$ is
1057: constant on $\scrj$. We shall denote it simply by $c_1(X)$. Set
1058: \begin{equation}
1059: \mu \deff \la c_1(X), [C] \ra.
1060: \eqqno(1.3.1a)
1061: \end{equation}
1062:
1063:
1064: \smallskip
1065: Consider the subset $\scrp\subset \scrs\times \scrj_S \times \scrj$
1066: consisting of all triples $(u,J_S,J)$ with $u$ being $(J_S,J)$-holomorphic,
1067: \begin{equation}
1068: \scrp = \{(u, J_S, J)\in \scrs\times \scrj_S\times \scrj: du +
1069: J\scirc du\scirc J_S = 0 \}.
1070: \eqqno(1.3.2)
1071: \end{equation}
1072:
1073: Let $\nabla$ be some symmetric connection on $TX$. Covariant differentiation
1074: of \eqqref(1.1.1) gives the equation for the tangent space to $\scrp$.
1075: Namely, a vector $(v,\dot J_S,\dot J)$ is tangent to $\scrp$ at the point
1076: $(u,J_S,J)$ if it satisfies the equation
1077: \begin{equation}
1078: \nabla v + J\scirc \nabla v\scirc J_S + (\nabla_vJ)\scirc (du\scirc J_S) +
1079: J\scirc du\scirc \dot J_S + \dot J\scirc du\scirc J_S = 0. \eqqno(1.3.3)
1080: \end{equation}
1081:
1082: \newdefi{def1.3.1}\.{\bf a)} For a complex bundle $E$ over $S$ let
1083: \begin{equation}
1084: L^p_{(0,1)}(S, E)\deff L^p(S, E\otimes \Lambda^{(0,1)}S)
1085: \end{equation}
1086: denote the Banach space of $L^p$-integrable $E$-valued (0,1)-forms on $S$.
1087:
1088: \statep b). Let $u$ be a $J$-holomorphic curve in $X$. Define the operator
1089: $D_{u, J}: L^{1,p}(S, E_u) \to L^p_{(0,1)}(S, E_u)$ as
1090: \begin{equation}
1091: D_{u, J}(v) \deff \nabla v + J\scirc\nabla v\scirc J_S
1092: + (\nabla_vJ) \scirc du\scirc J_S
1093: \eqqno(1.3.4)
1094: \end{equation}
1095:
1096: \statep c). Define complex Banach bundles $\scre$ and $\scre'$ over $\scrs
1097: \times \scrj_S \times \scrj$ by
1098: \begin{equation}
1099: \scre_{(u,J_S,J)} \deff L^{1,p}(S, E_u)
1100: \quad\hbox{and}\quad
1101: \scre'_{(u,J_S,J)} \deff L^p_{(0,1)}(S, E_u).
1102: \eqqno(1.3.5)
1103: \end{equation}
1104: These bundles are $C^\ell$-smooth and the formula \eqqref(1.3.4) defines a
1105: $\rr$-linear homomorphism $D=D_{u,J_S,J}: \scre \to \scre'$ which is $C^{\ell-1}
1106: $-smooth. The bundle $\scre$ is essentially the tangent bundle to $\scrs$, whereas
1107: $\scre'$ appears as the space where the equation \eqqref(1.1.1) ``lives''.
1108: More precisely, \eqqref(1.1.1) defines a section $\sigma_\dbar$ of
1109: $\scre'$,
1110: \begin{equation}
1111: \sigma_\dbar: (u,J_S,J) \in \scrs \times \scrj_S \times \scrj
1112: \;\mapsto \;
1113: (du + J \scirc du \scirc J_S) \in \scre'_{(u,J_S,J)},
1114: \eqqno(1.3.6)
1115: \end{equation}
1116: such that the equation \eqqref(1.1.1) reads $\sigma_\dbar(u, J_S, J)=0$.
1117: The space $\scr P$ appears then as the zero set of the section $\sigma_\dbar$.
1118: \end{defi}
1119:
1120: \state Remark. Here and thereafter we use the normalization ${\d \over \d\bar z}
1121: = {\d\over \d x} + \isl {\d\over \d y}$, deviating from the usual convention
1122: ${\d \over \d\bar z} = \mathbf{\half} \cdot ({\d\over \d x} +\isl {\d\over\d y})$.
1123: The same normalization is used for all operators with Cauchy-Riemann symbol.
1124:
1125:
1126: \newlemma{lem1.3.1}
1127: Let $\scrx$ be a Banach manifold, $\scre\to \scrx$
1128: and $\scre'\to\scrx$ $C^1$-smooth Banach bundles over $\scrx$, $\nabla$ and
1129: $\nabla'$ linear connections in $\scre$ and $\scre'$ respectively, $\sigma$
1130: a $C^1$-smooth section of $\scre$, and $D: \scre \to \scre'$ a $C^1$-smooth
1131: bundle homomorphism.
1132:
1133: \sli If $\sigma(x)=0$ for some $x\in \scrx$, then the map $\nabla\sigma_x:
1134: T_x\scrx \to \scre_x$ is independent of the choice of the connection $\nabla$
1135: in $\scre$;
1136:
1137: \slii Set $K_x\deff \ker(D_x:\scre_x\to \scre'_x)$ and $Q_x\deff \coker(D_x:
1138: \scre_x\to \scre'_x)$ with the corresponding imbedding $i_x:K_x \to \scre_x$
1139: and projection $p_x:\scre'_x \to Q_x$. Let $\nabla^\hom$ be the connection in
1140: $\hom(\scre,\scre')$ induced by the connections $\nabla$ and $\nabla'$. Then
1141: the map
1142: \begin{equation}
1143: p_x \circ(\nabla^\hom\!\! D_x)\scirc i_x:
1144: T_x\scrx\to \hom(K_x,Q_x)
1145: \end{equation}
1146: is independent of the choice of connections
1147: $\nabla$ and $\nabla'$.
1148: \end{lem}
1149:
1150: \state Remark. Taking this lemma into account, we shall use the following
1151: notation. For $\sigma\in \Gamma(\scrx, \scre)$, $D\in \Gamma(\scrx,
1152: \hom(\scre, \scre'))$ and $x\in \scrx$ as in the hypothesis of the lemma we
1153: shall denote by $\nabla\sigma_x:T_x\scrx \to \scre_x$ and ${\nabla\!D}:
1154: T_x\scrx\times \ker D_x \to \coker D_x$ the corresponding operators without
1155: pointing out which connections were used to define them.
1156:
1157: \proof \sli Let $\wt\nabla$ be another connection in $\scre$. Then
1158: $\wt\nabla$ has the form $\wt\nabla=\nabla +A$ for some
1159: $A\in\Gamma(\scrx,\hom(T\scrx,\endo(\scre)))$. So for $\xi\in T_x\scrx$
1160: we obtain $\wt\nabla_\xi\sigma -\nabla_\xi\sigma = A(\xi,\sigma(x))=0$.
1161:
1162: \slii Similarly, let $\wt\nabla'$ be another connection in $\scre'$, and let
1163: $\wt\nabla^\hom$ be the connection in $\hom(\scre,\scre')$ induced by
1164: $\wt\nabla$ and $\wt\nabla'$. Then $\wt\nabla'$ also has the form $\wt\nabla=
1165: \nabla +A'$ for some $A'\in\Gamma(\scrx,\hom(T\scrx, \endo(\scre')))$.
1166: So for $\xi\in T_x\scrx$ we obtain $\wt\nabla_\xi^\hom D -\nabla_\xi^\hom D=
1167: A'(\xi)\scirc D_x - D_x \scirc A(\xi)$. The statement of the lemma now follows
1168: from the identities $p_x\scirc D_x=0$ and $D_x\scirc i_x=0$.
1169: \qed
1170:
1171: \smallskip
1172: \state Remark. The operator $D_{u, J}$ is the linearization of the equation
1173: \eqqref(1.1.1). Thus \lemma{lem1.3.1} shows that the definition of $D_{u, J}$
1174: is independent of the choice of $\nabla$. In particular, one can also use
1175: non-symmetric connections, \eg those compatible with $J$, as it is done in
1176: \cite{Gro}. However, it is convenient to have a fixed
1177: connection considering varying almost complex structures $J$ on $X$. But this
1178: is impossible for $\nabla$ compatible with $J$. On the other hand, with a
1179: symmetric connection computations become simpler.
1180:
1181: \smallskip
1182: The operator $D_{u, J}$, as well as the equation \eqqref(1.1.1) itself, is
1183: elliptic
1184: of order 1 with the Cauchy-Riemann symbol. This implies standard regularity
1185: properties for $D_{u, J}$. In particular, the kernel and the cokernel are of
1186: finite dimension. The Riemann-Roch formula gives the index of $D_{u, J}$:
1187: \begin{equation}
1188: \dim_\rr \ker D_{u, J} - \dim_\rr \coker D_{u, J} = 2\cdot
1189: \bigl(\mu + n (1-g) \bigr),
1190: \eqqno(1.3.7)
1191: \end{equation}
1192: where $\mu \deff c_1(X) \cdot [u(S)]$, $g$ is the genus of $S$, and $n$ the
1193: {\sl complex} dimension of $X$, \ie $n \deff \half \dim_\rr X$. The factor
1194: 2 appears because we compute {\sl real, not complex} dimensions of the
1195: (co)kernel.
1196:
1197:
1198: \newsubsection[1.4]{Holomorphic structure on the induced bundle}
1199: Now we want to understand the structure of the operator $D_{u,J}$ in more
1200: detail. Note that the pulled-back bundle $E_u = u^*TX$ carries a complex
1201: structure, namely $J$ itself, or more accurately $u^*J$. However, the
1202: operator $D\deff D_{u,J}$ is only $\rr$-linear. So we decompose it into
1203: $J$-linear and $J$-antilinear parts. Namely, for $v\in
1204: L^{1,p}(S, E)$ we write $D v = \half \bigl( D v - JD(Jv)\bigr) +
1205: \half (D v + JD (Jv)\bigr) = \dbar_{u, J}v+ R(v)$.
1206:
1207: \newdefi{def1.4.1} The $J$-linear part $\dbar_{u, J}$ of the operator $D_{u, J}$
1208: is called the {\sl $\dbar$-operator associated with a $J$-holomorphic map $u$}.
1209: \end{defi}
1210:
1211: \smallskip
1212: By the definition, the operator $\dbar_{u, J} : L^{1,p}(S, E_u) \to L^p
1213: _{(0,1)}(S, E_u)$ is $J$-linear. The following statement is well
1214: known in the smooth case.
1215:
1216: \newlemma{lem1.4.1A} Let $S$ be a Riemann surface with a complex
1217: structure $J_S$ and $E$ a $L^{1,p}$-smooth complex vector bundle of rank $r$
1218: over $S$. Let also $\dbar_E : L^{1,p}(S, E) \to L^p_{(0,1)}(S,E)$ be a complex
1219: linear differential operator satisfying the condition
1220: \begin{equation}
1221: \dbar_E(f\xi) = \dbar_S f \otimes \xi
1222: + f\cdot \dbar_E\xi, \eqqno(1.3.6a)
1223: \end{equation}
1224: where $\dbar_S$ is the Cauchy-Riemann operator on $S$ associated to $J_S$.
1225: Then the sheaf
1226: \begin{equation}
1227: U\subset S \mapsto \scro(E)(U) := \{\,\xi \in L^{1,p}(U, E) \, :\,
1228: \dbar_E\xi=0 \,\} \eqqno(1.3.7a)
1229: \end{equation}
1230: is coherent and locally free of rank $r$. This defines a holomorphic
1231: structure on $E$ for which $\dbar_E$ is the associated Cauchy-Riemann
1232: operator.
1233: \end{lem}
1234:
1235:
1236: \state Remark. The condition $\eqqref(1.3.6a)$ means that $\dbar_E$ is of
1237: order 1 and has the Cauchy-Riemann symbol. For the proof we refer to
1238: \cite{Iv-Sh-1} and \cite{Iv-Sh-2} for the general case, or to \cite{H-L-S}
1239: for the case of line bundles.
1240:
1241:
1242: \smallskip
1243: Thus, according to \lemma{lem1.4.1A}, the operator $\dbar_{u, J}$ defines
1244: a holomorphic structure on the bundle $E_u$. We shall denote by $\scro(E_u)$
1245: the sheaf of holomorphic sections of $E_u$. The tangent bundle $TS$ to our
1246: Riemann surface also carries a natural holomorphic structure. We shall
1247: denote by $\scro(TS)$ the~corresponding coherent sheaf.
1248:
1249: \medskip
1250: Denote by $N_J(v, w)$ the Nijenhuis torsion tensor of the almost complex
1251: structure $J$, (see \eg \cite{Ko-No}, Vol.II., p.123.)\.\footnote{\.
1252: Note that in \cite{Ko-No} another normalization constant is used. However,
1253: this is not essential for our purpose.}
1254:
1255: \newlemma{lem1.4.1}
1256: \sli The $J$-antilinear part $R$ of $D_{u,J}$ is related to $u$ and $J$ by
1257: the formula
1258: \begin{equation}
1259: R(v)(\xi ) = N_J(v, du(\xi ))\qquad \xi \in TS.
1260: \eqqno(1.4.1)
1261: \end{equation}
1262: Thus $R$ is a continuous $J$-antilinear operator from $E$ to $\Lambda^{0,1}S
1263: \otimes E_u$ of order zero which satisfies $R\scirc du \equiv 0 $, \ie
1264: $R( du(\eta), \xi )=0$ for all $\eta ,\xi \in TS$.
1265:
1266: \smallskip
1267: \slii If $u$ is non-constant, then $du$ defines an injective analytic
1268: morphism of coherent sheaves
1269: \begin{equation}
1270: 0\lrar \scro(TS) \buildrel du \over \lrar \scro(E_u). \eqqno(1.4.2)
1271: \end{equation}
1272: \end{lem}
1273:
1274: \proof \sli Formula \eqqref(1.4.1) can be found in \cite{McD-2}. The rest of
1275: part \sli follows from the well-known fact that $N_J(v, w)$ is skew-symmetric
1276: and $J$-antilinear in both arguments.
1277:
1278: \noindent \slii The fact that $du: TS \to E_u$ defines a morphism between
1279: coherent sheaves $\scro(TS)$ and $\scro(E_u)$ means that $du$ is
1280: a {\sl holomorphic} section of $T^*S \otimes E_u$. This is equivalent to
1281: relation
1282: \begin{equation}
1283: du \scirc \dbar_S = \dbar_{u,J} \scirc du.
1284: \end{equation}
1285: For the proof of this fact we refer to \cite{Iv-Sh-1} and \cite{Iv-Sh-2}.
1286:
1287: Injectivity of the sheaf homomorphism $du$ is equivalent to its
1288: nondegeneracy which is the case in our context. \qed
1289:
1290:
1291:
1292: \smallskip
1293: The zeros of the analytic morphism $du : \scro(TS) \to \scro (E_u)$ are
1294: isolated. So we obtain
1295:
1296: \newcorol{cor1.4.2} {\rm(\cite{Mi-Wh}, \cite{Sk-1})}
1297: The set of critical points of a $J$-holo\-mor\-phic map is discrete,
1298: provided $J$ is of class $C^1$.
1299: \end{corol}
1300:
1301: \newdefi{def1.4.2} By the {\sl order of zero $\ord_p du$} of the
1302: differential $du$ at a point $p\in S$ we shall understand the order of
1303: vanishing at $p$ of the holomorphic morphism $du : \scro(TS)\to \scro(E_u)$.
1304: \end{defi}
1305:
1306: It follows from \lemma{lem1.4.1} that $\ord_p du$ is a well-defined
1307: non-negative integer.
1308:
1309:
1310: \newsubsection[1.5]{The normal sheaf of a pseudoholomorphic curve}
1311: From \eqqref(1.4.2) we obtain the following short exact sequence of coherent
1312: sheaves
1313: \begin{equation}
1314: 0\lrar \scro(TS) \buildrel du \over\lrar \scro(E_u)
1315: \lrar \scrn_u\lrar 0,
1316: \eqqno(1.5.1)
1317: \end{equation}
1318: where $\scrn_u\deff \scro(E)/du(\scro(TS))$ is the quotient sheaf. It follows from
1319: \lemma{lem1.4.1} \slii that there is a decomposition $\scrn_u= \scro(N_u)\oplus
1320: \scrn_u\sing$ where $N_u$ is a holomorphic
1321: vector bundle and $\scrn_u\sing = \bigoplus_{z\in S} \cc_z^{\ord_z du}$ is a
1322: discrete sheaf with support in the set of critical points $a_i$ of $u$
1323: with the stalk $\cc^{n_i}$ of dimension $n_i \deff \ord_{a_i}du$ at every
1324: such point $a_i$.
1325:
1326: \newdefi{def1.5.1} The quotient sheaf $\scrn_u \deff \scro(E)/du(TS)$ is called
1327: the {\sl normal sheaf of a $J$-curve $u:S \to X$}, $N_u$ the {\sl normal
1328: bundle to the $J$-curve $u:S \to X$}, and $[A] \deff \sum n_i[a_i]$
1329: the {\sl branching divisor of the $J$-curve $u:S \to X$}.
1330: \end{defi}
1331:
1332:
1333: Denote by $\scro([A])$ the sheaf of meromorphic functions on $S$ having poles
1334: of order at most $n_i$ at $a_i$. Then \eqqref(1.5.1) gives rise to the exact
1335: sequence of coherent sheaves
1336: \begin{equation}
1337: 0\lrar \scro(TS)\otimes \scro([A])\buildrel{du}\over
1338: {\lrar} \scro(E_u)
1339: \lrar \scro(N_u)\lrar 0.
1340: \eqqno(1.5.2)
1341: \end{equation}
1342:
1343: The holomorphic structure in $N_u$ defines the Cauchy-Riemann operator
1344: $\dbar_N :L^{1,p}(S, N_u) \allowbreak \lrar L^p_{(0,1)}(S, N_u)$.
1345: \lemma{lem1.4.1} implies that
1346: the homomorphism $R: E_u \to E_u \otimes \Lambda^{(0,1)}S$
1347: induces a $J$-antilinear bundle homomorphism $R_N : N_u \to N_u \otimes
1348: \Lambda^{(0,1)}S$. Define the operator
1349: \begin{equation}
1350: D_{u, J}^N : L^{1,p}(S, N_u) \allowbreak \lrar L^p_{(0,1)}(S, N_u)
1351: \quad\hbox{by}\quad
1352: D_{u, J}^N \deff\dbar_N + R_N.
1353: \eqqno(1.5.3)
1354: \end{equation}
1355:
1356:
1357: \smallskip
1358: \newdefi{def1.5.2} Let $E$ be a holomorphic vector bundle over a compact Riemann
1359: surface $S$ of genus $g$ and let $D:L^{1,p}(S, E)\to L^p(S, \Lambda^{0,1}S \otimes
1360: E)$ be an operator of the~form $D=\dbar + R$ where $R\in L^p\bigl(S,\,\hom_\rr(E,
1361: \,\Lambda^{0, 1}S\otimes E) \bigr)$ with $2<p<\infty$. Define $\sfh^0_D(S, E)\deff
1362: \ker D$ and $\sfh^1_D(S, E) \deff \coker D$. The groups $\sfh^i_D(S, E)$ are
1363: referred to as {\sl $D$-cohomology groups of $E$}.
1364: \end{defi}
1365:
1366: \medskip
1367: The Riemann-Roch formula gives the {\sl index} of $D$,
1368: \begin{equation}
1369: \ind_\rr D \deff \dimr\sfh^0_D(S, E) - \dimr\sfh^1_D(S, E)=
1370: 2\bigl( c_1(E) + \rank(E) (1-g)\bigr).
1371: \eqqno(1.5.?4)
1372: \end{equation}
1373:
1374:
1375: \state Remark. Taking into account the elliptic regularity of the
1376: Cauchy-Riemann operator, for given $S$, $E$ and $R\in L^p$, $2<p<\infty$, one
1377: can define $\sfh^i_D(S, E)$ as the (co)kernel of the operator $\dbar +R:
1378: L^{1,q}(S, E) \to L^q(S,\,\Lambda^{0, 1}S\otimes E)$ for any $q\in\; ]1, p]$.
1379: So the definition is independent of the choice of the functional spaces. Note also
1380: that the $\sfh^i_D(S, E)$ are of finite dimension provided that $S$ is closed.
1381: For details, see \eg \cite{Iv-Sh-1}.
1382:
1383: \smallskip
1384: The following lemmas contain main properties of $D$-cohomologies which will
1385: be used later. For complete proofs we refer to \cite{Iv-Sh-1} and \cite{Iv-Sh-2}.
1386:
1387: \newlemma{lem1.5.1} {\sl (Serre duality for $D$-cohomologies.)}
1388: Let $E$ be a holomorphic vector bundle over a~compact Riemann
1389: surface $S$ and let $D:L^{1,p}(S, E)\to L^p_{(0,1)}(S, E)$ be
1390: an operator of the~form $D=\dbar + R$, where $R\in L^p\bigl(S,\,
1391: \hom_\rr(E,\,\Lambda^{0, 1}S\otimes E) \bigr)$ with $2<p <\infty$. Let
1392: $K_S\deff \Lambda^{1, 0}S$ be the~canonical holomorphic line
1393: bundle of $S$. Then there exists the~naturally defined operator
1394: \begin{equation}
1395: D^*=\dbar- R^*: L^{1,p}(S, E^* \otimes K_S)
1396: \to L^p_{(0,1)}(S, E^* \otimes K_S)
1397: \end{equation}
1398: with $R^* \in L^p\bigl(S,\,\homr(E^*\otimes K_S,\,
1399: \Lambda^{0, 1}S \otimes E^*\otimes K_S) \bigr)$ and the~natural isomorphisms
1400: \begin{equation}
1401: \matrix
1402: \sfh^0_D(S,\, E)^* &\cong& \sfh^1_{D^*}(S,\, E^*\otimes K_S),
1403: \cr
1404: \sfh^1_D(S,\, E)^* &\cong& \sfh^0_{D^*}(S,\, E^*\otimes K_S),
1405: \endmatrix
1406: \eqqno(1.5.4)
1407: \end{equation}
1408: induced by the pairings
1409: \begin{equation}
1410: \begin{matrix}%\format\l\ \ &\l\ \ &\l\\
1411: \phi \in \sfh^0_D(S,\, E), \quad
1412: \psi \in L^p_{(0,1)}(S,\, E^*\otimes K_S)&
1413: \mapsto \<\phi,\psi\> &
1414: \deff \re \int_S \psi \scirc \phi
1415: \cr
1416: \psi \in \sfh^0_D(S,\, E^*\otimes K_S), \quad
1417: \phi \in L^p_{(0,1)}(S,\, E)&
1418: \mapsto \<\phi,\psi\> &
1419: \deff \re \int_S \psi \scirc \phi
1420: \end{matrix}
1421: \eqqno(1.5.5)
1422: \end{equation}
1423: If, in addition, $R$ is $\cc$-antilinear, then $R^*$ is also
1424: $\cc$-antilinear.
1425: \end{lem}
1426:
1427:
1428: \state Remark. The lemma expresses the well-known relation $\ker(D^*) =(\im D)
1429: ^\perp$ between a linear operator $D$ and its adjoint. It is worth observing
1430: that the spaces themselves and the duality are defined only over the real numbers
1431: $\rr$ and not over $\cc$.
1432:
1433:
1434: \smallskip
1435: \newlemma{lem1.5.2}
1436: {\rm(\cite{H-L-S}, {\sl Vanishing theorem for $D$-cohomologies.})}
1437: Let $S$ be a closed Riemann surface of genus $g$ and $L$
1438: a~holomorphic {\sl line} bundle over $S$, equipped with a~differential
1439: operator $D=\dbar + R$ with $R\in L^p\bigr(S, \homr(L,\, \Lambda^{0,
1440: 1}S\otimes L) \bigl)$, $p>2$. If $c_1(L)<0$, then $\sfh^0_D (S,\, L)=0$. If
1441: $c_1(L)>2g-2$, then $\sfh^1_D (S,\, L)=0$.
1442: \end{lem}
1443:
1444:
1445:
1446: \medskip
1447: The importance of the operator $D=\dbar + R$ lies in the fact that we can
1448: associate with the short exact sequence \eqqref(1.5.1) the long exact
1449: sequence of $D$-cohomo\-logies. Note, that due to \lemma{lem1.4.1} we obtain
1450: the short exact sequence of complexes
1451: \begin{equation}
1452: \eqqno(1.5.6)
1453: \def\normalbaselines{\baselineskip20pt\lineskip3pt \lineskiplimit3pt }
1454: \setbox1=\hbox{$\lrar$}
1455: \def\mapright#1{\!\!\!\smash{\mathop{{\hbox to
1456: \wd1{\hss\hbox{$\displaystyle\longrightarrow$}\hss}}}\limits^{#1}}\!\!\!}
1457: %
1458: \def\mapdown#1{\Big\downarrow\rlap{$\vcenter{\hbox{$\scriptstyle#1$}}$}}
1459: %
1460: \begin{matrix}
1461: \llap{$0\lrar$} L^{1,p}(S, TS)&
1462: \mapright{du}&
1463: L^{1,p}(S, E)&
1464: \mapright{\barr\pr}&
1465: L^{1,p}(S, E)\bigm/ du(L^{1,p}(S,
1466: \rlap{$TS))\lrar0$}
1467: \\
1468: \mapdown{\barr \partial_S}\qquad&& \mapdown{D}\qquad&&
1469: \mapdown{\barr D}\qquad\qquad\qquad\qquad
1470: \\
1471: \llap{$0\lrar$} L^p_{(0,1)}(S, TS)&
1472: \mapright{du}&
1473: L^p_{(0,1)}(S, E)&
1474: \mapright{\barr\pr}&
1475: L^p_{(0,1)}(S, E)\bigm/
1476: du(L^p_{(0,1)}(S,
1477: \rlap{$TS)) \lrar0$}
1478: \end{matrix}
1479: \end{equation}
1480: where $\barr D$ is induced by $D\equiv D_{u, J}$.
1481:
1482:
1483: \newlemma{lem1.5.3} For $\barr D$ as in \eqqref(1.5.6), $\ker \barr D =
1484: \sfh^0_D (S, N_u) \oplus \sfh^0(S, \scrn_u\sing)$ and $\coker \barr D
1485: \allowbreak= \sfh^1_D (S, N_u)$.
1486: \end{lem}
1487:
1488:
1489: \proof For an open set $U \subset S$ let $\Gamma_D(U, E_u) \deff \{ v\in L^{1,p}
1490: _\loc(U, E_u) \;:\; Dv = 0\}$. Use the analogous notation for $N_u$. Consider the
1491: sheaves $U \mapsto \Gamma\, (U, \scro(TS)\,)$, $U \mapsto \Gamma_D (U, E_u)$, and
1492: $U \mapsto \Gamma_D(U, N_u) \oplus \Gamma(U, \scrn_u \sing)$. It is easy to show
1493: that the first two columns of the diagram \eqqref(1.5.6)
1494: define fine resolutions of the sheaves $\scro(TS)$ and $\Gamma_D (\, \cdot\,,
1495: E_u)$. Moreover, $du$ defines injective homomorphisms between these sheaves
1496: and between their resolutions. An explicit computation shows that $\Gamma_D(\,
1497: \cdot \,, N_u) \oplus \Gamma(\,\cdot\,, \scrn_u\sing)$ is the
1498: corresponding quotient sheaf and that the third column of \eqqref(1.5.6) is its
1499: resolution. For details, see \cite{Iv-Sh-1}. \qed
1500:
1501: \newcorol{cor1.5.4} The~short exact sequence $\eqqref(1.5.1)$ induces
1502: the~long exact sequence of $D$-cohomologies
1503: \begin{equation*}
1504: \def\normalbaselines{\baselineskip20pt\lineskip3pt \lineskiplimit3pt }
1505: \def\mapright#1{\;\smash{\mathop{\longrightarrow}\limits^{#1}}\;}
1506: \def\mapdown{\Big\downarrow}
1507: %
1508: \matrix%\format\c&\c&\c&\c&\c&\c&\l&\c&\l\\
1509: 0& \mapright{}& \sfh^0(S, TS) &\mapright{} & \sfh^0_D(S, E)
1510: & \mapright{}& \sfh^0_D(S, N_u)\oplus \sfh^0(S, \scrn_u\sing)
1511: &\mapright{\delta} &\vphantom{\mapdown}\\
1512: %
1513: & \mapright{}& \sfh^1(S, TS) &\mapright{} & \sfh^1_D(S, E)
1514: & \mapright{}& \sfh^1_D(S, N_u) &\mapright{} &0.
1515: \endmatrix
1516: \end{equation*}
1517: \end{corol}
1518:
1519: \medskip
1520: \newsection[2]{The total moduli space of pseudoholomorphic curves}
1521:
1522: \newsubsection[2.1]{Transversality}
1523: Any deformation of a given $J$-holomorphic map $u: S \to X$ defines a path in
1524: the space $\scrp$ of pseudoholomorphic maps. Thus to construct such a deformation
1525: we want to equip the space $\scrp$ with a structure of a smooth Banach manifold.
1526:
1527: Note that by definition the set $\scrp$ is the zero set of the
1528: section $\sigma_\dbar$ of the bundle $\scre' $, \ie the intersection of
1529: the images of $\sigma_\dbar$ and the zero-section $\sigma_0$.
1530: Thus we are interested in which points these sections meet
1531: transversally. The analysis of the problem leads to the following
1532:
1533: \newdefi{def2.1.1} Let $\scrx$, $\scry$, and $\scrz$ be Banach manifolds
1534: with $C^\ell$-smooth maps $f:\scry \to \scrx$ and $g:\scrz\to\scrx$, $\ell
1535: \ge1$. Define the {\sl fiber product} $\scry\times_\scrx \scrz $ by setting
1536: $\scry\times_\scrx \scrz \deff \{\,(y,z)\in \scry\times\scrz \;:\; f(y)= g(z)
1537: \,\}$. The map $f$ is called {\sl transversal to $g$} at a point $(y,z)\in
1538: \scry\times_\scrx \scrz$ with $x\deff f(y)=g(z)$, and $(y,z)$ is called a
1539: {\sl transversality point}, if the map $df_y\oplus dg_z: T_y\scry \oplus
1540: T_z\scrz \to T_x\scrx$ is {\sl surjective} and admits a {\sl closed
1541: complement} to its kernel. The set of transversality points $(y,z)\in \scry
1542: \times_\scrx \scrz$ will be denoted by $\scry\times^\trans_\scrx \scrz$, with
1543: $\trans$ symbolizing the transversality condition.
1544:
1545: We say that $f:\scry \to \scrx$ is {\sl transversal} to $g:\scrz\to\scrx$ if
1546: every point $(y,z) \in \scry\times_\scrx \scrz$ is a transversality point. In
1547: particular, if $\scry$ consists of a point $x \in \scrx$ and the imbedding $\{
1548: x\} \hook \scrx$ is transversal to $g$, we call $x$ a {\sl regular value of
1549: $g$}. Note that by this definition any $x\in \scrx\bs g(\scrz)$ is a regular
1550: value of $g$.
1551:
1552: In the special case when the map $g:\scrz \to \scrx$ is a closed imbedding,
1553: the fiber product $\scry\times_\scrx \scrz$ is simply the preimage
1554: $f\inv\scrz$ of $\scrz\subset\scrx$. In particular, every point
1555: $(y,z)\in\scry\times_\scrx \scrz$ is completely defined by its component
1556: $y\in\scry$, $z=f(y)\in\scrz\subset \scrx$. In this case we simply say that
1557: $f:\scry \to \scrx$ is {\sl transversal to $\scrz$ at $y\in\scry$},
1558: \iff $(y,f(y))$ is a transversal point of $\scry\times_\scrx \scrz\cong
1559: f\inv\scrz$.
1560: \end{defi}
1561:
1562:
1563: \newlemma{lem2.1.1}
1564: The set $\scry\times^\trans_\scrx \scrz$ is open in $\scry\times_\scrx \scrz$
1565: and is a $C^\ell$-smooth Banach manifold with tangent space
1566: \begin{equation}
1567: T_{(y,z)}\scry\times^\trans_\scrx \scrz = \ker\bigl(
1568: df_y\oplus d(-g_z): T_y\scry \oplus T_z\scrz \to T_x\scrx \bigr).
1569: \end{equation}
1570: \end{lem}
1571:
1572:
1573: \proof Fix $w_0\deff(y_0,z_0)\in \scry\times^\trans_\scrx \scrz$
1574: and set $K_0 \deff \ker( df_{y_0}\oplus dg_{z_0}:
1575: T_{y_0}\scry \oplus T_{z_0}\scrz \to T_x\scrx \bigr)$. Let $Q_0$ be a closed
1576: complement to $K_0$. Then the map $df_{y_0}\oplus dg_{z_0}:Q_0\to T_x\scrx$
1577: is an isomorphism.
1578:
1579: Due to the choice of $Q_0$, there exists a neighborhood $V\subset
1580: \scry\times\scrz$ of $(y_0,z_0)$ and $C^\ell$-smooth maps $w':V\to K_0$ and
1581: $w'':V\to Q_0$, such that $dw'_{w_0}$ (resp.\ $dw''_{w_0}$) is the projection
1582: from $T_{y_0}\scry \oplus T_{z_0}\scrz$ onto $K_0$ (resp.\ onto $Q_0$),
1583: so that $(w',w'')$ are coordinates in some smaller neighborhood $V_1\subset
1584: \scry\times\scrz$ of $w_0=(y_0,z_0)$. It remains to consider the equation
1585: $f(y)=g(z)$ in new coordinates $(w',w'')$ and apply the implicit function
1586: theorem.
1587: %
1588: \qed
1589:
1590: \medskip
1591: Due to \lemma{lem2.1.1}, the set $\scrp$ is a Banach manifold at those points
1592: $(u,J_S,J)\in\scrp$ where $\sigma_\dbar$ is transversal to $\sigma_0$.
1593: However, at any point $(u,J_S,J;0)$ on the zero section $\sigma_0$ of
1594: $\scre'$ we have the natural decomposition
1595: \begin{equation}
1596: T_{(u,J_S,J;0)}\scre' = d\sigma_0\bigl( T_{(u,J_S,J)}
1597: (\scrs\times \scrj_S\times \scrj) \bigr)
1598: \oplus \scre'_{(u,J_S,J)},
1599: \end{equation}
1600: where the first component is the tangent space to the zero section of $\scre'$
1601: and the second one is the tangent space to the fiber $\scre'_{(u,J_S,J)}$.
1602:
1603: Let $p_2$ denote the projection on the second component. Then the
1604: transversality $\sigma_\dbar$ and $\sigma_0$ is equivalent to the surjectivity
1605: of the map $p_2\scirc d\sigma_\dbar: T_{(u,J_S,J)}(\scrs\times \scrj_S \times
1606: \scrj)\to \scre'_{(u,J_S,J)}$, \ie to the surjectivity of the operator
1607: %\begin{equation}
1608: \begin{align*}
1609: \nabla\sigma_\dbar:\;&
1610: T_uL^{1,p}(S,X)\oplus T_{J_S}\ttt_g \oplus T_J\scrj \lrar
1611: \scre'_{(u,J_S,J)}
1612: \\
1613: \nabla\sigma_\dbar:\;&(v,\dot J_S,\dot J) \longmapsto
1614: D_{(u,J)}v + J\scirc du \scirc \dot J_S
1615: + \dot J\scirc du \scirc J_S.
1616: \end{align*}
1617: %\end{equation}
1618: By \refdefi{def1.5.2}, the quotient of $\scre'_{(u,J_S,J)}$ with respect
1619: to the image of $D_{u,J}$ is $\sfh^1_D(S,E_u)$. The induced map
1620: \begin{equation}
1621: T_{J_S}\scrj_S \ni \dot J_S \mapsto J\scirc du \scirc \dot J_S
1622: \in \sfh^1_D(S,E_u)
1623: \eqqno(2.1.1)
1624: \end{equation}
1625: is also easy to describe. Recall that for a given complex
1626: structure $J_S$ on $S$ one has the Dolbeault isomorphism
1627: \begin{equation}
1628: \sfh^1(S,TS)= \coker \bigl(\,\dbar :
1629: C^{2,\alpha}(S, TS) \lrar C^{1,\alpha}_{(0,1)}(S, TS) \,\bigr)
1630: \end{equation}
1631: with the operator $\dbar$ associated to $J_S$. Recall also that $C^{1,\alpha}
1632: _{(0,1)}(S, TS)$ is the tangent space $T_{J_S}\scrj_S$. This shows that the map
1633: \eqqref(2.1.1) is the same as the map $J\scirc du: \sfh^1(S,TS)
1634: \to \sfh^1_D(S,E_u)$ and, due to identity $J\scirc du = du\scirc J_S$ and
1635: \refcorol{cor1.5.4}, its cokernel is $\sfh^1_D(S,N_u)$.
1636:
1637: It remains to study the image of $T_J\scrj$ in $\sfh^1_D(S,N_u)$.
1638:
1639: \newdefi{def2.1.2} For $(u,J_S,J)\in \scrp$ we define $\Psi=
1640: \Psi_{(u,J)}:T_J\scrj \to \scre'_{(u,J_S,J)}$ by setting $\Psi_{(u,J
1641: )}( \dot J) \deff \dot J\scirc du \scirc J_S$. Let $\barr\Psi=\barr\Psi_{(u,
1642: J)}:T_J\scrj \to \sfh^1_D(S,N_u)$ be induced by $\Psi$. Finally, define
1643: $\scrp^* \deff \{ (u,J_S,J) \in \scrp: u \hbox{ is injective in generic }z
1644: \in S\}$.
1645: \end{defi}
1646:
1647: \state Remark. One can show that $\scrp\bs \scrp^*$ consists of {\sl multiple
1648: curves} for which the map $u:(S,J_S) \to (X,J)$ admits a factorization
1649: $u= u' \scirc g$ for some non-trivial holomorphic branched covering $g:(S,J_S)
1650: \to (S',J'_S)$ and a $J$-holomorphic map $u':(S',J'_S) \to X$. On the other
1651: hand, for any $(u,J_S,J) \in \scrp^*$ the map $u$ is a smooth imbedding outside
1652: finitely many points in $S$. For details see \cite{Mi-Wh} or \cite{Iv-Sh-1}.
1653:
1654:
1655: \newlemma{lem2.1.2} {\sl(Infinitesimal transversality)}.
1656: Let $(u,J_S,J)\in \scrp^*$. Then the operator $\barr\Psi:T_J\scrj \to
1657: \sfh^1_D(S, N_u)$ is surjective.
1658: \end{lem}
1659:
1660:
1661: \proof Choose some nonempty open set $V\subset S$, such that
1662: $u\ogran_V$ is an imbedding. Use Serre duality (\lemma{lem1.5.1}) and
1663: find a basis $\psi_1,\ldots \psi_l\in \sfh^0_D(S, N^*\otimes K_S)\cong
1664: \sfh^1_D(S, N)^*$.
1665:
1666: Note that $\psi_i$ satisfy the equation $D \psi_i =0$, where the operator
1667: $D= D_{N^*\otimes K_S}$ is of the form $\dbar + R$. One can show
1668: (see \eg \cite{Iv-Sh-1} or \cite{H-L-S}) that any solution $v$ of the equation
1669: $(\dbar + R)v=0$ is $L^{1,p}$-smooth and furthermore such a $v$ is either
1670: identically zero or has isolated zeros.
1671:
1672: This implies that there exist $I_1,\ldots I_l\in C^\ell(S, N \otimes
1673: \Lambda^{(0,1)} )$ with supports $\supp(I_i)$ in $V$ such that the
1674: matrix $\bigl(\re\int_S \psi_i \scirc I_j\bigr)_{i,j=1} ^l$ is non-degenerate.
1675: Since $u\ogran_V$ is a $C^{\ell+1}$-smooth imbedding, any such $I_i$ can be
1676: represented in the form $I_i = \dot J_i \scirc du \scirc J_S$ with some
1677: $\dot J_i \in C^\ell(X, \endo(TX))$ with $ J \scirc \dot J_i + \dot J_i\scirc
1678: J =0$. The latter relation means that $\dot J_i \in T_J\scrj$. \qed
1679:
1680:
1681: \smallskip
1682: \newcorol{cor2.1.3} $\scrp^*$ is a $C^\ell$-smooth Banach manifold
1683: with the tangent space
1684: \begin{equation}
1685: T_{(u,J_S,J)} \scrp^* = \bigl\{ (v,\dot J_S, \dot J) \;:\;
1686: D_{u,J}v + J \scirc du \scirc \dot J_S + \dot J \scirc du \scirc J_S
1687: =0 \;\bigr\}.
1688: \eqqno(2.1.2)
1689: \end{equation}
1690: \end{corol}
1691:
1692: \state Remark. $\scrp^*$ is in general smaller than the set $\scrp^\trans
1693: \deff \sigma_\dbar \times^\trans_{\scre'} \sigma_0$ of all transversality
1694: points of $\scrp$. On the other hand, it is sufficient for applications to
1695: consider only the space $\scrp^*$.
1696:
1697:
1698:
1699: \newsubsection[2.2]{Moduli space of pseudoholomorphic curves}
1700: The space $\scrp$ (see \eqqref(1.3.2)) of all pseudoholomorphic maps is too big.
1701: Indeed, one has a natural right action of the group $\diff _+(S)$ of the
1702: orientation preserving $C^{2,\alpha}$-smooth diffeomorphisms of $S$ on the
1703: product $\scrs \times \scrj_S \times \scrj$ given by formula
1704: \begin{equation}
1705: (u,J_S,J) \in \scrs \times \scrj_S \times \scrj, g\in \diff_+(S)
1706: \; \lrar \; (u,J_S,J)\cdot g \deff (u\scirc g, J_S \scirc g, J),
1707: \end{equation}
1708: such that $\scrp$ and $\scrp^*$ are invariant \wrt this action. It is natural
1709: to consider $(u,J_S,J) \in \scrp$ and $(u,J_S,J) \cdot g$ as two
1710: parameterization of the same a $J$-curve. In other words, we are interested
1711: in the quotient space $\scrp/ \diff_+(S)$ rather than the space $\scrp$
1712: itself. As in Yang-Mills theory one can treat the group
1713: $\diff_+(S)$ as the gauge group of the problem and the quotient as
1714: the corresponding moduli space. Again as in Yang-Mills theory it is useful to
1715: know in which points the quotient spaces $\scrp/\diff_+(S)$ is
1716: a Banach manifold.
1717:
1718: \smallskip
1719: First we consider the action of the group $\diff_+(S)$ on the space $\scrj_S$.
1720: Note that if $J'_S, J''_S \in \scrj_S$ are related by $J''_S = J'_S \scirc g$
1721: for some $C^1$-diffeomorphism $g:S \to S$, then $g$ is $(J'_S, J''_S)
1722: $-holomorphic. Since $J'_S$ and $J''_S$ are $C^{1,\alpha}$-smooth, elliptic
1723: regularity implies that $g$ is $C^{2,\alpha}$-smooth.
1724:
1725: Further, it is known that the action of $\diff_+(S)$ on $\scrj_S$
1726: admits a global finite-dimen\-sional slice. To describe this slice we recall
1727: some standard facts from Teichm\"uller theory.
1728:
1729: \smallskip
1730: Denote by $\ttt_g$ the Teichm\"uller space of marked complex structures on $S$.
1731: This is a complex manifold of dimension
1732: \begin{equation}
1733: \dim_\cc \ttt_g = \cases 0 &\text{ if $g=0$;} \\
1734: 1 &\text{ if $g=1$;} \\
1735: 3g-3 &\text{ if $g\ge2$;}
1736: \endcases
1737: \end{equation}
1738: which can be completely characterized in the following way.
1739:
1740: \newprop{prop2.2.A1} The product
1741: $S\times \ttt_g$ admits a (non-unique) complex (\ie holomorphic) structure
1742: $J_{S\times \ttt}$ such that:
1743:
1744: \sli The natural projection $\pi_\ttt: S\times \ttt_g \to \ttt_g$ is
1745: holomorphic, so that for any $\tau\in \ttt_g$ the identification $S \cong S
1746: \times \{\tau\}$ induces the complex structure $J_S(\tau)\deff J_{S\times
1747: \ttt}\ogran_{S\times \{\tau\}}$ on $S$;
1748:
1749: \slii For any complex structure $I_S$ on $S$ there exist a uniquely
1750: defined $\tau\in\ttt_g$ and a diffeomorphism $f: S\to S$ homotopic to the
1751: identity map $\id_S: S\to S$ such that $I_S = f^*J_S(\tau)$;
1752:
1753: \sliii Moreover, for any finite-dimen\-sional manifold $Y$ and any smooth
1754: map $H: Y \to \scrj_S$ there exist maps $F:Y \to \diff_+(S)$ and $h:Y \to
1755: \ttt$ such that $H(y) = (F(y))^* h(y)$;
1756:
1757: \sliv The group $\bfg$ of automorphisms of $S\times \ttt_g$ preserving the
1758: projection onto $\ttt_g$ is
1759: \begin{equation}
1760: \bfg=\cases \mathbf{PGl}(2,\cc) &\text{ for $g=0$,} \\
1761: \mathbf{Sp}(2,\;\;\zz) \ltimes T^2 &\text{ for $g=1$,} \\
1762: \hbox{\rm discrete} &\text{ for $g\ge2$;}
1763: \endcases
1764: \eqqno(2.2.1)
1765: \end{equation}
1766:
1767: \slv The tangent space to $\ttt_g$ at $\tau$ is canonically isomorphic to
1768: $\sfh^1(S, TS)$ where $S$ is equipped with the structure $J_S(\tau)$. The
1769: group $\sfh^0(S, TS)$ is canonically isomorphic to the Lie algebra of
1770: $\bfg$.
1771: \end{prop}
1772:
1773: \smallskip
1774: We shall assume that such a structure $J_{S \times \ttt}$ is fixed. Then we
1775: obtain an imbedding $\ttt \hook \scrj_S$ given by $\tau \in \ttt \mapsto
1776: J_S(\tau) \in \scrj_S$. Using it, we identify $\ttt$ with its image in
1777: $\scrj_S$. For any $J_S \in \ttt_g$ this induces a monomorphism
1778: $T_{J_S}\ttt_g \hook T_{J_S}\scrj_S=C^{1,\alpha}_{(0,1)}(S, TS)$. Now, the
1779: isomorphism $T_{J_S}\ttt_g \cong \sfh^1(S, TS)$ mentioned in
1780: \slv is obtained as the composition
1781: \begin{equation}
1782: T_{J_S}\ttt_g \hook T_{J_S}\scrj_S=C^{1,\alpha}_{(0,1)}(S, TS) \lrar
1783: C^{1,\alpha}_{(0,1)}(S, TS)/\dbar\bigl(C^{2,\alpha}(S, TS)\bigr) =
1784: \sfh^1(S, TS).
1785: \eqqno(2.2.2)
1786: \end{equation}
1787:
1788: \smallskip
1789: By our construction, any $\diff_+(S)$-orbit in $\scrj_S$ intersects
1790: $\ttt$. This implies that instead of $\scrp/\diff_+(S)$ we can consider the
1791: quotient $\scrp \cap (\scrs \times \ttt \times \scrj_S)$ by the action of
1792: $\bfg$.
1793:
1794: \newdefi{def2.2.1} Let $\whcalm \deff \scrp^* \cap (\scrs \times
1795: \ttt \times \scrj_S)$ and use the same the notations to the restrictions of
1796: the bundles $\scre$ and $\scre'$ onto $\whcalm$ and for the induced
1797: operator $D: \scre\to \scre'$. The quotient $\scrm \deff \whcalm/ \bfg$ is
1798: the {\sl total moduli space of parameterized pseudoholomorphic curves}. It
1799: is equipped with the projection $\pr_{\!\!\scrj}:\scrm \to \scrj$. Elements of
1800: $\scrm$ are denoted by $[u,J]$. To indicate the surface $S$, the ambient manifold
1801: $X$, and the homology class $[C]\in \sfh_2(X, \zz)$ involved in the definition
1802: of $\scrm$ we shall also use the notation $\scrm(S, X, [C])$. The same meaning
1803: have the notations $\whcalm(S, X, [C])$ and $\scrp(S, X, [C])$.
1804: \end{defi}
1805:
1806: \newlemma{lem2.2.1}.
1807: \sli The projection $\wh\pr:\whcalm \lrar \scrm$ is a principal $\bfg$-bundle.
1808:
1809: \slii The bundles\/ $\scre$ and $\scre'$ over $\whcalm$ admit a natural
1810: $C^\ell$-smooth $\bfg$-action such that $D:\scre \to \scre'$ is
1811: $\bfg$-invariant.
1812:
1813: \sliii For any $J \in \scrj$ and any non-multiple $J$-holomorphic map $u: S
1814: \to X$ there exists a diffeomorphism $\phi: S \to S$ such that $[u\scirc \phi,
1815: J]$ lies in $\scrm$. Moreover, such element of $\scrm$ is unique.
1816: \end{lem}
1817:
1818:
1819: \state Remarks.~1. Part \slii of the lemma is equivalent to the existence
1820: of $C^\ell$-smooth bundles $\scre_\scrm$ and $\scre'_\scrm$ over $\scrm$ and
1821: a $C^{\ell-1}$-smooth bundle homomorphism $D_\scrm: \scre_\scrm \to \scre'_\scrm$
1822: which lift to the corresponding objects over $\whcalm$. Later on we drop the
1823: sub-index $_\scrm$, so that, \eg $\scre$ will denote also the corresponding
1824: bundle over $\scrm$.
1825:
1826: \state 2. Our main interest is the space $\scrm$. However, in the proofs below
1827: we shall mostly work with $\whcalm$. The reason is that an element $(u, J_S,
1828: J) \in \whcalm$ fixes a parameterization of a pseudoholomorphic curve, whereas
1829: $[u,J] \in \scrm$ defines only an appropriate equivalence class of
1830: parameterizations.
1831:
1832:
1833:
1834:
1835: \proof Part \slip.
1836:
1837: {\sl Case $g\ge2$}. It is known that in this case $\bfg$ acts properly
1838: discontinuously on $\ttt_g$. This implies that the same is true for the action
1839: of $\bfg$ on $\whcalm$. Moreover, it is clear that $\bfg$ acts freely on $\whcalm$.
1840: %Moreover, the definition of $\whcalm$ provides that $\bfg$ acts freely on
1841: %this space.
1842: Consequently, the map $\whcalm\lrar \scrm = \whcalm/\bfg$ is simply an (unbranched)
1843: covering.
1844:
1845: \smallskip
1846: {\sl Case $g=0$}. In this case $S=S^2$, $\ttt_0=\{J\st\}$, and the action of
1847: $\bfg$ on $S$ is generated by holomorphic vector fields, \ie by the space
1848: $\sfh^0(S,TS)$. One can show that the action of $\bfg$ on
1849: $\whcalm$ is generated by vector fields
1850: \begin{equation}
1851: (u, J\st,J)\in \whcalm \mapsto
1852: (du(v), 0,0) \in T_{(u, J\st, J)}\whcalm
1853: \quad\hbox{with $v\in \sfh^0(S,TS)$ fixed}.
1854: \end{equation}
1855: In particular, the action is {\sl smooth} or, more precisely, $C^\ell$-smooth.
1856:
1857: Consequently, for a given $(u^0,J\st,J^0)\in \whcalm$ we can find a closed
1858: complementing space $\scrv \subset T_{(u^0,J\st,J^0)}\whcalm$ to $(u^0 (\sfh
1859: ^0(S,TS)),0,0)$. Represent it as the tangent space of a submanifold $\scrw
1860: \subset \whcalm$ through $(u^0,J\st,J^0)$, $T_{(u^0,J\st,J^0)} \scrw=\scrv$.
1861: If $\scrv$ is chosen sufficiently small, then it intersects every orbit $\bfg
1862: \cdot (u,J\st,J)$ transversally in exactly one point. Moreover, we have a
1863: $\bfg$-invariant diffeomorphism $\bfg\cdot \scrw \cong \bfg
1864: \times \scrw$, so that $\scrw$ is a local slice of $\bfg$-action at $(u^0,
1865: J\st,J^0)$. This equips the quotient $\whcalm/\bfg$ with a structure
1866: of a smooth Banach manifold such that the projection $\scrw \to \scrm=
1867: \whcalm/\bfg$ is a $C^\ell$-smooth chart.
1868:
1869: \smallskip
1870: {\sl Case $g=1$} is a combination of the above two cases. First we consider
1871: the action of $T^2 \triangleleft \bfg$. The existence of a local $T^2$-slice
1872: $\scrw$ through any given $(u^0,J_S^0, J^0)\in \whcalm$ can be
1873: shown by copying the construction of {\sl Case $g=0$}. This implies that
1874: $\whcalm \lrar \whcalm/T^2$ is a principle $T^2$-bundle. Then we repeat
1875: the argument of {\sl Case $g\ge2$} and show that $\whcalm/T^2 \to
1876: \whcalm/\bfg$ is an unbranched covering with the group $\mathbf{Sp}(2,\zz) =
1877: \bfg/ T^2$. This completes the proof of part \slip.
1878:
1879: \medskip {\sl Part} \sliip.
1880: The action of $\bfg$ extends in a natural way to an action on
1881: $\scrz\deff S \times \scrs \times \ttt \times \scrj$. The evaluation map
1882: $\ev: \scrz \to X$, $\ev(z, u, J_S, J) \deff u(z)$ is $\bfg$-equivariant.
1883: Consequently, the bundle $E\deff \ev^* TX$ over $\scrz$ is equipped
1884: with the natural $\bfg$-action. The action of $\bfg$ on $E$ induces the
1885: actions on section spaces $\scre$ and $\scre'$. Since all constructions are
1886: natural, $D: \scre \to \scre'$ is $\bfg$-invariant.
1887:
1888: Finally, it remains to note that the action of $\bfg$ on the bundles $\scre$
1889: and $\scre'$ over $\whcalm$ is $C^\ell$-smooth.
1890:
1891: \smallskip\noindent
1892: {\sl Part} \sliii of the lemma states the universality property of
1893: $\scrm$ which easily follows from the definitions.
1894: \qed
1895:
1896: \newcorol{cor2.2.2} $\scrm$ is a $C^\ell$-smooth Banach
1897: manifold and $\pi_{\!\!\scrj}:\scrm \to \scrj$ is a Fredholm map. For
1898: $[u,J]\in\scrm$ there exist natural isomorphisms
1899: %\begin{equation}
1900: \begin{align*}
1901: \ker(d\pi_{\!\!\scrj}:T_{[u,J]}\scrm \to T_J\scrj)&
1902: \;\cong\; \sfh^0_D(S, \scrn_u),
1903: \\
1904: \coker(d\pi_{\!\!\scrj}:T_{[u,J]} \scrm \to T_J\scrj)
1905: &\;\cong\; \sfh^1_D(S, \scrn_u).
1906: \end{align*}
1907: %\end{equation}
1908: In particular, the index of $\pi_{\!\!\scrj}$ is equal to
1909: \begin{equation}
1910: \ind_\rr \pi_{\!\!\scrj} = \chi_\rr(\scrn_u)= 2\bigl(\mu + (n-3)(1-g) \bigr),
1911: \eqqno(2.2.3)
1912: \end{equation}
1913: where $\mu = \la c_1(X), [C] \ra$.
1914: \end{corol}
1915:
1916: \proof The $C^\ell$-smooth structure on $\scrm$ is is the quotient structure
1917: defined by the the $C^\ell$-smooth $\bfg$-action on $\whcalm$.
1918: %provided by the fact that the action on $\bfg$ on $\whcalm$ is $C^\ell$-smooth.
1919:
1920: \smallskip
1921: Using \refcorol{cor2.1.3} we see that the tangent space
1922: to $\whcalm$ is
1923: \begin{equation}
1924: T_{(u,J_S,J)} \whcalm = \bigl\{ (v,\dot J_S, \dot J) \,:\,
1925: \dot J_S \in T_{J_S}\ttt,\;\,
1926: D_{u,J}v + J \scirc du \scirc \dot J_S + \dot J \scirc du \scirc J_S
1927: =0 \,\bigr\}.
1928: \eqqno(2.2.4)
1929: \end{equation}
1930: Consider the natural projection $\pi_\scrp: \scrp^* \to \scrj$, $(u,J_S,J)
1931: \mapsto J$ with the differential $d\pi_\scrp: T_{(u,J_S,J)}\scrp^* \to
1932: T_J\scrj$ given by $(v,\dot J_S,\dot J)\in T_{(u,J_S,J)}\scrp^*\mapsto
1933: \dot J\in T_J\scrj$.
1934:
1935: The kernel $\ker(d\pi_\scrp)$ consists of solutions $v\in \scre_{(u,J)}$
1936: of the equation
1937: \begin{equation}
1938: D_{u,J}v + J\scirc du \scirc \dot J_S =0
1939: \end{equation}
1940: with $\dot J_S\in T_{J_S}\ttt$. Since the map
1941: $\wh\pi: \wh\scrm\to \scrm$ is a principle $\bfg$-bundle, the kernel
1942: $\ker( d\pi_{\!\!\scrj}:T_{(M,J)}\scrm \to T_J\scrj)$ is obtained from $\ker(d\pi)$
1943: by taking the quotient by the tangent space to the fiber $\bfg
1944: \cdot (u,J_S,J)$ which is equal to $du(\sfh^0(S,TS))$. Using the relations
1945: $\sfh^0(S,TS)= \ker(\dbar_{TS}:L^{1,p}(S,TS)\to L^p(S,TS\otimes
1946: \Lambda^{(0,1)}S)$, $T_{J_S}\ttt_g\cong\sfh^1(S,TS)=\coker(\dbar_{TS})$, and
1947: $du\scirc \dbar_{TS}= D_{(u,J)}\scirc du$, we conclude that the space
1948: $\ker(d\pi_{\!\!\scrj})$ is isomorphic to the quotient
1949: \begin{equation}
1950: \{v\in L^{1,p}(S,E_u)\;:\; Dv=du(\phi) \hbox{ for some }\phi\in
1951: L^p(S,TS\otimes\Lambda^{(0,1)}S)\}
1952: \!\!\bigm/\!\!
1953: du\bigl(L^{1,p}(S,TS)\bigr).
1954: \end{equation}
1955: Hence, by \lemma{lem1.5.3}, $\ker(d\pi_{\!\!\scrj}:T_{[u,J]}\scrm\to T_J\scrj)
1956: \cong \sfh^0_D(S,\scrn_u)$. In particular, $\ker(d\pi_{\!\!\scrj})$ is finite
1957: dimensional.
1958:
1959: \smallskip
1960: Similarly, the image of $d\pi_{\!\!\scrj}$ consists of those $\dot J$ for which
1961: the equation
1962: \begin{equation}
1963: D_{u,J}v + J\scirc du \scirc \dot J_S + \dot J\scirc
1964: du \scirc J_S =0
1965: \end{equation}
1966: has a solution $(v,\dot J_S)$. Using \eqqref(1.5.6) and \lemma{lem1.5.3}
1967: we obtain the relations $\im(d\pi_{\!\!\scrj})=\ker\barr\Psi$ and
1968: $\coker(d\pi) \cong \sfh^1_D(S,N_u)$.
1969:
1970: This implies the Fredholm property for the projection $\pi_{\!\!\scrj}:\scrm \to
1971: \scrj$ and the formula $\ind(d\pi_{\!\!\scrj}) = \ind (\scrn_u)$.
1972: \qed
1973:
1974: \medskip
1975: We conclude the paragraph with a description of the deformations of non-closed
1976: pseudoholomorphic curves.
1977:
1978: \newdefi{def2.2.2} Let $\barr S=S \cup \d S$ be a compact {\sl non-closed}
1979: oriented surface with the boundary $\d S$ consisting of finitely many circles.
1980: Denote by $\scrj_S$ the (Banach) space of complex structure on $S$ which are
1981: compatible with the orientation of $S$ and $C^\ell$-smooth up to boundary
1982: $\d S$. As usual let
1983: \begin{equation}
1984: \eqqno(2.2.11)
1985: \scrp(S, X) \deff \{ (u,J_S,J) \in L^{1,p}(S, X) \times \scrj_S \times
1986: \scrj: \dbar_{J_S,J} u=0\},
1987: \end{equation}
1988: the space of pseudoholomorphic maps equipped it with the natural projections
1989: $\pr_{\scrj_S}: \scrp(S, X) \to \scrj_S$ and $\pr_{\!\!\scrj}: \scrp(S, X) \to
1990: \scrj$. The fibers of the projections are denoted by $\scrp(S, X, J)=\pr_{\!\!\scrj}
1991: \inv(J)$, $\scrp(S,J_S, X)=\pr_{\scrj_S}\inv(J_S)$, and $\scrp(S,J_S, X,J)=
1992: \scrp(S,J_S, X) \cap \scrp(S, X, J)$ respectively.
1993: \end{defi}
1994:
1995: \newlemma{lem2.2.4} \sli Let $S$ be a non-closed oriented surface. Then
1996:
1997: \sli the space $\scrp(S, X)$ is a Banach submanifolds of $L^{1,p}(S, X)
1998: \times \scrj_S\times \scrj$;
1999:
2000: \slii For any $(u,J_S, J)\in \scrp(S, X)$, the operators $d\pr_{\!\!\scrj}: T_{(u,
2001: J_S,J)}\scrp(S, X) \to T_J\scrj$ and $d\pr_{\scrj_S}: T_{(u,J_S,J)} \scrp(S,
2002: X) \to T_{J_S}\scrj$ are surjective and split. In particular, $\scrp(S; X,J)$
2003: and $\scrp(S,J_S; X)$ are are Banach submanifolds of $\scrp(S, X)$;
2004:
2005: \sliii For any $(u,J_S, J)\in \scrp(S, X)$, the submanifolds $\scrp(S; X,J)$
2006: and $\scrp(S,J_S; X)$ are transversal in $(u,J_S,J)$; in particular, $\scrp(S, J_S;
2007: X,J)= \scrp(S; X,J) \cap \scrp(S,J_S; X)$ is also a Banach submanifold;
2008:
2009: \sliv The the tangent spaces are given by
2010: \begin{align}
2011: T_{(u,J_S,J)}\scrp(S, X) &= \bigl\{ (v,\dot J_S,\dot J) \in T_uL^{1,p}(S, X)
2012: \times T_{J_S}\scrj_S \times T_J\scrj :
2013: \notag\\
2014: &\qquad\quad D_{u,J}v + \dot J \scirc du \scirc J_S +
2015: J \scirc du \scirc \dot J_S =0\bigr\};
2016: \eqqno(2.2.12)\\
2017: T_{(u,J_S)}\scrp(S; X,J)& = \bigl\{ (v,\dot J_S,\dot J) \in
2018: T_{(u,J_S,J)}\scrp(S, X): \dot J=0 \bigr\};
2019: \eqqno(2.2.13)\\
2020: T_{(u,J)}\scrp(S,J_S; X)& = \bigl\{ (v,\dot J_S,\dot J) \in
2021: T_{(u,J_S,J)}\scrp(S, X): \dot J_S=0 \bigr\};
2022: \eqqno(1.2.14)\\
2023: T_u\scrp(S,J_S; X,J)& = \bigl\{ (v,\dot J_S,\dot J) \in
2024: T_{(u,J_S,J)}\scrp(S, X): \dot J_S=0 = \dot J \bigr\}.
2025: \eqqno(2.2.15)
2026: \end{align}
2027: \end{lem}
2028:
2029: \proof The lemma is obtained by the transversality techniques of
2030: \refsubsection{2.1} using the following claim: {\sl For any $(u,J_S,J) \in
2031: \scrp(S,X)$ the operator $D_{u,J} : L^{1,p}(C, E_u) \to L^p_{(0,1)}(C, E_u)$
2032: is surjective}. Since $D_{u,J}$ is elliptic, this is a standard fact following
2033: from compactness and non-closedness of $S$. \qed
2034:
2035:
2036: \newsubsection[2.3]{Transversality I$\!$I}
2037: Before stating further results, we introduce some new notation. Here $S$ is a
2038: {\sl closed} real surface.
2039:
2040: \newdefi{def2.3.1}
2041: Let $Y$ be a $C^\ell$-smooth finite-dimen\-sional manifold, possibly with
2042: non-empty $C^\ell$-smooth boundary $\d Y$, and $h:Y \to \scrj$ a $C^\ell
2043: $-smooth map. Define the {\sl relative Moduli space}
2044: \begin{equation}
2045: \scrm_h \deff Y\times_{\!\!\scrj} \scrm \cong
2046: \{\,(u,J_S,y)\in \scrs\times \ttt_g\times Y\,:\, (u,J_S, h(y))\in \scrp^*
2047: \,\}/\bfg
2048: \end{equation}
2049: with the natural projection $\pi_h:\scrm_h \to Y$. In the special case
2050: $Y=\{J\}\hook \scrj$, we obtain the Moduli space of $J$-holomorphic curves
2051: $\scrm_J\deff \pi_{\!\!\scrj}\inv(J)$. The projection $\pi_h:\scrm_h \to Y$
2052: is a fibration with a fiber $\pi_h\inv(y)=\scrm_{h(y)}$. We shall denote
2053: elements of $\scrm_h$ by $[u,y]$, where $u:S\to X$ is a $h(y)$-holomorphic map.
2054: \end{defi}
2055:
2056: \medskip
2057: The next two lemmas follow from the transversality theory.
2058:
2059: \newlemma{lem2.3.1}
2060: Let $Y$ be a $C^\ell$-smooth finite-dimen\-sional
2061: manifold, and $h:Y \to \scrj$ a $C^\ell$-smooth map. Then $\scrm_h$ is a
2062: $C^\ell$-smooth manifold in some neighborhood of a point $[u,y] \in
2063: \scrm_h$ with $J \deff h(y)$ \iff the map $\barr\Psi_{u, J} \scirc dh
2064: : T_uY \to \sfh^1_D(S, N_u)$ is surjective.
2065: In this case the tangent space to $\scrm_h$ is
2066: \begin{equation}
2067: T_{[u,y]}\scrm_h=\ker\bigl(D\;\oplus\; \Psi\scirc dh:
2068: \scre_{u,h(y)}\oplus T_yY \lrar \scre'_{u,h(y)} \bigr)
2069: \bigm/ du(\sfh^0(S, TS))
2070: \eqqno(2.3.1)
2071: \end{equation}
2072: \end{lem}
2073:
2074:
2075: \proof We reformulate the transversality condition and
2076: use \lemma{lem2.1.1}.
2077: \qed
2078:
2079: \newlemma{lem2.3.2}
2080: \sli There exists a Baire subset $\scrj\reg \subset
2081: \scrj$ such that any $J \in \scrj$ is a regular value of $\pi_{\!\!\scrj} :\scrm
2082: \to \scrj$.
2083:
2084: \slii There exists a Baire subset $\scrv$ in the space $C^\ell([0,1], \;
2085: \scrj)$, such that any map $h:[0,1]\to \scrj$ from $\scrv$ is transversal to
2086: $\pi_{\!\!\scrj} :\scrm \to \scrj$ and both $h(0)$ and $h(1)$ are
2087: regular values of $\pi_{\!\!\scrj}$.
2088: \end{lem}
2089:
2090:
2091: \state Remark. In general, for any finite-dimen\-sional manifold $Y$ with
2092: boundary $\d Y$ there exists a Baire subset $\scrv \subset C^\ell(Y,\; \scrj)$
2093: such that any $h \in \scrv$, as well as its restriction $h\ogran_{\d Y}$ are
2094: transversal to $\pi_{\!\!\scrj}$. The proof uses the Sard lemma.
2095:
2096: \newlemma{lem2.3.3}
2097: Suppose that $S$ is the sphere $S^2$ and $\dim_\rr(X) =4$.
2098: Then the exists a {\sl connected} Baire subset $\scrj\reg \subset
2099: \scrj$ such that any $J \in \scrj$ is a regular value of $\pi_{\!\!\scrj} :\scrm
2100: \to \scrj$. Moreover, any $J_0, J_1 \in \scrj\reg$ can be connected by
2101: a smooth path $h: [0,1] \to \scrj\reg$.
2102: \end{lem}
2103:
2104:
2105: \proof By \lemma{lem2.3.2}, there exists a Baire subset $\scrj\reg
2106: \subset\scrj$ such that any $J \in \scrj\reg$ is a regular value of
2107: $\pi_{\!\!\scrj}$. Further, any $J_0, J_1 \in \scrj\reg$ can be connected by a
2108: smooth path $h: [0,1] \to \scrj$, transversal to $\pi_{\!\!\scrj}$.
2109:
2110: For any such path $h: [0,1] \to \scrj$ and any $[u,t] \in \scrm_h$ with
2111: $h(t)=J$ the map $\barr\Psi_{u,J}\scirc dh: T_t[0,1]\cong\rr \; \lrar
2112: \sfh^1(S, N_u)$ is surjective by \lemma{lem2.3.1}. Consequently,
2113: for such $u$ and $J$ we have $\dimr \sfh^1_D(S, \scrn_u)= \dimr \sfh^1_D(S,
2114: N_u) \le 1$.
2115:
2116: Recall that the difference $\dimr\sfh^0_D(S, N_u) -\dimr\sfh^1_D(S, N_u)$ is
2117: {\sl even}, see \eqqref(1.5.?4). Hence, if $\dimr \sfh^1(S, N_u)= 1$ then
2118: $\sfh^1_D(S, N_u)$ should also be nontrivial. On the other hand, the condition
2119: $\dimr(X)=4$ implies that $N_u$ is a {\sl line bundle}. But, in view of {\sl
2120: Lemma 1.5.2}, on the sphere $S=S^2$ one of the spaces $\sfh^i_D(S, N_u)$ must
2121: be trivial.
2122:
2123: Thus, we see that for any such path $h$ and any $[u,t]\in \scrm_h$ one has
2124: $\sfh^1_D(S, \scrn_u)=0$. This means that $h$ takes values in $\scrj\reg$.
2125: \qed
2126:
2127: \smallskip
2128: For higher genus $g(S) \ge1$ we have a similar, but weaker result.
2129:
2130: \newlemma{lem2.3.4}
2131: Assume that $\dim_\rr X =4$ and $g\deff g(S) \ge1$
2132: and set $\mu \deff c_1(X)[u(S)]$. Then
2133: \begin{equation}
2134: \mu \le \dim_\cc \sfh^0(S, \scrn\sing_u) \le \mu + g-1
2135: \eqqno(2.3.2)
2136: \end{equation}
2137: for any $[u,J] \in \scrm$ with $\sfh^1(S, N_u)\cong \rr$.
2138: \end{lem}
2139:
2140:
2141: \proof The condition $\dim_\rr X =4$ means that $N_u$ is a {\sl line} bundle.
2142: Thus, by \lemma{lem1.5.2}, $\sfh^1(S, N_u)\cong \rr$ implies $c_1(N_u) \le 2g
2143: -2$. Further, since $\dim_\rr \sfh^1(S, N_u)=1$ and $\ind_\rr D^N_{u,J} = \dimr
2144: \sfh^0_D(S, N_u) -\dimr\sfh^1_D(S, N_u)$ is {\sl even}, see \eqqref(1.5.?4), we
2145: conclude that $\dimr\sfh^0_D(S, N_u) \ge 1$. Consequently, $\ind_\rr D^N_{u,J}
2146: \ge 0$. The formula \eqqref(1.5.?4) for $\ind_\rr D^N_{u,J}$ yields $c_1(N_u)
2147: \ge g-1$. Finally, the definition of $\scrn_u$ yields the relation $\mu= c_1(X)
2148: [u(S)] = c_1(E_u) = c_1(TS) + c_1(N_u) + \dimc \sfh^0(S, \scrn\sing_u)$. \qed
2149:
2150: \newsubsection[2.4]{Pseudoholomorphic curves through fixed points} In this
2151: paragraph we consider the total moduli space of pseudoholomorphic curves passing
2152: through given fixed points $x_1, \ldots, x_m \allowbreak
2153: \in X$. Gromov in \cite{Gro} proposed a method to reduce
2154: this problem to the one of pseudoholomorphic curves
2155: without such constraints. The idea is to blow up $X$ in the points $x_1, \ldots,
2156: x_m$ and consider the curves on the blown-up space $\ti X$. He has shown that a
2157: $C^\ell$-smooth almost complex structure $J$ on $X$ lifts to a $C^{\ell-1}$-smooth
2158: almost complex structure $\ti J$ on $\ti X$ such that the natural
2159: projection $\pr: \ti X \to X$ is holomorphic and such that every $J$-holomorphic
2160: curve $C$ in $X$ passing through $x_1, \ldots, x_m$ lifts to a unique $\ti
2161: J$-holomorphic curve $\ti C$ in $\ti X$ with $C= \pr(\ti C)$.
2162:
2163:
2164:
2165: \smallskip
2166: Our aim here is to make an explicit construction for the moduli space of
2167: pseudoholomorphic curves passing through fixed points. Since the construction is
2168: simply a modification of the case $m=0$ where no points are marked, we shall
2169: mostly skip or merely indicate proof of claims.
2170:
2171: We begin by introducing some notation. Denote by $\mbfx=(x_1, \ldots x_m)$ the
2172: tuple of fixed points on $X$, which are supposed to be pairwise distinct.
2173: Also fix a tuple $\mbfz=(z_1, \ldots, z_m)$ of pairwise distinct points on the
2174: surface $S$. Define
2175: $$
2176: \scrs(\mbfz, \mbfx) \deff \{ u \in \scrs=L^{1,p}(S, X): u(z_i)=x_i \};
2177: $$
2178: $$
2179: \scrp(S,\mbfz; X, \mbfx) \deff \{ (u, J_S, J) \in \scrp: u\in \scrs(\mbfz,
2180: \mbfx) \};
2181: $$
2182: $$
2183: \scrp^*(S,\mbfz; X, \mbfx) \deff \scrp(S,\mbfz; X, \mbfx) \cap \scrp^*(S,X).
2184: $$
2185: The linearization of the conditions $u(z_i)=x_i$ yields the equations $v(z_i)=0$
2186: for $v \in T_u\scrs= L^{1,p}(S, E_u)$. Denote as above $E=E_u \deff u^*TX$, and
2187: set $E_i= E_{u,i} \deff (E_u)_{z_i} = T_{u(z_i)}X$. Then we obtain the bundle
2188: $E_\mbfz$ over $\scrs$ with a fiber $(E_\mbfz)_u\deff \oplus E_{u,i}$ equipped with
2189: the natural {\sl evaluation homomorphism} $\ev_\mbfz: \scre \to E_\mbfz$
2190: $$
2191: \ev_{\mbfz}: v \in \scre_u=L^{1,p}(S, E) \mapsto (v(z_1), \ldots, v(z_m)) \in
2192: E_\mbfz.
2193: $$
2194: It is easy to see that $\ev_\mbfz: \scre \to E_\mbfz$ is surjective. This means
2195: that the equations $u(z_i)=x_i$ are transversal and implies that $\scrs(\mbfz,
2196: \mbfx)$ is a Banach submanifold of $\scrs$ with the tangent space
2197: \begin{equation}\eqqno(2.4.1)
2198: T_u\scrs(\mbfz, \mbfx) = \{ v \in T_u\scrs= L^{1,p}(S, E_u): v(z_i)=0 \}.
2199: \end{equation}
2200: The same argument shows that $\scrp^*(S,\mbfz; X, \mbfx)$ is also a Banach
2201: submanifold of $\scrp^*(S,X)$ with the tangent space
2202: $$
2203: T_u\scrp^*(S,\mbfz; X, \mbfx) = \{ (v, \dot J_S, \dot J) \in
2204: T_u\scrp^*(S,X): v(z_i)=0 \}.
2205: $$
2206:
2207: \smallskip
2208: Let $\diff_+(S, \mbfz)$ be the subgroup of those $g\in \diff_+(S)$ which fix the
2209: marked points $z_1,\ldots, z_m$. Then $\diff_+(S, \mbfz)$ leaves the subsets
2210: $\scrs(\mbfz, \mbfx) \subset \scrs$ and $\scrp^*(S,\mbfz; X, \mbfx) \subset
2211: \scrp(S,X)$ invariant. So we can define the {\sl total moduli space of
2212: pseudoholomorphic curves through the given points $x_1, \ldots, x_m$} as the
2213: quotient $\scrm(\mbfx) \deff \scrp^*(S,\mbfz; X, \mbfx) /\diff_+(S, \mbfz)$. This
2214: space is equipped with the natural projection $\pi_\scrj: \scrm(\mbfx) \to \scrj$
2215: defined in an obvious way.
2216:
2217: \smallskip
2218: The smooth structure on $\scrm(\mbfx)$ is constructed in the same way as it was
2219: for $\scrm$. First, one constructs a global slice on the action of $\diff_+
2220: (S, \mbfz)$ on $\scrj_S$. To do this, we consider the action of the component
2221: of $\diff_0(S, \mbfz)$ the group $\diff_+(S, \mbfz)$ containing the identity. The
2222: quotient $\scrj_S/ \diff_0(S, \mbfz)$ is the {\sl Teichm\"uller space $\ttt
2223: _{g,m}$ of complex structures on a Riemann surface of genus $g=g(S)$ with $m$
2224: punctures}. The marked points $z_1,\ldots, z_m$ are the positions of punctures.
2225:
2226: As in the case $m=0$, one can imbed $\ttt_{g,m}$ in $\scrj_S$ in such a way that
2227: the composition $\ttt_{g,m} \hook \scrj_S \twoheadrightarrow \scrj_S/ \diff_0(S,
2228: \mbfz) \cong \ttt_{g,m}$ is the identity map. This imbedding $\ttt_{g,m} \hook
2229: \scrj_S$ is the desired slice. The choice of such imbedding $\ttt_{g,m} \hook
2230: \scrj_S$ is equivalent to the choice of the complex structure $J_{\ttt, S, \mbfz}$
2231: on the product $\ttt_{g,m} \times S$. One considers $\ttt_{g,m} \times S$ with
2232: this complex structure and with the holomorphic projection $\pr: \ttt_{g,m}
2233: \times S \to \ttt_{g,m}$ as the universal family corresponding to $\ttt_{g,m}$.
2234:
2235:
2236: After the choice of the slice $\ttt_{g,m} \hook \scrj_S$, $\scrm(\mbfx)$ is
2237: obtained as the quotient of the space
2238: $$
2239: \wh\scrm(\mbfx) \deff \{ (u, J_S, J) \in \scrp^*(S,\mbfz; X, \mbfx):
2240: J_S \in \ttt_{g,m} \}
2241: $$
2242: by the group $\bfg$ of biholomorphisms of $\ttt_{g,m} \times S$ preserving the
2243: projection $\pr: \ttt_{g,m} \times S \to \ttt_{g,m}$. It is discrete except the
2244: cases where $g=0$ and $m=1$ or $2$. In the case $m=1$, $S\bs \{z_1\}$ is the
2245: complex plane $\cc$ and $\bfg$ is its automorphism group $\cc^* \ltimes \cc$.
2246: Similarly, in the case $m=2$, $S\bs \{z_1, z_2\}$ is the punctured complex plane
2247: $\cc^*$ and $\bfg= \zz_2 \ltimes \cc^*$ is likewise its automorphism group. In
2248: either case one can construct a local slice for the action of $\bfg$ on $\wh\scrm
2249: (\mbfx)$ by repeating the arguments of \refsubsection{2.2}. So the quotient
2250: $\whcalm(\mbfx)/\bfg$ is a $C^\ell$-smooth Banach manifold.
2251:
2252: \medskip
2253: Now we define the notion of the normal sheaf of a pseudoholomorphic curve passing
2254: through fixed points on $X$. In this new situation, the linearization of
2255: $\dbar$-equations leads to the operator
2256: \begin{equation}\eqqno(2.4.1a)
2257: D=D_{u,J}: \{ v \in L^{1,p}(S,E_u): v(z_i)=0 \text{ for }i=1,\ldots,m\}
2258: \to L^p_{(0,1)}(S, E),
2259: \end{equation}
2260: which is the usual Gromov operator $D=D_{u,J}$, but now considered with
2261: a new domain of definition
2262: \begin{equation*}
2263: \scre_{u,\mbfx} \deff \{ v \in L^{1,p}(S,E_u): v(z_i)=0
2264: \text{ for }i=1,\ldots,m\}.
2265: \end{equation*}
2266: The space $\scre_{u,\mbfx}$ is the kernel of the evaluation homomorphism
2267: $\ev_z: \scre_u \to E_\mbfz$ and is the tangent plane to $\scrs(\mbfz, \mbfx)$,
2268: see \eqqref(2.4.1).
2269:
2270: We now describe the structure of the operator \eqqref(2.4.1a). Recall that we have
2271: the
2272: decomposition $D_{u,J}= \dbar_{u,J} + R_{u,J}$, see \refsubsection{1.4}. Observe
2273: that the sheaf $\scro(E_u)[-\mbfz]$ of holomorphic sections of $\scro(E_u)$
2274: vanishing at the points $z_1, \ldots, z_m \in S$ is locally free and hence
2275: corresponds to a holomorphic bundle. Let us denote this bundle by $E_{u, -\mbfz}$.
2276:
2277: \newlemma{lem2.4.1} \sli The (co)kernel of the operator
2278: \begin{equation}\eqqno(2.4.2a)
2279: \dbar_{u,J}: \{ v \in L^{1,p}(S,E_u): v(z_i)=0 \text{ for }i=1,\ldots,m\}
2280: \to L^p_{(0,1)}(S, E),
2281: \end{equation}
2282: is canonically isomorphic to the cohomology groups $\sfh^0_\dbar(S, E_{u,-\mbfz})$
2283: and $\sfh^1_\dbar(S, E_{u,-\mbfz})$.
2284:
2285: \slii The operator $D_{u,J}$ induces the operator
2286: $$
2287: D_{u,-\mbfz, J}: L^{1,p}(S, E_{u,-\mbfz}) \to L^p_{(0,1)}(S, E_{u,-\mbfz})
2288: $$
2289: which is of the form $D_{u,-\mbfz, J}= \dbar_{u,-\mbfz, J} + R_{u,-\mbfz, J}$, where
2290: $\dbar_{u,-\mbfz, J}$ is the Cauchy-Riemann operator corresponding to the natural
2291: holomorphic structure in $E_{u,-\mbfz}$ and $R_{u,-\mbfz, J}$ is a $\cc$-antilinear
2292: $L^\infty$-bounded bundle homomorphism, \ie
2293: $$
2294: R_{u,-\mbfz, J} \in L^\infty\bigl(S, \barr\hom_\cc(E_{u,-\mbfz},
2295: E_{u,-\mbfz} \otimes \Lambda^{(0,1)}S)\bigr).
2296: $$
2297:
2298: \sliii The (co)kernel of the operator
2299: \begin{equation}\eqqno(2.4.3a)
2300: D_{u,J}: \{ v \in L^{1,p}(S,E_u): v(z_i)=0 \text{ for }i=1,\ldots,m\}
2301: \to L^p_{(0,1)}(S, E)
2302: \end{equation}
2303: is canonically isomorphic to the cohomology groups $\sfh^0_D(S, E_{u,-\mbfz})$
2304: and $\sfh^1_D(S, E_{u,-\mbfz})$ corresponding to the operator $D_{u,-\mbfz, J}$.
2305: \end{lem}
2306:
2307: \proof Fix local holomorphic coordinates $\zeta_i$ on $S$, each centered at
2308: the corresponding marked point $z_i$. Consider the natural inclusions
2309: \begin{align}
2310: \eqqno(2.4.4a)
2311: j^0:L^{1,p}(S, E_{u,-\mbfz}) &\hook
2312: \{ v \in L^{1,p}(S,E_u): v(z_i)=0 \text{ for }i=1,\ldots,m\},
2313: \\
2314: \eqqno(2.4.5a)
2315: j^1: L^p_{(0,1)}(S, E_{u,-\mbfz}) &\hook L^p_{(0,1)}(S, E_u).
2316: \end{align}
2317: Observe that $v\in L^{1,p}(S,E_u)$ with $v(z_i)=0$ belongs to $L^p_{(0,1)}(S,
2318: E_{u,-\mbfz})$ \iff locally near every $z_i$ it has the form $v(\zeta_i)=
2319: \zeta_i w(\zeta_i)$ for some (uniquely defined!) $L^{1,p}$-section $w(\zeta_i)$
2320: of $E_{u,-\mbfz}$. This is equivalent to the condition $\zeta_i\inv \dbar_{u,J}
2321: v(\zeta_i) \in L^p$ as well as to the condition $\zeta_i\inv D_{u,J} v(\zeta_i)
2322: \in L^p$. Consequently, $D_{u,J}$ restricted to $L^{1,p}(S,E_{u,-\mbfz})$ takes
2323: values in $L^p_{(0,1)}(S, E_{u,-\mbfz})$. This yields the operator $D_{u,-\mbfz,
2324: J}$. Moreover, $D_{u,-\mbfz, J}$ is of order 1 and has the Cauchy-Riemann symbol.
2325: Consequently, it has the form $D_{u,-\mbfz, J}= \dbar_{u,-\mbfz, J} +R_{u,-\mbfz,
2326: J}$, where $\dbar_{u,-\mbfz, J}$ is the Cauchy-Riemann operator corresponding
2327: to the holomorphic structure in $E_{u,-\mbfz}$, and $R_{u,-\mbfz, J}$ is the
2328: $\cc$-antilinear part of $D_{u,-\mbfz, J}$.
2329:
2330: Let $v_1(\zeta_i), \ldots, v_n(\zeta_i)$ be a local holomorphic frame of $E_u$ in
2331: a neighborhood of $z_i$ and $R_{\alpha\beta}(\zeta_i)$ the matrix of $R_{u,J}$ in
2332: this frame. Then $\zeta_iv_1(\zeta_i), \ldots, \zeta_iv_n(\zeta_i)$ is a local
2333: frame of $E_{u,-\mbfz}$. From $\cc$-antilinearity of $R_{u,J}$ we obtain
2334: $$
2335: \textstyle
2336: R_{u,J}(\zeta_i\,v_\alpha(\zeta_i))=
2337: \sum_\beta R_{\alpha\beta}(\zeta_i)\bar \zeta_i v_\beta(\zeta_i)=
2338: \sum_\beta \msmall{\bar \zeta_i \over \zeta_i} R_{\alpha\beta}(\zeta_i)
2339: \cdot \zeta_i v_\beta(\zeta_i).
2340: $$
2341: This shows that ${\bar \zeta_i \over \zeta_i} R_{\alpha\beta}(\zeta_i)$ is the
2342: matrix of $R_{u, -\mbfz, J}$ in the frame $\zeta_iv_1(\zeta_i), \ldots, \zeta_i
2343: v_n(\zeta_i)$. Recall that $R_{u,J}$ is a continuous bundle homomorphism (see
2344: \lemma{lem1.4.1}, \slip). So we see that $R_{u, -\mbfz, J}$ is also continuous
2345: outside the marked points $z_i$ and has singularities of the form ${\bar
2346: \zeta_i \over \zeta_i} R_{u,J}$ at $z_i$. In particular, $R_{u, -\mbfz, J}$ is of
2347: type $L^\infty$, but is not continuous in general.
2348:
2349:
2350: \smallskip
2351: The equality of the kernels of the operators in \sli and \slii with the
2352: corresponding 0-cohomology groups follows directly from the definition of the
2353: operators $\dbar_{u,-\mbfz,J}$ and $D_{u,-\mbfz,J}$. The equality for
2354: 1-cohomology groups will be shown only for the operator $D_{u,-\mbfz,J}$, the
2355: other one is carried out in the same manner. So let $\phi \in L^p_{(0,1)}(S, E_{u,
2356: -\mbfz})$. If $\phi = D_{u,-\mbfz,J}(v)$ for $v \in L^{1,p}(S, E_{u, -\mbfz})$,
2357: then $v \in L^{1,p}(S, E_u)$ and $j^1\phi = D_{u,J}(v)$, or more precisely $j^1
2358: \phi = D_{u,J}(j^0(v))$. This shows that the inclusion $j^1$ in \eqqref(2.4.5a)
2359: induces a well-defined homomorphism from $\sfh^1_D(S, E_{u, -\mbfz})$ to the
2360: cokernel of \eqqref(2.4.3a). Moreover, $\phi \in L^p_{(0,1)}(S, E_{u, -\mbfz})$
2361: induces the zero class in the cokernel of \eqqref(2.4.3a) \iff $j^1(\phi) =
2362: D_{u,J}(v)$ for some $v \in L^{1,p}(S, E_u)$ with $v(z_i)=0$. But then locally
2363: $$
2364: D_{u,J}(\zeta_i\inv v(\zeta_i)) =
2365: \zeta_i\inv D_{u,-\mbfz,J}(v(\zeta_i)) =
2366: \zeta_i\inv j^1 \phi(\zeta_i) \in L^p
2367: $$
2368: by the definition of the inclusion $E_{u, -\mbfz} \hook E_u$. This implies
2369: that $v \in L^{1,p}(S, E_{u-\mbfz})$. Thus the homomorphism induces by $j^1$
2370: is injective.
2371:
2372: Further, for any $\phi \in L^p_{(0,1)}(S, E_u)$ there exists $v \in L^{1,p}(S,
2373: E_u)$ which vanishes at all $z_i$ and solves the equation $\phi = D_{u,J}(v)$
2374: in a neighborhood of every $z_i$. Then $\phi - D_{u,J}(v)$ represents the same
2375: class in the cokernel of \eqqref(2.4.3a) and is of the form $\phi - D_{u,J}(v)
2376: = j^1(\psi)$ for some $\psi \in L^p_{(0,1)}(S, E_{u, -\mbfz})$. This finishes
2377: the proof of the claim \sliiip. \qed
2378:
2379:
2380: \medskip
2381: Now we define the normal sheaf of a pseudoholomorphic curve passing through fixed
2382: points. The construction is completely analogous to that in the case of no fixed
2383: points. Here, instead of the tangent bundle $TS$ we use the bundle related
2384: the new situation. This is the bundle $TS_{-\mbfz}$ associated to the locally
2385: free coherent sheaf $\scro(TS)[-\mbfz]$ of local holomorphic sections of $TS$
2386: vanishing at the points $z_i$. One can prove the analog of \lemma{lem2.4.1}
2387: for $TS_{-\mbfz}$. Observe however, that such a result follows immediately
2388: from that lemma if we set $X=S$ and $u=\id_S$.
2389:
2390: As in \refsubsection{1.5}, we obtain the sheaf homomorphism $du: \scro(TS
2391: _{-\mbfz}) \to \scro(E_{u, -\mbfz})$, which is injective for non-constant $u:
2392: S \to X$. Now, the {\sl normal sheaf to curve $C=u(S)$ passing through the
2393: points $\mbfx=(x_1,\ldots, x_m)$} is defined as the quotient $\scrn_{u,\mbfx}
2394: \deff \scro(E_{u, -\mbfz}) / du\bigl(\scro(TS_{-\mbfz}) \bigr)$ together with
2395: the exact sequence
2396: \begin{equation}\eqqno(2.4.7)
2397: 0 \lrar \scro(TS_{-\mbfz}) \buildrel du \over \lrar \scro(E_{u, -\mbfz})
2398: \lrar \scrn_{u,\mbfx} \lrar 0.
2399: \end{equation}
2400: The sheaf $\scrn_{u,\mbfx}$ can be decomposed into its {\sl regular part} $\scrn
2401: \reg_{u,\mbfx}$ and its {\sl singular part} $\scrn\sing_{u,\mbfx}$, where
2402: $\scrn \reg_{u,\mbfx}$ is locally free and $\scrn\sing_{u, \mbfx}$ is a torsion
2403: sheaf. Then $\scrn \reg_{u,\mbfx}$ is a sheaf of local holomorphic sections of
2404: the {\sl normal bundle $N_{u,\mbfx}$ to curve $C=u(S)$ passing through the points
2405: $\mbfx=(x_1,\ldots, x_m)$}, so that $\scrn\reg_{u,\mbfx}= \scro(N_{u,\mbfx})$.
2406:
2407: As in \refsubsection{1.5}, we also obtain the exact sequence
2408: \begin{equation}\eqqno(2.4.8)
2409: 0 \lrar \scro(TS_{-\mbfz}) \otimes \scro([A])
2410: \buildrel du \over \lrar \scro(E_{u, -\mbfz})
2411: \scro(N_{u,\mbfx})
2412: \lrar 0.
2413: \end{equation}
2414: where $[A]$ is the branching divisor of $du$ (see \refdefi{def1.5.1}). This
2415: implies that the regular part $\scro(N_{u,\mbfx})$ is the quotient
2416: $$
2417: \scro(N_{u,\mbfx})=
2418: \scro(E_{u,-\mbfz}) / du\bigl(\scro(TS_{-\mbfz}) \otimes \scro([A])\bigr).
2419: $$
2420: From the definition of $E_{u,-\mbfz}$ and $TS_{-\mbfz}$ we obtain the isomorphism
2421: $$
2422: \scro(N_{u,\mbfx})\cong \scro(N_u) \otimes \scro([A]).
2423: $$
2424: On the other hand, the singular part remains the same as is the case without
2425: constraints:
2426: $$
2427: \scrn\sing_{u,\mbfx}\cong \scrn\sing_u \cong \scro /\scro(-[A]).
2428: $$
2429:
2430: Further, we observe that the operators $\dbar$ on $TS_{-\mbfz}$ and $D_{u,-\mbfz,
2431: J}$ in $E_{u,-\mbfz}$ commute with the homomorphism $du: TS_{-\mbfz} \to E_{u,
2432: -\mbfz}$. Consequently, $D_{u,-\mbfz, J}$ induces the operator
2433: $$
2434: D^N_{u,-\mbfz, J}: L^{1,p}(S, N_{u, -\mbfx}) \to L^p_{0,1)}(S, N_{u, -\mbfx})
2435: $$
2436: with the properties similar to ones of \eqqref(1.5.3). Further, as in
2437: \lemma{lem1.5.3} and \refcorol{cor1.5.4} we obtain a long exact sequence of
2438: $D$-cohomologies.
2439:
2440: \newprop{prop2.4.2} The~short exact sequence $\eqqref(2.4.7)$ induces
2441: the~long exact sequence of $D$-cohomologies
2442: \begin{equation*}
2443: \def\normalbaselines{\baselineskip20pt\lineskip3pt \lineskiplimit3pt }
2444: \def\mapright#1{\smash{\mathop{\longrightarrow}\limits^{#1}}}
2445: \def\mapdown{\Big\downarrow}
2446: %
2447: \matrix%\format\c&\c&\c&\c&\c&\c&\l&\c&\l\\
2448: 0& \mapright{}& \sfh^0(S, TS_{-\mbfz}) &\mapright{}
2449: & \sfh^0_D(S, E_{u,-\mbfz})
2450: & \mapright{}& \sfh^0_D(S, N_{u,-\mbfz})\oplus \sfh^0(S, \scrn_u\sing)
2451: &\mapright{\delta} &\vphantom{\mapdown}\\
2452: %
2453: & \mapright{}& \sfh^1(S, TS_{-\mbfz}) &\mapright{}
2454: & \sfh^1_D(S, E_{u,-\mbfz})
2455: & \mapright{}& \sfh^1_D(S, N_{u,-\mbfz}) &\mapright{} &0.
2456: \endmatrix
2457: \end{equation*}
2458: \end{prop}
2459:
2460: \medskip
2461: Finally, we note that the results of {\sl Paragraphs \ref{sec:2.2}\/} and
2462: {\sl\ref{sec:2.3}\/} remain valid, after an appropriate modification,
2463: also for curves passing through fixed points. We state without the proof
2464: the summary of results which will be used later.
2465:
2466: \newthm{thm2.4.3} \sli The total moduli space $\scrm_\mbfx$ of pseudoholomorphic
2467: curves in a given homology class $[C] \in \sfh_2(X, \zz)$ passing through fixed
2468: pairwise distinct points $\mbfx=(x_1, \ldots, \allowbreak
2469: x_m)$ on $X$ is a $C^\ell$-smooth
2470: Banach submanifold of $\scrm$ of real codimension $2m$. In particular, the
2471: projection $\pi_{\!\!\scrj}: \scrm_\mbfx \to \scrj$ is a $C^\ell$-smooth Fredholm
2472: map of index
2473: $$
2474: 2(c_1(X)[C] + (n-3)(1-g) -m).
2475: $$
2476:
2477: \slii For a generic $J \in \scrj$ and a generic $C^\ell$-smooth path $h:[0,1] \to
2478: \scrj$ the fiber
2479: $$
2480: \scrm_{J,\mbfx} \deff \pi_\scrj\inv(J)
2481: $$
2482: and the {\sl relative moduli space}
2483: $$
2484: \scrm_{h,\mbfx} \deff [0,1] \times_\scrj \scrm_\mbfx
2485: $$
2486: are $C^\ell$-smooth manifolds of expected dimension $2(c_1(X)[C] + (n-3)(1-g) -m)$
2487: and $2(c_1(X)[C] + (n-3)(1-g) -m) +1$ respectively.
2488: \end{thm}
2489:
2490:
2491:
2492: \medskip
2493: \newsection[4]{Cusp-curves in the moduli space.}
2494: In this section we study the problem of deformation of pseudoholomorphic curves
2495: with prescribed singularities and develop the techniques required for controlling
2496: their singularities under deformation. As the main result of this section we show
2497: that the locus of pseudoholomorphic curves with a prescribed type of singularity
2498: is a smooth
2499: Banach submanifold of expected codimension in the total moduli space of
2500: pseudoholomorphic curves. This improvement of the result of
2501: Micallef and White (see \lemma{lem1.2.1}) plays a crucial role below in
2502: \refsection{3} in the proof of the {\sl saddle point property}.
2503:
2504: Recall that our moduli space $\scrm$ consists of parameterized non-multiple
2505: pseudoholomorphic curves, \ie pseudoholomorphic maps from a fixed real surface
2506: $S$ modulo reparameterizations.
2507:
2508: \newdefi{def4.0.1}
2509: A point $z \in S$ on a $J$-holomorphic curve $u: S \to X$ is a {\sl cusp}, or a
2510: {\sl cuspidal point}, if $\ord_z du >0$.
2511: The number $\ord_z du$ is called the {\sl order of the cusp} of $u$ at $z$.
2512: A $J$-holomorphic curve $u: S \to X$ containing cuspidal points is called
2513: a {\sl cusp curve}.
2514: \end{defi}
2515:
2516: Note that in the literature on pseudoholomorphic curves the notion ``cusp curve''
2517: has a different meaning. Our terminology agrees rather with the one used in
2518: algebraic geometry where the notion ``cusp'' means a ``peak'', \ie an irreducible
2519: singularity. This describes the situation at hand more accurately.
2520: %In the author's opinion, such terminology is more correct.
2521:
2522:
2523: \newsubsection[4.1]{Deformation of pseudoholomorphic maps}
2524: Explicit construction of deformations is needed to obtain local charts for
2525: subspaces of curves with prescribed singularities.
2526:
2527: \newlemma{lem4.1.1} Let $B \subset \rr^{2n} \cong \cc^n$ be the unit ball,
2528: $\scry$ a Banach manifold, $\{J_{\eta,t}\}_{\eta\in \scry, t\in [0,1]}$ a
2529: family of homotopies of almost complex structures in $B$ with parameterized by $
2530: \scry$ and depending $C^{\ell-1}$-smoothly on $(\eta,t)\in \scry\times [0,1]$.
2531: Further, let $u_{\eta,0}: \Delta \to B$, $\eta\in \scry$, be a $C^{\ell-1}
2532: $-smooth family of $J_{\eta,0}$-holomorphic map, such that $u_{\eta,0}(\Delta)
2533: \subset B(\half)$, and $\ord_0(du_{\eta,t})= \mu$.
2534:
2535: Then for any family $v_\eta\in \rr^{2n}$ depending $C^{\ell-1}$-smoothly on
2536: $\eta \in \scry$ and any $\nu \in \nn$ there exists $t^*=t^*(J_t, u_0,v, \mu,
2537: \nu)>0$, a neighborhood $U_\scry$ of a given $\eta^*\in \scry$, and a
2538: $C^{\ell-1}$-smooth family of homotopies $\{ w_{\eta,t} \}_{\eta\in U_\scry,
2539: t\in [0, t^*]}$ with $w_{\eta,t}\in L^{1,p}(\Delta, \rr^{2n})$ such that the
2540: maps $u_{\eta,t}: \Delta \to B$ given by
2541: \begin{equation}
2542: u_{\eta,t}(z) = u_{\eta,0}(z) + z^\nu (t\,v_\eta + w_{\eta,t}(z))
2543: \eqqno(4.1.1)
2544: \end{equation}
2545:
2546: \noindent
2547: \sli are $J_{\eta,t}$-holomorphic if $\nu \le 2\mu +1$, and
2548:
2549: \noindent
2550: \slii are $J_{\eta,0}$-holomorphic if $\nu >2\mu +1$.
2551:
2552: Moreover, for $z\not =0$ the function $w_{\eta,t}(z)$ depends
2553: $C^{\ell-1}$-smoothly on $(\eta, t, z)$.
2554: \end{lem}
2555:
2556: \state Remarks.~1.
2557: In other words, there exists a pseudoholomorphic deformation $u_t$ of a given map
2558: $u_0$ in a given direction ${d \over dt}u_t\ogran_{t=0}=z^\nu v+ O(|z|^{\nu+
2559: \alpha})$; and moreover, for smaller $\nu$ it is possible to deform
2560: simultaneously the almost complex structure. Furthermore, if the initial data
2561: depend smoothly on the parameter $\eta$, then the corresponding constructions give
2562: a smooth dependence of the maps on $\eta$.
2563:
2564: \state 2. The loss of smoothness from $C^\ell$ to $C^{\ell-1}$ is due to the fact
2565: that the Gromov operator $D_{u,J}$ depends only $C^{\ell-1}$-smoothly
2566: on $u$. Indeed, $D_{u,J}$ is the derivative of the $\dbar$-operator $u \mapsto
2567: \dbar_J u$ in the $u$-direction, which is only $C^\ell$-smooth.
2568:
2569: \proof We give only a sketch. First, we fix a family $\phi_{\eta,t}$ of affine
2570: transformations of $\rr^{2n}$ with depend $C^{\ell-1}$-smoothly on $(\eta,t)$
2571: such that $\phi_{\eta,t} \scirc u_{\eta,0} (0) =0 \in B$ and $\phi_{\eta,t}\scirc
2572: J_{\eta,t} \scirc \phi_{\eta,t}\inv$ coincide with $J\st$ in $u_{\eta,0}(0)$.
2573: Setting $\ti u_{\eta,0} \deff \phi_{\eta,t} \scirc u_{\eta,0}$ and $\ti J_{\eta,t}
2574: \deff \phi_{\eta,t} \scirc J_{\eta,t} \scirc \phi_{\eta,t}\inv$ we reduce the
2575: problem to the case where $\ti u_{\eta,0}(0) = 0$ and $\ti J_{\eta,t}(0) = J\st$.
2576: %the standard structure in $\rr^{2n} \cong \cc^n$.
2577:
2578: Now we assume that there is no dependence on the parameter and drop the index
2579: $\eta$. Using \eqqref(4.1.1) one writes the equation $\dbar_{J_t} u_t =0$ in the
2580: form
2581: \begin{equation}
2582: (x + y J\st)^{-\nu} \dbar_{J_t}
2583: \bigl(u_0(z) + (x + y J\st)^\nu (t\,v + w_t(z)) \bigr)=0
2584: \eqqno(4.1.2)
2585: \end{equation}
2586: with $x +\isl y =z$ the standard coordinates on $\Delta$, and considers
2587: \eqqref(4.1.2) as an equation for $w_t(z)$. Then one shows that under the
2588: hypotheses of the lemma the linearization of \eqqref(4.1.2) has the form
2589: \begin{equation}
2590: (\dbar^{(\nu)}_{u_t, J_t} + R^{(\nu)}_{u_t, J_t}) \dot w_t(z) =
2591: \psi^{(\nu)}_{u_t, J_t}(\dot J_t)(z),
2592: \eqqno(4.1.3)
2593: \end{equation}
2594: where $\dot w_t(z)= {d \over dt}w_t(z)$ and $\psi^{(\nu)}_t(\dot J_t)(z) \in
2595: L^\infty(\Delta, \cc^n)$. Thus it is sufficient to find a right inverse
2596: $T^{(\nu)}_{u_t, J_t}$ of the Gromov type operator $D^{(\nu)}_{u_t, J_t} =
2597: \dbar^{(\nu)}_{u_t, J_t} + R^{(\nu)}_{u_t, J_t}$ with an additional condition
2598: $\dot w_t(0)=0$. We refer to \cite{Iv-Sh-1}, {\sl Lemma 3.3.1}, for the
2599: explicit construction of such a right inverse $T^{(\nu)}_{u, J}$. Moreover,
2600: the operator $T^{(\nu)}_{u, J}$ and the inhomogeneity term $\psi^{(\nu)}_{u,J}$
2601: depend smoothly on $u$ and $J$. As a consequence, the solution $w_t$ of
2602: \eqqref(4.1.2) depends $C^{\ell-1}$-smoothly on the parameter $\eta \in \scry$.
2603: \qed
2604:
2605:
2606: \newdefi{def4.1.1} Let $B \subset \rr^{2n} \cong \cc^n$ be a ball, $J_0$ a
2607: $C^\ell$-smooth almost complex structure in $B$, $u_0: \Delta \to B$ a
2608: $J_0$-holomorphic map and $\nu\ge 1$ an integer exponent. Denote
2609: by $\dfrm_\nu (u, J; v)$ the a map depending $C^{\ell-1}$-smoothly on
2610:
2611: \begin{itemize}
2612: \item
2613: a $C^\ell$-smooth almost complex structure $J$ in $B$, sufficiently close to $J_0$;
2614:
2615: \item a $J$-holomorphic map $u$, sufficiently close to $u_0$;
2616:
2617: \item a vector $v \in \rr^{2n}$, sufficiently close to $0$;
2618: \end{itemize}
2619:
2620: \noindent
2621: such that $\ti u\deff \dfrm_\nu (u, J; v)$ is a $J$-holomorphic map of the form
2622: $\ti u(z)= u(z) + z^\nu v + O(|z|^{\nu + \alpha})$. Note that the choice of such
2623: a map $\dfrm_\nu$ is not unique.
2624: \end{defi}
2625:
2626:
2627: \smallskip
2628: \lemma{lem4.1.1} allows us to construct local deformations of pseudoholomorphic
2629: maps with appropriate types of singularities. To obtain a global deformation,
2630: we use
2631:
2632: \newlemma{lem4.1.2} Let $u_0: S \to X$ be a non-multiple $J_0$-holomorphic
2633: map, $z_1, \ldots, z_m$ fixed points on $S$, and $U_1, \ldots, U_m \subset S$
2634: disjoint neighborhoods of these points. Further, let $\{J_t\}_{t\in [0,1]}$
2635: be a given $C^{\ell-1}$-smooth homotopy of almost complex structures on $X$, and
2636: $\{u_{i,t}\}_{t\in [0,1]}$ given $C^{\ell-1}$-smooth homotopies of
2637: $J_t$-holomorphic maps $u_{i,t}: U_i \to X$.
2638:
2639: Then there exist $t^*>0$, a $C^{\ell-1}$-smooth homotopy $\{\ti J_t\}_{t
2640: \in [0,t^*]}$ of almost complex structures on $X$, a $C^{\ell-1}$-smooth homotopy
2641: $\{\ti u_t\}_{t\in [0,t^*]}$ of $\ti J_t$-holomorphic maps $\ti u_t: S \to X$
2642: such that $u_t$ coincides with each $u_{i,t}$ in some (possibly smaller)
2643: neighborhood of $z_i$ and $\ti J_t$ coincides with $J_t$ in some
2644: neighborhood of each $x_i\deff u_0(z_i)$.
2645: \end{lem}
2646:
2647: The proof of the lemma is left to the reader.
2648:
2649: \smallskip
2650: Refining the result of \lemma{lem4.1.1} we show that the condition $u_1(z)-
2651: u_2(z)= o(|z|^k)$ of \lemma{lem1.2.4} defines a submanifold in the spaces of
2652: pairs of pseudoholomorphic maps.
2653:
2654:
2655: \newdefi{def4.2b.1} Define {\sl the spaces of pairs of pseudoholomorphic maps
2656: coinciding up to order $k$ at $z=0$} as $\scrpp_k(\Delta, X) \deff $
2657: \begin{equation}
2658: \bigl\{ (u', u'', J)\in
2659: L^{1,p}(\Delta, X) \times L^{1,p}(\Delta, X) \times \scrj:
2660: \dbar_Ju'=0=\dbar_Ju'', u'(z) - u''(z) = o(|z|^k) \bigr\},
2661: \eqqno(4.2b.1)
2662: \end{equation}
2663: where the condition $u'(z) - u''(z) = o(z^k)$ is related to any local
2664: coordinate system on $X$ in a neighborhood of the point $u'(0)= u''(0) \in X$.
2665: \end{defi}
2666:
2667: The structure of $\scrpp_m(\Delta, X)$ for the cases $k=0$ and $k=1$ is easily
2668: obtained from transversality techniques. In general we have
2669:
2670: \newthm{thm4.2b.1} Assume that $\scrj$ consists of $C^\ell$-smooth structures
2671: with $\ell\ge2$. Then the space $\scrpp_k(\Delta, X)$ is a $C^{\ell-1}
2672: $-submanifold of the fiber product $\scrp(\Delta, X) \times_{\!\!\scrj}
2673: \scrp(\Delta, X)$ of codimension of $2n(k+1)$ with the the tangent space
2674: \begin{multline}
2675: T_{(u',u'',J)}\scrpp_k(\Delta, X) =
2676: \\
2677: \bigl \{ (v',v'',\dot J) \in
2678: T_{(u',u'',J)}\bigl(\scrp(\Delta, X) \times_{\!\!\scrj} \scrp(\Delta, X)
2679: \bigr) :j^k(v' -v'') = 0 \bigr\}.
2680: \eqqno(4.2b.2)
2681: \end{multline}
2682:
2683: Moreover, for $k=0$ and $k=1$ the space $\scrpp_k(\Delta, X)$ is well-defined
2684: and $C^\ell$-smooth also for $\ell \ge1$.
2685: \end{thm}
2686:
2687: \proof It follows from \lemma{lem2.2.4} that $\scrp(\Delta, X)
2688: \times_{\!\!\scrj} \scrp(\Delta, X)$ is a $C^\ell$-smooth Banach manifold with
2689: the tangent space
2690: \begin{multline}
2691: T_{(u',u'',J)}\bigl(\scrp(\Delta, X) \times_{\!\!\scrj} \scrp(\Delta, X)
2692: \bigr)=
2693: \\
2694: \bigl\{ (v', v'', \dot J) : (v',\dot J) \in T_{(u',J)}\scrp(\Delta, X),
2695: (v'',\dot J) \in T_{(u'',J)}\scrp(\Delta, X) \bigr\},
2696: \eqqno(4.2b.3)
2697: \end{multline}
2698: so that $D_{u',J}v' + \dot J \scirc du' \scirc J_\Delta =0$ and similarly for
2699: $v''$.
2700:
2701: Fix $(u'_0, u''_0, J_0) \in \scrpp_0(\Delta, X)$ and local coordinates $(w_i)$ in
2702: a neighborhood $U \subset X$ of $x^* \deff u'_0(0)= u''_0(0) \in X$. Then there
2703: exists $r>0$ such that for any pair $(u', u'')$ of $L^{1,p}(\Delta, X)$-maps
2704: sufficiently close to $(u'_0, u''_0)$ we have $u'(\Delta(r)) \subset U$ and
2705: $u''(\Delta(r)) \subset U$. The coordinates in $U$ induce the linear structure.
2706: Thus we can consider the difference $u'(z)- u''(z)$ having in mind that it
2707: is well-defined only for $z\in \Delta(r)$.
2708:
2709: The subspace $\scrpp_0(\Delta, X)$ is defined by the condition $u'(0)=u''(0)$
2710: for $(u',u'',J) \in \scrp(\Delta, X) \times_{\!\!\scrj} \scrp(\Delta, X)$. Setting
2711: $F(u', u'', J) \deff u'(0) -u''(0)$ we obtain a $C^\ell$-smooth function, which
2712: is well-defined in a neighborhood of $(u'_0, u''_0, J_0)$ and is a
2713: local defining function for $\scrpp_0(\Delta, X)$. The differential of $F$ in
2714: $(u',u'',J) \in \scrpp_0(\Delta, X)$,
2715: $$
2716: dF: T_{(u',u'',J)}\bigl(\scrp(\Delta, X) \times_{\!\!\scrj} \scrp(\Delta, X)
2717: \bigr) \to T_{u'(0)} X,
2718: $$
2719: is given by the formula $dF(v',v'', \dot J) = v'(0)-v''(0)$ and is
2720: a surjective map. Thus $\scrpp_0(\Delta, X)$ is a $C^\ell$-smooth submanifold
2721: of $\scrp(\Delta, X) \times_{\!\!\scrj} \scrp(\Delta, X)$ of codimension $2n =
2722: \dimr X$.
2723:
2724: \smallskip
2725: Considering the $C^\ell$-smooth map $\ev_0: \scrpp_0(\Delta, X) \to X$ with
2726: $$\ev_0(u',u'', J) \deff u'(0) = u''(0),$$ we obtain a $C^\ell$-smooth bundle
2727: $E^{(0)}$ over $\scrpp_0(\Delta, X)$ with fiber $E^{(0)}_{(u',u'', J)}= {u'}^*
2728: T_{u'(0)}X$. The formulas $\sigma'(u',u'', J) \deff du'(0)$ and $\sigma''(u',u'',
2729: J) \deff du''(0)$ define $C^\ell$-smooth sections of the bundle $T^*_0\Delta
2730: \otimes E^{(0)}$ over $\scrpp_0(\Delta, X)$. Thus the condition $du'(0)= du''(0)$
2731: is equivalent to the vanishing of $\sigma' -\sigma''$. Consequently, $\scrpp_1
2732: (\Delta, X)$ is a $C^\ell$-smooth submanifold of $\scrpp_0(\Delta, X)$ of
2733: codimension $2n$.
2734:
2735: \smallskip
2736: We proceed further by induction using the case $k=0$ as the base. Our notation is
2737: as follows. For a triple $(u',u'', J)\in \scrpp_0(\Delta,X)$ we
2738: consider the (integrable) complex structure $J\st$ in $U$ with coincides with
2739: $J$ at the point $u'(z)= u''(z)$ and is constant \wrt the coordinates in $U$.
2740: Note that $J\st$ depends $C^\ell$-smoothly on $(u',u'', J) \in \scrpp_k(\Delta,
2741: X)$. Thus we can regard $U$ as an open subset in $\cc^n$.
2742:
2743: For a pair $(u',u'')$ of $J$-holomorphic maps with values in $U \subset X$ we
2744: obtain
2745: \begin{align}
2746: 0& = \dbar_J u' -\dbar_J u'' = \bigl( (\d_x u' - J(u')\cdot \d_y u') -
2747: (\d_x u'' - J(u'')\cdot \d_y u'')
2748: \notag\\
2749: &= \d_x(u'-u'') + J(u')\cdot \d_y(u' -u'') +
2750: \bigl(J(u') -J(u'')\bigr)\cdot \d_y u''
2751: \notag\\
2752: &= \dbar_{J(u')}(u' -u'') +\bigl(J(u') -J(u'')\bigr)\cdot \d_y u'' .
2753: \eqqno(4.2b.4)
2754: \end{align}
2755: Consequently,
2756: \begin{align}
2757: \llap{$\dbar_{J\st}$}(u'-u'')&
2758: = \dbar_{J\st}(u'-u'') - (\dbar_J u' -\dbar_J u'')
2759: \notag\\
2760: &=\bigl(J\st - J(u')\bigr)\cdot \d_y (u'-u'')-
2761: \bigl(J(u') -J(u'')\bigr)\cdot \d_y u'' .
2762: \eqqno(4.2b.5)
2763: \end{align}
2764: Let us denote the last expression by $H_{u',u'',J}(z)$
2765:
2766: \smallskip
2767: Now suppose that $(u',u'', J)$ varies in $\scrpp_k(\Delta,X)$ with $k\ge1$.
2768: We can assume by induction that $\scrpp_k(\Delta,X)$ is a $C^{\ell-1}$-smooth
2769: manifold. We claim that for any $p<\infty$
2770: \begin{equation}
2771: \eqqno(4.2b.7)
2772: \llap{$f_k(z) \deff $}z^{-(k+1)}(u'(z) - u''(z))
2773: \end{equation}
2774: is a well-defined $L^{1,p}(\Delta(r), \cc^n)$-valued function depending $C^{\ell-
2775: 1}$-smoothly on $(u',u'',J) \in \scrpp_k(\Delta,X)$. The claim implies the
2776: theorem. Indeed, the function $F_k$ given by $F_k: (u',u'', J) \in \scrpp_k(
2777: \Delta, X) \mapsto f_k(0) \in \cc^n$ is then a local defining function for
2778: $\scrpp_{k-1}(\Delta,X)$ inside $\scrpp_k(\Delta,X)$, whereas non-degeneracy
2779: of $dF_k$ can be easily obtained from \lemma{lem4.1.1}.
2780:
2781: Again by induction, we can suppose that $f_{k-1}(z)= z^{-k}(u'(z) - u''(z))$ is a
2782: well-defined $L^{1,p}(\Delta(r), \cc^n)$-valued function depending $C^{\ell-1}
2783: $-smoothly on $(u',u'',J) \in \scrpp_{k-1}(\Delta,X)$. Note that $f_{k-1}(0)$
2784: vanishes identically on $\scrpp_k(\Delta,X)$. Further, for any exponents $p, p'$
2785: with $2<p'<p <\infty$ the map $f(z)\in L^{1,p}(\Delta, \cc^n) \mapsto z\inv (f(z)
2786: -f(0)) \in L^{p'}(\Delta, \cc^n)$ is linear and bounded. Consequently, for any
2787: $p<\infty$ the function $f_k(z) = z\inv f_{k-1}(z)$ lies in $L^p(\Delta, \cc^n)$
2788: and depends $C^{\ell-1}$-smoothly on $(u',u'',J) \in \scrpp_k(\Delta,X)$ \wrt
2789: the $L^p$-topology.
2790:
2791: \smallskip
2792: Without loss of generality we may assume that $U$ is convex. The identity
2793: $$
2794: J(w) = J(w^*) + \int_{t=0}^1 \d_t J(w^* +t(w-w^*)) dt
2795: $$
2796: for $(w,w^*) \in U \times U$ implies the relation $J(w)= J(w^*) +\sum_i (w_i-
2797: w^*_i) S_i(w,w^*) = S(w,w^*; \allowbreak w-w^*)$ with the function $S(w,w^*;
2798: \ti w)$ depending $C^\ell$-smoothly on $J \in \scrj$, $C^{\ell-1}$-smoothly on
2799: $(w,w^*) \in U\times U$ and $\rr$-linearly on $\ti w\in \cc^n$. Substituting
2800: $u''(z) = u'(z) + z^{k+1} f_k(z)$ in $(J(u') -J(u''))\cdot \d_y u''$ we obtain
2801: $$
2802: (J(u'(z)) -J(u''(z)))\cdot \d_y u''(z) = S\bigl(u''(z), u'(z); z^{k+1}f_k(z)
2803: \bigr)\cdot \d_y u''(z)
2804: $$
2805: By apriori regularity estimates, for $r<1$ we can consider $du''(z)$ as a map from
2806: $\scrpp_{\!k}(\Delta,X)$ to $C^0(\Delta(r),\cc^n)$ which depends $C^\ell$-smoothly
2807: on $(u',u'',J)$. Thus we have represented the term $(J(u') -J(u''))\cdot \d_y u''$
2808: as a composition of the $C^{\ell-1}$-smooth map
2809: $$
2810: (u',u'',J)\in \scrpp_k(\Delta,X) \;\mapsto\; S(u''(z),u'(z); f_k(z)) \cdot
2811: \d_y u''(z) \in L^p(\Delta(r), \cc^n)
2812: $$
2813: and the linear bounded map
2814: $$
2815: S(u''(z),u'(z); f_k(z)) \d_y u''(z)\;\mapsto\; z^{-(k+1)} \cdot
2816: S(u''(z),u'(z); z^{k+1} \cdot f_k(z)) \d_y u''(z).
2817: $$
2818: Thus $(J(u') -J(u''))\cdot \d_y u''$ depends $C^\ell$-smoothly on $(u',u'',J)\in
2819: \scrpp_k(\Delta,X)$ \wrt the norm topology in $L^p(\Delta(r), \cc^n)$.
2820: Consequently, the formula
2821: $$
2822: (u',u'', J) \in \scrpp_k(\Delta,X) \;\mapsto \;
2823: z^{-(k+1)} \cdot (J\st - J(u'(z))) \cdot \d_y (u'(z)-
2824: u''(z))
2825: $$
2826: defines a $L^p(\Delta(r), \cc^n)$-valued map depending $C^{\ell-1}
2827: $-smoothly on $(u',u'', J) \in \scrpp_k(\Delta,X)$.
2828:
2829: Similar estimates can be be carried out for the first term $(J\st - J(u'))\cdot
2830: \d_y(u'-u'')$ in \eqqref(4.2b.5). Together, this implies that $h_k(z) \deff z^{-k}
2831: H_{u',u'',J}(z)$ lies in $L^p(\Delta(r),\cc^n)$ and depends $C^{\ell-1}$-smoothly
2832: on $(u',u'', J) \in \scrpp_k(\Delta,X)$ \wrt $L^p$-topology. Now let $f_{\dbar,
2833: k}(z)$ be a solution of the equation $\dbar_{J\st}f_{\dbar, k}(z) = h_k(z)$
2834: depending $C^{\ell-1}$-smoothly on $(u',u'', J) \in \scrpp_k(\Delta,X)$ \wrt
2835: the $L^{1,p}$-topology. Then $(u'(z) -u''(z)) - z^{k+1} f_{\dbar, k}(z)$ is
2836: a holomorphic $\cc^n$-valued function, depending $C^{\ell-1} $-smoothly on
2837: $(u',u'', J) \in \scrpp_k(\Delta,X)$ \wrt $L^{1,p}$-topology and vanishing in
2838: $z=0$ up to order $k+1$. Consequently,
2839: $$
2840: (u'(z) -u''(z)) - z^{k+1} f_{\dbar, k}(z)= z^{k+1}f_{\scro, k}(z)
2841: $$
2842: and $f_k(z) = f_{\scro, k}(z) + f_{\dbar, k}(z)$ possesses the property claimed
2843: above. \qed
2844:
2845:
2846: \smallskip
2847:
2848: \newsubsection[4.2a]{Curves with prescribed cusp order}
2849: In this paragraph we show that $J$-curves with cusps of given order form a
2850: Banach submanifold of the moduli space and compute its codimension.
2851:
2852:
2853: %\nobreak
2854: \newdefi{def4.2.1} For a given natural $m$ we denote by $\mbfk$ an $m$-tuple
2855: $(k_1, \ldots, \allowbreak
2856: k_m)$ with $k_i \ge1$ and set $|\mbfk| \deff \sum_i
2857: k_i$. The $m$-tuple $(1,\ldots,1)$ is denoted $\bfone_m$.
2858: Define the {\sl moduli space $\scrm_\mbfk$ of pseudoholomorphic curves
2859: with a given cusp order $\mbfk$} as the set of classes $[u,J, \mbfz]$ such
2860: that $[u,J] \in \scrm$ and $u$ has $m$ (marked) cusp-points $\mbfz= \{z^*_1,
2861: \ldots, z^*_m \}$ with $\ord_{z^*_i} \ge k_i$. Two triples $(u,J, \mbfz)$
2862: and $(\ti u, \ti J, \ti\mbfz)$ define the same class $[u,J, \mbfz] = [\ti u,
2863: \ti J, \ti\mbfz] \in \scrm_\mbfk$ \iff there exists $g \in \bfg$ such that
2864: $\ti u = u \scirc g$ and $\ti z^*_i = g(z^*_i)$.
2865: \end{defi}
2866:
2867:
2868: \smallskip
2869: The main result of this paragraph is
2870:
2871: \newthm{thm4.2.1} The set $\scrm_\mbfk$ is a $C^\ell$-smooth
2872: manifold and the natural map $\scrm_\mbfk \lrar \scrm$ given by $[u,J, \mbfz]
2873: \mapsto [u,J]$ of $\scrm$ is an immersion of codimension $2\, (n\, |\mbfk|
2874: -m)$, where $n= \dimc X$ and $m$ is the number of marked cusp-points.
2875: \end{thm}
2876:
2877: \smallskip
2878: We divide the proof in several steps. First we consider the corresponding
2879: problem for $\whcalm$. The reason is that it is more convenient to work
2880: with maps, \ie elements of
2881: $\whcalm$, than with parameterized curves, \ie elements of $\scrm$. This
2882: means that we are interested in the set
2883: \begin{equation}
2884: \whcalm_\mbfk \deff \left\{ (u,J_S, J; z^*_1, \ldots z^*_m) \in
2885: \whcalm \times (S)^m :
2886: \msmall{ \matrix %\format\c\\
2887: z^*_i\text{ are pairwise distinct, }\cr
2888: \ord_{z^*_i} du \ge k_i% \text{\rmnine has a cusp in $z^*_i$ of order}\;
2889: \endmatrix}
2890: \right\},
2891: \eqqno(4.2.1)
2892: \end{equation}
2893: where $(S)^m= S \times \cdots \times S$ is the $m$-fold product of $S$.
2894: Obviously, the projection from $\whcalm \times (S)^m$ onto $\whcalm$ and then
2895: onto $\scrm$ maps $\whcalm_\mbfk$ onto $\scrm_\mbfk$. In our proof of
2896: \refthm{thm4.2.1} we shall show that this map $\whcalm_\mbfk \to
2897: \scrm_\mbfk$ is a principle $\bfg$-bundle.
2898:
2899: \newdefi{def4.2.1n} Set
2900: \begin{equation}
2901: \whcalm^{(m)} \deff \{ (u,J_S, J; z^*_1, \ldots z^*_m) \in
2902: \whcalm \times (S)^m : z^*_i \not = z^*_j \text{ for every } i\not=j\,\}
2903: \eqqno(4.2.2n)
2904: \end{equation}
2905: denoting by $S_i$ the $i$-th factor in $(S)^m$. Equip $\whcalm^{(m)}$ with the
2906: maps $\ev_i: \whcalm^{(m)} \to X^m$ defined by $\ev_i(u,J_S, J; z^*_1, \ldots,
2907: z^*_m) \deff u(z^*_i)$. Denote by $E_i$ the pulled-back bundles $\ev_i^*TX$ and
2908: $\ev^{(m)}{}^* T(X^m)$ over $\whcalm^{(m)}$. The fiber of $E_i$ over $(u, J_S,
2909: J; \mbfz)$ is $(E_i)_{(u, J_S, J; \mbfz)}= T_{u(z^*_i)}X$.
2910: \end{defi}
2911:
2912: Obviously, the space $\whcalm^{(m)}$ is a $C^\ell$-smooth Banach manifold,
2913: $\ev_i: \whcalm^{(m)} \to X^m$ are $C^\ell$-smooth maps, and $E_i$ are
2914: $C^\ell$-smooth bundles over $\whcalm^{(m)}$. Note that we also have line
2915: bundles $TS_i$ and $T^*S_i$ over $\whcalm^{(m)}$ which are defined in an obvious
2916: way as the (co)tangent bundles to each $S_i$.
2917:
2918: \newlemma{lem4.2.2n} The formula $\yps(u,J_S, J; z^*_1, \ldots z^*_m) \deff
2919: (du(z^*_1), \ldots, du(z^*_m)) \in \bigoplus_i T^*S_i \otimes E_i$ defines
2920: a $C^\ell$-smooth section of\/ $\bigoplus_i T^*S_i \otimes E_i$ over
2921: $\whcalm^{(m)}$, transversal to the zero section. The zero-set of $\yps$
2922: coincides with the space $\whcalm_{\bfone_m}$ of maps having cups in each
2923: marked $z^*_i$. Thus $\whcalm_{\bfone_m}$ is a $C^\ell$-smooth Banach
2924: submanifold of $\whcalm^{(m)}$ of codimension $2nm$.
2925: \end{lem}
2926:
2927: Before starting the proof we introduce some new notation.
2928:
2929: \newdefi{def4.2.2n} Let $\scry$ be a $C^\ell$-smooth Banach manifold and $f:
2930: \scry \to \ttt_g \times S$ a $C^\ell$-smooth map of the form $f(y)= (J_S(y),
2931: z^*(y))$. Set $F(y) \deff (y, z^*(y))$ so that $F: \scry \to \scry \times S$ is
2932: an imbedding. A {\sl local $J_S(y)$-holomorphic coordinate (or simply
2933: a {\sl$J_S$-holomorphic coordinate}) on $\scry \times S$ centered at $z^*$}
2934: is a $C^\ell$-smooth $\cc$-valued function $z$ defined in some neighborhood $U
2935: \subset \scry \times S$ of $F(\scry)$ which vanishes along $F(\scry)$ and is
2936: $J_S(y)$-holomorphic along each $\{y\} \times S$. One can use \lemma{lem4.1.1}
2937: for a proof of the existence of such a local holomorphic coordinate.
2938: \end{defi}
2939:
2940: \statep Proof of. \lemma{lem4.2.2n}. It is obvious that $\yps$ is
2941: well-defined. To show the $C^\ell$-smoothness of $\yps$, for any $i=1,\ldots,m$,
2942: we fix some local coordinate $z_i$ on $\whcalm^{(m)}$ which is
2943: $J_S$-holomorphic along $S_i$ and centered at $z^*_i \in S_i$. Now we can
2944: find a local frame $\bfxi=(\xi_1, \ldots, \xi_n)$ of $T^*S_i \otimes E_i$ which
2945: depends $C^\ell$-smoothly on $(u,J_S, J) \in \whcalm$ and
2946: holomorphically on the coordinate $z_i$. The existence of such a frame follows
2947: from \refdefi{def1.4.1} and a parametric version of \lemma{lem1.4.1A}.
2948: The coefficients of $du \in T^*S_i \otimes E_i$ \wrt such a frame
2949: $\bfxi$ depend $C^\ell$-smoothly on $(u,J_S, J) \in \whcalm$ and holomorphically
2950: on $z_i$. Consequently, the $du(z^*_i)$ depend $C^\ell$-smoothly on $(u,J_S, J;
2951: \mbfz) \in \whcalm^{(m)}$. Thus $\yps$ is $C^\ell$-smooth.
2952:
2953: The transversality of $\yps$ to the zero-section of $\bigoplus_i T^*S_i \otimes
2954: E_i$ follows immediately from results of \refsubsection{4.1}. In particular,
2955: $\whcalm^{(m)}_{\bfone_m}$ is the $C^\ell$-smooth Banach submanifold of
2956: $\whcalm^{(m)}$. The corresponding codimension is $\rank_\rr\left(
2957: \bigoplus_i T^*S_i \otimes E_i\right) =2nm$. \qed
2958:
2959:
2960: \newdefi{def4.2.3n} For (finite-dimensional) complex vector spaces $V$, $W$,
2961: and $k\in \nn$ denote by $j^k(V,W)$ the vector space of polynomial maps
2962: $f:V \to W$ of degree $\deg f\le k$ with $f(0)=0$, considered as the space of
2963: $k$-jets of holomorphic maps $F: V \to W$. For $l\ge k$ the natural projection
2964: $\pr: j^l(V,W) \to j^k(V,W)$ is well-defined. Let $j^{k,l}(V,W)$ denote
2965: its kernel. Similar notation for complex bundles is used. Note that
2966: $j^1(V,W) = \hom(V,W) = V^* \otimes W$.
2967: \end{defi}
2968:
2969: \newlemma{lem4.2.3n} \sli For any $(u, J_S, J; \mbfz) \in \whcalm_\mbfk$ the
2970: jet $j^{2k_i+1}u(z^*_i)$ is a well-defined element of $j^{2k_i+1} (T_{z^*_i},
2971: T_{u(z^*_i)}X) = j^{2k_i+1} (TS_i, E_i)_{(u, J_S, J; \mbfz)}$.
2972:
2973: \slii Moreover, $j^{2k_i+1}u(z^*_i) \in j^{k_i+1, 2k_i+1} (T_{z^*_i},
2974: T_{u(z^*_i)}X)= j^{k_i+1, 2k_i+1} (TS_i, E_i)_{(u, J_S, J; \mbfz)}$.
2975:
2976: \sliii Set $\yps_\mbfk(u, J_S, J; \mbfz) \deff \bigl(j^{k_1+1, 2k_1+1} u(z^*_1),
2977: \ldots, j^{k_m+1, 2k_m+1}u(z^*_m) \bigr)$. Then $\yps_\mbfk: \whcalm_\mbfk \to
2978: \bigoplus_i j^{k_i+1, 2k_i+1} (TS_i, E_i)$ is a section which is $C^\ell$-smooth
2979: and transversal to the zero-section.
2980: \end{lem}
2981:
2982:
2983: \proof Assertions \sli and \slii follow essentially from \lemma{lem1.2.4}.
2984: The nontrivial points here are the following. First, the jet
2985: $j^{2k_i+1}u(z^*_i)$ is defined even if the structure $J$ is $C^\ell$-smooth
2986: with $\ell<2k_i$ and the map $u$ is $C^{\ell+1}$-smooth, since in general there
2987: are no higher smoothness for $u$. Second, the jet $j^{2k_i+1}
2988: u(z^*_i)$ is a {\sl complex} polynomial. Finally, the jet $j^{2k_i+1}u(z^*_i)$
2989: is independent of the choice of the integrable structure $J\st$ and
2990: $J\st$-holomorphic coordinates in a neighborhood of $u(z^*_i)$ used in
2991: \lemma{lem1.2.4} for definition of the jet. Let us give a proof of the latter
2992: property.
2993:
2994: Let $J'$ and $J''$ be integrable complex structures in a neighborhood of
2995: $u(z^*_i)$ such that $J'(u(z^*_i)) = J''(u(z^*_i)) = J(u(z^*_i))$. Find local
2996: complex coordinate systems $\mbfw'= (w'_1, \ldots, \allowbreak
2997: w'_n)$ and $\mbfw''= (w''_1, \ldots,w''_n)$ which are centered in $u(z^*_i)$ and
2998: holomorphic \wrt $J'$ and
2999: $J''$ respectively. Without loss of generality we may assume that the frames
3000: $({\d \over \d w'_1}, \ldots, {\d \over \d w'_n})$ and $({\d \over \d w''_1},
3001: \ldots, {\d \over \d w''_n})$ coincide in $u(z^*_i)\in X$. Consequently,
3002: we can express one system by another using the formula $\mbfw''= \mbfw' +
3003: F(\mbfw')$ with
3004: \begin{equation}
3005: F(\mbfw') = O(|\mbfw'|^2) \qquad \text{and} \qquad
3006: dF(\mbfw') = O(|\mbfw'|).
3007: \eqqno(4.2.3n)
3008: \end{equation}
3009: Let $u'(z)$ and $u'(z)$ be the local expressions of $u: S \to X$ in the local
3010: coordinate systems $\mbfw'$ and $\mbfw''$ respectively. Then $u''(z)= u'(z) +
3011: F(u'(z))$. So from \eqqref(4.2.3n) and $u'(z)= O(|z|^{k_i+2})$ we see that
3012: coefficients of polynomials $j^{2k_i+1}u'(z)$ and $j^{2k_i+1}u''(z)$ coincide.
3013:
3014: \smallskip
3015: To show the smoothness of the section $\yps_\mbfk$ we fix an element
3016: $(u_0,J_{S,0}, J_0; \mbfz_0)\in \whcalm_\mbfk$, $\mbfz_0=(z^*_{1,0}, \ldots,
3017: z^*_{m,0})$, and a sufficiently small neighborhood $\scry \subset \whcalm
3018: _\mbfk$ of $y_0 \deff (u_0, J_{S,0}, J_0; \mbfz_0)$.
3019: In what follows, for any $i=1, \ldots,m$, we fix families of certain structures
3020: on various spaces. We assume that the members of the families are parameterized
3021: by and depend $C^\ell$-smoothly on $y=(u,J_S, J; \mbfz)\in \scry$. The
3022: families are:
3023: \begin{enumerate}
3024: \item integrable complex structures $J'_i$ in a neighborhood
3025: of each $u(z^*_{i,0})$ such that each $J'_i$ coincides with $J$ in $u(z^*_i)$;
3026: \item local complex coordinate systems $\mbfw'_i= (w'_{i,1}, \ldots,
3027: w'_{i,n})$ on $X$ centered in $u(z^*_i)$ and holomorphic \wrt $J'_i$;
3028: \item local frames $\bfxi_i= (\xi_{i,1},\ldots ,\xi_{i,n})$ of the bundles
3029: $E_i$ which are defined in a neighborhood of $z^*_i$ and holomorphic along
3030: $S_i$;
3031: \item local $J_S$-holomorphic coordinates $z_i$ on $S_i$ centered in
3032: $z^*_i$.
3033: \end{enumerate}
3034: Further, we assume that every coordinate $z_i$ has image the whole disc $\Delta$.
3035: Note that pulling back the frames $({\d \over \d w'_{i,1}}, \ldots, {\d \over
3036: \d w'_{i,n}} )$ we obtain local frames $\left(u^*({\d \over \d w'_{i,1}}), \ldots,
3037: u^*({\d \over \d w'_{i,n}}) \right)$ of $E_i$ which depend $C^\ell$-smoothly
3038: on $y=(u,J_S, J; \mbfz)\in \scry$. Now, the expression of $u(z)$
3039: in the local coordinate system $\mbfw'_i$ yields an element $u'_i(z_i) \in
3040: L^{1,p}(\Delta, \cc^n)$ which depends $C^\ell$-smoothly on $y\in \scry$
3041: \wrt the standard smooth structure in $L^{1,p}(\Delta, \cc^n)$. Deriving, we
3042: obtain an element $du'_i(z_i)\in L^p(\Delta, \cc^n\otimes_\rr T^*\Delta)$ which
3043: depends $C^\ell$-smoothly on $y\in \scry$ \wrt the standard smooth structure
3044: in $L^p(\Delta, \cc^n\otimes_\rr T^*\Delta)$.
3045:
3046: Consider now $du'_i$ as a
3047: section of $E_i\otimes T^*S_i$, and its coefficients of $du'_i$ in the frame
3048: $\left(u^*({\d \over \d w'_{i,1}}) \otimes dz_i, \ldots, u^*({\d \over
3049: \d w'_{i,n}}) \otimes dz_i \right)$ as $L^p(\Delta, \cc)$-functions. Thus we
3050: can conclude that the coefficients of $du'_i$ depend $C^\ell$-smoothly on
3051: $y\in \scry$ \wrt the standard smooth structure in $L^p(\Delta, \cc)$.
3052: Consequently, the same is true for the coefficients of $du'_i$ in the frame
3053: $(\xi_{i,1}\otimes dz_i,\ldots, \xi_{i,n}\otimes dz_i)$. Since the latter frame
3054: is holomorphic, the coefficients of the jet $j^{2k_i}du(z^*_i)$ depend
3055: $C^\ell$-smoothly on $y\in \scry$. This provides the desired smoothness
3056: property of $\yps_\mbfk$.
3057:
3058: \smallskip
3059: Finally, note that the transversality of $\yps_\mbfk$ to the zero-section
3060: follows from results of \refsubsection{4.1}. \qed
3061:
3062:
3063: \newcorol{cor4.2.4n} For any $\mbfk=(k_1,\ldots, k_m)$ with $k_i \ge 1$
3064: the space $\whcalm_\mbfk$ is a $C^\ell$-submanifold of $\whcalm^{(m)}$ of
3065: codimension $2|\mbfk|n$.
3066: \end{corol}
3067:
3068: \proof Assume that for a given $\mbfk=(k_1,\ldots, k_m)$ with
3069: $k_i \ge 1$ the claim holds. Fix some $\mbfk^+=(k^+_1,\ldots, k^+_m)$ with
3070: $k_i \le k^+_i \le 2 k_i$ and consider truncated section $\yps_{\mbfk,\mbfk^+}
3071: : \whcalm_\mbfk \to \bigoplus_i j^{k_i+1, k^+_i+1} (TS_i, E_i)$ given
3072: by
3073: $$
3074: \yps_{\mbfk,\mbfk^+}(u, J_S, J; \mbfz) \deff \bigl(j^{k_1+1, k^+_1+1}
3075: u(z^*_1), \ldots, j^{k_m+1, k^+_m+1}u(z^*_m) \bigr).
3076: $$
3077: Then $\whcalm_{\mbfk^+}$ is identified with the zero set of $\yps_{\mbfk,
3078: \mbfk^+}$. By \lemma{lem4.2.3n}, $\yps_{\mbfk,\mbfk^+}$ is transversal to
3079: the zero-section. Thus $\whcalm_{\mbfk^+}$ is a $C^\ell$-smooth submanifold
3080: of $\whcalm_\mbfk$ of codimension equal to $\rank_\rr \bigoplus_i j^{k_i+1,
3081: k^+_i+1} (TS_i, E_i) = 2n(|\mbfk^+| -|\mbfk|)$. So we can apply the induction.
3082: \qed
3083:
3084: \smallskip
3085: \newlemma{lem4.2.5n} The natural projection $\wh\pr_\mbfk : \whcalm_\mbfk
3086: \to \whcalm$ given by the formula $\wh\pr_\mbfk(u, J_S, J; \mbfz)
3087: \allowbreak
3088: \deff (u, J_S, J)$ is an immersion of codimension $2(|\mbfk|n -m)$.
3089: \end{lem}
3090:
3091: \proof The differential of the projection $\wh\pr_\mbfk$ is given by
3092: $$
3093: d\wh\pr_\mbfk: (v, \dot J_S, \dot J; \dot \mbfz)\in
3094: T_{(u, J_S, J; \mbfz)}\whcalm_\mbfk \mapsto
3095: (v, \dot J_S, \dot J) \in T_{(u, J_S, J)}\whcalm.
3096: $$
3097: Thus the kernel $\ker d\wh\pr_\mbfk$ consists of vectors of the form $(0,0,0;
3098: \dot \mbfz)$ with $\dot \mbfz_i = (\dot z^*_1, \ldots, \dot z^*_m) \in
3099: \bigoplus T_{z^*_i}S_i$ and we must show that $\ker d\wh\pr_\mbfk$ is trivial.
3100: Intuitively this is obvious, since elements of the kernel correspond to
3101: deformations leaving $(u,J_S,J)$ unchanged but moving cusp-points $z^*_i$ on
3102: $S$ and this is impossible.
3103:
3104: For a rigorous proof we use conclusions of the proof of \lemma{lem4.2.3n}.
3105: Consider $du(z_i)$ as a holomorphic section of $T^*S_i
3106: \otimes E_i$. Then $du(z_i)$ vanishes in $z^*_{i,t}$ up to the order $\ge k_i$
3107: and there are no other zeros of $du(z_i)$ in a neighborhood of $z^*_i$. Thus
3108: we can locally express $z^*_i$ as the zero set of $du(z_i)$. This implies
3109: that locally there exists $C^\ell$-smooth functions $F_i$ of $(u,J_S,J)\in
3110: \whcalm$ such that $F_i(u,J_S,J) =z^*_i$ for $(u,J_S,J)\in \whcalm_\mbfk$.
3111: Thus $\wh\pr_\mbfk : \whcalm_\mbfk \to \whcalm$ is an immersion.
3112:
3113: To compute the codimension of $\wh\pr_\mbfk: \whcalm_\mbfk \hook \whcalm$
3114: one represents $\wh\pr_\mbfk$ as the composition $ \whcalm_\mbfk \hook
3115: \whcalm^{(m)} \buildrel \pr \over \lrar \whcalm$.
3116: \qed
3117:
3118:
3119:
3120: \medskip
3121: Now we can finish
3122:
3123: \nobreak
3124: \statep Proof of. \refthm{thm4.2.1}. Consider the action of $\bfg$ on
3125: $\whcalm$ and the diagonal action of $\bfg$ on $\whcalm \times (S)^m$. The
3126: both actions are $C^\ell$-smooth, free, and commute with the projection $\pr:
3127: \whcalm \times (S)^m \to \whcalm$. Moreover, for every $\mbfk=(k_1,\ldots,
3128: k_m)$ with $k_i\ge 1$ the submanifold $\whcalm_\mbfk \hook \whcalm \times
3129: (S)^m$ is $\bfg$-invariant \wrt the diagonal action of $\bfg$.
3130: For the quotient $\scrm_\mbfk= \whcalm_\mbfk /
3131: \bfg$ one can construct a $C^\ell$-smooth atlas in the same way as it was
3132: done for $\scrm= \whcalm /\bfg$. The construction shows that the map
3133: $\scrm_\mbfk \to \scrm$ is a $C^\ell$-smooth immersion of codimension
3134: equal to the codimension of $\wh\pr_\mbfk: \whcalm_\mbfk \hook \whcalm$.
3135: \qed
3136:
3137: \medskip
3138: Summarizing the results and notation of this paragraph, we obtain
3139:
3140:
3141: \newcorol{cor4.2.6n} The maps $\ev_\mbfk : \scrm_\mbfk \to X^m$ and $\ev_i:
3142: \scrm_\mbfk \to X$ given by $\ev_\mbfk([u,J, \mbfz]) \deff (u(z^*_1), \ldots ,
3143: u(z^*_m))$ and $\ev_i([u,J, \mbfz]) \deff u(z^*_i)$ are well-defined and
3144: $C^\ell$-smooth. This yields $C^\ell$-smooth bundles $E_i \deff \ev_i^* TX$
3145: with a fiber $(E_i)_{[u,J, \mbfz]} = T_{u(z^*_i)}X$. The bundles $T_{z^*_i}
3146: S_i$ over $\whcalm_\mbfk$ induce $C^\ell$-smooth bundles $L_i$ over $\scrm
3147: _\mbfk$ with the fiber $(L_i) _{[u,J, \mbfz]}= T_{z^*_i} S_i$.
3148: The section $\yps_\mbfk: \whcalm_\mbfk \to \bigoplus_i j^{2k_i+1}(TS_i, E_i)$
3149: induces the section $\yps_\mbfk: \scrm_\mbfk \to \bigoplus_i j^{2k_i+1}
3150: (L_i, E_i)$ with $\yps_\mbfk([u,J, \mbfz]) \deff \yps_\mbfk(u,J_S,J; \mbfz)$.
3151: \end{corol}
3152:
3153:
3154: \proof The claim follows from the fact that all the constructions are
3155: compatible with $\bfg$-action.
3156: \qed
3157:
3158: \smallskip
3159: \newdefi{def4.2.4n} %{def4.3.2a}
3160: For a given $\mbfk=(k_1, \ldots, k_m)$ we set
3161: \begin{align}
3162: \whcalm_{=\mbfk} &\deff \bigl\{ (u, J_S, J; \mbfz) \in \whcalm_\mbfk
3163: \;:\; \ord_{z^*_i} du =k_i \,\bigl\};
3164: \\
3165: \scrm_{=\mbfk} &\deff \bigl\{ [u, J; \mbfz] \in \scrm_\mbfk
3166: \;:\; \ord_{z^*_i} du =k_i \,\bigl\}.
3167: \end{align}
3168: \end{defi}
3169:
3170: \newlemma{lem4.2.7n} %{lem4.3.3}
3171: \sli The set $\whcalm_{=\mbfk}$ is an open $C^{\ell-1}$-smooth submanifold
3172: of $\whcalm_\mbfk$ invariant \wrt the natural action of\/ $\bfg$ on
3173: $\whcalm_\mbfk$.
3174:
3175: \slii The image of the projection of $\whcalm_{=\mbfk}$ to $\whcalm$ is an
3176: {\sl imbedded} submanifold of $\whcalm$, and the projection is a non-ramified
3177: covering over the image.
3178:
3179: \sliii There exists a $C^{\ell-1}$-smooth bundle $N$ over $\whcalm_{=\mbfk}
3180: \times S$ whose restriction onto $\{(u, J_S,
3181: \allowbreak
3182: J; \mbfz)\} \times S$ coincides with $N_u$. The diagonal action
3183: of\/ $\bfg$ on $\whcalm_{=\mbfk} \times S$ lifts canonically to the action
3184: on the bundle $N$.
3185:
3186: \sliv The bundle $N$ induces Banach bundles $L^{1,p}(S, N)$ and $L^p_{(0,1)}
3187: (S, N)$ over $\whcalm_{=\mbfk}$ with fibers $L^{1,p}(S, N_u)$ and $L^p_{(0,1)}
3188: (S, N_u)$ over $(u, J_S, J, \mbfz) \in \whcalm_{=\mbfk}$ respectively. The
3189: operators $D^N_{u,J}: L^{1,p}(S, N_u) \to L^p_{(0,1)}(S, N_u)$ induce a
3190: $C^{\ell-1}$-smooth bundle homomorphism $D^N: L^{1,p}(S, N) \to L^p_{(0,1)}
3191: (S, N)$.
3192: \end{lem}
3193:
3194: \proof \sli The complement $\whcalm_\mbfk \bs \whcalm_{=\mbfk}$ is of the
3195: union of (the projections of) the spaces $\whcalm_{\mbfk'}$ such that either
3196: $\mbfk' =(k_1, \ldots, k_m, 1)$, or $\mbfk' =(k'_1, \ldots, k'_m)$ with $k'_i
3197: \ge k_i$ and $k'_{i_0} > k_{i_0}$ for some $i_0$. In other words, we have
3198: either at least one additional cusp-point or a higher order cusp in at least
3199: one point. Obviously, these conditions define closed subsets in
3200: $\whcalm_\mbfk$. The $\bfg$-invariance of $\whcalm_{=\mbfk}$ follows from
3201: the definition.
3202:
3203: \slii The set $\whcalm_{=\mbfk}$ admits a finite transformation group $\aut(
3204: \mbfk)$ generated by transpositions of marked cusp-points $z^*_i$ and
3205: $z^*_j$ with $k_i = k_j$. The rest of part \slii follows.
3206:
3207: \sliii Let $z_i$ be a local $J_S$-holomorphic coordinate on $\whcalm_{=\mbfk}
3208: \times S$ centered at $z^*_i$ as in \refdefi{def4.2.2n}. It follows from the
3209: proof of \lemma{lem4.2.3n} that $z_i^{-k_i} du(z_i)$ is a well-defined
3210: {\sl non-vanishing} local section of $\hom(TS, E_u)$,
3211: which depends $C^{\ell-1}$-smoothly on $(u,J_S, J)$ and holomorphically on
3212: $z_i$. This provides the existence on $N$ with the stated property,
3213: at least locally in a neighborhood of $(u,J_S, J; z^*_i)$. The globalization
3214: of $N$ is trivial. Since the constructions involved are natural, the
3215: $\bfg$-action admits the desired lift.
3216:
3217: \sliv One uses the fact that the constructions of the bundles $L^{\!1,p}
3218: \!(S, N)$, $L^p_{(0,1)}\!(S, N)$, and the operator $D^N$ are natural. This
3219: implies $C^{\ell-1}$-smoothness of the obtained objects. \qed
3220:
3221:
3222: \smallskip
3223: \state Remark. One could explain the meaning of \lemma{lem4.2.7n} as follows.
3224: First, we note that for the globalization of normal bundles $N_u$ to $\scrm$
3225: we should use not the Cartesian product $\scrm \times S$, but the
3226: $\bfg$-twisted product $\scrm \ltimes S$, \ie
3227: $\whcalm \times_\bfg S \deff \bigl( \whcalm \times S\bigr)/\bfg$. Second, we
3228: must choose a stratification of $\scrm$ by strata where $N_u$ does not
3229: ``jump''. By the definition of $N_u$ such strata are exactly $\scrm_{=\mbfk}
3230: = \whcalm_{=\mbfk} / \bfg$ where there is no ``jump'' of
3231: the cusp-order.
3232:
3233:
3234:
3235: \medskip
3236: Another application of the techniques used in the proof of \refthm{thm4.2.1}
3237: is a local version of the theorem. Below $\scrp(\Delta, X)$ denotes the Banach
3238: space of pseudoholomorphic maps between the unit disc $\Delta$ with the standard
3239: structure $J\st$ and $X$, \ie $\scrp(\Delta, X) = \{ (u,J) \in L^{1,p}(\Delta,
3240: X) \times \scrj: \dbar_{J\st,J} u=0 \}$.
3241:
3242: \newlemma{lem4.2.8n} \sli For any given integer $k\ge1$ the set
3243: \begin{equation}
3244: \eqqno(4.2.4n)
3245: \llap{$\scrp_k($}\Delta,0; X) \deff
3246: \{ (u,J) \in \scrp(\Delta, X) : \ord_{z=0}(du) \ge k \}
3247: \end{equation}
3248: is a $C^\ell$-smooth submanifold of $\scrp(\Delta, X)$ of
3249: real codimension $2kn$, $n\deff \dimc X= \half \dimr X$, with tangent
3250: space
3251: \begin{equation}
3252: \eqqno(4.2.5n)
3253: T_{(u,J)}\scrp_k(\Delta,0; X) =
3254: \{ (v,\dot J) \in T_uL^{1,p}(\Delta, X)\times T_J\scrj :
3255: D_{u,J}v=0, j^k(v(z)-v(0)) =0 \}.
3256: \end{equation}
3257: \end{lem}
3258:
3259:
3260: \newsubsection[4.2s]{Curves with prescribed secondary cusp index}
3261: %In this paragraph we prove some claims to be used later.
3262: Recall that by \lemma{lem1.2.3} for a pseudoholomorphic map $u:(S,J_S) \to (X,J)$
3263: with cusp order $k$ at $z^*\in S$ the jet $j^{2k+1}u(z^*)$ is well-defined. As we
3264: shall see, the part of the jet $j^{2k+1}u(z^*)$ invariant under reparameterization
3265: plays an important role for determining the type of critical points on
3266: moduli spaces (see \refsubsection{3.3}).
3267:
3268: \newdefi{def4.2s.1} \sli Let $u:(S,J_S) \to (X,J)$ be pseudoholomorphic map with
3269: a cusp of order $k\deff \ord_{z^*} du$ at $z^*\in S$, $\pr_N: E_u \to N_u$
3270: the projection to the normal bundle, and $z$ a local holomorphic coordinate
3271: on $S$ centered at $z^*$. Define the {\sl secondary cusp index
3272: $l$ of $u$ at $z^*\in S$} by setting $l\deff k$ if $\pr_N \scirc j^{2k+1}
3273: u(z^*)$ is zero polynomial and $l\deff \ord_{z=0} \pr_N \scirc j^{2k+1}
3274: u(z^*) -k-1$ otherwise.
3275:
3276: \slii For a given $m$-tuple $\mbfk=(k_1, \ldots, k_m)$ of prescribed
3277: orders of cusps we consider $m$-tuples $\mbfl= (l_1, \ldots, l_m)$
3278: with $0 \le l_i \le k_i$ and set $|\mbfl| \deff \sum_i l_i$.
3279: Define the {\sl moduli space $\scrm_{\mbfk, \mbfl}$ of pseudoholomorphic maps
3280: with cusps of given order and secondary index $(\mbfk, \mbfl)$} as the set of
3281: $[u,J, \mbfz] \in \scrm_\mbfk$ such that $\ord_{z^*_i} du =k_i$
3282: and the secondary cusp index of $u$ at $z^*_i$ is at least $l_i$. Set
3283: \begin{equation}
3284: \eqqno(4.2s.1)
3285: \whcalm_{\mbfk, \mbfl} \deff \{ (u, J_S, J, \mbfz) \in \whcalm_{=\mbfk}
3286: \;:\; [u, J, \mbfz] \in \scrm_{\mbfk, \mbfl} \}.
3287: \end{equation}
3288: \end{defi}
3289:
3290:
3291: \smallskip
3292: \newthm{thm4.2s.1} The space $\scrm_{\mbfk, \mbfl}$ is a closed $C^{\ell-1}
3293: $-smooth submanifold of $\scrm_{=\mbfk}$ of codimension $2(n-1)|\mbfl|$.
3294: \end{thm}
3295:
3296: \state Remark.
3297: The meaning of the notion of secondary cusp index can be explained as follows.
3298: One expects that for a $J$-holomorphic map $u:S \to X$ with a cusp of order $k=
3299: \ord_{z^*}du$ at $z^* \in S$ the polynomial $\pr_N \scirc j^{2k+1} u(z^*)$ has
3300: vanishing order $k+1$. Thus the secondary cusp index $l$ is the order of deviation
3301: from this condition. The content of \refthm{thm4.2s.1} is that for generic
3302: $[u,J;\mbfz] \in \scrm_{=\mbfk}$ there is no deviation and that the space of
3303: curves with cusps of prescribed degeneration order is of expected codimension.
3304:
3305: We note also that the range $0\le l_i \le k_i=\ord_{z^*_i}du$ is the maximal
3306: one where the secondary cusp index is well-defined: The higher order terms
3307: of $\pr_N \scirc du$, as well as the coefficients of $du$ (considered as a
3308: holomorphic section of $T^*S \otimes E_u$), depend on the choice of the local
3309: holomorphic coordinate $z_i$ centered at $z^*_i\in S$.
3310:
3311: \proof We maintain the notation of \lemma{lem4.2.3n}. Now, for any $(u,J_S,J;
3312: \mbfz)\in \whcalm_{=\mbfk}$, $\mbfz = (z^*_1,\ldots, z^*_m)$, the jets
3313: $j^{2k_i+1} u(z^*_i) \in j^{2k_i+1} (TS_i, E_i)_{(u,J_S, J; \mbfz)}$ are
3314: well-defined and depend $C^{\ell-1}$-smoothly on $(u,J; \mbfz)$. By
3315: \lemma{lem4.2.7n}, for any $i=1,\ldots,m$ the formula $(N_i)_{(u,J_S, J;
3316: \mbfz)} \deff (N_u)_{z^*_i}$ defines a $C^{\ell-1}$-smooth bundle $N_i$ over
3317: $\whcalm_{=\mbfk}$ with the projection $\pr_N : E_i \to N_i$.
3318: This yields the compositions $\pr_N \scirc j^{2k_i+1}u(z^*_i) \in j^{2k_i+1}
3319: (TS_i, N_i)_{(u,J_S, J; \mbfz)}$ which depend $C^{\ell-1}$-smoothly on
3320: $(u,J_S,J; \mbfz)$. Thus we obtain a $C^{\ell-1}$-smooth bundle
3321: $$
3322: \bigoplus\nolimits_{i=1}^m j^{k_i+1, k_i+l_i+1}(TS_i, N_i)
3323: _{(u,J_S, J; \mbfz)}
3324: $$
3325: over $\whcalm_{=\mbfk}$ of rank $2(n-1)|\mbfl|$ and a $C^{\ell-1}$-smooth
3326: section
3327: $$
3328: \yps^N_{\mbfk,\mbfl} \deff
3329: (\pr_N \scirc j^{k_i+1, k_i+l_i+1} u(z^*_i))_{i=1}^m.
3330: $$
3331: Observe that $\whcalm_{\mbfk,\mbfl}$ is defined in $\whcalm_{=\mbfk}$ as the
3332: zero set of $\yps^N_{\mbfk,\mbfl}$. It follows from \lemma{lem4.2.3n} that
3333: $\yps^N_{\mbfk,\mbfl}$ is transversal to the zero section. Consequently,
3334: $\whcalm_{\mbfk,\mbfl}$ is a submanifold of $\whcalm_{=\mbfk}$ of codimension
3335: $2(n-1)|\mbfl|$. The claim of the theorem follows now by taking the
3336: $\bfg$-quotient.\qed
3337:
3338:
3339:
3340:
3341: \newsubsection[4.3a]{Curves with cusps of prescribed type}
3342: In this paragraph we give a construction of $J$-curves of any given cusp type,
3343: completing the result of Micallef and White. In particular, we obtain a
3344: more direct and constructive proof of \lemma{lem1.2.1} without referring
3345: to local structure of minimal surfaces, as is done in \cite{Mi-Wh}. Then
3346: we show that the set of cusp-curves with prescribed cusp type is a Banach
3347: submanifold of the total moduli space and compute its codimension.
3348:
3349: \smallskip
3350: Let $J$ be an almost complex structure on the ball $B \subset \cc^n$ such that
3351: $J(0) =J\st(0)$. We assume that $J$ is $C^\ell$-smooth with $\ell \ge 2$.
3352: First we consider the local structure of multiple maps.
3353:
3354: \newlemma{lem4.3a.1} Let $u: \Delta \to B$ be a non-constant $J$-holomorphic
3355: map with $u(0)=0 \in B$. Then there exist a radius $r>0$, a uniquely defined
3356: $\nu\in \nn$, and a non-multiple $J$-holomorphic map $u': \Delta(r^\nu) \to B$
3357: such that $u(z) = u'(z^\nu)$ for $z\in \Delta(r)$.
3358: \end{lem}
3359:
3360: \proof By \lemma{lem1.2.4}, $u(z) = v \cdot z^\mu + O(|z|^{\mu + \alpha})$
3361: with some $v\in T_0B=\cc^n$, positive $\mu \in\nn$, and $\alpha>0$. If $\mu
3362: =1$, then $u$ is already non-multiple in some $\Delta(r)$ and there are
3363: nothing to prove. Thus we may assume that $\mu \ge 2$.
3364:
3365: Take a sufficiently small $\rho_0>0$ and consider $U\deff u\inv(B(\rho_0))$.
3366: By the first part of \lemma{lem1.2.4}, $U$ is a disc and $u$ is an immersion
3367: in $U\bs \{0\}$. Using the second part of \lemma{lem1.2.4} it is not difficult
3368: to show that $u(U \bs \{0\})$ is an immersed $J$-holomorphic punctured disc
3369: in $B$. Therefore the restriction $u\ogran_U$ is a composition
3370: of a non-multiple $J$-holomorphic map and a covering branched only in
3371: $0\in U$.\qed%\hfill$\square$
3372:
3373: \medskip
3374: It is known that any non-multiple holomorphic map $u: \Delta \to \cc^n$,
3375: in appropriate holomorphic coordinates on $\Delta$ and $\cc^n$,
3376: has locally the form
3377: \begin{equation}
3378: \notag%\eqqno(4.3a.0a)
3379: u(z) = \sum_{i=0}^l v_i z^{p_i},
3380: \end{equation}
3381: with the following properties. $p_0 = \ord_0(du) +1$, $v_0 \not=0$, the vectors
3382: $v_i \in \cc^n$ are linearly independent of $v_0$ for $i>0$, and $\gcd(p_0,\ldots,
3383: p_l)=1$. We want to establish a similar result for pseudoholomorphic curves,
3384: replacing the operation $u_{i-1}(z) \mapsto u_{i-1}(z) + v_i z^{p_i}$ by
3385: $u_{i-1}(z) \mapsto \dfrm_{p_i}(u_{i-1}, v_i)$.
3386:
3387: \newlemma{lem4.3a.2} Let $B \subset \cc^n$ be the unit ball, $J$ a $C^\ell$-smooth
3388: almost complex structure on $B$ with $J(0)= J\st$, and $u: \Delta \to B$ a
3389: non-multiple $J$-holomorphic map such that $u(z)= v_0 z^{p_0} +o(z^{p_0})$ for
3390: some $p_0>1$ and $v_0 \not=0 \in \cc^n$. Take a divisor $d>1$ of $p_0$ and
3391: denote by $\eta$ a primitive $d$-th root of unity. Then there exist an integer
3392: $q>0$, a vector $v\in \cc^n$, and a complex polynomial $\psi(z)$ such that
3393:
3394: \sli $q$ is {\sl not} a multiple of $d$;
3395: %\lineeqqno(4.3a.1a)
3396:
3397: \slii $v$ is $\cc$-linearly independent of $v_0;$ in particular, $v \not=0;$
3398:
3399: \sliii $\psi(z)= z + o(z)$ and $\deg\psi(z)\le q;$
3400:
3401: \sliv $u(\eta z) = u( \psi(z)) + z^q \cdot v + o(z^q).$
3402: \lineeqqno(4.3a.2a)
3403:
3404: %\slv Moreover, for any complex polynomial $\phi(z)$ with $\phi(z)=z +o(z)$
3405: %\begin{equation}
3406: %\eqqno(4.3a.2b)
3407: %u(\eta \phi(z))= u( \psi(z)) + z^q \cdot v + o(z^q)
3408: %\end{equation}
3409: \end{lem}
3410:
3411: \proof Denote by $v_0^\perp\subset \cc^n$ a complex orthogonal complement to
3412: $v_0$, and by $B^\perp(\rho)$ the ball of radius $\rho$ in $v_0^\perp$.
3413: Note that we can canonically identify the space $v_0^\perp$ with the fiber
3414: $(N_u)_{z=0}$ of the normal bundle $N_u$ of $u$ (see \refdefi{def1.5.1}).
3415: Fixing a holomorphic frame $w_1(z), \ldots w_{n-1}(z)$ of $N_u$ we can
3416: identify $\Delta \times v_0^\perp$ with the total space of $N_u$ over $\Delta$
3417: and use $(z, w_1, \ldots w_{n-1})$ as coordinates in $\Delta \times v_0^\perp$.
3418: Denote by $J\st$ the standard integrable complex structure in $\Delta \times
3419: v_0^\perp$. It coincides with the canonical holomorphic structure in $N_u$. Set
3420: $$
3421: U_{r,\rho} \deff \Delta(r) \times B^\perp(\rho)
3422: $$
3423: Fix a holomorphic splitting $F_0: N_u \to E_u$ of the projection $\pr_N : E_u
3424: \to N_u$. We shall identify $N_u$ as a subbundle of $E_u$ by means of $F_0$.
3425: Define the map $F: U_{r,\rho} \to \cc^n$ as the composition
3426: $$
3427: (z,w) \mapsto F_0(z)(w) \in (E_u)_z = T_{u(z)}B=\cc^n
3428: \mapsto F(z,w) \deff u(z) + F_0(z)(w).
3429: $$
3430: It is not difficult to see that for sufficiently small $r$ and $\rho$ the map
3431: $F=F(z,w)$ takes values in $B$ and has the following properties:
3432:
3433: \begin{itemize}
3434: \item $F(z,w)$ is $C^1$-smooth;
3435: \item $F(z,0) = u(z)$ and $\nabla_{\dot w}F(z,0)=\dot w$; or more precisely,
3436: $\nabla_{\dot w}F(z,0)=F_0(z)(\dot w)$;
3437: \item the pulled-back structure $\ti J \deff F^*J$ coincides with $J\st$ along
3438: the set $\check\Delta \times \{0\}$, \ie
3439: \begin{equation}
3440: \ti J(z,0) = J\st(z,0).
3441: \eqqno(4.3a.3a)
3442: \end{equation}
3443: \end{itemize}
3444:
3445:
3446: From \eqqref(4.3a.3a) we obtain a uniform estimate
3447: \begin{equation}
3448: \eqqno(4.3a.4a)
3449: \bigl|\ti J(z,w) - J\st(z,w)\bigr| \le C \cdot |w|.
3450: \end{equation}
3451: Further, $\eta^{p_0}=1$ obviously gives $u(\eta z) -u(z)= o(z^{p_0}) = o(z^d)$.
3452: This implies that for sufficiently
3453: small $r'$ we can represent $u(z)$ in the form $u(z) = F(\zeta(z), \ti w(z))$
3454: with uniquely defined $C^1$-smooth $\zeta(z): \Delta(r') \to \Delta(r)$ and
3455: $\ti w(z): \Delta(r') \to B^\perp(\rho)$ fulfilling the condition $\zeta(z)= z
3456: +o(z)$. Further, $\ti w(z)= o(z^d)$.
3457:
3458: Set $\ti u(z) \deff \bigl( \zeta(z), \ti w(z) \bigr)$. We obtain a $C^1$-smooth
3459: map $\ti u(z): \Delta(r') \to U_{r,\rho}$, for which
3460: $$
3461: |J\st(\ti u(z)) - \ti J(\ti u(z)) | =
3462: |J\st (\zeta(z), \ti w(z))- \ti J(\zeta(z), \ti w(z)) |
3463: \le C'\cdot |\ti w(z)|.
3464: $$
3465: Consequently
3466: $$
3467: \bigl|\dbar_{J\st} \ti u(z) \bigr|=
3468: \bigl|\dbar_{J\st} \ti u(z) - \dbar_{\ti J} \ti u(z) \bigr|=
3469: \bigl|\bigl(J\st(\ti u(z)) - \ti J(\ti u(z)) \bigr) \d_y\ti u(z)\bigr|
3470: \le C''\cdot |\ti w(z)|,
3471: $$
3472: or explicitly for components $\zeta(z)$ and $\ti w(z)$
3473: \begin{align}
3474: \bigl|\dbar_{J\st} \ti w(z) \bigr| & \le C''\cdot |\ti w(z)|;
3475: \eqqno(4.3a.5a)\\
3476: \bigl|\dbar_{J\st} \zeta(z) \bigr| & \le C''\cdot |\ti w(z)|.
3477: \eqqno(4.3a.6a)
3478: \end{align}
3479: Observe that $\ti w(z)$ is not identically zero. Otherwise we would obtain
3480: that $u(\eta z) = u(\zeta(z))$, which would contradict the condition of
3481: non-multiplicity of $u(z)$.
3482:
3483: Hence, by \lemma{lem1.2.1}, $\ti w(z) = z^q v + o(z^q)$ and $\zeta(z)= \psi(z)
3484: + o(z^q)$ for some $q>0$, non-zero $v\in v_0^\perp$, and a complex polynomial
3485: $\psi(z)$ of degree $\le q$. Substituting these relations in $u(z)=
3486: F(\ti u(z))$ we obtain \eqqref(4.3a.2a).
3487:
3488: Finally, the identity $\sum_{j=1}^d \bigl(u(\eta^j z) - u(\eta^{j-1} z) \bigr)
3489: \equiv 0$ together with \eqqref(4.3a.2a) implies that $\sum_{j=1}^d (\eta^{j-1}
3490: z)^q \cdot v =0$. Thus $\sum_{j=1}^d \eta^{jq} =0$ which is possible \iff $q$
3491: is not a multiple of $d$. \qed
3492:
3493: \medskip
3494: Iterating the construction of \lemma{lem4.3a.2}, we obtain
3495:
3496: \newcorol{cor4.3a.3} Let $B \subset \cc^n$ be the unit ball, $J$ a $C^\ell
3497: $-smooth almost complex structure in $B$ with $J(0)= J\st$, and $u: \Delta \to B$
3498: a non-multiple $J$-holomorphic map with $u(0)=0$.
3499:
3500: Then there exist uniquely defined sequences $(p_0, p_1, \ldots, p_l)$ and
3501: $(d_0, d_1, \ldots, d_l)$ of positive integers with the following properties:
3502:
3503: \sli $p_0 = \ord_0 du +1$, so that $u(z) = z^{p_0} v_0 + o(z^{p_0})$ with
3504: non-zero $v_0 \in \cc^n$;
3505:
3506: \slii $d_i = \gcd(p_0, \ldots, p_i);$
3507:
3508: \sliii $p_i< p_{i+1}$, $d_i > d_{i+1}$, and $d_l=1;$
3509: in particular, $p_{i+1}$ is not a multiple of $d_i$;
3510: \lineeqqno(4.3a.7a)
3511:
3512: \sliv if $\eta_i$ is the primitive $d_i$-th root of unity, then
3513: $$
3514: u(\eta_i z) = u(\psi_i(z)) \cdot v_0 + z^{p_{i+1}} \cdot v_{i+1}
3515: + o(z^{p_{i+1}})
3516: $$
3517: for appropriate complex polynomials $\psi_i(z)$ with $\psi_i(z)=z+o(z)$, and
3518: vector $v_{i+1}\in \cc^n$, $\cc$-linearly independent of $v_0$.
3519: \end{corol}
3520:
3521: %\proof The claim follows immediately from \lemma{lem4.3a.2}. \qed
3522:
3523: \newdefi{def4.3a.1} \sli To any increasing sequence of positive integers
3524: $1\le p_0 < p_1 < \cdots < p_l$ we associate the {\sl sequence of divisors}
3525: $d_i \ge d_1 \ge \cdots \ge d_l$ defined by $d_i = \gcd(p_0, \ldots, p_i)$.
3526: In particular, $d_0 =p_0$.
3527:
3528: \slii A sequence $\vec p= (p_0,p_1,\ldots, p_l)$ of positive integer exponents
3529: is called a {\sl cusp type} if $p_i$ and the associate divisors $d_i = \gcd
3530: (p_0, \ldots, p_i)$ satisfy the condition \eqqref(4.3a.7a). In the situation of
3531: \refcorol{cor4.3a.3}, the sequence $\vec p= (p_0,p_1, \ldots, p_l)$ is called
3532: the {\sl cusp type of $u$ at $z=0$}, $p_i$ the {\sl critical exponents of $u$
3533: at $z=0$}, and $\vec d =(d_i)$ the {\sl sequence of divisors of $u$ at $z=0$}.
3534:
3535: \sliii For a given cusp type $\vec p= (p_0,p_1, \ldots, p_l)$, an integer $p'$
3536: is called an {\sl admissible exponent} if $p'$ equals $p_l$ or is of the
3537: form $p'= p_i+ j\cdot d_i$ for some $i=0,\ldots,l-1$ and $j=0,\ldots, l_i$,
3538: $l_i \deff \left[ {p_{i+1} - p_i \over d_i } \right]$. Thus all critical
3539: exponents are admissible and there are exactly $l_i$ non-critical admissible
3540: exponents between $p_i$ and $p_{i+1}$. Denote by $\vpp = (p'_0, \ldots,
3541: p'_{l'})$ the sequence of the admissible exponents ordered by growth. Its
3542: length is $l' = l + \sum_{i=0}^{l-1} l_i = l + \sum_{i=0}^{l-1}\left[
3543: {p_{i+1} - p_i \over d_i } \right]$.
3544:
3545: Note that the corresponding sequence of divisors $d'_j \deff \gcd(p'_0,
3546: \ldots, p'_j)$ consists of divisors $d_i$ of critical exponents, repeated $l_i+
3547: 1$ times. Vice versa, an admissible exponent $p'_j>p'_0=p_0$ is critical \iff
3548: $d'_j < d'_{j-1}$.
3549: \end{defi}
3550:
3551: \smallskip
3552: \newthm{thm4.3a.4} Let $B \subset \cc^n$ be the unit ball, $J$ an almost complex
3553: on $B$ with $J(0)= J\st$, and $u: \Delta \to B$ a non-multiple $J$-holomorphic map
3554: such that $u(0)= 0$. Further, let $\vec p= (p_0, \ldots, p_l)$ and $\vpp= (p_0,
3555: \ldots, p_{l'})$ be the sequences of critical and resp.\ admissible exponents
3556: of\/ $u$ at\/ $z=0$, and $\vdp =(d'_0, \ldots, d'_{l'})$ the corresponding
3557: sequence of divisors.
3558:
3559: Then there exist a sequence $(v_0, \ldots, v_{l'})$ of vectors in $\cc^n$ (one
3560: $v_j$ for each $p'_j$), a complex polynomial $\phi(z)$, and a radius $r>0$, such
3561: that the following holds.
3562:
3563: \sli $u(z)= z^{p_0}\cdot v_0 + o(z^{p_0})$; $v_0 \not=0$, $v_1, \ldots, v_{l'}$
3564: are complex orthogonal to $v_0;$
3565: \lineeqqno(4.3a.2)
3566:
3567:
3568: \slii $\phi(z)=z + o(z)$ and $\deg \phi(z) \le p_l-p_0 +1;$
3569: \lineeqqno(4.3a.3)
3570:
3571:
3572: \sliii for appropriately chosen maps $\dfrm$, the recursive formula
3573: \begin{equation}
3574: \eqqno(4.3a.4)
3575: u_j(z) \deff \dfrm_{p'_j/d'_j}\bigl(u_{j-1}(z^{d'_{j-1}/d'_j}), J; v_j\bigr)
3576: \quad j=0,1,\ldots,l'
3577: \end{equation}
3578: beginning from $u_{-1}(z) \equiv 0$ yields a sequence of well-defined
3579: $J$-holomorphic maps $u_j: \Delta(r^{d'_j}) \to B$ with the property
3580: \begin{equation}
3581: \eqqno(4.3a.5)
3582: u(\phi(z)) - u_j(z^{d'_j}) = v_{j+1} z^{p'_{j+1}} + o(z^{p'_{j+1}}).
3583: \end{equation}
3584:
3585: \noindent
3586: Moreover, such $v_j$ and $\phi(z)$ are uniquely defined. Further, $v_j$ is
3587: non-zero if $p'_j$ is critical.
3588: \end{thm}
3589:
3590: \proof The choice of the maps $\dfrm_d $ ensuring that at each recursive step
3591: the right hand side of \eqqref(4.3a.5) is well-defined will be made below. Now
3592: we assume that for given $j<l'$ we have constructed a $J$-holomorphic map $u_j:
3593: \Delta(r^{d'_j}) \to X$ and a polynomial $\phi(z)$ such that $\phi(z)= z+o(z)$
3594: and $u(\phi(z)) = u_j(z^{d'_j}) +o(z^{p'_j})$. Then by \lemma{lem1.2.4},
3595: \begin{equation}
3596: \eqqno(4.3a.6)
3597: u(\phi(z)) = u_j(z^{d'_j}) + z^q w + o(z^q)
3598: \end{equation}
3599: for some non-zero $w \in \cc^n$ and $q>p'_j$. Represent $w\in \cc^n$ in the form
3600: $w= a \cdot v_0 + w'$ and replace $\phi$ by $\phi'(z) \deff \phi(z)- {a\over p_0}
3601: \cdot z^{q-p_0 +1}$. The relations $u(z) = z^{p_0} v_0 + o(z^{p_0})$ and
3602: \eqqref(4.3a.6) yield
3603: \begin{equation}
3604: \eqqno(4.3a.6')
3605: u(\phi'(z)) = u_j(z^{d'_j}) + z^q w' + o(z^q).
3606: \end{equation}
3607: If $w' =0$, we can consider \eqqref(4.3a.6) with some $q'>q$. Thus we may assume
3608: that $w' \not=0$. Moreover, we see that $\phi(z)$ is defined uniquely by
3609: \eqqref(4.3a.5) up to degree $p'_j -p_0 +1$.
3610:
3611: Denote by $\eta_j$ the primitive $d'_j$-th root of unity. Then
3612: by \lemma{lem4.3a.2},
3613: \begin{equation}
3614: \eqqno(4.3a.7)
3615: u(\eta_j z) = u(\psi_j(z)) + v z^p + o(z^p)
3616: \end{equation}
3617: for an appropriate polynomial $\psi_j(z) =z +o(z)$ and $v$ linearly independent
3618: of $v_0$. Moreover, $p$ is the first critical exponent after $p'_j$ in the
3619: sequence $\vpp$ of the admissible exponents of $u(z)$ at $z=0$. In particular, $p$
3620: is not a multiple of $d'_j$. Set $\hat\phi_{j+1}(z) \deff \eta_j\inv\phi_{j+1}
3621: (\eta_j z)$. Then we obtain $\hat\phi_{j+1}(z) = z+ o(z)$ and
3622: \begin{align}
3623: u(\phi_{j+1}(\eta_j z)) &= u(\eta_j\hat\phi_{j+1}(z)) =
3624: u(\psi_j(\hat\phi_{j+1}(z))) + z^p v + o(z^p)
3625: \notag\\
3626: &= u(\phi_{j+1}(\hat\psi_j(z))) + z^p v + o(z^p),
3627: \eqqno(4.3a.8)
3628: \end{align}
3629: where $\hat\psi_j(z)$ is a polynomial with $\hat\psi_j(z)= z +o(z)$ and
3630: $\hat\psi_j(\hat\phi_{j+1}(z)) = \phi_{j+1}(\hat\psi_j(z)) +o(z^p)$.
3631: Substitution of \eqqref(4.3a.6') in \eqqref(4.3a.8) together with the identity
3632: $\eta_j^{d'_j}=1$ yields
3633: \begin{equation}
3634: \eqqno(4.3a.9)
3635: u_j( z^{d'_j}) + \eta_j^q z^q w' =
3636: u_j\bigl( \hat\psi_j^{d'_j}(z) \bigr)
3637: + z^q w' + v z^p + o(z^{\min(p,q)}).
3638: \end{equation}
3639: Further, since $\hat\psi_j(z)= z +o(z)$, we can find a polynomial $\ti\psi_j(z)$
3640: with the properties $\ti\psi_j(z)= z +o(z)$ and $\hat\psi_j^{d'_j}(z)= \ti\psi_j
3641: (z^{d'_j})$. For such $\ti\psi_j(z)$, the relation \eqqref(4.3a.9) transforms to
3642: \begin{equation}
3643: \eqqno(4.3a.10)
3644: u_j( z^{d'_j}) =
3645: u_j\bigl( \ti\psi_j(z^{d'_j}) \bigr)
3646: + (1 - \eta_j^q) z^q w' + v z^p + o(z^{\min(p,q)}).
3647: \end{equation}
3648: Assume that $q<p$. Then $q$ is a multiple of $d'_j$. In particular, $q\ge
3649: p'_{j+1}$. In the case $q> p'_{j+1}$ we simply set $v_{j+1} \deff 0$ and obtain
3650: the relation \eqqref(4.3a.5). In the case $q= p'_{j+1}$ we set $v_{j+1} \deff w'$
3651: and come to the relation \eqqref(4.3a.5) again. The case $p<q$ is impossible since
3652: $p$ is not a multiple of $d'_j$.
3653:
3654: In the remaining case $q=p$ we have two subcases, $p'_{j+1} <p$ and $p'_{j+1}
3655: =p$. Then we set $v_{j+1} \deff 0$ or respectively $v_{j+1} \deff w'$ and obtain
3656: \eqqref(4.3a.5) from \eqqref(4.3a.6').
3657:
3658:
3659: \medskip
3660: Now we construct the maps $\dfrm_{p'_j/d_j}$ with the desired properties. The idea
3661: is to rescale the maps $u_j$ making the norms $\norm{du_j}_{L^2}$ sufficiently
3662: small and obtaining a recursive apriori estimate on $v_j$. For this fix some
3663: $r \in\; ]0,1[$ and maps $\wt \dfrm_p$ with the properties listed in
3664: \refdefi{def4.1.1}. Then the substitutions $\ti u(z) \deff u(r z)$,
3665: $\ti u_j(z) \deff u_j(r^{d'_j} z)$,
3666: $\ti v_j \deff r^{p'_j}v_j$, and $ \ti \phi_j(z) \deff r\inv \phi_j(r z)$
3667: transform \eqqref(4.3a.4) and \eqqref(4.3a.5) into recursive relations
3668: \begin{equation}
3669: \eqqno(4.3a.a6)
3670: \llap{$\ti u_j(z)= $}\;\wt\dfrm_{p'_j/d'_j}
3671: \bigl(\ti u_{j-1}(z^{d'_{j-1}/d'_j}), J;
3672: \ti v_j\bigr),
3673: \end{equation}
3674: \begin{equation}
3675: \eqqno(4.3a.a7)
3676: \ti u(\ti \phi(z)) - \ti u_j(z^{d'_j})=
3677: \ti v_{j+1} z^{p'_{j+1}} + o(z^{p'_{j+1}}).
3678: \end{equation}
3679: for $J$-holomorphic maps $\ti u_j: \Delta \to B$. Note that $\norm{d\ti u}
3680: _{L^2(\Delta)} = \norm{du}_{L^2(\Delta(r))}$ will be arbitrarily small for $r$
3681: small enough. Choosing an appropriate $r\ll1$ and using induction, on can obtain
3682: sufficiently small upper bounds on $\ti v_j$, ensuring that \eqqref(4.3a.a6) is
3683: well-defined for $j=0,\ldots, l$. For such $r$, we define $\dfrm_{p'_j/d_j}$
3684: by the reverse substitutions in \eqqref(4.3a.a6).
3685: \qed
3686:
3687:
3688: \smallskip
3689: \state Remark. For almost complex surface, \ie in the case $n=2$, the critical
3690: exponents determine a topological type of a cusp. In particular, under
3691: hypotheses of \refthm{thm4.3a.4}, the intersection of the image $u(\Delta)$
3692: with the sphere $S^3_r$ of a sufficiently small radius $r>0$ is an iterated
3693: toric knot $\gamma$ transversal to the 2-plane distribution $\xi$ on $S^3_r$
3694: given by $\xi_x \deff T_xS^3_r \cap J(x) T_xS^3_r$. Thus the Bennequin index
3695: $\beta(\gamma, \xi)$ is well-defined. We refer to \cite{Iv-Sh-1} for the proof
3696: of the formula $\beta= 2\delta -1$ relating the Bennequin index $\beta$ of
3697: $\gamma$ and the nodal number $\delta$ of $u(\Delta)$ in $0\in B$. On the
3698: other hand, $\delta$ can be computed by the formula
3699: \begin{equation}
3700: \eqqno(4.3a.11a)
3701: \delta = \sum_{i=1}^m (d_{i-1}- d_i)(p_i -1),
3702: \end{equation}
3703: see \cite{Rf} or \cite{Mil}. In the higher dimensional setting,
3704: \ie for $n\ge3$, the topological type of the
3705: cusp $u(\Delta)$ is not determined by the critical exponents and depends on
3706: additional information encoding further linear relations between $v_j$. For
3707: example, the condition {\sl $v_2$ and $v_1$ are linearly dependent} defines a
3708: proper subset in $\scrp_{\vec p}(\Delta, 0;B)$. Moreover, using the techniques
3709: of this paragraph one can show that this subset is a $C^{\ell-1}$-smooth
3710: submanifold in $\scrp_{\vec p}(\Delta, 0;B)$. Details can be recovered by an
3711: interested reader.
3712:
3713:
3714:
3715:
3716: \smallskip
3717: \newthm{thm4.3a.5} Let $B \subset \cc^n$ be the unit ball, $\vec p= (p_0,\ldots,
3718: p_l)$ a cusp type, $\vpp= (p_0, \ldots, p_{l'})$ the corresponding sequences of
3719: admissible exponents, and $\vdp =(d'_0, \ldots, d'_{l'})$ the sequence of
3720: divisors associated with $\vpp$. Then the set
3721: \begin{equation}
3722: \eqqno(4.3a.11)
3723: \llap{$\scrp_{\vec p} (\Delta, 0$};B) \deff
3724: \bigl\{ (u,J) \in \scrp_{p_0-1} (\Delta, 0; B):
3725: u \text{ \rm has a cusp type $\vec p$ in $z=0$}\; \bigr\}
3726: \end{equation}
3727: is a $C^{\ell-1}$-smooth submanifold of $\scrp_{p_0-1} (\Delta, 0;B)$ of real
3728: codimension $2(n-1)(p_l -p_0-l')$.
3729: \end{thm}
3730:
3731: Note that $p_0=p'_0$ and $p_l= p'_{l'}$.
3732:
3733: \proof Let $u:\Delta(r) \to B$ be a $J$-holomorphic map, $r>0$, and let
3734: \begin{equation}
3735: \scrp_k(\Delta, u; B,J) \deff
3736: \{ u'\in \scrp(\Delta; B, J) : u(z) -u'(z) = o(z^k) \}
3737: \notag
3738: \end{equation}
3739: By \refthm{thm4.2b.1}, $\scrp_k(\Delta, u; B,J)$ is a $C^{\ell-1} $-smooth
3740: submanifold of $\scrp(\Delta; B, J)$ of codimension $2n(k+1)$. For $l>k$ it
3741: follows that $\scrp_l(\Delta, u; B,J)$ has codimension $2n(l-k)$ in $\scrp_k(
3742: \Delta, u; B,J)$. Moreover, if $J_y$ is a $C^\ell$-smooth family of almost complex
3743: structures in $B$ parameterized by a (Banach) manifold $\scry$ and $u_y\in
3744: L^{1,p}(\Delta(r), B)$ a $C^{\ell-1}$-smooth family of $J_y$-holomorphic maps,
3745: then $\cup_{y\in \scry} \scrp_k(\Delta, u_y; B,J_y)$ is a $C^{\ell-1} $-smooth
3746: manifold.
3747:
3748: \smallskip
3749: For a given $(u^*, J^*) \in \scrp_{\vec p} (\Delta, 0;B)$, let $v^*_0, \ldots,
3750: v^*_{l'}$ and $\phi^*(z) = z + \phi^*_2 z^2 + \phi^*_3 z^3+ \cdots$ be the
3751: parameters of $u^*$ constructed in \refthm{thm4.3a.4}. Define $\scry$ to be
3752: the space of small deformations of $v^*_j$ and $\phi^*_i$. This means that $y
3753: \in \scry$ is a tuple $(v_0, \ldots, v_{l'}; \phi_2, \ldots, \phi_{p_l-p_0+1})$
3754: with $v_j \in \cc^n$ and $\phi_i \in \cc$ satisfying $|v_j - v^*_j| <\epsi$
3755: and $|\phi_i - \phi^*_i| < \epsi$ with $\epsi$ sufficiently small. Further,
3756: let $U$ denote a sufficiently small neighborhood of $J^*$ in the space of
3757: $C^\ell$-smooth almost complex structures in $B$. For $y=(v_0, \ldots;\phi_2,
3758: \ldots) \in \scry$ and $J\in U$ we construct the maps $u_{y,J; j}$, $j=0,
3759: \ldots, l'$, using the recursive relation \eqqref(4.3a.4) and set $\phi_y(z)
3760: \deff z + \phi_2 z^2 + \cdots +\phi_{p_l -p_0 +1} z^{p_l -p_0 +1}$. Then for
3761: $|z| < r' \ll1$ the inverse map $\phi_y\inv(z)$ is well-defined and
3762: holomorphic. Define $u_{y,J}(z) \deff u_{y,J; l'}(\phi\inv(z))$. We obtain a
3763: $C^{\ell-1}$-smooth family of pseudoholomorphic maps $u_{y,J}: \Delta(r') \to B$
3764: parameterized by $(y,J) \in \scry\times U$.
3765:
3766: Note that by \refthm{thm4.3a.4} every $(u,J) \in \scrp_{\vec p} (\Delta, 0;B)$
3767: sufficiently close to $(u^*, J^*)$ lies in $\scrp_{p_l}(\Delta, u_{y,J}; B,J)$
3768: for an appropriate $y\in \scry$, and such $y\in \scry$ is uniquely defined.
3769: Thus the union $\cup_{(y,J)\in \scry\times U} \scrp_{p_l}(\Delta, u_{y,J}; B,
3770: J)$ is a local $C^{\ell-1}$-smooth chart for $\scrp_{\vec p} (\Delta, 0;B)$.
3771:
3772: Finally, note that the union $\cup_{(y,J)\in \scry\times U} \scrp_{p_0-1}
3773: (\Delta, u_{y,J}; B,J)$ is naturally isomorphic to $\scrp_{p_0}(\Delta, 0; B)
3774: \times \scry$. Computing the number of parameters we obtain the codimension of
3775: the imbedding $\scrp_{\vec p} (\Delta, 0;B) \hook \scrp_{p_0} (\Delta, 0;B)$.
3776: \qed
3777:
3778: Globalizing \refthm{thm4.3a.5} we obtain
3779:
3780: \newcorol{cor4.3a.6} Let $\vec \mbfp =(\vec p_1, \ldots, \vec p_m)$ be a
3781: sequence of cusp types, $\vec p_i= (p_{i,0}, \ldots p_{i, l_i})$. Set
3782: $k_i \deff p_{i,0} -1$ and $\mbfk\deff (k_1, \ldots, k_m)$. Then the space
3783: $$
3784: \scrm_{\vec \mbfp} \deff \{ [u,J; z^*_1, \ldots z^*_m] \in \scrm_\mbfk:
3785: u \text{ \sl has cusp type $\vec p_i$ in }z^*_i\; \}
3786: $$
3787: is a $C^{\ell-1}$-smooth submanifold of $\scrm_\mbfk$ of codimension
3788: $2(n-1)\sum_{i=0}^m (p_{i,l_i} -p_{i,0}-l'_i)$.
3789: \end{corol}
3790:
3791:
3792: \medskip
3793: \newsection[3]{Saddle points in the moduli space}
3794:
3795: \newsubsection[3.1]{Critical and saddle points in the moduli space}
3796: In application of the continuity method for constructing $J$-holomorphic
3797: curves two main difficulties occur. The first one appears in the
3798: proof of the ``closedness part'', when one tries to extend a deformation $[u_t,
3799: J_t] \in \scrm, t\in [0,t'[\;$ into the endpoint $t'$. This difficulty is
3800: connected with the fact that the projection $\pi_{\!\!\scrj}: \scrm \to \scrj$
3801: is, in general, not proper. In particular, for a path $J_t\in \scrj, t\in [0,t']$
3802: there may not exist a lift $[u_t, J_t]$ to $\scrm$, and the fibers
3803: $\scrm_{\!\!\scrj} = \pi_J\inv (J)$ can be non-compact. Gromov's compactness
3804: theorem (\cite{Gro}, see also \cite{Iv-Sh-3}) gives a fiberwise topological
3805: compactification of $\scrm$ by adding certain degenerate curves $C$.
3806: However, for the moment we neglect this difficulty assuming we can avoid it in
3807: our case.
3808:
3809: The second main difficulty appears in the proof of the ``openness part'' when one
3810: tries to extend a lift $[u_t, J_t] \in \scrm$, $t \in[0, t']$, of a
3811: path of $J_t\in \scrj, t\in [0, 1]$ to a bigger interval $t \in[0, t''[$ with
3812: some $t''>t'$. Obviously, this difficulty can appear only if $[u_{t'}, J_{t'}]$
3813: is a {\sl critical point} of $\pi _{\!\!\scrj}$, \ie when the differential
3814: of the projection $d\pi_{\!\!\scrj}$ is not surjective in $[u_{t'}, J_{t'}]$.
3815: Thus it is desirable to find conditions on the critical points of
3816: $\pi_{\!\!\scrj}$ which ensure the existence of such a lift.
3817:
3818: \smallskip
3819: Assume additionally that the given path $h: [0,1] \to \scrj$, $h(t)\deff
3820: J_t$, is $C^2$-smooth and transversal to $\pi_h: \scrm \to \scrj$, \ie
3821: $\scrm_h$ is a manifold. Let $\pi_h: \scrm_h \to I$, $I\deff [0,1]$, be the
3822: projection. Then we have a well-defined $C^1$-smooth bundle homomorphism
3823: $d\pi_h: \scrm_h \to \pi_h^*(TI) \cong \rr$. Further, $d\pi_h$ vanishes
3824: exactly at critical points of $\pi_h$ and, by \lemma{lem1.3.1}, at each such
3825: point $p\deff [u,h(t)]$ we have a well-defined quadratic form $\nabla
3826: d\pi_h(p): T_p\scrm_h \to \rr\cong T_tI$. For our purpose it is
3827: sufficient to show that each critical point $p$ is a {\sl saddle}, \ie
3828: the quadratic form $\nabla d\pi_h(p)$ has at least one positive and one
3829: negative eigenvalue.
3830:
3831:
3832: \smallskip
3833: It turns out that this condition depends only on the geometry of the projection
3834: $\pi_{\!\!\scrj}:\scrm \to \scrj$ at $p=[u,h(t)]$, and not on the particular choice
3835: of a transversal map $h:I \to \scrj$. In more detail, the situation is as
3836: follows.
3837:
3838:
3839: \smallskip
3840: First, since $\scrm$ is $C^\ell$-smooth with $\ell\ge2$, the map $\pi_{\!\!\scrj}:
3841: \scrm \to \scrj$ defines a $C^1$-smooth homomorphism of Banach bundles $d\pi_{\!\!
3842: \scrj}: T\scrm \to \pi_{\!\!\scrj}^*(T\scrj)$. \refcorol{cor2.2.2} relates
3843: the (co)kernel of $d\pi_{\!\!\scrj}$ for a given $[u,J] \in \scrm$ with $\sfh^i(S,
3844: \scrn_u)$, and \lemma{lem1.3.1} provides a well-defined bilinear map
3845: \begin{equation}
3846: \Phi=\Phi_{[u,J]} \deff \nabla d \pi_{\!\!\scrj}:
3847: T_{[u,J]}\scrm \times \sfh^0(S, \scrn_u)
3848: \to \sfh^1(S, \scrn_u).
3849: \eqqno(3.1.1)
3850: \end{equation}
3851:
3852: The situation remains essentially the same if we consider a relative moduli
3853: space $\scrm_h= Y\times_h\scrm$ with a $C^\ell$-smooth map $h: Y \to \scrj$
3854: transversal to $\pi_{\!\!\scrj}$. Indeed, one can easily see that for the natural
3855: projection $\pi_h: \scrm_h \to Y$ and a point $[u,y] \in \scrm_h$ with $h(y)
3856: \ddef J$ one has the natural isomorphisms
3857: \begin{equation}
3858: \matrix%\format\r\;&\c\;&\r\;&\c&\l\\
3859: \ker(d\pi_h :T_{[u,y]}\scrm_h \to T_yY)
3860: &\cong& \ker(d\pi_{\!\!\scrj} :T_{[u,J]}\scrm \to T_J\scrj)
3861: &\cong \sfh^0_D(S, \scrn_u),
3862: \cr
3863: \coker(d\pi_h :T_{[u,y]}\scrm_h \to T_yY)
3864: &\cong& \coker(d\pi_{\!\!\scrj} :T_{[u,J]}\scrm \to T_J\scrj)
3865: &\cong \sfh^1(S, \scrn_u).
3866: \endmatrix
3867: \eqqno(3.1.2)
3868: \end{equation}
3869: Further, the relation between $\Phi=\nabla d \pi_{\!\!\scrj}$ and $\nabla d \pi_h$
3870: is given by the following
3871:
3872: \newlemma{lem3.1.1}
3873: \sli The isomorphism $\coker(d\pi_h)\cong \sfh^1(S,
3874: \scrn_u)$ is induced by composition $T_yY \buildrel dh \over \lrar T_J\scrj
3875: \buildrel \barr \Psi_{u,J} \over {\relbar\!\!\relbar\!\!\lrar}
3876: \sfh^1(S, \scrn_u)$.
3877:
3878: \smallskip
3879: \baselineskip=14pt
3880: \slii The bilinear map $\nabla d \pi_h: T_{[u,y]}\scrm_h \times \sfh^0_D(S,
3881: \scrn_u) \to \sfh^1(S, \scrn_u)$ is induced by the composition $T_{[u,y]}
3882: \scrm_h \hook T_{[u,J]}\scrm \oplus T_yY \twoheadrightarrow T_{[u,J]}\scrm$
3883: and the bilinear map $\Phi: T_{[u,J]}\scrm \times \sfh^0_D(S, \scrn_u) \to
3884: \sfh^1(S, \scrn_u)$.
3885: \end{lem}
3886:
3887:
3888: \medskip
3889: Summing up, we obtain the following situation in the most important case $Y=I$.
3890:
3891: \newlemma{lem3.1.2}
3892: For a map $h: I \to \scrj$ transversal to $\pi_J$,
3893: the singular points of the projection $\pi_h: \scrm_h \to I$ are exactly those
3894: $[u,t] \in \scrm_h$ for which $\sfh^1(S, \scrn_u)=\rr$.
3895:
3896: For such $[u,t] \in \scrm_h$ with $J\deff h(t)$ one has the equality
3897: $T_{[u,t]} \scrm_h = \sfh^0_D(S, \scrn_u)$ and the isomorphism $\barr\Psi
3898: _{[u,J]}: T_tI \buildrel \cong \over \lrar \sfh^1(S, \scrn_u)$. Moreover,
3899: the quadratic form $\Phi_{[u,J]}: \sfh^0_D(S, \scrn_u)\to \sfh^1(S, \scrn_u)$
3900: equals to the composition of the quadratic form $\nabla d\pi_h: T_{[u,t]}
3901: \scrm_h \to T_tI$ with $\barr\Psi_{[u,J]}: T_tI \to \sfh^1(S, \scrn_u)$.
3902: \end{lem}
3903:
3904:
3905: \newcorol{cor3.1.3} The nullity, rank and signature of $\Phi_{[u,J]}$ and
3906: $\nabla d\pi_h$ coincide.
3907: \end{corol}
3908:
3909: \newdefi{def3.1.1} Let $Q$ be a quadratic form defined on a
3910: (finite-dimen\-sional) vector space $V$ and taking values in a vector space $W$
3911: with $\dimr W=1$. Define the {\sl saddle index of $Q$} by $\sind Q \deff
3912: \min \{ \ind_+ Q, \ind_- Q \}$, where $\ind_\pm Q$ are respectively the
3913: positive and negative indices of $Q$ \wrt some (in fact, any) orientation of
3914: $W$. For a critical point $[u,J]\in \scrm$ with $\sfh^1(S, \scrn_u) \cong \rr$,
3915: call $\sind \Phi_{[u,J]}$ the {\sl saddle index of $[u,J]$}. A point
3916: $[u,J]\in \scrm$ is a {\sl saddle point} of the moduli space $\scrm$ \iff
3917: $\sind \Phi_{[u,J]}$ is strictly positive.
3918: \end{defi}
3919:
3920:
3921:
3922: \newsubsection[3.2]{Second variation of the $\dbar$-equation}
3923: To find saddle points of $\scrm$ we need to find an explicit formula for the
3924: form $\Phi$ in \eqqref(3.1.1). Note that, since the space $\scrp$ appears as
3925: the zero-set of the $\dbar$-equation \eqqref(1.1.1), the description of the
3926: tangent space $T\scrp$ is given by the variation of the $\dbar$-equation.
3927: Similarly, we show that the form $\Phi$ is essentially the part of the second
3928: variation of the $\dbar$-equation invariant \wrt the choice of a connection
3929: being used.
3930:
3931: \smallskip
3932: Let $[u,J]\in\scrm$ be represented by $(u,J_S,J)\in\whcalm$. Recall the
3933: description of $T_{(u,J_S,J)}\whcalm$ given in \eqqref(2.2.4). Moreover, since
3934: $du$ is non-vanishing at a generic point, $\dot J_S$ is determined by $v$ and
3935: $\dot J$. Note that the tangent space to an orbit $\bfg\cdot(u,J) \subset\whcalm$
3936: can be identified with $du(\sfh^0(S,TS)) \subset \scre_{(u,J_S,J)}$. This
3937: defines a subbundle of $\scre$ which we also denote by $du(\sfh^0(S,TS))$.
3938: Thus we obtain the isomorphism
3939: \begin{equation}
3940: T_{[u,J]}\scrm\cong T_{(u,J_S,J)}\whcalm/du(\sfh^0(S,TS)).
3941: \eqqno(3.2.1)
3942: \end{equation}
3943: Explicitly, the tangent space $T_{[u,J]}\scrm$ consists of triples
3944: $([v], \dot J_S, \dot J)$ for which $(v, \dot J_S,\dot J) \in T_{(u,J_S,J)}
3945: \whcalm$ and $[v]$ is the equivalence class $v + du\bigl(\sfh^0(S,TS) \bigr)$.
3946:
3947: \newdefi{def3.2.1} Set
3948: \begin{equation}
3949: \wh\scre_{(u,J_S,J)} \deff
3950: \bigl(\scre_{(u,J_S,J)} /du(\sfh^0(S,TS) \bigr) \oplus \sfh^1(S,TS).
3951: \eqqno(3.2.2)
3952: \end{equation}
3953: Recall the canonical isomorphism $T_{J_S}\ttt_g \cong \sfh^1(S,TS)$ given
3954: by \eqqref(2.2.2). For $(u,J_S,J)\allowbreak\in \whcalm$ define the operator
3955: \begin{equation}
3956: \wh D=\wh D_{u,J}: \wh\scre_{(u,J_S,J)} \to \scre'_{(u,J_S,J)}
3957: \qquad
3958: \wh D ([v], [I_S]) \deff Dv + J\scirc du \scirc I_S.
3959: \eqqno(3.2.3)
3960: \end{equation}
3961: \end{defi}
3962:
3963:
3964: \newlemma{lem3.2.1}
3965: Formula \eqqref(3.2.2) defines a $C^\ell$-smooth Banach bundle $\wh\scre$
3966: over $\scrm$ with the fiber $\wh\scre_{(u,J_S,J)}$
3967: over $[u,J]\in \scrm$. The tangent bundle $T\scrm$ can be included in the
3968: following exact sequence of bundles over $\scrm$
3969: \begin{equation}
3970: 0\to T\scrm \buildrel \alpha \over\lrar
3971: \wh\scre \oplus \pi_{\!\!\scrj}^*T\scrj
3972: \buildrel \beta \over\lrar \scre' \to0,
3973: \eqqno(3.2.4)
3974: \end{equation}
3975: where the homomorphisms $\alpha=(\alpha_1,\alpha_2)$ and $\beta=(\beta_1,
3976: \beta_2)$ are given by
3977: \begin{equation}
3978: \matrix%\format \r& \;\c\;& \l \\
3979: \alpha_1([v],\dot J_S, \dot J) &\deff&
3980: ([v], [\dot J_S]) \in \wh\scre=
3981: \scre/du\bigl(\sfh^0(S,TS)\bigr)\bigoplus \sfh^1(S,TS)
3982: \cr
3983: \alpha_2 &\deff& d\pi_{\!\!\scrj}: T_{[u,J]}\scrm \to T_J\scrj
3984: \cr
3985: \beta_1 &\deff& \wh D_{u,J}: \wh\scre_{(u,J_S,J)}\to\scre'_{(u,J_S,J)}
3986: \cr
3987: \beta_2 &\deff& \Psi_{u,J}: T_J\scrj\to\scre'_{(u,J_S,J)}
3988: \endmatrix
3989: \end{equation}
3990: \end{lem}
3991:
3992: \proof It is easy to show that $\sfh^0(S,TS)$ and $\sfh^1(S,TS)$ can
3993: be considered as smooth bundles over $\whcalm$ equipped with the natural
3994: $\bfg$-action. Then $du$ defines a $\bfg$-equivariant homomorphism between the
3995: bundles $\sfh^0(S, TS)$ and $\scre$. Hence, using formula \eqqref(3.2.2), we
3996: can construct a bundle $\wh\scre_{\whcalm}$ over $\whcalm$ with the induced
3997: $\bfg$-action. By \lemma{lem2.2.1}\.\slip, this is equivalent to the first
3998: assertion of the lemma.
3999:
4000: The exactness of \eqqref(3.2.4) follows from relations \eqqref(2.2.4) and
4001: (\ref{eq3.2.1}--\ref{eq3.2.3}).
4002: \qed
4003:
4004: \smallskip
4005: \newlemma{lem3.2.2}
4006: The homomorphisms $\alpha_1$ and $\beta_2$ yield isomorphisms
4007: \begin{equation}
4008: \sfh^0_D(S, \scrn_u) \cong \ker \alpha_2 \buildrel
4009: \alpha_1 \over \cong \ker \beta_1
4010: \quad\text{and}\quad
4011: \sfh^1_D(S, \scrn_u) \cong \coker \alpha_2 \buildrel
4012: \beta_2 \over \cong \coker \beta_1,
4013: \eqqno(3.2.5)
4014: \end{equation}
4015: inducing the identity
4016: \begin{equation}
4017: \Phi_{u,J}= -\nabla \wh D: T_{[u,J]}\scrm \times \sfh^0_D(S, \scrn_u)
4018: \to \sfh^1_D(S, \scrn_u).
4019: \eqqno(3.2.6)
4020: \end{equation}
4021: \end{lem}
4022:
4023:
4024: \proof The isomorphisms \eqqref(3.2.5) follow from definitions and
4025: \refcorol{cor2.2.2}. Moreover, we can identify $\sfh^0_D(S, \scrn_u)$ with
4026: $\ker\bigl( \wh D_{u,J}: \wh\scre_{(u, J_S, J)} \to \scre'_{(u, J_S, J)}
4027: \bigr)$.
4028:
4029: Let $i: \sfh^0_D(S, \scrn_u) \to T_{[u,J]}\scrm$ and $p:\scre'_{(u,J_S,J)} \to
4030: \sfh^1_D(S, \scrn_u)$ denote the corresponding inclusion and projection. Fix
4031: some connections on $T\scrm$, $\pi^*_{\!\!\scrj} T\scrj$, $\wh\scre$, and $\scre'$,
4032: and denote all of them simply by $\nabla$. Covariant differentiation of the
4033: relation $\beta_1 \scirc \alpha_1 + \beta_2 \scirc \alpha_2=0$ gives
4034: \begin{equation}
4035: \nabla\beta_1 \scirc \alpha_1 + \nabla\beta_2 \scirc \alpha_2 +
4036: \beta_1 \scirc \nabla\alpha_1 + \beta_2 \scirc \nabla\alpha_2 = 0,
4037: \end{equation}
4038: which together with $\alpha_2\scirc i=0$ and $p\scirc \beta_1=0$ yields
4039: \begin{equation}
4040: p\scirc \nabla\beta_1 \scirc \alpha_1\scirc i=
4041: - p\scirc \beta_2 \scirc \nabla \alpha_2\scirc i.
4042: %\eqNno\square
4043: \end{equation}
4044: \qed
4045:
4046: \newdefi{def3.2.2} Using the isomorphisms $\ker \bigl( \wh D_{u,J}: \wh
4047: \scre_{(u,J_S,J)} \to \scre'_{(u,J_S,J)} \bigr) \cong \sfh^0_D(S, \scrn_u)$
4048: from \eqqref(3.2.5) and $\sfh^1(S, TS) \cong T_{J_S} \ttt_g$ from \eqqref(2.2.2),
4049: redefine
4050: \begin{equation}
4051: \sfh^0_D(S, \scrn_u) \deff \bigl\{\, ([v], I_S)\in \wh\scre_{(u,J_S,J)}
4052: \oplus T_{J_S} \ttt_g\;:\; Dv + J \scirc du \scirc I_S =0 \;\bigr\}.
4053: \eqqno(3.2.7)
4054: \end{equation}
4055: Then the projection $\sfh^0_D(S, \scrn_u) \to \sfh^0_D(S, N_u)$ is given by
4056: the formula $([v], I_S) \mapsto \pr_N(v)$ with $\pr_N: E_u \to N_u$
4057: defined by \eqqref(1.5.2).
4058: \end{defi}
4059:
4060: \smallskip
4061: Now assume that some symmetric connections on $TX$ and $TS$ are fixed.
4062: They induce connections on $\scre$ and $\scre'$, on the tangent bundles
4063: $T\whcalm$ and $T\scrm$, and so on. We shall use the same notation $\nabla$
4064: for all these connections. Further, denote by $R^X(\cdot,\cdot;\cdot)$ the
4065: curvature operator of the connection $\nabla$ on $X$.
4066:
4067: \newlemma{lem3.2.3}
4068: For $([v], \dot J_S, \dot J) \in T_{[u,J]}\scrm$
4069: and $([w], I_S) \in \sfh^0_D(S, \scrn_u) \subset \wh\scre_{(u, J_S, J)}\!\!$
4070:
4071: \smallskip
4072: $\qquad
4073: \bigl(\nabla_{([v],\dot J_S, \dot J)}\wh D\bigr) ([w],I_S)=
4074: \term1{R^X(v,du ;w)} + \term2{J \scirc R^X(v, du\scirc J_S;w)} +
4075: $
4076: \lineeqqno(3.2.8)
4077:
4078: \smallskip
4079: $\qquad
4080: + \term3{\nabla_vJ\scirc \nabla w \scirc J_S} +
4081: \term4{\nabla^2_{v,w}J \scirc du \scirc J_S} +
4082: \term5{\nabla_wJ\scirc \nabla v \scirc J_S} +
4083: \term6{\dot J\scirc \nabla w \scirc J_S} +
4084: $
4085:
4086: \smallskip
4087: $\qquad
4088: + \term7{\nabla_w\dot J\scirc du \scirc J_S} +
4089: \term8{J\scirc \nabla w \scirc \dot J_S} +
4090: \term9{\nabla_wJ\scirc du \scirc \dot J_S} +
4091: \term{10}{\nabla_vJ \scirc du \scirc I_S} +
4092: $
4093:
4094: \smallskip
4095: $\qquad
4096: + \term{11}{J \scirc \nabla v \scirc I_S} +
4097: \term{12}{\dot J \scirc du \scirc I_S}.
4098: $
4099: \end{lem}
4100:
4101:
4102: \state Remark. The numerical subscripts on the various terms are for future
4103: reference.
4104: %The numeration of particular terms is done for convenience.
4105:
4106: \proof Consider the bundle $\wt\scre\deff \scre \oplus \sfh^1(S,TS)$
4107: over $\whcalm$ with the bundle homomorphism $\wt D: \wt\scre \to \scre'$,
4108: $\wt D(w, [I_S]) \deff Dw + J \scirc du \scirc I_S$. We claim that for the
4109: covariant derivative $\bigl(\nabla_{([v],\dot J_S, \dot J)}\wt D\bigr)(w,
4110: I_S)$ we obtain the same expression as in the statement of the lemma. Obviously,
4111: this would imply the lemma.
4112:
4113: The only nontrivial point here is to compute the derivative of the operator of
4114: covariant differentiation $\nabla^\op\deff \nabla: L^{1,p}(S, u^*TX) \to L^p(S,
4115: u^*TX \otimes T^*S)$ in the direction given by some $v \in T_u L^{1,p}(S, X)=
4116: L^{1,p}(S, u^*TX)$. To do this, we fix a smooth vector field $\xi$ on $S$ and
4117: a local section $\mbfw$ of $\scre$. Then $\bigl( \nabla^\op\mbfw \bigr) (\xi)=
4118: \nabla_\xi \mbfw$ is a local section of a Banach bundle with the fiber $L^p(S,
4119: u^*TX)$ over $u\in L^{1,p}(S, X)$.
4120:
4121: Differentiation in the direction $v$ yields
4122: \begin{equation}
4123: \nabla _v \bigl( \nabla^\op \mbfw \bigr) (\xi)
4124: = \nabla^2 _{v,\xi} \mbfw=
4125: R^X(v, du(\xi); \mbfw) + \nabla^2_{\xi,v}\mbfw =
4126: \end{equation}
4127: \begin{equation}
4128: R^X(v, du; \mbfw)(\xi) + \bigl(\nabla^\op_\xi(\nabla \mbfw) \bigr)(v).
4129: \end{equation}
4130: Thus we obtain the formula $\nabla_v (\nabla^\op)= R^X(v, du; \cdot)$. Besides,
4131: we have the relation $\nabla_v du= \nabla v$, which was already used for
4132: deriving \eqqref(1.3.3) from \eqqref(1.1.1). Now, the proof of the lemma
4133: can be completed by explicit calculations. \qed
4134:
4135: Using \eqqref(3.2.8) we can describe in more detail the structure of $\pi
4136: _{\!\!\scrj} :\scrm \to \scrj$ at critical points with $\sfh^1_D(S, N_u) \cong \rr$.
4137: Note that the term $[4]$ in \eqqref(3.2.8) is the only one that depends on
4138: second order derivatives of $J$. Further, the operator $D=D_{u,J}$ is also
4139: independent of second order derivatives. Thus, deforming $J$ and preserving
4140: the order one jet $j^1\!J|_{u(S)}$, the map $u:S \to X$ remains $J$-holomorphic
4141: with same the $D$-cohomology groups $\sfh^i(S, \scrn_u)$. The result of such
4142: changes of $J$ is given by
4143:
4144: \newlemma{lem3.2.4} Let $[u,J] \in \scrm$ with $\sfh^1_D(S, N_u)\cong
4145: \rr$ and a quadratic form $\ti \Phi: \sfh^0_D(S, N_u)\to \sfh^1_D(S, N_u)$ be
4146: given. Then there exists a $C^1$-small perturbation $\ti J \in \scrj$ of $J$
4147: such that $j^1\! J|_{u(S)} = j^1\! \ti J|_{u(S)}$ and the restriction of
4148: $\Phi_{u, \ti J}$ to $\sfh^0_D(S, N_u)$ equals the given $\ti \Phi$.
4149: Moreover, such a perturbation $\ti J$ of $J$ can be realized in an arbitrarily
4150: small neighborhood of a given point $x \in u(S)$.
4151: \end{lem}
4152:
4153:
4154: \proof Let $U\subset X$ be a neighborhood of the given $x$. Find
4155: $U' \subset U$ such that $u\inv(U') \not= \emptyset$ and $u$ is an imbedding
4156: on $u\inv(U')$. Obviously, it is sufficient to find an appropriate jet $j^2\!
4157: \ti J|_{u(S)}$ which differs from $j^2\! J|_{u(S)}$ only in $U'\cap u(S)$.
4158: Then $j^2\! \ti J|_{u(S)}$ can be extended to $\ti J$ with the desired
4159: properties.
4160:
4161: Covariant differentiation of the identity $J^2 = -\id$ gives the relations
4162: $\nabla_v J \scirc J + J \scirc \nabla_v J=0$ and
4163: \begin{equation}
4164: \nabla^2 _{v_1,v_2}J \scirc J + \nabla_{v_1} J \scirc \nabla_{v_2} J +
4165: \nabla_{v_2} J \scirc \nabla_{v_1}J + J \scirc \nabla^2 _{v_1,v_2}J =0,
4166: \qquad v_1,v_2 \in T_xX.
4167: \end{equation}
4168: Consequently, we have the following description of the possible choice for
4169: $j^2\! \ti J|_{u(S)}$ with $j^1\! \ti J|_{u(S)}= j^1\! J|_{u(S)}$. The tensor
4170: field $u(S) \ni x \mapsto \Theta_x$ defined by
4171: \begin{equation}
4172: v_1, v_2, w \in T_xX \mapsto \Theta_x(v_1, v_2; w)
4173: \deff \nabla^2 _{v_1,v_2}(\ti J -J) (w)\in T_xX
4174: \end{equation}
4175: must be supported in $U'$, symmetric%
4176: \footnote{\.Obviously, $\nabla^2 _{v_1,v_2}(\ti J -J)- \nabla^2 _{v_2,
4177: v_1} (\ti J -J)$ can be expressed via $\ti J -J$ and the curvature tensor
4178: $R^X(\cdot,\cdot; \cdot)$ of $\nabla$. But $\ti J -J$ vanishes on $u(S)$.}
4179: %
4180: in $v_1$ and $v_2$, $J$-antilinear in $w$, and zero for $v_1,v_2\in T_x(u(S))
4181: \subset T_xX$. Vice versa, any tensor field $\Theta$ with these properties has
4182: the form $\Theta(v_1, v_2; w) = \nabla^2 _{v_1,v_2}(\ti J -J) (w)$ for an
4183: appropriate $\ti J$ with $j^1\! \ti J|_{u(S)}= j^1\! J|_{u(S)}$.
4184:
4185: The condition that $\Theta_x(v_1, v_2; w)$ vanishes for $v_1, v_2\in T_x(u(S)
4186: )$ means that for $x\in U'\cap u(S)$ we can consider $\Theta_x(v_1, v_2;w)$
4187: as a tensor with arguments $v_1, v_2$ varying in the normal bundle $N_u$.
4188: Now, for $\ti J$ as above, $v\in \sfh^0_D(S, N_u)$, and $\psi \in \sfh^0_D(S,
4189: N_u \otimes K_S) \allowbreak \cong \sfh^1(S, N_u)^*$ we obtain the relation
4190: \begin{equation}
4191: \la\psi, \Phi_{u,\ti J}(v,v) \ra = \la\psi, \Phi_{u,J}(v,v) \ra +
4192: \re \int_S \psi \scirc \Theta(v,v; du).
4193: \end{equation}
4194: Finally, observe that any quadratic form on a {\sl finite dimensional}
4195: space $\sfh^0_D(S, N_u)$ can be realized as $\re \int_S \psi \scirc \Theta(v,v;
4196: du)$ with $\Theta$ satisfying the conditions stated above. \qed
4197:
4198:
4199:
4200: \newsubsection[3.3]{Second variation at cusp-curves}
4201: Our aim in this paragraph is to find conditions ensuring that a critical
4202: point $[u,J] \in \scrm$ with $\sfh^1(S, N_u)\cong \rr$ is a saddle point.
4203: \lemma{lem3.2.4} shows that such critical points with $\scrn\sing_u \cong 0$
4204: are ``hopeless'' from this point of view. Hence, we need to understand in more
4205: detail the structure of the bilinear operator $\Phi$ on the component $\sfh^0
4206: (S, \scrn_u \sing) \subset \sfh^0_D(S, \scrn_u)$.
4207:
4208: %%\newdefi{def3.3.1}
4209: Recall that by the definition of the normal sheaf the stalk $(\scrn_u\sing)_z$
4210: at $z \in S$ is non-trivial exactly when $z$ is a cusp-point of $u: S \to X$
4211: and in this case $\dimc (\scrn_u\sing)_z = \ord_z du$. Thus we want to
4212: understand the structure of the moduli space at critical points corresponding
4213: to cusp-curves. The following two lemmas contain technical results needed for
4214: this purpose. Recall that the holomorphic line bundle $\scro([A])$ was
4215: introduced in \refdefi{def1.5.1}. We maintain the notation $\nabla$ and
4216: $R^X( \cdot{,} \cdot{;} \cdot)$ from \lemma{lem3.2.3}. In particular, we
4217: have $\nabla_\xi J= \nabla_{du(\xi )}J$, $\nabla^2_{\xi, \eta}v - \nabla^2
4218: _{\eta,\xi} v = R^X(du(\xi), du(\eta); v)$, and other similar relations.
4219: Further, we assume that $\nabla J_S=0$.
4220:
4221:
4222: \newlemma{lem3.3.1} An element $([w], I_S) \in \sfh^0_D(S, \scrn_u)$
4223: lies in $\sfh^0(S, \scrn\sing_u)$ \iff $w=du(\ti w)$ for some $\ti w \in
4224: L^{1,p}_\loc (S\bs \supp(\scrn_u\sing), TS)$ that extends to
4225: $\ti w \in L^{1,p}(S, TS \otimes \scro([A])\mkern .5mu)$.
4226:
4227: \smallskip
4228: In this case, outside the zero-set of $du$ one has
4229: \begin{equation}
4230: \dbar \ti w \equiv \nabla\ti w +J_S \scirc \nabla\ti w \scirc J_S
4231: = - J_S \scirc I_S
4232: \eqqno(3.3.1)
4233: \end{equation}
4234: and
4235: \begin{align}
4236: 0\equiv &\nabla_{\ti w}(D_{u,J}v + \dot J \scirc du \scirc J_S
4237: + J \scirc du \scirc \dot J_S) =\term{D}{D_{u,J}(\nabla_{\ti w}v) }+
4238: \eqqno(3.3.2) \\
4239: &\term{1'}{R^X(w,du ;v)} + \term{2'}{J \scirc R^X(w, du\scirc J_S;v)} +
4240: \term5{\nabla_wJ\scirc \nabla v \scirc J_S} +
4241: \term{4'}{\nabla^2_{w,v}J \scirc du \scirc J_S} +
4242: \notag\\
4243: &\term{3'}{\nabla_vJ\scirc \nabla_{\ti w}du \scirc J_S} +
4244: \term{6'}{\dot J\scirc \nabla_{\ti w}du \scirc J_S} +
4245: \term7{\nabla_w\dot J\scirc du \scirc J_S} +
4246: \term{8'}{J\scirc \nabla_{\ti w}du \scirc \dot J_S} +
4247: \notag\\
4248: &\term9{\nabla_wJ\scirc du \scirc \dot J_S} +
4249: \term{13}{J\scirc du \scirc \nabla_{\ti w} \dot J_S} -
4250: \term{14}{\nabla v \scirc \nabla \ti w} -
4251: \term{15}{J \scirc \nabla v \scirc \nabla \ti w \scirc J_S}.
4252: \notag
4253: \end{align}
4254: \end{lem}
4255:
4256:
4257: \proof Relation \eqqref(3.3.1) follows from the equality
4258: %\begin{equation}
4259: \begin{align*}
4260: 0&= D_{u,J}(du(\ti w))+ J \scirc du \scirc I_S
4261: = du(\dbar\ti w)+ du \scirc J_S \scirc I_S =
4262: \cr
4263: &=du \Bigl( ( \nabla \ti w + J_S \scirc \nabla \ti w\scirc J_S)
4264: + J_S \scirc I_S \Bigr).
4265: \end{align*}
4266: %\end{equation}
4267: Using $J_S^2 = -\id$ we can write the relation in the form
4268: $I_S = J_S \scirc \nabla \ti w - \nabla \ti w \scirc J_S$.
4269:
4270: \medskip
4271: To show \eqqref(3.3.2) we start with the computation of
4272: $D_{u,J}(\nabla _{\ti w} v)$:
4273: \begin{gather}
4274: \eqqno(3.3.2a)
4275: D_{u,J}(\nabla_{\ti w}v) = \nabla(\nabla_{\ti w}v) +
4276: J \scirc \nabla(\nabla_w v) \scirc J_S +
4277: \nabla J({\nabla_w v}, du \scirc J_S)=
4278: \\
4279: \notag
4280: \term{16}{\nabla^2_{\cdot, \ti w}v} \!+
4281: \term{14}{\nabla v \scirc \nabla \ti w \mathstrut} \!+
4282: \term{17}{J \scirc \nabla^2_{\cdot,\ti w}v \scirc J_S} \!+
4283: \term{15}{J \scirc \nabla v \scirc \nabla \ti w \scirc J_S \mathstrut}
4284: \!+
4285: \term{18}{\nabla J({\nabla_{\ti w} v}, du \scirc J_S)} .
4286: \end{gather}
4287: \pagebreak[3]
4288: Similarly,
4289: \begin{align}
4290: \eqqno(3.3.2b)
4291: &\nabla_{\ti w}(D_{u,J}v + \dot J \scirc du \scirc J_S
4292: + J \scirc du \scirc \dot J_S)=
4293: \\
4294: &\nabla_{\ti w}( \nabla v + J\scirc \nabla v \scirc J_S +
4295: \nabla_v J \scirc du \scirc J_S +
4296: \dot J \scirc du \scirc J_S + J \scirc du \scirc \dot J_S)=
4297: \notag\\
4298: &\term{16'}{\nabla^2_{\ti w,\cdot}v} +
4299: \term{17'}{J \scirc \nabla^2_{\ti w,\cdot}v \scirc J_S} +
4300: \term{5}{\nabla_wJ \scirc \nabla v \scirc J_S} +
4301: \term{4'}{\nabla^2_{w,v}J \scirc du \scirc J_S} +
4302: \notag\\
4303: &\term{18}{\nabla J(\nabla_{\ti w}v; du \scirc J_S)} +
4304: \term{3'}{\nabla_v J \scirc \nabla_{\ti w}du \scirc J_S} +
4305: \term{7}{\nabla_w\dot J \scirc du \scirc J_S} +
4306: \notag\\
4307: &\term{6'}{\dot J \scirc \nabla_{\ti w}du \scirc J_S} +
4308: \term{9}{\nabla_w J \scirc du \scirc \dot J_S} +
4309: \term{8'}{J \scirc \nabla_{\ti w}du \scirc \dot J_S} +
4310: \term{13}{J \scirc du \scirc \nabla_{\ti w}\dot J_S}.
4311: \notag
4312: \end{align}
4313: Comparing the terms [16] and $[16']$, it follows that $\nabla^2_{\cdot, \ti w}v -
4314: \nabla^2_{\ti w,\cdot}v = R^X(du, w;v)$. A similar relation holds for the
4315: terms [17] and $[17']$. The equality \eqqref(3.3.2) of the lemma is obtained
4316: by subtracting \eqqref(3.3.2a) from \eqqref(3.3.2b).
4317: %
4318: \qed
4319:
4320: \newlemma{lem3.3.2}
4321: \sli In the situation of \lemma{lem3.3.1}, let $z^*
4322: \in S$ be a cusp-point. Consider $\ti w$ as a section of $TS$ with poles. Set
4323: $k \deff \ord_{z^*}du = \dimc (\scrn_u\sing)_{z^*}$ and choose a local complex
4324: coordinate $z$ on $S$ centered in $z^*$. Fix additionally $([v], \dot J_S, \dot
4325: J) \in T_{[u, J]} \scrm$ and $\psi \in \sfh^0_D(S, N_u^* \otimes K_S)$. Then
4326: locally in a neighborhood of $z^*$
4327: \begin{equation}
4328: \matrix%\format \r &\;\c \;& \l &\; \l & \; \l & \; \l\\
4329: % w(z) = t0 t1 tk t*
4330: z^k \cdot \ti w(z) &=&
4331: w_0 & + z\cdot w_1 &+ \cdots + z^k\cdot w_k &+ z^k\cdot w^*(z),
4332: \cr
4333: v(z) &=&
4334: v_0 & + z\cdot v_1 &+ \cdots +z^k \cdot v_k &+ z^k \cdot v^*(z),
4335: \cr
4336: \psi(z) &=&
4337: \psi_0 & + z\cdot \psi_1 &+ \cdots +z^k \cdot \psi_k &
4338: + z^k \cdot \psi^*(z),
4339: \endmatrix
4340: \eqqno(3.3.2a1)
4341: \end{equation}
4342: where $w^*(z)$, $v^*(z)$, and $\psi^*(z)$ are $L^{1,p}$-smooth local sections
4343: of the corresponding bundles vanishing at $z=0$.
4344:
4345: \slii The polynomials in \eqqref(3.3.2a1) can be considered as the order
4346: $k$ jets of
4347: the following local holomorphic objects: a section of $TS$ for $w_0 + \cdots +
4348: z^k\cdot w_k$, a $(E_u)_{z^*}$-valued function for $v_0+ \cdots +z^k \cdot
4349: v_k$, and resp.\ $(N_u^*)_{z^*}$-valued $(0,1)$-form for $\psi_0 +\cdots +z^k
4350: \cdot \psi_k$. In particular, the coefficients can be considered as
4351: well-defined elements
4352: \begin{equation}
4353: \matrix%\format \r & \;\c\;& \r &\; \c\; & \; \l \\
4354: % w_i = \na \in T^*S
4355: w_i &=& \left({\d \over \d z}\right)^i(z^k \ti w(z))\ogran_{z=0}
4356: &\in& (T^*_{z^*} S)^{\otimes i-k} \otimes T_{z^*} S,
4357: \cr
4358: v_i &=& \nabla^i(v(z))\ogran_{z=0} &\in&
4359: (T^*_{z^*} S)^{\otimes i} \otimes (E_u)_{z^*},
4360: \cr
4361: \psi_i &=& \nabla^i(\psi(z))\ogran_{z=0} &\in&
4362: (T^*_{z^*} S)^{\otimes i} \otimes (N^*_u \otimes K_S)_{z^*}.
4363: \cr
4364: \endmatrix
4365: \eqqno(3.3.2b1)
4366: \end{equation}
4367: \end{lem}
4368:
4369:
4370: \proof \sli It follows from \refdefi{def1.5.1} that $du$, considered as a
4371: {\sl holomorphic} section of the bundle $\hom_\cc(TS, E_u)$, locally has the
4372: form $du(z)= z^k s(z)$ for some local holomorphic non-vanishing section $s$.
4373: Consequently, $\scro([A]) = \scro( k\cdot [z^*])$ in a neighborhood of $z^*$.
4374: Thus by \lemma{lem3.3.1} $\ti w$ can be locally
4375: represented in the form $\ti w(z) = z^{-k} \cdot \hat w(z)$ for some local
4376: $L^{1,p}_\loc$-smooth section of $TS$. Equation \eqqref(3.3.1) is equivalent
4377: to $\dbar \hat w = -\isl\cdot z^k \cdot I_S$ and implies the estimate $|\dbar
4378: \hat w(z)| \le |z^k|\cdot |I_S(z)|$. Now we use \lemma{lem1.2.3}.
4379:
4380: The same argument applies to $v(z)$ and $\psi(z)$. Indeed, equation
4381: \eqqref(2.2.4) on $v$ and relation \eqqref(1.4.1) imply the inequality
4382: $|\dbar v(z)| \le c\cdot |z^k|$ with some constant $c$. A similar inequality for
4383: $\psi(z)$ follows from \eqqref(1.5.3).
4384:
4385: \smallskip
4386: \slii This part of the lemma can be reformulated in terms of the transformation
4387: of coefficients $w_i$, $v_i$, and $\psi_i$ under the change of a local
4388: holomorphic coordinate $z$ on $S$ and local coordinates on $X$. The claim
4389: concerning the change of $z$ is obvious.
4390:
4391: Considering changes of coordinates on $X$ we make the following observation.
4392: If $\mbfx'=(x'_1, \ldots, x'_{2n})$ and $\mbfx''=(x''_1, \ldots,
4393: x''_{2n})$ are two systems of coordinates on $X$ centered in $u(z^*)$, then
4394: $\mbfx''=L(\mbfx') + Q(\mbfx') + \cdots$, where $L: \rr^{2n} \to
4395: \rr^{2n}$ (resp. $Q: \rr^{2n} \to \rr^{2n}$) is an appropriate linear (resp.\
4396: quadratic) map, and so on. In particular, $\mbfx'' - L(\mbfx')= O(
4397: |\mbfx'|^2)$. Consequently, for local frames $\d_{\mbfx'}= (\d_{x'_1},
4398: \ldots,\d_{x'_{2n}})$ and $\d_{\mbfx''}= (\d_{x''_1}, \ldots,\d_{x''_{2n}}
4399: )$ of $TX$ we obtain the relation $\d_{\mbfx''}(\mbfx') - L^{\sf t}(
4400: \d_{\mbfx''}) (\mbfx') = O(|\mbfx'|)$. Thus, for the pulled-back
4401: frames $u^*\d_{\mbfx'}$ and $u^*\d_{\mbfx''}$ of $E_u$ we have
4402: $u^*\d_{\mbfx'}(z) - L^{\sf t}(u^*\d_{\mbfx''})(z)= O(|z|^{k+1})$.
4403: This implies that the change of local coordinates on $X$ induces only a linear
4404: transformation of the $k$-jet of $v$, \ie the $k$-jet of $v$ behaves like a
4405: $k$-jet of a $T_{u(z^*)}$-valued function. The same argumentations
4406: can be applied to the $k$-jet of $\psi$.
4407: \qed
4408:
4409: \newlemma{lem3.3.3}
4410: For $[u,J] \in \scrm\!\!$, $([w], I) \in \sfh^0_D(S,
4411: \scrn_n\sing)$, $([v], \dot J_S, \dot J) \in T_{[u,J]}\scrm\!\!$, and
4412: $\psi \in \sfh^0_D(S, N_u \otimes K_S) \cong \sfh^1_D(S, N_u)^*$
4413: it follows that %one has the relation
4414: \begin{equation}
4415: \Bigl\la \psi ,
4416: \Phi_{u,J}\bigl( ([v], \dot J_S, \dot J),\; ([w],I_S) \bigr) \Bigr\ra=
4417: \re \res_S(\psi\scirc \nabla_{\ti w}v),
4418: \eqqno(3.3.3)
4419: \end{equation}
4420: where $\res_S(\psi\scirc \nabla_{\ti w}v)$ denotes the residual type sum
4421: \begin{equation}
4422: \res_S(\psi\scirc \nabla_{\ti w}v) \deff
4423: \mathop{\textstyle\sum}\limits_{du(z^*_i)=0}\;
4424: \lim\limits_{\epsi \lrar 0}\;
4425: \int_{|z - z^*_i|=\epsi} \psi\scirc \nabla_{\ti w}v
4426: \eqqno(3.3.4)
4427: \end{equation}
4428: over all cusp-points $z^*_i\in S$ of $u$.
4429:
4430: Moreover, if $\dot J=0$ and $([v], \dot J_S) \in \sfh^0(S, \scrn_u\sing)$,
4431: then $v=du(\ti v)$ with $\ti v \in L^{1, p}(S, TS \otimes \scro([A])\,)$
4432: and
4433: \begin{equation}
4434: \Bigl\la \psi ,
4435: \Phi_{u,J}\bigl( ([v], \dot J_S, 0),\; ([w],I_S) \bigr) \Bigr\ra=
4436: \re \res_S(\psi\scirc \nabla du(\ti w, \ti v)).
4437: \eqqno(3.3.5)
4438: \end{equation}
4439: \end{lem}
4440:
4441:
4442: \proof First, we note that by \lemma{lem3.3.2} the formulas
4443: (\ref{eq3.3.3}--\ref{eq3.3.5}) are well-defined.
4444:
4445: Now, compute the subtraction of \eqqref(3.3.2) from \eqqref(3.2.8). The terms
4446: $[5]$, $[7]$, and $[9]$ cancel. To simplify further terms we use the Bianchi
4447: identity and antisymmetry of $R^X(\cdot {,} \cdot{;}\cdot)$ in the first two
4448: arguments. The difference of terms $[1] +[2] +[4] -[1'] -[2'] -[4']$ is zero:
4449: \begin{equation}
4450: \term1{R^X( v, du; w)} - \term{1'}{R^X( w, du; v)} +
4451: \term2{J \scirc R^X( v, du \scirc J_S; w)} -
4452: \end{equation}
4453: \begin{equation}
4454: \term{2'}{J \scirc R^X( w, du\scirc J_S; v)} +
4455: \term4{\nabla^2_{v,w}J \scirc du \scirc J_S} -
4456: \term{4'}{\nabla^2_{w,v}J \scirc du \scirc J_S} =
4457: \end{equation}
4458: \begin{equation}
4459: = R^X( v, du; w) + R^X( du, w; v) +
4460: J \scirc R^X( v, du \scirc J_S; w) +
4461: \end{equation}
4462: \begin{equation}
4463: J \scirc R^X( du\scirc J_S ,w ; v) +
4464: R^X(v,w; J \scirc du \scirc J_S) -
4465: J \scirc R^X(v,w; du \scirc J_S) =
4466: \end{equation}
4467: \begin{equation}
4468: =R^X( v, du; w) + R^X( du, w; v) + R^X(w,v; du) +
4469: \end{equation}
4470: \begin{equation}
4471: J \scirc R^X( v, du \scirc J_S; w) +
4472: J \scirc R^X( du\scirc J_S, w; v) +
4473: J \scirc R^X( w, v; du \scirc J_S) = 0.
4474: \end{equation}
4475:
4476: \smallskip
4477: In the differences $[3] -[3']$, $[6] -[6']$, and $[8] -[8']$ respectively,
4478: we use the relation
4479: \begin{equation}
4480: \nabla(w)=\nabla(du(\ti w))= \nabla_{\ti w}du + du\scirc
4481: \nabla \ti w.
4482: \eqqno(3.3.6)
4483: \end{equation}
4484: This yields
4485: \begin{equation}
4486: \term3{\nabla_vJ \scirc \nabla w \scirc J_S} -
4487: \term{3'}{\nabla_vJ \scirc \nabla_{\ti w}du \scirc J_S} =
4488: \term{3''}{\nabla_vJ \scirc du \scirc \nabla \ti w \scirc J_S},
4489: \end{equation}
4490: and similar equalities for $[6] -[6']$ and $[8] -[8']$. Thus we obtain
4491: \begin{equation}
4492: \nabla_{([v],\dot J_S, \dot J)}(\wh D)([w],I_S) =
4493: \term{3''}{\nabla_vJ\scirc du\scirc \nabla\ti w \scirc J_S} +
4494: \term{6''}{\dot J\scirc du\scirc \nabla\ti w \scirc J_S} +
4495: \end{equation}
4496: \begin{equation}
4497: \term{8''}{J\scirc du\scirc \nabla\ti w \scirc \dot J_S}
4498: + \term{10}{\nabla_vJ \scirc du \scirc I_S}
4499: + \term{11}{J \scirc \nabla v \scirc I_S}
4500: + \term{12}{\dot J \scirc du \scirc I_S}
4501: \end{equation}
4502: \begin{equation}
4503: - \term{13}{J\scirc du \scirc \nabla_{\ti w} \dot J_S} +
4504: \term{14}{\nabla v \scirc \nabla \ti w} +
4505: \term{15}{J \scirc \nabla v \scirc \nabla \ti w \scirc J_S}
4506: - \term{D}{D_{u,J}(\nabla_{\ti w}v) }.
4507: \end{equation}
4508:
4509: \smallskip
4510: Further simplification uses the relation $I_S = J_S \scirc \nabla \ti w -
4511: \nabla \ti w \scirc J_S$. This gives
4512: \begin{equation}
4513: \term{3''}{\nabla_vJ \scirc du\scirc \nabla\ti w \scirc J_S} +
4514: \term{10}{\nabla_vJ \scirc du \scirc I_S} =
4515: \nabla_vJ\scirc du\scirc \nabla\ti w \scirc J_S +
4516: \end{equation}
4517: \begin{equation}
4518: \nabla_vJ \scirc du \scirc
4519: (J_S \scirc \nabla \ti w - \nabla \ti w \scirc J_S)
4520: = \term{3'''}{\nabla_vJ \scirc du\scirc J_S \scirc \nabla \ti w},
4521: \end{equation}
4522: and similarly,
4523: \begin{equation}
4524: \term{6''}{\dot J\scirc du\scirc \nabla\ti w \scirc J_S} +
4525: \term{12}{\dot J \scirc du \scirc I_S} =
4526: \term{6'''}{\dot J\scirc du\scirc J_S \scirc \nabla \ti w},
4527: \end{equation}
4528: \begin{equation}
4529: \term{11}{J \scirc \nabla v \scirc I_S} +
4530: \term{15}{J \scirc \nabla v \scirc \nabla \ti w \scirc J_S}
4531: = \term{15'}{J \scirc \nabla v \scirc J_S \scirc \nabla \ti w}.
4532: \end{equation}
4533:
4534: \smallskip
4535: Now we put together the terms $[3''']$, $[6''']$, $[14]$, and $[15']$.
4536: Because of the relation
4537: \begin{equation}
4538: \nabla v + J \scirc \nabla v \scirc J_S + \nabla_vJ \scirc du\scirc J_S +
4539: \dot J \scirc du\scirc J_S + J\scirc du\scirc \dot J_S =0
4540: \end{equation}
4541: this yields
4542: \begin{equation}
4543: \term{3'''}{\nabla_vJ \scirc du\scirc J_S \scirc \nabla \ti w}+
4544: \term{6'''}{\dot J\scirc du\scirc J_S \scirc \nabla \ti w} +
4545: \term{14}{\nabla v \scirc \nabla \ti w} +
4546: \term{15'}{J \scirc \nabla v \scirc J_S \scirc \nabla \ti w}=
4547: \end{equation}
4548: \begin{equation}
4549: \bigl( \nabla_vJ \scirc du\scirc J_S +
4550: \dot J\scirc du\scirc J_S +
4551: \nabla v +
4552: J \scirc \nabla v \scirc J_S \bigr)\scirc \nabla \ti w =
4553: - \term{8'''}{J\scirc du\scirc \dot J_S \scirc \nabla \ti w}.
4554: \end{equation}
4555:
4556: \smallskip
4557: Finally, we conclude that outside the zero-set of $du$ one has
4558: \begin{equation}
4559: \nabla_{([v],\dot J_S, \dot J)}(\wh D)([w],I_S) =
4560: \end{equation}
4561: \begin{equation}
4562: J\scirc du\scirc \bigl( \term{8''}{\nabla\ti w \scirc \dot J_S}
4563: - \term{8'''}{\dot J_S \scirc \nabla \ti w}
4564: - \term{13}{\nabla_{\ti w} \dot J_S} \bigr)
4565: - \term{D}{D_{u,J}(\nabla_{\ti w}v) }.
4566: \end{equation}
4567: Note that $\psi \scirc J \scirc du = \psi \scirc du \scirc J_S =0$, since
4568: $\psi$ vanishes on $du(TS)$. Consequently,
4569: \begin{equation}
4570: \Bigl\la \psi , \Phi_{u,J}\bigl( ([v], \dot J_S, \dot J),
4571: \; ([w],I_S) \bigr) \Bigr\ra=
4572: \re \int_S \psi \scirc
4573: \Bigl( - \nabla _{([v], \dot J_S, \dot J)} \wh D\Bigr)([w], I_S)=
4574: \end{equation}
4575: \begin{equation}
4576: \re \lim\limits_{\epsi \to 0} \int_{S \bs \cup \Delta(z_i, \epsi)}
4577: \psi \scirc \Bigl( -\nabla _{([v], \dot J_S, \dot J)} \wh D\Bigr)([w], I_S)
4578: =
4579: \re \lim\limits_{\epsi \to 0} \int_{S \bs \cup \Delta(z_i, \epsi)}
4580: \psi \scirc D_{u,J}(\nabla_{\ti w}v).
4581: \end{equation}
4582: Integrating by parts and using $D \psi=0$ we obtain the desired formula
4583: \eqqref(3.3.3).
4584:
4585: To obtain \eqqref(3.3.5) we use \eqqref(3.3.6) and relation $\psi \scirc du =0$.
4586: \qed
4587:
4588: \smallskip
4589: Now we can describe the structure of $\Phi_{u,J}$ for cusp-curves. Here we
4590: restrict ourselves to the case when $(X,J)$ is an almost complex surface. The
4591: point is that, unlike to the higher dimensional situation, in this dimension
4592: there are {\sl topological} reasons for the existence of cusp-curves, see
4593: \lemma{lem2.3.4}.
4594:
4595: Other than the restriction on dimension, our setting is as follows. $[u,J] \in
4596: \scrm$ is a $J$-holomorphic curve with $\sfh^1(S, N_u) \cong \rr$, $z^*\in S$ a
4597: cuspidal point, $k\deff \ord _{z^*} du$, $z$ a local complex coordinate centered
4598: at $z^*$, $J^*$ a local (integrable) complex structure in a neighborhood of
4599: $u(z^*)$, and $(w^1, w^2)$ a local system of $J^*$-holomorphic coordinates on
4600: $X$ centered at $u(z^*)$. Finally, we fix some {\sl non-zero} $\psi \in
4601: \sfh^0(S, N_u^* \otimes K_S) \cong \sfh^1(S, N_u)^*$ and denote by $\ord_{z^*}
4602: \psi$ the order of vanishing of $\psi$ at $z^*$.
4603:
4604: \newlemma{lem3.3.4}
4605: \sli After a polynomial transformation of the coordinates $(w^1,
4606: \allowbreak
4607: w^2)$, chosen above, the map $u$ will have the form
4608: \begin{equation}
4609: u(z)=\bigr( z^{k+1} P_1(z), z^{k+l +2} P_2(z) \bigl) + z^{2k+1} g(z),
4610: \eqqno(3.3.7)
4611: \end{equation}
4612: such that $0\le l \le k$, $P_1$ is a polynomial of degree $\le k$ with $P_1(0)
4613: \not =0$, $P_2$ is a polynomial of degree $\le k-l-1$, trivial if $l=k$ or with
4614: $P_2(0) \not =0$ otherwise, and $g(z)$ is an $L^{1,p}$-smooth $\cc^2$-valued
4615: function.
4616:
4617: \slii The integers $\ord_{z^*} \psi$ and $l$ do not depend on the particular
4618: choice of coordinates $(w^1, w^2)$ and $\psi\in \sfh^0(S, N_u^* \otimes K_S)$.
4619: For the restriction of $\Phi_{u,J}$ on the stalk $(\scrn_u\sing)_{z^*}
4620: \subset \sfh^0 (S, \scrn\sing_u)$, it follows that
4621: \begin{equation}
4622: \matrix %\format\r& \,\c\,&\r& \,\c\, \cr \\
4623: \ind_+ \bigl( \Phi_{u,J}\ogran_{(\scrn_u\sing)_{z^*} } \bigr)&=&
4624: \ind_- \bigl( \Phi_{u,J}\ogran_{(\scrn_u\sing)_{z^*} } \bigr)=\cr
4625: \sind \bigl( \Phi_{u,J}\ogran_{(\scrn_u\sing)_{z^*} } \bigr)&=&
4626: \max\,( 0, k-l -\ord_{z^*} \psi).
4627: \endmatrix
4628: \eqqno(3.3.8)
4629: \end{equation}
4630:
4631: \sliii If $z^*_1$ and $z^*_2$ are distinct cusp-points of $u:S \to X$,
4632: then the stalks $(\scrn_u\sing)_{z^*_1}$ and $(\scrn_u\sing)_{z^*_2}$ are
4633: $\Phi$-orthogonal, \ie
4634: \begin{equation}
4635: \Phi_{u,J}\bigl( (\scrn_u\sing)_{z^*_1}, (\scrn_u\sing)_{z^*_2} \bigr)=0.
4636: \eqqno(3.3.9)
4637: \end{equation}
4638: \end{lem}
4639:
4640: \proof {\sl Part} \sli follows immediately from \lemma{lem1.2.4}. It
4641: simply says that if $\ord_{z^*}du=k$, then the jet $j^{2k+1}u$ is well-defined
4642: and {\sl holomorphic}, \ie can be represented by a complex polynomial.
4643: Note that the theorem of \cite{Mi-Wh} (see \lemma{lem1.2.1}) says that {\sl
4644: topologically} one can also define higher terms which determine the whole
4645: behavior of $u$ at $z^*$.
4646:
4647: {\sl Part} \sliii can be easily obtained from \eqqref(3.3.4) and
4648: \eqqref(3.3.5). It remains to consider
4649:
4650: {\sl Part \sliip}. First, we observe that the integer $l$ is the secondary cusp
4651: index of $u$ at $z^*$ (see \refdefi{def4.2s.1}). It follows then from the
4652: results of \refsubsection{4.2s} that this integer is well defined and
4653: independent of the choice of $(w^1, w^2)$. The independence of $\ord_{z^*} \psi$ of
4654: the choice of $(w^1, w^2)$ and $\psi$ is obvious.
4655:
4656: Let $J^*$ and $(w^1, w^2)$
4657: be a complex structure and $J^*$-holomorphic coordinates in a neighborhood
4658: of $u(z^*)$, such that $J^*(u(z^*)) = J(u(z^*))$ and $u$ has the local form
4659: \eqqref(3.3.7). Differentiating \eqqref(3.3.7) we see that in the coordinates
4660: $(w^1, w^2)$
4661: \begin{equation}
4662: du(z)=\bigr( z^k P'_1(z), z^{k+l +1} P'_2(z) \bigl) + z^{2k} g'(z),
4663: \eqqno(3.3.10)
4664: \end{equation}
4665: with polynomials
4666: \begin{equation}P'_1(z)= (k+1)P_1(z) + z {\textstyle{d \over dz}} P_1(z)
4667: \quad\text{and}\quad
4668: P'_2(z)= (k+l+2)P_2(z) + z {\textstyle{d \over dz}} P_2(z),
4669: \eqqno(3.3.11)
4670: \end{equation}
4671: of degree $\le k$ and $\le k-l-1$ respectively
4672: and with $g'(z) = (2k+1) g(z) dz + z dg(z)$ being $L^p$-bounded.
4673:
4674: From the definition of the Nijenhuis torsion tensor $N_J$ of $J$ we obtain a
4675: pointwise estimate $|\dbar_J w_\alpha| \le |N_J|$. Further,
4676: \begin{equation}
4677: \dbar(w_\alpha \scirc u) = (dw_\alpha \scirc du)^{(0,1)} =
4678: \dbar_J w_\alpha \scirc du,
4679: \end{equation}
4680: since $u$ is $J$-holomorphic. Consequently, we obtain a pointwise estimate
4681: \begin{equation}
4682: |\dbar(w_\alpha \scirc u)(z)| \le c\cdot |z^k|
4683: \eqqno(3.3.12)
4684: \end{equation}
4685: with some constant $c$. Let $\{ e^*_\alpha\}_{\alpha=1,2}$ be the local
4686: $J^*$-complex frame of $T^*X$ dual to the frame $\{ dw_\alpha\}_{\alpha=1,2}$.
4687: Then there exists a local $J$-complex frame $\{ e_\alpha\}_{\alpha=1,2}$ of
4688: $T^*X$ with pointwise estimates
4689: \begin{equation}
4690: |e^*_\alpha(w) -e_\alpha(w)| \le c\cdot |w|
4691: \quad\text{and}\quad
4692: |\nabla e^*_\alpha(w) - \nabla e_\alpha(w)| \le c,
4693: \eqqno(3.3.13)
4694: \end{equation}
4695: where $|w|^2 = |w_1|^2 + |w_2|^2$ and $c$ is some constant.
4696: Using (\ref{eq3.3.10}--\ref{eq3.3.13}) and the estimates
4697: $|u(z)| \le c\cdot |z^{k+1}|$ and $|du(z)|
4698: \le c\cdot |z^k|$ we conclude that
4699:
4700: {\sl a)\.} $\mbfe_\alpha \deff u^* e_\alpha$ is a local complex frame of
4701: $E_u = u^*TX$ with a pointwise estimate
4702: \begin{equation}
4703: |\dbar_{u,J} \mbfe_\alpha(z)| \le c\cdot |z^k|;
4704: \end{equation}
4705:
4706: {\sl b)\.} $du$, considered as a section of $E_u\otimes T^*S$ with the frame
4707: $\mbfe_\alpha\otimes dz$, has local form \eqqref(3.3.10), possibly with another
4708: $g'(z) \in L^p$. Moreover, since $du$ is a holomorphic section and
4709: $\mbfe_\alpha$ are sufficiently regular, this new $g'(z)$ is $C^1$-smooth.
4710: Further, since $z dg(z)= g'(z) - (2k+1)zg(z) dz$ is continuous and $dg(z) \in
4711: L^p$ with $p>2$, we conclude that $z dg(z)$ vanishes at $z=0$. This gives
4712: an additional relation $g'(0)=0$.
4713:
4714: Differentiating \eqqref(3.3.10) we obtain that in the frame
4715: $\mbfe_\alpha\otimes dz^2$
4716: \begin{equation}
4717: \nabla du(z)=\bigr( z^{k-1} P''_1(z), z^{k+l } P''_2(z) \bigl) +
4718: z^{2k-1} g''(z),
4719: \eqqno(3.3.14)
4720: \end{equation}
4721: with polynomials
4722: \begin{equation}
4723: P''_1(z)= k\,P'_1(z) + z {\textstyle{d \over dz}} P'_1(z)
4724: \quad\text{and}\quad
4725: P''_2(z)= (k+l+1)P_2(z) + z {\textstyle{d \over dz}} P'_2(z),
4726: \eqqno(3.3.15)
4727: \end{equation}
4728: of degree $\le k$ and $\le k-l-1$ respectively and with
4729: \begin{equation}
4730: g''(z) = (2k+1) g'(z) \otimes dz + z \nabla g'(z)
4731: \end{equation}
4732: continuous and vanishing at $z=0$. By our construction, $P''_1(0)= (k+1)k
4733: P_1(0) \not =0$ and $P''_2(0)= (k+l+2)(k+l+1) P_1(0)$ vanishes \iff $l=k$.
4734:
4735: Since the projection $\pr: E_u \to N_u$ is obtained as the quotient \wrt
4736: the image of $z^{-k} du \sim (P'_1(z), z^{l+1} P'_2(z))$, we have the following
4737: form for the composition:
4738: \begin{equation}
4739: \pr \scirc \nabla du(z)= P'''(z) + g'''(z),
4740: \end{equation}
4741: where $P'''(z)$ is a polynomial $P'''(z)$ of degree $\le k-l-1$ given by
4742: the relation
4743: \begin{equation}
4744: z^{k+l}P'''(z) = z^{k+l} P''_2(z) -
4745: {z^{k+l+1} P'_2(z) \cdot z^{k-1} P''_1(z) \over
4746: z^k P'_1(z)} + o(z^{2k-1}).
4747: \end{equation}
4748: In particular, $P'''(0)= (k+l+1)(l+1)P_2(0)$ vanishes \iff $l=k$.
4749:
4750: Denoting $\nu \deff \ord_{z^*} \psi$ we obtain that
4751: \begin{equation}
4752: \psi \scirc \nabla du(z) = a z^{k+l+\nu} + o(z^{k+l+\nu})
4753: \end{equation}
4754: with $a$ vanishing \iff $l=k$. The proof of part \slii of the lemma can be now
4755: finished using the following algebraic result. \qed
4756:
4757: \newlemma{lem3.3.5} For a given polynomial $P(z)= a_0 + a_1 \,z
4758: + \cdots + a_{ k-l-1} z^{ k-l-1}$ with $a_0 \not=0$ and $0\le l <k$
4759: the quadratic form
4760: \begin{equation}
4761: (w_0,\ldots,w_k) \in \cc^{k+1}
4762: \mapsto \re\res_{z=0} \left( { z^{k+l}\, P(z)\,
4763: \bigl(\sum_{i=0}^k w_i\, z^i\bigr)^2 \over z^{2k} } dz \right) \in \rr
4764: \end{equation}
4765: is equivalent to the quadratic form
4766: \begin{equation}
4767: (w_0,\ldots,w_k) \in \cc^{k+1}
4768: \mapsto \re\res_{z=0} \left( { z^{k+l}\, a_0\,
4769: \bigl(\sum_{i=0}^k w_i\, z^i\bigr)^2 \over z^{2k} } dz \right)\in \rr
4770: \end{equation}
4771: and satisfies the index relations
4772: \begin{equation}
4773: \ind_+ Q = \ind_- Q = \sind Q = k-l.
4774: \eqqno(3.3.16)
4775: \end{equation}
4776: \end{lem}
4777:
4778:
4779: %\newdefi{def3.3.2} In the situation of \lemma{lem3.3.4} the integer $l$
4780: %is called the {\sl secondary cusp index of $u$ at point $z^*$}.
4781: %\end{defi} ***!!!
4782:
4783:
4784: \newsubsection[4.4]{Critical points and cusp-curves in the moduli space}
4785: Recall that in \lemma{lem3.3.4} we found two obstructions for existence of
4786: saddle points. They are encoded in the secondary cusp-indices $l_i$ of
4787: cusp-points $z^*_i$ of $u$ (see \refdefi{def4.2s.1}) and the vanishing
4788: order at $z^*_i$ of a generic $\psi \in \sfh^0_D(S, N_u \otimes K_S) \cong
4789: \bigl(\sfh^1_D(S, N_u) \bigr) ^*$. The behavior $l_i$ under deformation was
4790: studied in \refsubsection{4.2s}.
4791: In this paragraph we describe the behavior of $\sfh^0_D(S, N_u \otimes K_S)$.
4792: Our main interest is, of course, $[u,J] \in \scrm$ with $\dimr\sfh^1_D (S, N_u)
4793: =1$, because these are candidates for saddle points. We start with
4794:
4795: \newlemma{lem4.4.1} Let $\mbfk=(k_1, \ldots, k_m)$ and $h^1\in \nn$
4796: be given. Then the set
4797: \begin{equation}
4798: \whcalm_{=\mbfk, h^1} \deff \bigl\{ (u, J_S, J; \mbfz) \in \whcalm
4799: _{=\mbfk} \;:\; \dimr \sfh^1_D(S, N_u) = h^1\,\bigr\}
4800: \subset \whcalm_{=\mbfk}
4801: \end{equation}
4802: is a $C^{\ell-1}$-smooth submanifold of codimension $h^0{\cdot} h^1$ where
4803: $h^0 \deff \dimr \sfh^0_D(S, N_u)$.
4804:
4805: The set $\whcalm_{=\mbfk, h^1}$ is $\bfg$-invariant and the projection $\pr:
4806: \whcalm_{=\mbfk, h^1} \lrar \scrm_{=\mbfk, h^1} \deff \whcalm_{=\mbfk, h^1}
4807: /\bfg$ is a $C^{\ell-1}$-smooth principle $\bfg$-bundle.
4808: \end{lem}
4809:
4810:
4811: \state Remark. The definition \eqqref(1.5.1) of $\scrn_u$ and the index formula
4812: \eqqref(2.2.3) imply that $h^0= h^1 +2(\mu +(g-1)(3-n) -|\mbfk|)$, where $n=\half
4813: \dimr X$ and $\mu \deff \la c_1(X), [u(S)] \ra$. So $h^0=\dimr \sfh^0_D(S,
4814: N_u)$ is constant along $\whcalm_{=\mbfk, h^1}$ and
4815: \begin{equation}
4816: \whcalm_{=\mbfk} = \mathop{\textstyle\bigsqcup\,}\nolimits
4817: _{h^1 =0} ^\infty \;\whcalm_{=\mbfk, h^1}
4818: \end{equation}
4819: is a stratification of $\whcalm_{=\mbfk}$ indexed by $h^1 =\dimr \sfh^1_D(S,
4820: N_u)$. Taking the $\bfg$-quotients, we obtain a similar stratification of $\scrm
4821: _{=\mbfk}$. Another stratification, more interesting for our purpose, is
4822: \begin{equation}
4823: \scrm_{h^1} = \mathop{\textstyle\bigsqcup\,}\nolimits
4824: _{\mbfk}\scrm_{=\mbfk, h^1}
4825: \end{equation}
4826: with $h^1=1$.
4827: Note that if for given $\mbfk$ and $h^1$ the expected value of $h^0$ is
4828: negative, then $\whcalm_{=\mbfk, h^1}$ is empty.
4829:
4830: \proof Consider the Banach bundles $L^{1,p}(S, N)$, $L^p_{(0,1)}(S, N)$
4831: over $\whcalm_{=\mbfk}$, and the bundle homomorphism $D^N: L^{1,p}(S, N) \to
4832: L^p_{(0,1)}(S, N)$ constructed in \lemma{lem4.2.7n}. Then $\sfh^i_D(S, N_u)$
4833: is the (co)kernel of $D^N$. From \lemma{lem1.3.1} we obtain the map
4834: \begin{equation}
4835: \nabla_{(v, \dot J_S, \dot J)}D^N: \sfh^0_D(S, N_u) \to \sfh^1_D(S, N_u)
4836: \eqqno(4.4.1),
4837: \end{equation}
4838: which is bilinear in $(v, \dot J_S, \dot J) \in T_{(u, J_S, J)}\whcalm_{=\mbfk}$
4839: and $w \in \sfh^0_D(S, N_u)$. It is not difficult to see that the map
4840: \eqqref(4.4.1) can be computed using \eqqref(3.2.8) and that it coincides with
4841: the restriction of $\Phi$ from \eqqref(3.1.1) to the corresponding spaces.
4842:
4843: The key point of the proof is to show the surjectivity of the induced map
4844: \begin{equation}
4845: \Phi: T_{(u, J_S, J)}\whcalm_{=\mbfk} \;\lrar\; \homr \bigl(
4846: \sfh^0_D(S, N_u),\; \sfh^1_D(S, N_u) \bigr)
4847: \end{equation}
4848: Then the claim of the lemma will follow from the implicit function theorem.
4849:
4850: Fix bases $(w_1, \ldots, w_{h^0})$ of $\sfh^0_D(S, N_u)$ and $(\psi_1, \ldots,
4851: \psi_{h^1})$ of $\sfh^0_D(S, N_u \otimes K_S) \cong \bigl(\sfh^1(S,N_u)
4852: \bigr)^*$. The last isomorphism is the Serre duality from \lemma{lem1.5.1}.
4853: We must find tangent vectors $(v_{ij}, \dot J_{S,ij}, \dot J_{ij}) \in T
4854: _{(u, J_S, J)}\whcalm_{=\mbfk}$, $i=1,\ldots, h^0$, $j=1,\ldots, h^1$
4855: obeying the relation
4856: \begin{equation}
4857: \bigl\la \psi_{j'},\,
4858: \Phi\bigl( (v_{ij}, \dot J_{S,ij}, \dot J_{ij}), w_{i'} \bigr) \bigr\ra
4859: = \delta_{ii'} \delta_{jj'}
4860: \eqqno(4.4.2)
4861: \end{equation}
4862: with $\la\cdot\,, \cdot \ra$ denoting the pairing from \eqqref(1.5.5).
4863:
4864: The main idea is to find solutions of \eqqref(4.4.2) in the special form such that
4865: $v_{ij}$ and $\dot J_{S,ij}$ are identically zero, and $\dot J_{ij}$ vanish
4866: along $u(S)$ and in a neighborhood of all cusp-points on $u(S)$. This
4867: assumption implies that all the terms in \eqqref(3.2.8) except $[7]$ vanish.
4868: Thus \eqqref(4.4.2) reduces to
4869: \begin{equation}
4870: \re \int_S \psi_{j'} \scirc \nabla_{w_{i'}} \dot J_{ij}\scirc du
4871: \scirc J_S = \delta_{ii'} \delta_{jj'}.
4872: \end{equation}
4873: From this point we can use the arguments either of \lemma{lem2.1.2} or {\sl
4874: Lemma 3.2.4}. Note that we can arrange $\dot J_{ij}$ to have support in
4875: any given open subset $U \subset X$ with $U\cap u(S) \not = \emptyset$. \qed
4876:
4877: \smallskip
4878: The big freedom in the choice of $\dot J_{ij}$ implies the following
4879:
4880: \newcorol{cor4.4.2} Let $\dimr X =4$, \ie $X$ is an almost complex
4881: surface. Then the intersection of $\whcalm_{=\mbfk, \mbfl}$ and
4882: $\whcalm_{=\mbfk, h^1}$ is transversal, so that the set
4883: \begin{equation}
4884: \whcalm_{=\mbfk, \mbfl, h^1}\deff \whcalm_{=\mbfk, \mbfl} \cap
4885: \whcalm_{=\mbfk, h^1}
4886: \end{equation}
4887: is a $C^{\ell-1}$-smooth submanifold of $\whcalm_{=\mbfk}$ of codimension
4888: $2|\mbfl| + h^0{\cdot}h^1$. A similar result also holds for
4889: $\scrm_{=\mbfk, \mbfl, h^1} \deff \whcalm_{=\mbfk, \mbfl, h^1} /\bfg =
4890: \scrm_{=\mbfk, \mbfl} \cap \scrm_{=\mbfk, h^1}$.
4891: \end{corol}
4892:
4893:
4894: \medskip
4895: Now we will study the behavior of zeros of a non-trivial $\psi \in \sfh^0_D(S,
4896: N_u \otimes K_S)$ for $[u, J] \in \scrm_{=\mbfk, h^1=1}$. Note that, modifying
4897: the construction from \lemma{lem4.4.3}, we obtain a bundle $N^* \otimes K_S$
4898: over $\whcalm_{=\mbfk} \times S$, $C^{\ell-1}$-smooth Banach
4899: bundles $L^{1,p}(S, N^* \otimes K_S)$ and $L^p_{(0,1)}(S, N^* \otimes K_S)$
4900: over $\whcalm_{=\mbfk}$, and a $C^{\ell-1}$-smooth bundle homomorphism
4901: \begin{equation}
4902: (D^N)^*: L^{1,p}(S, N^* \otimes K_S) \to L^p_{(0,1)}(S, N^* \otimes K_S).
4903: \end{equation}
4904: Since the kernel of $(D^N)^*$ is of constant dimension on each $\whcalm_{=\mbfk,
4905: h^1}$, we obtain a $C^{\ell-1}$-smooth bundle $\sfh^0_D(S, N^* \otimes K_S)$ of
4906: $\rank_\rr =h^1$ on $\whcalm_{=\mbfk, h^1}$. This means that there exists a
4907: (local) frame $\psi_1, \ldots, \psi_{h^1}$ of $\sfh^0_D(S, N^* \otimes K_S)$
4908: which depends $C^{\ell-1}$-smoothly on $(u, J_S, J) \in \whcalm_{=\mbfk, h^1}$.
4909:
4910: In the particular case $h^1=1$ we obtain a (local) $C^{\ell-1}$-smooth family of
4911: non-trivial $\psi \in \sfh^0_D(S,N^* \otimes K_S)$ such that for every $(u,
4912: J_S, J)$ the corresponding $\psi$ is defined uniquely up to a constant factor.
4913: \lemma{lem1.2.3} ensures that the zero-divisor of such $\psi$ is well-defined
4914: and has degree
4915: $c_1(N^* \otimes K_S)$. By \lemma{lem2.3.4}, the possible range for $c_1(N^*
4916: \otimes K_S)$ is the interval between 0 and $g-1$. We are interested in the
4917: distribution of the zeros of $\psi$, especially at cusp-points of $u$. For
4918: a given $\mbfk=(k_1, \ldots, k_m)$ we consider $m$-tuples $\bfnu=(\nu_1,
4919: \ldots, \nu_m)$ with $0 \le \nu_i \le k_i$, $i=1, \ldots, m$. Denote
4920: $|\bfnu| \deff \sum_{i=1}^m \nu_i$.
4921:
4922: \newlemma{lem4.4.3}
4923: \sli The set $\whcalm_{=\mbfk, \bfnu} \deff \bigl\{ (u, J_S, J) \in
4924: \whcalm_{=\mbfk, h^1=1} \;:\; \ord_{z^*_i} \psi \ge \nu_i \,\bigr\} \subset
4925: \whcalm_{=\mbfk, h^1=1}$ is a $C^{\ell-1}$-smooth submanifold of codimension
4926: $2(n-1)|\bfnu|$, $n=\half\,\dimr X$.
4927:
4928: \slii Let $n=2$, \ie $X$ is an almost complex surface. Then the intersection of
4929: $\whcalm_{=\mbfk, \mbfl}$ and $\whcalm_{=\mbfk, \bfnu}$ is transversal, so
4930: that the set
4931: \begin{equation}
4932: \whcalm_{=\mbfk, \mbfl, \bfnu} \deff \whcalm_{=\mbfk, \mbfl, h^1=1} \cap
4933: \whcalm_{=\mbfk, \bfnu} \;\subset\; \whcalm_{=\mbfk, h^1=1}
4934: \end{equation}
4935: is a $C^{\ell-1}$-smooth submanifold of codimension $2|\mbfl| + |\bfnu|$.
4936:
4937:
4938: \sliii Similar results hold for $\scrm_{=\mbfk, \mbfl, \bfnu} \deff
4939: \whcalm_{=\mbfk, \mbfl, \bfnu}/\bfg= \scrm_{=\mbfk, \mbfl,h^1=1} \cap
4940: \scrm_{=\mbfk, \bfnu}$.
4941: \end{lem}
4942:
4943:
4944: \proof \slip.
4945: Fix $(u_0, J_{S,0}, J_0) \in \whcalm_{=\mbfk, \bfnu}$. Let $z_i$
4946: be a local $J_S$-holomorphic coordinate on $\whcalm_{=\mbfk, h^1=1}$ in the
4947: sense of \refdefi{def4.2.2n}, centered at the cusp-point $z^*_i$ of $(u,
4948: J_S, J)$, $i=1, \ldots, m$. Further, let $\psi$ be a local $C^{\ell-1}$-smooth
4949: family of non-trivial elements of $\sfh^0_D(S,N^* \otimes K_S)$. Then by
4950: \lemma{lem3.3.2}, for each $i=1, \ldots, m$ and each $(u, J_S, J) \in
4951: \whcalm_{=\mbfk, h^1=1}$ we can construct the jets $j^{k_i\!}\psi =
4952: \sum_{j=0} ^{k_i} \psi_{i,j} \cdot z_i^{\;j}$ of $\psi$ at $z^*_i$.
4953:
4954: Repeating the arguments used in the proof of \lemma{lem4.2.3n} %%?? Check!
4955: we can show
4956: that the coefficients $\psi_{i,j} \in (T^*_{z^*_i})^{\otimes j} \otimes (N^*
4957: \otimes K_S)_{z^*_i}$ depend $C^{\ell-1}$-smoothly on $(u, J_S, J) \in \whcalm
4958: _{=\mbfk, h^1=1}$. This means that $\whcalm_{=\mbfk, \bfnu}$ is the zero set
4959: of the (locally defined) function $\yps^\psi_\bfnu$ on $\whcalm_{=\mbfk,
4960: h^1=1}$ given by the first $\nu_i$ coefficients of each $j^{k_i\!}\psi$,
4961: $i=1, \ldots, m$, \ie
4962: \begin{equation}
4963: \yps^\psi_\bfnu(u, J_S, J) =
4964: (\psi_{1,0}, \ldots, \psi_{1, \nu_1 -1}, \ldots, \psi_{m,0}, \ldots,
4965: \psi_{m, \nu_m -1}).
4966: \end{equation}
4967:
4968: Consequently, it is sufficient to show the surjectivity of the differential
4969: $d\yps^\psi_\bfnu$ at the fixed $(u_0, J_{S,0}, J_0)$. But first we must compute
4970: $d\yps^\psi_\bfnu$ for a given $(v, \dot J_S, \dot J) \in T_{(u_0, J_{S,0}, J_0)}
4971: \whcalm_{=\mbfk, h^1=1}$. Let $\gamma(t)= (u_t, J_{S,t}, J_t)$ be a curve in
4972: $\whcalm_{=\mbfk, h^1=1}$ which starts at $(u_0, J_{S,0}, J_0)$ and has the
4973: tangent vector $(v, \dot J_S, \dot J)$ at $t=0$. Then we obtain a family $\psi_t$
4974: of non-trivial $\psi_t \in \sfh^0_D(S, N_{u_t}^* \otimes K_S)$. In particular,
4975: for each $t$ we obtain the relation $D^*_t \psi_t=0$, where $D^*_t$ denotes
4976: the operator $D^{N^* \otimes K_S}$ corresponding to $(u_t, J_{S,t}, J_t)$.
4977:
4978: Fix some symmetric connections on $X$ and $S$. As in \refsubsection{3.2}, we
4979: obtain induced connections for all (usual and Banach) bundles involved in our
4980: computations. We use the same notation $\nabla$ for all these connections,
4981: in particular, for the connection in the bundle $L^{1,p}(S, N_u)$ with
4982: the fiber $L^{1,p}(S, N_{u_t})$ over $(u_t, J_{S,t}, J_t)$. Hence for any
4983: $w_0 \in L^{1,p}(S, N_{u_0})$ we can construct a family $w_t \in L^{1,p}(S,
4984: N_{u_t})$ which is covariantly constant. This yields a covariantly constant
4985: trivialization of the Banach bundle $L^{1,p}(S, N_u)$ along $\gamma$.
4986:
4987: For every such family $w_t$ we have the relation
4988: \begin{equation}
4989: \la w_t, D^*_t \psi_t \ra=0.
4990: \eqqno(4.4.3)
4991: \end{equation}
4992: Vice versa, a family $\psi_t \in L^{1,p}(S, N^*_{u_t} \otimes K_S)$ lies
4993: in $\sfh^1_D(S, N^*_{u_t} \otimes K_S)$ if \eqqref(4.4.3) holds. Rewrite
4994: \eqqref(4.4.3) in the form
4995: \begin{equation}
4996: \la \psi_t, D_t w_t \ra=0
4997: \eqqno(4.4.4)
4998: \end{equation}
4999: with $D_t$ denoting the operator $D^N_{u_t, J_t}: L^{1,p}(S, N_{u_t}) \to
5000: L^p_{(0,1)}(S, N_{u_t})$. After covariant differentiation in $t$ we obtain
5001: $\la \dot \psi_t, D_t w_t \ra + \la \psi_t, (\nabla_{(v_t, \dot J_{S,t},
5002: \dot J_t)} D_t) w_t \ra =0$. The latter is equivalent to
5003: \begin{equation}
5004: \la D^*_t \dot\psi_t, w_t \ra +
5005: \la \psi_t, (\nabla_{(v_t, \dot J_{S,t}, \dot J_t)} D_t) w_t \ra =0.
5006: \eqqno(4.4.5)
5007: \end{equation}
5008:
5009: Now we can give the description of $d\yps^\psi_\bfnu$ at $(u_0, J_{S,0}, J_0) \in
5010: \whcalm_{=\mbfk, h^1=1}$. For a given tangent vector $(v, \dot J_S, \dot J)$ we
5011: find $\dot \psi \in L^{1,p}(S, N_{u_0} \otimes K_S)$ such that \eqqref(4.4.5)
5012: holds for every $w \in L^{1,p}(S, N_{u_0} \otimes K_S)$. The existence of such
5013: $\dot \psi$ is equivalent to the condition that $(v, \dot J_S, \dot J)$ is
5014: tangent to $\whcalm_{=\mbfk, h^1=1}$. Such $\dot \psi$ is unique up to
5015: addition of $\psi \in \sfh^0_D(S, N_{u_0} \otimes K_S)$. The
5016: jets $j^{k_i\!} \dot \psi= \sum_{j=0} ^{k_i} \dot \psi_{i,j} \cdot z_i^{\;j}$
5017: of such $\dot \psi$ at $z^*_i$ are well-defined and
5018: \begin{equation}
5019: d\yps^\psi_\bfnu(v, \dot J_S, \dot J) =
5020: (\dot \psi_{1,0}, \ldots, \dot \psi_{1, \nu_1 -1}, \ldots,
5021: \dot \psi_{m,0}, \ldots,
5022: \dot \psi_{m, \nu_m -1}).
5023: \end{equation}
5024: $d\yps^\psi_\bfnu$ is independent of the choice of $\dot \psi$ provided
5025: $(u_0, J_{S,0}, J_0) \in \whcalm_{=\mbfk, \bfnu}$.
5026:
5027: \smallskip
5028: To show the surjectivity of $d\yps^\psi_\bfnu$ we must invert the construction
5029: above. Let $j^{\nu_i-1\!} \dot \psi$ be given jets. Extend them to jets
5030: $j^{k_i\!} \dot \psi$. Note that by definition the operator $D^*_0 = D^{N^*
5031: \otimes K_S} _{u_0, J_0}$ has the form $\dbar^{N^* \otimes K_S}_{u_0, J_0} +
5032: R^{N^* \otimes K_S}_{u_0, J_0}$ where
5033: \begin{equation}
5034: R^{N^* \otimes K_S}_{u_0, J_0}: N^* \otimes K_S \to
5035: N^* \otimes K_S \otimes \Lambda^{(0,1)}
5036: \end{equation}
5037: is a continuous bundle homomorphism. Consider the equations
5038: \begin{equation}
5039: z_i^{-k_i}\bigl(\dbar^{N^* \otimes K_S}_{u_0, J_0} +
5040: R^{N^* \otimes K_S}_{u_0, J_0} \bigr)
5041: \bigl(j^{k_i\!} \dot \psi + z_i^{k_i} \phi_i(z_i)\bigr)=0
5042: \eqqno(4.4.6)
5043: \end{equation}
5044: for unknown $\phi_i(z_i)$ defined in a neighborhood of $z^*_i$. Using
5045: \lemma{lem1.4.1} we obtain pointwise estimates $|R^{N^* \otimes K_S}_{u_0,
5046: J_0}(z_i)| \le C\cdot |z_i|^{k_i}$. Thus equation \eqqref(4.4.6) is equivalent to
5047: \begin{equation}
5048: \left(\dbar^{N^* \otimes K_S}_{u_0, J_0} +
5049: \left(\msmall{\bar z_i \over z_i}\right)^{k_i}
5050: R^{N^* \otimes K_S}_{u_0, J_0} \right) \phi_i(z_i)
5051: + z_i^{- k_i} R^{N^* \otimes K_S}_{u_0, J_0}j^{k_i\!}=0.
5052: \eqqno(4.4.7)
5053: \end{equation}
5054: The existence of solutions of \eqqref(4.4.7) can be deduced from the
5055: surjectivity of the operator $\dbar + R: L^{1,p}(\Delta,\cc^n) \to L^p(\Delta,
5056: \cc^n)$ with $R \in L^p$, $p>2$.
5057: We refer to \cite{Iv-Sh-1} for the construction of a right inverse for such
5058: $\dbar + R$. This implies the local existence of solutions $\phi_i(z_i)$
5059: of \eqqref(4.4.6).
5060:
5061: The regularity property of $R^{N^* \otimes K_S}_{u_0, J_0}$ implies that the $z_i
5062: ^{k_i} \phi_i(z_i)$ are $C^{\ell-1}$-smooth. Thus we can construct a $\dot \psi
5063: \in C^{\ell-1}(S, N^*_{u_0} \otimes K_S)$ which locally near $z^*_i$ has the form
5064: $\dot \psi(z_i) = j^{k_i\!} \dot \psi + z_i^{k_i} \phi_i(z_i)$ and satisfies
5065: \eqqref(4.4.6). Now, the surjectivity of $\yps^\psi_\bfnu$
5066: will follow from the existence of $(v, \dot J_S, \dot J)\in T_{(u_0, J_{S,0},
5067: J_0)} \whcalm_{=\mbfk, h^1=1}$ such that for the constructed $\dot \psi$ and
5068: a fixed non-zero $\psi_0 \in \sfh^0_D(S, N^*_{u_0} \otimes K_S)$
5069: the relation \eqqref(4.4.5) holds for any $w \in L^{1,p}(S, N_{u_0})$.
5070:
5071: Now observe that we can use \eqqref(3.2.8) to compute $\nabla_{ (v, \dot J_v,
5072: \dot J)} D^N_{u_0, J_0}$. This implies that we can use the trick from the proof
5073: of \lemma{lem4.4.1}. Namely, we look for the desired $(v, \dot J_S, \dot J)$
5074: in the special form, such that $v$ and $\dot J_S$ vanish identically, and
5075: $\dot J$ vanishes along $u_0(S)$ and in some neighborhoods of cusp-points of
5076: $u(S)$. Now all terms in \eqqref(3.2.8) except $[7]$ vanish, and
5077: \eqqref(4.4.5) is equivalent to
5078: \begin{equation}
5079: D^{N^* \otimes K_S}_{u_0, J_0} \dot \psi +
5080: \psi_0 \scirc \nabla \dot J \scirc du_0 \scirc J_S =0.
5081: \end{equation}
5082: To finish the construction of $\dot J$ we use the fact that $D^{N^* \otimes
5083: K_S}_{u_0, J_0} \dot \psi$ vanishes in a neighborhood of each cusp-point
5084: $z^*_i$. This yields the surjectivity of $\yps^\psi_\bfnu$ and
5085: the first assertion of the lemma.
5086:
5087: \smallskip
5088: The second and third assertions follow from previous considerations. \qed
5089:
5090:
5091:
5092: \newsubsection[4.5]{(Non)existence of saddle points in the moduli space}
5093: The results obtained above in this section allow us to prove the main technical
5094: result of the paper. Let $X$ be a manifold of dimension $2n$,
5095: $\scrj$ an open connected set in the space of $C^\ell$-smooth almost complex
5096: structures on $X$ with $\ell >2$ non-integer, $S$ a closed surface of genus
5097: $g\ge1$, and $[C] \in \sfh_2(X,\zz)$ a homology class.
5098:
5099: \newdefi{def4.5.1} A pseudoholomorphic map $u:S \to X$ has an {\sl
5100: ordinary cusp} at $z^*\in S$ if for appropriate coordinates $z$ on $S$ and
5101: $(w_1, w_2)$ on $X$
5102: \begin{equation}
5103: u(z)=\bigl(z^2 +O(|z|^3), z^3 + O(|z|^{3+\alpha}) \bigr).
5104: \end{equation}
5105: This property is equivalent to the condition that $u$ has a cusp of order 1
5106: and the secondary cusp-index 0 at $z^*$.
5107: \end{defi}
5108:
5109: \newthm{thm4.5.1} Let $h(t)=J_t$, $t\in I=[0,1]$, be a {\sl generic}
5110: path in $\scrj$ and $\scrm_h$ the corresponding relative moduli space of
5111: parameterized pseudoholomorphic curves of genus $g\ge1$ in the homology class
5112: $[C]$.
5113:
5114: \sli If $n\ge 3$, then every critical point of the projection $\pi_h: \scrm_h
5115: \to I$ is represented by an {\sl imbedded} curve $C=u(S)$, $u: S \to X$;
5116:
5117: \slii If $n=2$, then every critical point of the projection $\pi_h: \scrm_h
5118: \to I$ is represented by a curve $C=u(S)$ such that:
5119: \begin{itemize}
5120: \item the only singularities on $C$ are nodes or ordinary cusps;
5121: %nodal an ordinary cuspidal ones;
5122: \item the possible number of cuspidal points $\vkappa$ on $C$ is
5123: \begin{equation}\eqqno(4.5.2)
5124: \mu \le \vkappa \le \mu +g-1,
5125: \end{equation}
5126: where $\mu \deff \la c_1(X), [C] \ra$ and $g$ is the (geometric) genus of $C$,
5127: $g=g(S)$;
5128: \item the saddle index of $d^2\pi_h$ at $C$ is at least $\vkappa$, \ie
5129: $$
5130: \sind_C d^2\pi_h \ge \vkappa \ge \mu .
5131: $$
5132: \end{itemize}
5133:
5134: \sliii In the case when the inequality \eqqref(4.5.2) is a contradiction,
5135: the claim \slii has the following meaning:
5136: \begin{itemize}
5137: \item If $g=0$, then $\pi_h$ has no critical points;
5138: \item If $\mu +g -1<0$, then the space $\scrm_h$ is empty for generic $h$.
5139: \end{itemize}
5140: \end{thm}
5141:
5142: Before giving the proof we must specify the meaning of the notion {\sl generic
5143: path}. One of the most reasonable conditions is that any two {\sl regular}
5144: almost complex structures $J_0,\; J_1 \in \scrj$ (see \S\.{\sl2.3}) can be
5145: connected by a path $\{J_t\}_{t\in I=[0,1]}$ with the property stated in
5146: the theorem. To ensure this we need the following easy
5147:
5148: \newprop{prop4.5.2} Let $F: \scrx \to \scry$ be a $C^1$-smooth
5149: Fredholm map between separable Banach manifolds. Assume that $\scry$ is
5150: connected and that the index of $F$ is at most $-2$. Then the set $\scry \bs
5151: F(\scrx)$ is path-connected.
5152: \end{prop}
5153:
5154: \state Remark. The proposition generalizes the obvious fact that submanifolds
5155: of codimension at least 2 do not divide the ambient manifold. Note that one can
5156: have at most countably many connected components of $\scrx$ and that on these
5157: components the index of $F$ can vary from component to component.
5158:
5159: \statep Proof. \refthm{thm4.5.1}. We already know from \refsection{2} that for
5160: a generic path $h(t) =J_t$ in $\scrj$ the set $\scrm_h$ is a manifold. In previous
5161: paragraphs of this section we have showed that critical points $[u,J]$ of the
5162: projection $\pi_h: \scrm_h \to I$ have an intrinsic description independent of the
5163: particular choice of the path $J_t$. Moreover, the quadratic form $d^2\pi_h$ at
5164: these points also admits a similar intrinsic description. Furthermore, we have
5165: found a stratification of the set of ``suspicious''
5166: points $[u,J] \in \scrm$ by submanifolds and estimated their codimension. It
5167: remains to find the strata with the Fredholm index $\le -2$ over $\scrj$ and
5168: apply \propo{prop4.5.2}.
5169:
5170: The ``suspicious'' points $[u,J]$ on $\scrm$ are those with $\sfh^1_D(S,
5171: \scrn_u) \cong \rr$. They can be separated into classes according to the
5172: structure of the normal sheaf $\scrn_u$. Since the singular part $\scrn_u\sing$
5173: of $\scrn_u$ reflects the cusp-curves we are led to the spaces $\scrm_{=\mbfk}$
5174: of curves with prescribed order of cusps.
5175:
5176: Denote by $\ind$ the index of the projection $\pr _{\!\!\scrj}: \scrm \to \scrj$,
5177: so that $\ind = 2(\la c_1(X), [C] \ra + (g-1)(3-n))$. If $\ind < 0$, then
5178: for a generic path $h(t)=J_t$ the set $\scrm_h$ is empty and the claim of the
5179: theorem holds. Thus we may assume that $\ind \ge0$. By \refthm{thm4.2.1}, we
5180: must ``pay'' at least $2(n-1)|\mbfk|$ dimensions to go to $\scrm_{=\mbfk}$. By
5181: \lemma{lem4.4.1}, we must ``pay'' further $\ind-2|\mbfk| +1$ dimensions to
5182: obtain the condition $\sfh^1_D(S, \scrn_u) \cong \rr$. Note that $2(n-1)|\mbfk|
5183: \ge 2|\mbfk| +2$ if $n\ge 3$ and $\mbfk$ is non-trivial. Thus in the case $n\ge3$
5184: we ``overdraw'' our ``credit'' $\ind$ at least by $3$. This means that for
5185: non-trivial $\mbfk$ the index of the projection from $\scrm_{=\mbfk,h^1=1}$ to
5186: $\scrj$ is at most $-3$ and we can apply \propo{prop4.5.2}. Thus for $n\ge3$ any
5187: critical point of $\scrm_h$ is represented by an immersion $u:S \to X$.
5188:
5189: \medskip
5190: In the case $n=2$ we can ``strike the balance'' in a similar way. Indeed, we
5191: come to the ``overdraw'' of at least $3$ dimensions in each of the following
5192: cases:
5193:
5194: {\sl a)} $\sfh^1_D(S, \scrn_u) \cong \rr$ and there exists at least one
5195: cusp-point of cusp-order$\ge2$;
5196:
5197: {\sl b)} $\sfh^1_D(S, \scrn_u) \cong \rr$ and there exists at least one
5198: cusp-point the secondary cusp-index $\ge1$;
5199:
5200: {\sl c)} $\sfh^1_D(S, \scrn_u) \cong \rr$ and a non-trivial $\psi \in
5201: \sfh^1_D(S, N_u^* \otimes K_S )$ vanishes in at least one cusp-point.
5202:
5203: \noindent
5204: Thus for generic $h(t)=J_t$ we can exclude all these possibilities. The
5205: remaining case admits only cusps of order 1 with the secondary cusp-index 0.
5206: This means that $u$ has only ordinary cusps. Since possibility
5207: {\sl c)} is excluded, each such cusp gives input 1 into the saddle index by
5208: \lemma{lem3.3.4}. Finally, we estimate the number of such cusps using
5209: \lemma{lem2.3.4}.
5210:
5211: \medskip
5212: Now we show that for $n\ge3$ and generic $h$ any critical point of $\scrm_h$ is
5213: represented by an {\sl imbedding} $u:S \to X$. Since this result will be not used
5214: in the sequel, we give only a sketchy proof.
5215:
5216: Denote by $\scrm_{\sf{imm}}$
5217: the total moduli space of {\sl immersed pseudoholomorphic curves} with the same
5218: topological data $g=g(S)$ and $[C] \in \sfh_2(X, \zz)$ as usual. In other words,
5219: $\scrm_{\sf{imm}}= \scrm_{=\mbfk}$ with trivial $\mbfk$. It follows easily from
5220: \refsection{4} that this is an open set in the whole space $\scrm$. The space
5221: $\scrm_{\sf{imm}}$ admits a natural stratification in which every stratum contains
5222: curves with the same number and type of multiple point on the image $C=u(S)$.
5223: Obviously, the biggest stratum is the subspace of {\sl imbedded} curves, and this
5224: is an open subset in $\scrm_{\sf{imm}}$. The next biggest stratum consists of
5225: curves with exactly one transversal double point on $C=u(S)$. Let us denote it by
5226: $\scrm_{\sf{imm}}^\times$ with the character $\times$ symbolizing a transversal
5227: self-intersection of exactly 2 branches of $C=u(S)$. %%here
5228:
5229: Locally, $\scrm_{\sf{imm}}^\times$ is defined by the condition $u(z_1) =u(z_2)$
5230: for some $z_1 \not= z_2 \in S$. Linearization of this condition is the equation
5231: $$
5232: \pr_{N^\times}(v(z_1) -v(z_2)) =0
5233: $$
5234: on $[v, \dot J_S, \dot J] \in T_{[u,J]}\scrm_{\sf{imm}}$, where $N^\times$ denotes
5235: the plane in $T_{x^\times}X$ normal to both branches of $C=u(S)$ at the
5236: point $x^\times=u(z_1)=u(z_2)$, \ie $N^\times \deff T_{x^\times}
5237: X /\bigl(du(T_{z_1}S) \oplus du(T_{z_2} S) \bigr)$. It is easy to see that this
5238: condition is transversal. Thus $\scrm_{\sf{imm}}^\times$ is a $C^\ell$-smooth
5239: submanifold of real codimension $2(n-2)$. Moreover, it follows from the proof of
5240: \lemma{lem4.4.1} that biggest stratum is transversal to the subspace $\scrm
5241: _{\sf{imm}, h^1=1}$ of immersed curves with $\sf{h}^1(S, \scrn_u)=1$. Consequently,
5242: the space $\scrm^\times_{\sf{imm}, h^1=1} \deff \scrm_{\sf{imm}}^\times \cap \scrm
5243: _{\sf{imm}, h^1=1}$ of immersed curves with $\sf{h}^1(S, \scrn_u)=1$ and with
5244: exactly one transversal self-intersection point is a $C^\ell$-smooth submanifold
5245: of real codimension $2(n-2)$ in $\scrm_{\sf{imm}, h^1=1}$, and of real codimension
5246: $\ind + 1+ 2(n-2)$ in $\scrm$. Since $2(n-2) \ge2$ for $n\ge3$, we can apply
5247: \propo{prop4.5.2}.
5248:
5249: The complementary strata $\scrm _{\sf{imm}}^{\mib{a}}$ of $\scrm_{\sf{imm}}$
5250: consist of curves having either several double points, or one double point with
5251: tangency of higher degree, or even more complicated multiple points, with the index
5252: $\mib{a}$ encoding the number and the type of multiple points. A similar argument
5253: shows that these strata $\scrm_{\sf{imm}}^{\mib{a}}$ are transversal to $\scrm
5254: _{\sf{imm},h^1 =1}$ and that the intersections $\scrm _{\sf{imm}, h^1=1}^{\mib{a}}
5255: \deff \scrm _{\sf{imm}}^{\mib{a}} \cap \scrm _{\sf{imm}, h^1=1}$ are transversal.
5256: The computation of the number of conditions shows that these strata have even
5257: higher codimension in $\scrm_{\sf{imm}}$. So \propo{prop4.5.2} still applies.
5258: This finishes the proof of the theorem. \qed
5259:
5260: \medskip
5261: In applications, one needs a version of \refthm{thm4.5.1} for the case of curves
5262: passing through given fixed points $\mbfx=(x_1,\ldots,x_m)$ on $X$. Recall that
5263: for a $C^\ell$-smooth map $h: I\deff[0,1] \to \scrj$ we denote by $\scrm_{h,\mbfx}$
5264: the relative moduli space of $J_t=h(t)$-holomorphic curves passing through $\mbfx=
5265: (x_1,\ldots,x_m)$ (see \refsubsection{2.4}). Let $\pi_{h,\mbfx}: \scrm_{h,\mbfx}
5266: \to I$ be the corresponding projection. We also assume that $\dimr X=4$.
5267:
5268: \newthm{thm4.5.3} For a generic $h$ every critical point of the projection
5269: $\pi_{h, \mbfx}: \scrm_{h,\mbfx} \to I$ is represented by a curve $C$ such that:
5270: \begin{itemize}
5271: \item the only singularities on $C$ are nodes or ordinary cusps;
5272: \item the marked points $(x_1,\ldots, x_m)$ are smooth points of $C=u(S)$;
5273: \item the possible number of cuspidal points $\vkappa$ on $C$ is
5274: \begin{equation}\eqqno(4.5.3)
5275: \mu -m \le \vkappa \le \mu -m +g-1
5276: \end{equation}
5277: where $g$ is the (geometric) genus of $C$;
5278: \item the saddle index of $d^2\pi_h$ at $C$ is at least $\vkappa$, \ie
5279: $$
5280: \sind_C d^2\pi_h \ge \vkappa \ge \mu -m.
5281: $$
5282: \end{itemize}
5283:
5284: In the case when the inequality \eqqref(4.5.3) is a contradiction
5285: the claim has the following meaning:
5286: \begin{itemize}
5287: \item If $g=0$, then $\pi_{h, \mbfx}$ has no critical points;
5288: \item If $\mu -m +g -1 <0$, then the space $\scrm_h$ is empty for generic $h$.
5289: \end{itemize}
5290:
5291:
5292: \end{thm}
5293:
5294: \proof The main observation in the proof is that after an appropriate modification
5295: all the results of this section remain valid also for curves passing through
5296: fixed points. In particular, the most important formulas \eqqref(3.3.3) and
5297: \eqqref(3.3.5) from \lemma{lem3.3.3} holds after replacing $\scrn\sing_u$ by
5298: $\scrn\sing_{u,\mbfx}$. To show this we note first that {\sl Lemmas
5299: \ref{lem3.2.3}, \ref{lem3.3.1}}, and {\sl \ref{lem3.3.2}} can be applied without
5300: any modification. After this, the proof of \lemma{lem3.3.3} applies with
5301: the only difference that the usual Gromov operator $D_{u,J}$ acting in $E$
5302: should be replaced by the operator $D_{u,-\mbfz,J}$ acting in $E_{-\mbfz}$.
5303: The validation of such a replacement is justified in \refsubsection{2.4}. Indeed,
5304: by the very definition, $D_{u,-\mbfz,J}$ is the restriction of $D_{u,J}$ to
5305: the subspace of sections of the subbundle $E_{u,-\mbfz} \subset E_u$.
5306: In a similar way one modifies the argumentation of \refsubsection{4.4}.
5307:
5308: Finally, we note that the condition of coincidence of some cusp point of $C=u(S)$
5309: with some of marked points $x_1,\ldots,x_m$ defines a subset in $\scrm_\mbfx$
5310: which has a natural stratification into submanifolds of codimension $\ge2$.
5311: Every such stratum is defined by the cusp order $\mbfk$ of $C=u(S)$ and indication
5312: of the those cuspidal points which pass through the marked points $x_1,\ldots,
5313: x_m$. This means that every such stratum is a submanifold of the space $\scrm
5314: _{=\mbfk}$. Moreover, the codimension of every such stratum in $\scrm_{=\mbfk}$ is
5315: $4a$, where $a$ is the number of cusps lying in the marked points. As in the case
5316: $m=0$ above, one can show that the intersection of such a stratum with the space
5317: $\scrm_{=\mbfk, h^1=1}$ is transversal and has the expected codimension.
5318: Hence we may conclude that for generic $h$ such a coincidence can not occur in
5319: the critical points of $\pi_{h,\mbfx}$. The same argument is applied to show that
5320: for generic $h$ there are no coincidence of the marked points $x_1,\ldots,x_m$
5321: with nodal points of the curve $C=u(S)$ representing a critical point of $\pi_h$.
5322: \qed
5323:
5324:
5325: \newsection[5]{Deformation of nodal curves}
5326:
5327: \newsubsection[5.2]{Nodal curves and Gromov compactness theorem}
5328: The total moduli space $\scrm$ constructed in {\sl Section 2} is not complete.
5329: More precisely, the projection $\pi_{\!\!\scrj}: \scrm \to \scrj$ is, in
5330: general, not proper. This means that there exists a sequence $[u_i, J_i] \in
5331: \scrm$ such that $J_i$ converges to $J_\infty \in \scrj$ but no subsequence of
5332: $\{ u_i \}$ converges in $L^{1,p}(S, X)$-topology, even after
5333: reparameterization. Gromov compactness theorem ensures that there still exists
5334: subsequence of $\{ u_i \}$ which converges \wrt the Gromov topology, which is
5335: weaker that the Sobolev $L^{1,p}$-topology.
5336:
5337:
5338: \smallskip
5339: In the literature one can find several non-equivalent definitions for Gromov
5340: topo\-logy. In this paper we shall use that one which is equivalent to the
5341: original definition of Gromov (\cite{Gro}). However, our version is more
5342: detailed in the sense that it is based on the notion of {\sl stable maps}.
5343: This notion for curves in a complex algebraic manifold $X$ was introduced by
5344: Kontsevich in \cite{K}, see also \cite{K-M}. Our definition of stable maps over
5345: $(X,J)$ is simply a translation of this notion to almost complex manifolds.
5346:
5347: \newdefi{def5.2.1} The {\sl standard node} is the complex analytic
5348: set
5349: \begin{equation}
5350: \scra_0 \deff \{ (z_1,z_2)\in \Delta^2 : z_1\cdot z_2 =0\}.
5351: \eqqno(5.2.1)
5352: \end{equation}
5353: A point on a complex curve is called a {\sl nodal point}, if has a
5354: neighborhood biholomorphic to the standard node. A {\sl nodal curve} $C$ is a
5355: complex analytic space of pure dimension 1 with only nodal points as
5356: singularities.
5357: \end{defi}
5358:
5359: \newdefi{def5.2.1a} An annulus $A$ with a complex structure $J$ has
5360: {\sl conformal radius} $R>1$ if $A$ is biholomorphic to $A(1,R) \deff \{ z\in
5361: \cc \,:\, 1 <|z| < R \}$. Define a {\sl cylinder} $Z(a,b) \deff S^1
5362: \times [a,b] = \{ (\theta, t) : 0 \le \theta \le 2\pi,\; a\le t \le b \}$, $a<
5363: b$, with the complex structure $J_Z ({\d\over \d\theta}) \deff {\d\over \d t}$.
5364: Obviously, $Z(a,b)$ is also an annulus $A$ of conformal radius $R= e^{b-a}$.
5365: Also denote $Z_k\deff Z(k, k+1)$.
5366: \end{defi}
5367:
5368:
5369: \smallskip
5370: In other terminology, nodal curves are called {\sl prestable}. We shall
5371: always suppose that $C$ is connected and has a ``finite topology", \ie $C$ has
5372: finitely many irreducible components, finitely many nodal points, and that
5373: $C$ has a smooth boundary $\d C$ consisting of finitely many smooth circles
5374: $\gamma_i$, such that $\barr C \deff C \cup \d C$ is compact.
5375:
5376:
5377: \newdefi{def5.2.2} A real oriented surface with boundary $(\Sigma,
5378: \d\Sigma)$ {\sl parameterizes} a complex nodal curve $C$ if there
5379: is a continuous map $\sigma :\barr\Sigma \to \barr C$ such that:
5380:
5381: \sli if $a\in C$ is a nodal point, then $\gamma_a = \sigma\inv(a)$ is a
5382: smooth imbedded circle in $\Sigma \bs \d \Sigma $, and if $a\not= b$ then
5383: $\gamma_a \cap \gamma_b= \emptyset$;
5384:
5385: \slii $\sigma :\barr\Sigma \bs \bigcup_{i=1}^N\gamma_{a_i}\to \barr C \bs \{
5386: a_1,\ldots ,a_N\} $ is a diffeomorphism, where $a_1,\ldots ,a_N$ are the
5387: nodes of $C$.
5388: \end{defi}
5389:
5390:
5391: \bigskip
5392: \vbox{\xsize=.54\hsize\nolineskip\rm
5393: \putm[.17][-.01]{\gamma_1}%
5394: \putm[.31][.27]{\gamma_2}%
5395: \putm[.53][0.275]{\gamma_3}
5396: %
5397: \putm[.71][.265]{\gamma_4}%
5398: \putm[.87][.22]{\gamma_5}%
5399: \putt[1.1][0]{\advance\hsize-1.1\xsize%
5400: \centerline{Fig.~1}\smallskip%\parindent=0pt
5401: Circles $\gamma_1,..., \gamma_5$ are contracted by the parameterization
5402: map $\sigma$ to nodal points $a_1, \ldots a_5$.
5403: }%
5404: \putm[.56][.37]{\bigg\downarrow\sigma}%
5405: \epsfxsize=\xsize\epsfbox{pic1.eps}%%
5406: \vskip.15\xsize
5407: %% second picture
5408: \putm[.20][.05]{a_1}%
5409: \putm[.36][.255]{a_2}%
5410: \putm[.49][0.27]{a_3}%
5411: \putm[.725][.23]{a_4}%
5412: \putm[.86][.19]{a_5}%
5413: \epsfxsize=\xsize\epsfbox{pic2.eps}
5414: }
5415:
5416:
5417: \smallskip
5418: Note that such a parameterization is not unique: if $g:\barr\Sigma \to
5419: \barr\Sigma$ is any orientation preserving diffeomorphism then $\sigma \scirc
5420: g: \barr\Sigma \to \barr C$ is again a parameterization.
5421:
5422: A parameterization of a nodal curve $C$ by a real surface can be considered as
5423: a method of ``smoothing'' of $C$. We shall also use an alternative method of
5424: ``smoothing'', the normalization.
5425: Consider the normalization $\hat C$ of $C$. Mark on each component of this
5426: normalization the pre-images (under the normalization map $\pi_C: \hat C \to
5427: C$) of nodal points of $C$. Let $\hat C_i$ be a component of $\hat C$. We can
5428: also obtain $\hat C_i$ by taking an appropriate irreducible component $C_i$,
5429: replacing nodes contained in $C_i$ by pairs of discs with marked points, and
5430: marking the remaining nodal points. Since it is convenient to consider components
5431: in this form, we make the following
5432:
5433: \newdefi{def5.2.3} A {\sl component $C'$} of a nodal curve $C$ is the
5434: normalization of an irreducible component of $C$ with marked points selected
5435: as above.
5436: \end{defi}
5437:
5438: \smallskip
5439: This definition allows us to introduce Sobolev and H\"older spaces of
5440: functions and (continuous) maps of nodal curves.
5441:
5442: \newdefi{def5.2.3a} A continuous map $u: C \to X$ is Sobolev $L^{1,p}
5443: $-smooth, $u\in L^{1,p}(C, X)$ if the induced maps $u_i \deff u\ogran_{C_i}
5444: : C_i \to X$ of all of its components $C_i$ are $L^{1,p}$-smooth.
5445: % if so are all its restrictions on components of $C$.
5446: The notion of $J$-holomorphic maps $u: C \to X$ is similarly defined.
5447: For $u \in L^{1,p}(C, X)$ define $E_u \deff u^*TX$. Thus $E_u$ is determined
5448: by restricting to each component and by identifying fibers over pairs $(z', z'')$
5449: of marked points corresponding to nodal points. An $L^{1,p}$-smooth section $v$ of
5450: $E_u$ over $C$ is given by a collection of sections $v_{C_i} \in L^{1,p}(C_i,
5451: E_u)$, one for every component $C_i$ of $C$, such that $v(z')=v(z'')$ for each pair
5452: $(z', z'')$ of marked points chosen as above. Denote by $L^{1,p}(C, E_u)$ the space
5453: of $L^{1,p}$-sections of $E_u$.
5454: \end{defi}
5455:
5456:
5457: \smallskip
5458: \newdefi{def5.2.4} The {\sl energy} or the {\sl area} of a
5459: continuous $L_\loc ^{1,2}$-smooth map \wrt a metric $h$ on $X$ is defined as
5460: \begin{equation}
5461: \area_h(u) \deff \norm{du}^2_{L^2(C)} = \int_C |du|^2_h
5462: \eqqno(5.2.2)
5463: \end{equation}
5464: \end{defi}
5465:
5466: This definition depends only on the complex structure on $C$ but not on the
5467: choice of a metric on $C$ in the given conformal class. If an $\omega$-tame
5468: almost complex structure $J$ is given, there is a prefered choice of a metric
5469: $h$ on $X$ defined by $h(v,v) \deff \omega(v, Jv)$ for $v\in TX$.
5470:
5471: \state Remark. Our definition of the area uses the following fact. Let $g$ be
5472: a Riemannian
5473: metric on $C$ compatible with $j_C$, $h$ a Riemannian metric on $X$, and $u:C
5474: \to X$ a $J$-holomorphic immersion. Then $\norm{du}^2_{L^2(C)}$ is independent
5475: of the choice of $g$ and coincides with the area of the image $u(C)$ \wrt
5476: the metric $h_J(\cdot, \cdot) \deff \half(h(\cdot, \cdot) + h(J\cdot, J\cdot)
5477: )$. The metric $h_J$ here can be seen as a ``Hermitization'' of $h$ \wrt
5478: $J$. It is well-known that $\norm{du}^2_{L^2(C)}$ is independent of the choice
5479: of a metric $g$ on $C$ in the same conformal class, see \eg \cite{S-U}.
5480: %The independence of $\norm{du}^2_{L^2(C)}$ of the choice of a metric $g$
5481: %on $C$ in the same conformal class is a well-known fact, see \eg \cite{S-U}.
5482: Thus we can use the flat metric $dx^2 + dy^2$ to compare area and energy.
5483: For a $J$-holomorphic map we obtain
5484: \begin{equation}
5485: \norm{du}^2_{L^2(C)}= \int_C |\d_x u|_h^2 + |\d_y u|_h^2 =
5486: \int_C |\d_x u|_h^2 + |J \d_x u|_h^2 = \int_C |du|_{h_J}^2 = \area_{h_J}(u(C)),
5487: \end{equation}
5488: where the last equality is another well-known result, see \eg \cite{Gro}. Since we
5489: consider varying almost complex structures on $X$, it is useful to know that
5490: we can use any Riemannian metric on $X$ having a reasonable notion of area.
5491:
5492: \smallskip
5493: \newdefi{def5.2.5} A {\sl stable curve over $(X,J)$} is a pair
5494: $(C,u)$, where $C$ is a nodal curve and $u: C\to X$ is a $J$-holomorphic map
5495: satisfying the following condition: If $C'$ is a closed component of $C$ such
5496: that $u$ is constant on $C'$, then there exist only finitely many biholomorphisms
5497: of $C'$ which preserve the marked points of $C$. In this case $u$ is
5498: called a {\sl stable map}.
5499: \end{defi}
5500:
5501:
5502: \state Remark. One can see that stability condition is nontrivial only in the
5503: following cases:
5504:
5505: \begin{itemize}
5506: \item[{\sl 1)}] some component $C'$ is biholomorphic to $\cp^1$ with 1 or 2
5507: marked points; in this case $u$ should be non-constant on any such component $C'$;
5508: \item[{\sl 2)}] some irreducible component $C'$ of $C$ is $\cp^1$ or a torus
5509: without nodal points.
5510: \end{itemize}
5511:
5512: \noindent
5513: Since we consider only connected nodal curves, case {\sl 2)} can occur only
5514: if $C$ irreducible, \ie $C'=C$. In this case $u$ must be non-constant on
5515: $C$.
5516:
5517: \newdefi{def5.2.6} A component $C'$ of a nodal curve $C$ is called
5518: {\sl non-stable} in the following cases:
5519:
5520: { 1)} $C'$ is $\cp^1$ and has one or two marked points;
5521:
5522: { 2)} $C'$ is $\cp^1$ or a torus and has no marked points.
5523:
5524: \noindent
5525: Let $u: C \to X$ be a pseudoholomorphic map. An irreducible component $C'$ of $C$
5526: is a {\sl ghost} component (\wrt $u$) if $u$ is constant on $C'$. The {\sl ghost
5527: part $C^{gh}$ of $C$ (\wrt $u$)} is the union of all ghost components. In this
5528: paper we shall deal only with the case when all ghost components are closed.
5529:
5530: A map $u$ is {\sl non-multiple} if, except finitely many points $z\in C$,
5531: one has $u\inv(u(z)) =\{z \}$. Note that this condition excludes also ghost
5532: components.
5533: \end{defi}
5534:
5535:
5536: \smallskip
5537: Now we are going to describe the Gromov topology on the space of stable curves
5538: over $X$ introduced in \cite{Gro}. Let $\{J_n\}$ be a sequence of continuous
5539: almost complex structures on $X$ which converges to $J_\infty$ in the $C^0
5540: $-topology. Furthermore, let $(C_n, u_n)$ be a sequence of stable curves over
5541: $(X, J_n)$, such that all $C_n$ are parameterized by the same real surface $S$.
5542:
5543: \newdefi{def5.2.7} We say that $(C_n,u_n)$ {\sl converges in the Gromov
5544: topology to a stable $J_\infty$-holomorphic curve $(C_\infty,u_\infty)$ over
5545: $X$} if the parameterizations $\sigma_n: \barr S \to \barr C_n$ and
5546: $\sigma_\infty: \barr S \to \barr C_\infty$ can be chosen in such a way
5547: that the following holds:
5548:
5549: \sli $u_n\scirc \sigma_n$ converges to $u_\infty\scirc \sigma_\infty$ in the
5550: $C^0(S, X)$-topology;
5551:
5552: \slii if $\{ a_k \}$ is the set of nodes of $C_\infty$ and $\{\gamma_k\}$ are
5553: the corresponding circles in $S$, then on any compact subset $K\comp
5554: S \bs \cup_k \gamma_k$ the convergence $u_n\scirc \sigma_n \to u_\infty
5555: \scirc \sigma_\infty$ is $L^{1,p}(K, X)$ for all $p< \infty$;
5556:
5557: \sliii for any compact subset $K\comp \barr S \bs \cup_k\gamma_k$ there
5558: exists $n_0=n_0(K)$ such that $ \sigma_n^{-1}(\{ a_k \}) \cap K= \emptyset$
5559: for all $n\ge n_0$ and the complex structures $\sigma_n^*j_{C_n}$ converge
5560: smoothly to $\sigma_\infty^*j_{C_\infty}$ on $K$;
5561:
5562: \sliv the structures $\sigma_n^*j_{C_n}$ are independent of $n$ near the
5563: boundary $\d S$.
5564: \end{defi}
5565:
5566:
5567: \medskip
5568: Condition \sliv is trivial if $S $ is closed, but it is useful when
5569: one considers the ``free boundary case'', \ie when $S$ (and thus all
5570: $C_n$) are not closed and no boundary condition is imposed.
5571:
5572: The reason for introducing the notion of a curve stable over $X$ is similar
5573: to the one for the Gromov topology. We are looking for a completion of the
5574: space of smooth imbedded pseudoholomorphic curves which has ``nice''
5575: properties, namely: \.1) such a completion should contain the limit of a
5576: subsequence of every sequence of smooth curves which is bounded in an appropriate
5577: sense; \.2) such a limit should also exist for every sequence in the
5578: completed space; \.3) such a limit should be unique. The Gromov's compactness
5579: theorem ensures us that the space of curves stable over $X$ has these nice
5580: properties.
5581:
5582:
5583: \newdefi{def5.2.8} Let $C_n$ be a sequence of nodal curves,
5584: parameterized by the same real surface $S$. We say that the complex
5585: structures on $C_n$ {\sl do not degenerate near boundary}, if there exist
5586: $R>1$, such that for any $n$ and any boundary circle $\gamma_{n, i}$ of $C_n$
5587: there exist an annulus $A_{n,i} \subset C_n$ adjacent to $\gamma_{n, i}$,
5588: such that all $A_{n,i}$ are mutually disjoint, do not contain nodal points of
5589: $C_n$, and have the same conformal radius $R$.
5590: \end{defi}
5591:
5592: Since the conformal radii of all $A_{n, i}$ are all the same, we can identify them
5593: with $A(1,R)$. This means that all changes of complex structures of $C_n$
5594: take place away from boundary. The condition is trivial if $C_n$ and $S$
5595: are closed, $\d S = \d C_n = \emptyset$.
5596:
5597: \state Remark. Changing our parameterizations $\sigma_n: S \to C_n$, we
5598: may suppose that for any $i$ the pre-image $\sigma_n\inv (A_{n,i} )$ is the
5599: same annulus $A_i$ independent of $n$.
5600:
5601: \medskip
5602: Now we state Gromov's compactness theorem for stable curves. Assume that
5603: $X$ is a compact manifold and fix some Riemannian metric $h$ on $X$.
5604:
5605: \newthm{thm5.2.1} Let $(C_n,u_n)$ be a sequence of stable
5606: $J_n$-holo\-mor\-phic curves over $X$ with parameterizations $\delta_n:
5607: S \to C_n$. Suppose that:
5608:
5609: \begin{itemize}
5610: \item[{\sl a)}] $\{J_n\}$ is a sequence of continuous almost complex structures
5611: on $X$, which converges to $J_\infty$ in the $C^0$-topology;
5612:
5613: \item[{\sl b)}] there is a constant $M$ such that $\area_h [u_n (C_n)]\le M$
5614: for all $n$;
5615:
5616: \item[{\sl c)}] complex structures on the $C_n$ do not degenerate near the
5617: boundary.
5618:
5619: \end{itemize}
5620:
5621: Then there is a subsequence $(C_{n_k},u_{n_k})$ and parameterizations $\sigma
5622: _{n_k}: S \to C_{n_k}$, such that $(C_{n_k}, u_{n_k}, \sigma_{n_k})$
5623: converges to a $J_\infty$-ho\-lo\-mor\-phic curve $(C_\infty,
5624: u_\infty, \sigma_\infty)$ stable over $X$.
5625:
5626: Moreover, the limit curve $(C_{n_k}, u_{n_k})$ is unique up to the choice
5627: of the parameterization $\sigma_\infty$.
5628:
5629: Furthermore, if the structures $\delta_n^*j_{C_n}$ are constant on the fixed
5630: annuli $A_i$, each adjacent to a boundary circle $\gamma_i$ of $S$, then
5631: the new parameterizations $\sigma_{n_k}$ can be taken equal to $\delta_{n_k}$
5632: on some subannuli $A'_i \subset A_i$, also adjacent to $\gamma_i$.
5633: \end{thm}
5634:
5635: A detailed proof of the theorem in the stated form can be found in \cite{Iv-Sh-3}.
5636: We also refer to the original proof of Gromov in \cite{Gro}.
5637:
5638: \smallskip
5639: Gromov's compactness theorem induce a natural completion of the moduli space
5640: $\scrm$. Let a closed real surface $S$ of genus $g$ and a homology class $A \in
5641: \sfh_2(X,\zz)$ be given.
5642:
5643: \newdefi{def5.2.9} Nodal $J$-holomorphic curves $u': C' \to X$ and $u'': C'' \to
5644: X$ are {\sl equivalent} if there exists a biholomorphism $\phi: C' \to C''$ with
5645: $u' = u'' \scirc \phi$. The {\sl total moduli space $\barm^{st}$ of stable nodal
5646: curves over $X$} is the set of equivalence classes $[C,u, J]$ with $J \in \scrj$
5647: and $u: C \to X$ a stable $J$-holomorphic curve representing a given class $A\in
5648: \sfh_2(X,\zz)$. The space $\barm^{st}$ is equipped with the {\sl Gromov topology}
5649: in which a sequence $[C_n,u_n, J_n]$ converges to $[C_\infty,u_\infty, J_\infty]$
5650: if $J_n$ converges to $J_\infty$ in the $C^\ell$-topology and $(C_n, u_n)$ to $(C
5651: _\infty, u_\infty,)$ in the sense of \refdefi{def5.2.7}. Denote by $\pr_{\!\!
5652: \scrj}^{st}$ the natural projection $\pr_{\!\!\scrj}^{st}: [C, u,J] \in \barm
5653: ^{Gr} \mapsto J \in \scrj$.
5654:
5655: Define the {\sl Gromov compactification} $\barm^{Gr}$ of the total moduli space
5656: $\scrm$ of pseudoholomorphic curves $X$ as the closure of $\scrm$ in $\barm^{st}$.
5657: \end{defi}
5658:
5659: Note that every fiber $\barm^{Gr}_J \deff \barm^{Gr} \cap \bigl(\pr_{\!\!\scrj}
5660: ^{st} \bigr)\inv (J)$ is compact. Note also that in general $\barm^{st} \not=
5661: \barm^{Gr}$. This means that there are stable curves $[C,u,J] \in \barm^{st}$
5662: which can not be reached from $\scrm$.
5663:
5664:
5665: \newsubsection[5.2c] {The cycle topology for pseudoholomorphic curves} The
5666: Gromov compactness theorem gives a precise description of the behavior of {\sl
5667: parameterized} pseudoholomorphic curves at ``infinity'' of the total moduli
5668: space. However, what we are really interested in is not a pseudoholomorphic map
5669: $u:C \to X$ itself but rather the image $u(C)\subset X$, \ie a {\sl
5670: non-pa\-ra\-me\-te\-rized} pseudoholomorphic curve. The natural space
5671: where non-pa\-ra\-me\-te\-rized curves ``live'' is the space $\scrz_2(X)$ of
5672: 2-currents on the ambient manifold $X$. Recall that $\scrz_2(X)$ is the dual space
5673: to the space $C^\infty(X, \Lambda^2X)$ of smooth 2-forms on $X$ (see \eg
5674: \cite{Gr-Ha}, {\sl Chapter 3}).
5675:
5676:
5677: \newdefi{def5.2c.1} Let $X$ be a manifold, $C$ an abstract nodal curve with the
5678: smooth boundary $\d C$ such that $\barr C \deff C \cup \d C$ is compact, and $u:
5679: C \to X$ a map which is $L^{1,p}$-smooth up to boundary. Define the {\sl cycle
5680: $u[C]$ associated with the map $u: C \to X$} as the current whose pairing with
5681: a smooth $2$-form $\phi$ on $X$ equals $\< u(C), \phi \> \deff \int_C u^*\phi$. In
5682: this case we also say that $u[C]$ is {\sl represented by the map $u: C \to X$}.
5683:
5684: If additionally $u: C \to X$ is $J$-holomorphic \wrt some an almost complex
5685: structure on $X$, we call $C' \deff u[C]$ an {\sl $J$-holo\-mor\-phic curve $C$
5686: \underline{\it in} $X$}. In this case we say that the $J$-holomorphic curve $(C,
5687: u)$ \underline{\it over} $X$ and the map $u:C\to X$ {\sl represent the curve
5688: $C'= u[C]$}. The set $u(C)$ is called the {\sl support of $C'$} and denoted
5689: by $\supp(C')$.
5690:
5691: A curve $C'$ in $X$ is {\sl non-multiple} if it can be represented by a
5692: non-multiple pseudoholomorphic map $u:C\to X$ (see {\sl Definitions
5693: \ref{def1.2.2}} and {\sl\ref{def5.2.6}}). In this case we identify the set $u(C)$
5694: and the current $u[C]$ and use the same notation $u(C)$.
5695:
5696: A sequence of cycles $u_n[C_n]$ {\sl converges to a cycle $u_\infty[C_\infty]$}
5697: if $\< u_n(C_n), \phi \>$ converges to $\< u_\infty(C_\infty), \phi \>$ for any
5698: smooth smooth $2$-form $\phi$ on $X$. In other words, the cycle topology is the
5699: topology induced from the space of currents $\scrz_2(X)$.
5700: \end{defi}
5701:
5702: \smallskip
5703: \newlemma{lem5.2c.1} \sli Let $(X,J)$ be an almost complex manifold
5704: and $(C, u)$, $(C', u')$ closed $J$-holomorphic curves over $X$. Assume that
5705: \begin{itemize}
5706: \item $C$ and $C'$ are parameterized by the same closed surface $S$;
5707: \item $(C, u)$ contains no multiple and ghost components;
5708: \item the associated cycles $u[C]$ and $u'[C']$ coincide.
5709: \end{itemize}
5710: Then $(C, u)$ are $(C', u')$ equivalent.
5711:
5712: \slii Let $J_n$ be a sequence of continuous almost complex
5713: structures on $X$ which converges to an almost complex structure $J_\infty$ in the
5714: $C^0$-topology, and $(C_n, u_n)$ a sequence of stable $J_n$-holomorphic closed
5715: curves over $X$ which converges to $(C_\infty, u_\infty)$ in the Gromov topology.
5716: Then $u_n[C_n]$ converges to $u_\infty[C_\infty]$ in the cycle topology.
5717:
5718: \sliii Let $J_n$ be a sequence of continuous almost complex
5719: structures on $X$ which converges to an almost complex structure $J_\infty$ in the
5720: $C^0$-topology, $(C_n, u_n)$ a sequence of stable $J_n$-holomorphic closed
5721: curves over $X$, and $(C_\infty, u_\infty)$ a parameterized $J_\infty$-holomorphic
5722: curve. Assume that
5723: \begin{itemize}
5724: \item $u_n[C_n]$ converges to $u_\infty[C_\infty]$ in the cycle topology;
5725: \item $C_n$ and $C_\infty$ are parameterized by the same closed surface $S$;
5726: \item $(C_\infty, u_\infty)$ contains no multiple and ghost components;
5727: \item $J_\infty$ is $C^1$-smooth.
5728: \end{itemize}
5729: Then $(C_n, u_n)$ converges to $(C_\infty, u_\infty)$ in the Gromov topology.
5730: \end{lem}
5731:
5732: \proof {\sl Part \slip}. The hypotheses on $(C, u)$ and $(C', u')$ imply
5733: that $(C', u')$ also contains no multiple and ghost components. The claim then
5734: follows from the unique continuation property of pseudoholomorphic curves
5735: (see \lemma{lem1.2.4}).
5736:
5737: {\sl Part \sliip}. This follows from the definition of the Gromov
5738: topology and the description of the convergence at nodes given in
5739: {\sl Step 0)} of the proof of \lemma{lem5.2.2} below.
5740:
5741: {\sl Part \sliiip}.
5742: Fix a $J_\infty$-Hermitian metric $h$ on $X$. Let $\omega$ be the associated
5743: 2-form, $\omega(v,w) \deff h(J_\infty v,w)$. Then the structures $J_n$ are
5744: $\omega$-tame for $n\gg1$.
5745: Note that even if $\omega$ is apriori only continuous and not closed, the notion
5746: of $\omega$-tameness is still meaningful. Moreover, the $\omega$-tameness provides
5747: a uniform bound of $h$-area of $u_n[C_n]$. Consequently, some subsequence
5748: $(C_{n'}, u_{n'})$ of $(C_n, u_n)$ converges in the Gromov topology to
5749: a stable $J_\infty$-holomorphic curve $(C'_\infty, u'_\infty)$. The hypotheses
5750: of the corollary imply that $(C'_\infty, u'_\infty)$ is equivalent to
5751: $(C_\infty, u_\infty)$ and the result follows.\qed
5752:
5753:
5754: \state Remark. The meaning of \lemma{lem5.2c.1} is that, in the absence of
5755: multiple and ghost components, the notions of pseudoholomorphic curves {\sl over}
5756: $X$ and {\sl in} $X$ essentially coincide. The same also holds for the Gromov
5757: and the cycle topologies. Note also that several authors (\cite{Ye},
5758: \cite{Pa-Wo}, \cite{Hum}) considered a weaker version of the Gromov compactness
5759: theorem where the cycle topology is used instead of the Gromov one.
5760:
5761:
5762: \newdefi{def5.2c.2} Define the {\sl cycle compactification $\barm$} of the total
5763: moduli space $\scrm$ as the set of pairs $(C,J)$, where $J \in \scrj$ and $C$ is a
5764: $J$-holomorphic curve {\it in} $X$ which considered as the cycle $u[C']$
5765: represented
5766: by some $J$-holomorphic map $u: C' \to X$. Equip the space $\barm$ with the
5767: {\sl cycle topology} in which a sequence $(C_n, J_n)$ converges to $(C_\infty,
5768: J_\infty)$ if $J_n$ converges to $J_\infty$ in the $C^\ell$-topology and $C_n$
5769: converges to $C_\infty$ in the sense of \refdefi{def5.2.7}. Denote by $\pr_{\!\!
5770: \scrj}$ the natural projection $\pr_{\!\!\scrj}: (C,J) \in \barm \mapsto J \in
5771: \scrj$. Define the natural projection $\pr^{Gr}:\barm^{Gr} \to \barm$ by
5772: $\pr^{Gr}: (C, u,J) \in \barm^{Gr} \mapsto u[C]\in \barm$.
5773: \end{defi}
5774:
5775:
5776: \newdefi{def5.2c.3} A {\sl normal parameterization} of a $J$-holomorphic curve
5777: $C$ in $X$ is given by a Riemann surface $S$ (possibly not connected) and a map $u:
5778: S \to X$ such that
5779: \begin{itemize}
5780: \item[1)] $u[S]=C$;
5781: \item[2)] $u$ is $J$-holomorphic and $L^{1,p}$-smooth up to boundary;
5782: \item[3)] the restriction of $u$ to every connected component of $S$ is
5783: non-multiple; in particular, there are no ghost components, \ie $u$ is
5784: non-constant on every connected component of $S$;
5785: \item[4)] the number of boundary circles of $S$ is as small as possible.
5786: \end{itemize}
5787: \end{defi}
5788:
5789: \state Remark. Without condition (3) one could add new ghost spheres to $S$
5790: and make the Euler characteristic $\Chi(S)$ arbitrarily large. Condition (4)
5791: excludes the possibility of dividing components of $S$ into pieces which also
5792: allows to increase $\Chi(S)$.
5793:
5794: \newlemma{lem5.2c.2} Let $J$ be a $C^1$-smooth almost complex structure on $X$,
5795: $C$ an abstract nodal curve, and $u: C\to X$ a $J$-holomorphic map which is an
5796: imbedding near the boundary $\d C$. Then, up to a diffeomorphism, there exists
5797: a unique normal parameterization $\ti u: S \to X$ of $u[C]$.
5798: \end{lem}
5799:
5800: \proof Let $C= \cup_i C_i$ be the decomposition of $C$ into irreducible
5801: components. Denote by $m_i$ the degree $u$ on $C_i$. This means that
5802: \begin{itemize}
5803: \item $m_i=0$ if $C_i$ is a ghost component, \ie $u$ is constant on $C_i$;
5804: \item $m_i$ is the number of points in the preimage $u\inv(x)$ for a
5805: generic $x\in u(C_i)$ otherwise.
5806: \end{itemize}
5807: For every non-zero $m_i$, denote by $S_i$ the normalization of the image $u(C_i)$.
5808: Denote by $\ti u_i: S_i \to u(C_i)$ the corresponding normalization maps. In
5809: particular, $m_i=1$ for every non-closed component $C_i$ and in this case $S_i$
5810: is the normalization of $C_i$. Define $S$ as the disjoint union of the surfaces
5811: $S_i$, each taken $m_i$ times. Let $\ti u: S \to X$ be the map which coincides
5812: with the composition $\ti u_i: S_i \to u(C_i) \hook X$ on every copy of $S_i$.
5813: One can see that $\ti u: S \to X$ is a normal parameterization of $u[C]$.
5814:
5815: The uniqueness of such a normal parameterization follows from \lemma{lem1.2.4}
5816: \qed
5817:
5818: \state Remark. Let us give an example showing that the condition on the behavior
5819: of $u$ at the boundary imposed in \lemma{lem5.2c.2} is necessary. Define curves
5820: $C'$ and $C''$ as the disjoint unions $C'\deff \{ z\in \cc: |z|<2\} \sqcup \{ z\in
5821: \cc: 1<|z|<3\}$ and $C''\deff \{ z\in \cc: |z|<3\} \sqcup \{z\in \cc:
5822: 1<|z|<2\}$. Let $u': C' \to \cc$ and $u'': C'' \to \cc$ be the maps which are the
5823: standard imbeddings on every component of $C'$ and $C''$. Then
5824: obviously $(C',u')$ and $(C'',u'')$ are not equivalent in the sense of
5825: \refdefi{def5.2.9} and define non-equivalent normal parameterizations
5826: of $u'[C'] =u''[C'']$.
5827:
5828:
5829: \newcorol{cor5.2c.3} Under the hypotheses of \lemma{lem5.2c.2}, the curve $u[C]$,
5830: considered as a current $u[C]\in \scrz_2(X)$, admits a unique representation in
5831: the form
5832: $$
5833: \textstyle
5834: u[C]= \sum_i m_i u_i[C_i],
5835: $$
5836: where the $u_i: C_i \to X$ are $J$-holomorphic maps and the $u_i(C_i)$ are
5837: the irreducible components of $\supp(u[C])$.
5838: \end{corol}
5839:
5840: The corollary ensures that the notions of an irreducible component and the
5841: multiplicity of a closed pseudoholomorphic curve {\it in} $X$ are well-defined.
5842:
5843: \medskip
5844: The importance of the notion of a normal parameterization lies in the fact that it
5845: allows us to define a natural stratification of the cycle compactification $\barm$
5846: of the total moduli space. Let $S$ be a given connected real surface $S$ of genus
5847: $g$ and $[C] \in \sfh_2(X, \zz)$ a homology class, and $\barm=\barm(S, X, [C])$
5848: the cycle compactification of the total space $\scrm=\scrm(S, X, [C])$ of
5849: irreducible pseudoholomorphic curves of genus $g$ in the homology class $[C]$.
5850: Take $(C,J) \in \scrm(S, X, [C])$ and consider a normal parameterization $u': S'
5851: \to X$ of $C$. Let $C= \sum_i m_i C_i$ be the decomposition into irreducible
5852: components in the sense of \refcorol{cor5.2c.3}. Restricting $u'$ to
5853: appropriate connected components we obtain normal parameterizations $u'_i: S'_i
5854: \to X$ of the corresponding $C_i$.
5855:
5856: \newdefi{def5.2c.4} The {\sl topological type of a component $C_i$} is the triple
5857: $(S'_i, m_i, [C_i])$, where $[C_i]$ denotes the homology class of $C_i$. The {\sl
5858: topological type $\bftau$ of a curve $(C,J) \in \scrm(S, X, [C])$} is the sequence
5859: of all topological types of components $(S'_i, m_i, [C_i])$ defined up to
5860: permutation.
5861: \end{defi}
5862:
5863: \newlemma{lem5.2c.4} \sli The space $\scrm_\bftau$ of pseudoholomorphic curves
5864: $(C,J) \in \barm(S, X, [C])$ of a given topological type $\tau$ is a
5865: $C^\ell$-smooth Banach manifold. The natural projection $\pr_{\!\!\scrj}: \scrm
5866: _\bftau \to \scrj$ is a $C^\ell$-smooth Fredholm map.
5867:
5868: \slii The space $\barm=\barm(S, X, [C])$ is the union of subspaces
5869: $\scrm_\bftau$.
5870: \end{lem}
5871:
5872: \proof
5873: The decomposition $C= \sum_i m_i C_i$ of every $(C,J) \in \scrm_\bftau$
5874: shows that $\scrm_\bftau$ is the fiber product of the spaces $\scrm(S'_i, X,
5875: [C_i])$ over all triples $(S'_i, m_i, [C_i]) \in \bftau$ taken over the space
5876: $\scrj \!\!$,
5877: $$
5878: \textstyle
5879: \scrm_\bftau=\prod_{(S'_i, m_i, [C_i]) \in \bftau }
5880: \scrm(S'_i, X, [C_i]) \bigm/ \!\!\!\scrj\!\!.
5881: $$
5882: Checking the transversality condition, one obtains the desired differentiable
5883: structure on $\scrm_\bftau$.
5884:
5885: The second assertion of the lemma is obvious. \qed
5886:
5887: \smallskip
5888:
5889:
5890: \newsubsection[5.2a]{Fine apriori estimates for convergence at a node}
5891: For the purpose of this paper we need a refined version of the {\sl Second
5892: apriori estimate} given in \cite{Iv-Sh-3}, {\sl Lemma 3.4}. This gives a
5893: precise description with estimates of the Gromov convergence in neighborhoods
5894: of the contracted circles.
5895:
5896: \newlemma{lem5.2.2} Let $X$ be a compact manifold $X$, $J^*$ a
5897: $C^{0,s}$-smooth almost complex structure on $X$ with $s>0$, and $h$ a metric on
5898: $X$. Then there exist constants $\epsi=\epsi(X,h, J^*, s)>0$ and $C <\infty$
5899: such that for any $C^{0,s}$-smooth almost complex structure $J$ with
5900: \begin{equation}
5901: \norm{J -J^*}_{C^{0,s}(X)} \le \epsi
5902: \eqqno(5.2a.1)
5903: \end{equation}
5904: and any $J$-holomorphic map $u: Z(0, l) \to X$ the condition
5905: \begin{equation}
5906: \norm{du}_{L^2(Z_k)} \le \epsi \quad \text{for any }k\in[0, l-1]
5907: \eqqno(5.2.3)
5908: \end{equation}
5909: implies the uniform estimate
5910: \begin{equation}
5911: \norm{du}^2_{L^2(Z_k)} \le
5912: C \cdot e^{-2k} \cdot \norm{du}^2_{L^2(Z(0, 2))} +
5913: C \cdot e^{-2 (l-k)} \cdot\norm{du}^2_{L^2(Z(l-2, l))}
5914: \eqqno(5.2.4)
5915: \end{equation}
5916: for any $k\in[1, l-2]$.
5917: \end{lem}
5918:
5919:
5920:
5921: \proof \step0.\. {\sl Lemma 3.3} in \cite{Iv-Sh-3} states that
5922: under hypotheses of the lemma one has a ``local'' estimate
5923: \begin{equation}
5924: \norm{du}^2_{L^2(Z_k)} \le \msmall{\gamma \over 2}
5925: \left( \norm{du}^2_{L^2(Z_{k-1})} +
5926: \norm{du}^2_{L^2(Z_{k+1})} \right ) \qquad \text{for any $k\in[1, l-2]$}
5927: \eqqno(5.2.5)
5928: \end{equation}
5929: with a universal constant $\gamma <1$. Then in {\sl Corollary 3.4} in
5930: \cite{Iv-Sh-3} it is shown that \eqqref(5.2.5) implies the estimate
5931: \begin{equation}
5932: \norm{du}^2_{L^2(Z_k)} \le
5933: e^{-2\alpha (k-1)} \cdot \norm{du}^2_{L^2(Z(0, 2))} +
5934: e^{-2\alpha (l-2-k)} \cdot\norm{du}^2_{L^2(Z(l-2, l))}
5935: \eqqno(5.2.6)
5936: \end{equation}
5937: for any $k\in[1, l-2]$ with a constant $\alpha >0$ related to $\gamma$
5938: by $\gamma = {1\over \cosh(2\alpha)}$.
5939:
5940: \state Remark. Note that in the proof of the estimates \eqqref(5.2.5) and
5941: \eqqref(5.2.6) are proven in \cite{Iv-Sh-3} under the following assumption:
5942: It is supposed that $J^*$ and $J$ in question are only continuous and that
5943: $\norm{J- J^*}_{C^0(X)} \le \epsi'$ for some $\epsi'=\epsi'(X,J^*,h)>0$
5944: independent of $J$.
5945:
5946: \medskip
5947: From the relation $\gamma = {1\over \cosh(2\alpha)}$ we see that the smaller the $
5948: \gamma$ we have the bigger the $\alpha$ in
5949: \eqqref(5.2.6) we obtain. For our purpose it would be sufficient to prove
5950: estimate \eqqref(5.2.5) with the parameter $\gamma^* \deff {1\over \cosh2}$.
5951: Note however that in the ``ideal'' case when $(X,J,h)$ is $\cc^n$ with the
5952: standard complex and Hermitian structures, $\gamma^*$ is exactly the best
5953: possible constant, see the proof of {\sl Lemma 3.3} in \cite{Iv-Sh-3}. Thus one
5954: can not expect that estimate \eqqref(5.2.5) holds with {\sl uniform} $\gamma \le
5955: \gamma^*$. The idea is to consider \eqqref(5.2.5) with parameters $\gamma_k$
5956: depending on $k$ and to estimate the difference $\gamma_k -\gamma^*$.
5957:
5958: \medskip\noindent
5959: \step1.\. {\sl Under hypotheses of the lemma, for any $k\in[1, l-2]$,
5960: one has the estimate
5961: \begin{equation}
5962: \norm{du}^2_{L^2(Z_k)} \le
5963: \msmall{\gamma_k \over 2}
5964: \cdot
5965: \left( \norm{du}^2_{L^2(Z_{k-1})} +
5966: \norm{du}^2_{L^2(Z_{k+1})} \right )
5967: \eqqno(5.2.7)
5968: \end{equation}
5969: for
5970: \begin{equation}
5971: \gamma_k \deff \gamma^* +
5972: C_1\cdot\left(e^{-\alpha s k} + e^{-\alpha s (l-k)} \right)
5973: \eqqno(5.2a.2)
5974: \end{equation}
5975: with the parameter $\alpha >0$ as in \step0 and some constant $C_1$
5976: depending only on $X$, $h$, $J^*$, and $s$.}
5977:
5978: \smallskip
5979: While proving this estimate we shall denote by $C$ a constant whose particular
5980: value is not important and which may not be the same in different formulas.
5981: The main condition is that these constants are {\sl uniform}, \ie independent
5982: of $J$, $u$, and $l$, and depend only on $X$, $h$, $J^*$, and $s$.
5983:
5984:
5985: Estimates \eqqref(5.2.3) and \eqqref(5.2.6) together with apriori
5986: estimates show that
5987: \begin{equation}
5988: \diam(u(Z(k-1, k+2)) \le C \cdot
5989: \left(e^{-\alpha k} + e^{-\alpha (l-k)} \right).
5990: \end{equation}
5991: Consequently, due to a uniform H\"older $C^{0,s}$-estimate on $J$,
5992: for the oscillation of $J$ on the image $u(Z(k-1, l-k))$ we obtain
5993: \begin{equation}\eqqno(5.2a.2a)
5994: \osc(J, u(Z(k-1, k+2))) \le C \cdot
5995: \left(e^{-\alpha s k} + e^{-\alpha s (l-k)} \right).
5996: \end{equation}
5997: This implies that in a neighborhood of each $u(Z(k-1, k+2))$ there exist an
5998: integrable structure $J\st$ and a flat (\ie Euclidean) metric $h\st$ such that
5999: \begin{equation}
6000: \norm{J -J\st}_{L^\infty(u(Z_k))} + \norm{h -h\st}_{L^\infty(u(Z_k))}
6001: \le C\cdot \left(e^{-\alpha s k} + e^{-\alpha s (l-k)} \right).
6002: \end{equation}
6003: Using this we obtain estimates
6004: \begin{align}
6005: \norm{\dbar\st u}_{L^2(Z_k)}
6006: = \norm{\dbar\st u - \dbar_J u}_{L^2(Z_k)}
6007: &\le \norm{J -J\st}_{L^\infty(u(Z_k))} \cdot \norm{du}_{L^2(Z_k)} \le
6008: \notag
6009: \\
6010: &\le C\cdot \left(e^{-\alpha s k} + e^{-\alpha s (l-k)} \right);
6011: \end{align}
6012: \begin{equation}
6013: \Bigl| \norm{du}_{L^2(Z_k), h}
6014: -\norm{du}_{L^2(Z_k), h\st} \Bigr| \le
6015: C\cdot \left(e^{-\alpha s k} + e^{-\alpha s (l-k)} \right)
6016: \cdot \norm{du}_{L^2(Z_k), h}.
6017: \end{equation}
6018: In particular, we can use $h\st$ instead of $h$ in our estimates.
6019:
6020: Now consider $U$ as a subset on $\cc^n$ with the standard $J\st$ and $h\st$.
6021: Then we can find $u_\dbar\in L^{1,2}(Z(k-1, k+2), \cc^n)$ such that $\dbar\st
6022: u_\dbar= \dbar\st u$ and
6023: \begin{equation}
6024: \norm{du_\dbar}_{L^2(Z(k-1, k+2))} \le C \norm{\dbar\st u}
6025: _{L^2(Z(k-1, k+2))}.
6026: \eqqno(5.2a.3)
6027: \end{equation}
6028: Set $u_\scro \deff u -u_\dbar$, so that $u_\scro$ is $J\st$-holomorphic.
6029: It follows that %Then we obtain
6030: \begin{equation}
6031: \norm{du_\scro}^2_{L^2(Z_k)} \le \msmall{\gamma^* \over2}
6032: \left( \norm{du_\scro}^2_{L^2(Z_{k-1})} +
6033: \norm{du_\scro}^2_{L^2(Z_{k+1})} \right).
6034: \eqqno(5.2.8)
6035: \end{equation}
6036: Together with the estimates on $u_\dbar$, \eqqref(5.2.8) implies
6037: \eqqref(5.2.7).
6038:
6039: \medskip\noindent
6040: \step2.\. {\sl There exist a uniform $k_0=k_0(X, h, J^*,s)$ and $A^\pm_k$,
6041: $k=k_0, \ldots,l-k_0 $ with the properties
6042:
6043: \sli $A^\pm_k$ are ``supersolutions'' of \eqqref(5.2.7), \ie
6044: \begin{equation}
6045: \eqqno(5.2a.4a)
6046: A^\pm_k \ge \msmall{\gamma_k \over2} (A^\pm_{k-1} + A^\pm_{k+1})
6047: \end{equation}
6048:
6049: \slii $A^\pm_k$ have the desired exponential decay}
6050: \begin{equation}
6051: \eqqno(5.2a.5a)
6052: A^+_k \le C\cdot e^{-2k},
6053: \qquad\qquad
6054: A^-_k \le C\cdot e^{-2(l-k)}.
6055: \end{equation}
6056:
6057: \sliii $\gamma_k <1$ for $k \in [k_0, l-k_0]$.
6058:
6059:
6060: \smallskip
6061: Fix $k^*\in \zz$ such that $l-1 \le k^* <l+1$, so that $k^* \approx {l\over2}$.
6062: Set
6063: \begin{align}
6064: A^+_k & \deff \cases
6065: e^{-2k -{1\over k}} & 0\le k\le k^* \\
6066: e^{-2k -{1\over k^*} + {1\over l-k}- {1\over l-k^*}}
6067: \hskip7pt& k^*\le k \le l
6068: \endcases \\
6069: A^- _k& \deff \cases
6070: e^{-2(l-k) -{1\over l-k^*} + {1\over k}-{1\over k^*}}
6071: & 0\le k\le k^* \\
6072: e^{-2(l-k) -{1\over l-k}} & k^*\le k \le l
6073: \endcases
6074: \end{align}
6075: Making the Taylor expansion in $k\inv$ we obtain
6076: $$
6077: \msmall{ 2 A^\pm_k \over A^\pm_{k-1} + A^\pm_{k+1}}=
6078: \cases
6079: \msmall { 1\over \cosh(2)}+
6080: \msmall { \sinh(2) \over \cosh^2(2)} \cdot k^{-2} + O(k^{-3})
6081: & \text{for }0< k \le k^*;
6082: \\
6083: \msmall { 1\over \cosh(2)}+
6084: \msmall { \sinh(2) \over \cosh^2(2)} \cdot (l-k)^{-2} + O((l-k)^{-3})
6085: & \text{for }k^*\le k <l;
6086: \endcases
6087: $$
6088: So the existence of the desired $k_0(s)$ follows from the asymptotic behavior
6089: $C_1 e^{-\alpha s k} = o(k^{-2})$ for $k\lrar \infty$.
6090:
6091:
6092: \medskip\noindent
6093: \step3. {\sl There exists a constant $C_2= C_2(X,h, J^*,s)$ such that
6094:
6095: %\begin{equation}
6096: $\qquad
6097: \norm{du}^2_{L^2(Z_k)} \le C_2\cdot \left(
6098: A^+_k \cdot \norm{du}^2_{L^2(Z(0,2))} +
6099: A^-_k \cdot\norm{du}^2_{L^2(Z(l-2,l))}
6100: \right)
6101: %\eqqno(5.2.9)
6102: $%\end{equation}
6103: \lineeqqno(5.2.9)
6104:
6105: \smallskip\noindent
6106: for any $k\in[k_0, l-k_0]$ with the uniform constant $k_0=k_0(s)$ chosen as
6107: above.}
6108:
6109:
6110: \smallskip
6111: Obviously, \eqqref(5.2.9) implies the claim of the lemma. Set
6112: $$
6113: A^*:= C_2 \cdot \left(A^+_k \cdot \norm{du}^2_{L^2(Z(0,2))} +
6114: A^-_k \cdot\norm{du}^2_{L^2(Z(l-2,l))} \right)
6115: $$
6116: and choose a constant $C_2$ so that
6117: $$
6118: A^*_{k_0} \ge \norm{du}^2_{L^2(Z_{k_0})}
6119: \qquad \text{and} \qquad
6120: A^*_{l-k_0} \ge
6121: \norm{du}^2_{L^2(Z_{l-k_0})}.
6122: $$
6123: Then by \eqqref(5.2.7) and \eqqref(5.2a.4a)
6124: \begin{equation}\notag
6125: \norm{du}^2_{L^2(Z_k)} - A^*_k \le
6126: \msmall{\gamma_k \over2} \cdot
6127: \left(\norm{du}^2_{L^2(Z_{k-1})} - A^*_{k-1}
6128: +\norm{du}^2_{L^2(Z_{k+1})} - A^*_{k+1} \right).
6129: \end{equation}
6130: Find $k_\max\in [k_0, l-k_0]$ realizing the maximum of
6131: $\norm{du}^2_{L^2(Z_k)} - A^*_k$. Then
6132: \begin{equation}
6133: \matrix%\format\l&\;\c\;&\l\\
6134: \norm{du}^2_{L^2(Z_{k_\max})} - A^*_{k_\max}
6135: &\le& \msmall{\gamma_{k_\max} \over2}
6136: \left(\norm{du}^2_{L^2(Z_{k_\max-1})} - A^*_{k-1}
6137: +\norm{du}^2_{L^2(Z_{k+1})} - A^*_{k_\max+1} \right)
6138: \cr\noalign{\vskip3pt}
6139: &\le&
6140: \gamma_{k_\max} \cdot( \norm{du}^2_{L^2(Z_{k_\max})} - A^*_{k_\max}).
6141: \endmatrix
6142: \end{equation}
6143: Since $\gamma_{k_\max}<1$, the last inequality holds only if
6144: $\norm{du}^2_{L^2(Z_{k_\max})} \le A^*_{k_\max}$. Thus
6145: \begin{equation}
6146: \norm{du}^2_{L^2(Z_k)} \le A^*_k
6147: \qquad\text{for any }k\in [k_0, l-k_0].
6148: \end{equation}
6149: This finishes the proof. \qed
6150:
6151:
6152: \smallskip
6153: \newthm{thm5.2.3} Let $J^*$ be a $C^{0,s}$-smooth almost complex structure on the
6154: ball $B\subset \rr^{2n}$ with $0<s<1$. Then there exists $\epsi^*= \epsi^* (J^*,
6155: s)$ with the following property. For any almost complex structure $J$ on $B$ with
6156: $\norm{J-J^*}_{C^{0,s}(B)} \le \epsi^*$ and any $J$-holomorphic map
6157: $u: Z(0,l) \to B(\half)$ with $l\ge3$ satisfying the condition
6158: $$
6159: \norm{du}_{L^2(Z_k)} \le \epsi^*\qquad\text{for any }k\in[1,l]
6160: $$
6161: there exist a {\sl linear} complex structure $J\st$ in $\rr^{2n}$ and vectors
6162: $v^+, v^0, v^- \in \rr^{2n}$ such that
6163: \begin{multline}\eqqno(5.2.10)
6164: \left\Vert
6165: u - \left( e^{-t+J\st\theta}v^+ + v^0 + e^{t-J\st\theta}v^- \right)
6166: \right\Vert^2_{L^{1,2}(Z_k)} \le
6167: \\
6168: \le C^* \cdot\bigl(k \;e^{-2(1+s)k}\norm{du}^2_{L^2(Z(0,2))} +
6169: (l-k)\;e^{-2(1+s)(l-k)}\norm{du}^2_{L^2(Z(l-2,l))}
6170: \bigr)\
6171: \end{multline}
6172: for any $k=1,\ldots, l-1$ with a constant $C^*=C^*(J^*,s) <\infty$ independent of
6173: $J$, $l$, and $u$.
6174: \end{thm}
6175:
6176:
6177: \proof In fact, we prove that for the constant $\epsi^*$ one can take the
6178: $\epsi$ from \lemma{lem5.2.2}. The proof also exploits the same ideas which were
6179: used in the proof of that lemma.
6180:
6181: \smallskip\noindent {\sl Step 1}.
6182: Let $J$ and $u: Z(0,l) \to B(\half)$ be as in the hypotheses of the theorem. For
6183: $k= 1,\ldots,l$, let $x_k \in B(\half)$ be the average value of $u$ on $Z_k$ \wrt
6184: the cylinder metric, \ie
6185: $$
6186: x_k \deff \bint_{Z_k} u \deff
6187: \msmall{1\over 2\pi}\int_{(t,\theta) \in Z_k} u(t,\theta)
6188: dt\,d\theta.
6189: $$
6190: Define the complex structures $J_k$ by $J_k\deff J(u(k,0))$. We consider every
6191: $J_k$ as a {\sl linear} complex structure in $\rr^{2n}$, \ie constant in $x\in
6192: \rr^{2n}$. Further, any $k=1,\ldots,l$ we define the metric $g_k$ setting $g_k(v,
6193: w) \deff \half (g\st(v,w) + g\st(J_kv, J_kw))$, where $g\st$ denotes the standard
6194: Euclidean metric in $\rr^{2n}$. Then $g_k$ are linear in the same sense as $J_k$.
6195: In computing various norms related to $Z_k$ or $Z(k-2, k+1)$, we shall use the
6196: metric $g_k$ without indicating this in the notation. Observe that all $g_k$ are
6197: equivalent since the $J_k$ are uniformly bounded. Further, convergence of
6198: $J_{k_\nu}$ implies convergence of $g_{k_\nu}$.
6199:
6200: For any $k=1,\ldots,l$ there exist uniquely defined vectors
6201: $v^+_k, v^0_k, v^-_k \in \rr^{2n}$ such that for the function
6202: $$
6203: v_k(t, \theta) \deff e^{-t+J_k\theta}v^+_k
6204: +v^0_k+ e^{t-J_k\theta}v^-_k
6205: $$
6206: the norm $\norm{u -v_k}_{L^{1,2}(Z_k)}$ (computed with $g_k$) attains the minimum.
6207:
6208: We claim that under the hypotheses of the theorem there exist a constant $C_1=
6209: C_1(J^*, s)$ and and an integer $k_0=k_0(J^*, s)$ such that for any integer
6210: $k=k_0,\ldots, l-k_0$
6211: \begin{multline}\eqqno(5.2.11)
6212: \norm{u -v_k}^2_{L^{1,2}(Z_k)} +
6213: \norm{v_{k-1} -v_k}_{L^{1,2}(Z_k)}^2 +
6214: \norm{v_{k+1} -v_k}_{L^{1,2}(Z_k)}^2 \le
6215: \\
6216: \le \gamma_s \cdot\left(
6217: \norm{u -v_{k-1}}^2_{L^{1,2}(Z_{k-1})} +
6218: \norm{u -v_{k+1}}^2_{L^{1,2}(Z_{k+1})} \right) +
6219: \\
6220: +C_1 \cdot \bigl(e^{-2(1+s)k} \norm{du}^2_{L^2(Z(0,2))} +
6221: e^{-2(1+s)(l-k)}\norm{du}^2_{L^2(Z(l-2,l))}\bigr) \
6222: \end{multline}
6223: with the parameter $\gamma_s \deff {1\over \cosh(2+2s)}\cdot\;$
6224: Assuming the contrary, there must exist sequences of %involved objects
6225: \begin{itemize}
6226: \item integers $l_\nu\lrar \infty$;
6227: \item integers $k_\nu \lrar \infty$ with $l_\nu -k_\nu \lrar \infty$;
6228: \item structures $J_\nu$ in $B$ with $\norm{J_\nu -J^*}_{C^{0,s}(B)} \le \epsi^*$;
6229: \item $J_\nu$-holomorphic maps $u_\nu: Z(0,l_\nu) \to B(\half)$ with
6230: $\sup_{k=0,\ldots,l_\nu} \norm{du_\nu}_{L^2(Z_k)} \lrar 0$
6231: \end{itemize}
6232: with the following property. For the points $x_{\nu,k} \deff \bint_{Z_k} u_\nu$,
6233: the linear complex structure $J_{\nu,k} \deff J_\nu(x_{\nu,k})$, the corresponding
6234: metrics $g_{\nu,k}$, and vectors $v^+_{\nu,k}, v^0_{\nu,k}, v^-_{\nu,k} \in \rr
6235: ^{2n}$ constructed as above for every $u_\nu \ogran_{Z_k}$ with $k=1,\ldots,
6236: l_\nu$, at the position $k=k_\nu$ we obtain the inequality in the opposite
6237: direction:
6238: \begin{multline}\eqqno(5.2.12)
6239: \norm{u_\nu -v_{\nu,k_\nu}}^2_{L^{1,2}(Z_{k_\nu})} +
6240: \norm{v_{k_\nu-1} -v_{k_\nu}}^2_{L^{1,2}(Z_{k_\nu})} +
6241: \norm{v_{k_\nu+1} -v_{k_\nu}}^2_{L^{1,2}(Z_{k_\nu})} \ge
6242: \\
6243: \ge \gamma_s \cdot\left(
6244: \norm{u_\nu -v_{\nu,k_\nu-1}}^2_{L^{1,2}(Z_{k_\nu-1})} +
6245: \norm{u_\nu -v_{\nu,k_\nu+1}}^2_{L^{1,2}(Z_{k_\nu+1})} \right) +
6246: \\
6247: +\nu \cdot \bigl(e^{-2(1+s)k_\nu} \norm{du_\nu}^2_{L^2(Z(0,2))} +
6248: e^{-2(1+s)(l-k_\nu)}\norm{du_\nu}^2_{L^2(Z(l_\nu-2,l_\nu))}\bigr). \
6249: \end{multline}
6250:
6251:
6252: \smallskip
6253: Let us estimate the behavior of $u_\nu -v_{\nu,k}$ in $Z_k$ for $k\approx k_\nu$.
6254: Set
6255: $$
6256: A_{\nu, k} \deff e^{-k} \norm{du_\nu}_{L^2(Z(0,2))} +
6257: e^{-(l-k)}\norm{du_\nu}_{L^2(Z(l_\nu-2,l_\nu))}.
6258: $$
6259: Then by \lemma{lem5.2.2} we have $\norm{du_\nu}_{L^2(Z_k)} \le C\cdot A_{\nu,k}$.
6260: This yields a similar estimate on the diameter: $\diam(u_\nu(Z_k)) \le C \cdot
6261: A_{\nu,k}$, possibly with a new constant $C$. Further, for a linear complex
6262: structure $J'$ with the corresponding operator $\dbar' \deff \dbar_{J'}$ we obtain
6263: the pointwise estimate
6264: \begin{multline*}%\eqqno(5.2.13)
6265: \left|\dbar' u_\nu \right| =
6266: \left|\dbar' u_\nu - \dbar_{J_\nu} u_\nu \right| =
6267: \left| (\d_x u_\nu - J' \cdot \d_y u_\nu) -
6268: (\d_x u_\nu - J_\nu(u_\nu) \cdot \d_yu_\nu)\right|
6269: \\
6270: \le \left|J' - J_\nu \scirc u_\nu\right| \cdot \left|du_\nu \right|.
6271: \qquad
6272: \end{multline*}
6273: For $J_{\nu,k}$ this yields the estimate
6274: \begin{equation}\eqqno(5.2.14)
6275: \norm{\dbar_{J_{\nu,k}} u_\nu}_{L^2(Z(k-2,k+1))} \le
6276: C \bigl(\diam(u_\nu(Z(k-2,k+1))) \bigr)^s \cdot
6277: \norm{du_\nu}_{L^2(Z(k-2,k+1))}
6278: \le C'\cdot A^{1+s}_{\nu,k}.
6279: \end{equation}
6280:
6281: By construction, $J_{\nu,k}$ are uniformly bounded. This implies that we can
6282: represent $u_\nu\ogran_{Z_k}$ in the form $u_\nu\ogran_{Z_k} = w_{\nu,k} +
6283: f_{\nu,k}$, where $w_{\nu,k}$ is $J_{\nu,k}$-holomorphic and $f_{\nu,k}$
6284: is estimated as
6285: $$
6286: \norm{f_{\nu,k}} _{L^{1,2}(Z_k)} \le C\cdot A^{1+s}_{\nu,k}.
6287: $$
6288:
6289: \smallskip
6290: Define the positive $\eta_\nu$ by the relation
6291: $$
6292: \eta_\nu^2= \norm{u_\nu -v_{\nu,k_\nu}}^2_{L^{1,2}(Z_{k_\nu})} +
6293: \norm{v_{k_\nu-1} -v_{k_\nu}}^2_{L^{1,2}(Z_{k_\nu})} +
6294: \norm{v_{k_\nu+1} -v_{k_\nu}}^2_{L^{1,2}(Z_{k_\nu})}
6295: $$
6296: and set
6297: $$
6298: \ti u_\nu(t, \theta) \deff
6299: \msmall{1\over \eta_\nu} u_\nu(t+k_\nu, \theta),
6300: \qquad\qquad
6301: \ti w_{\nu,k}(t, \theta) \deff
6302: \msmall{1\over \eta_\nu} w_{\nu,k+k_\nu}(t+k_\nu, \theta),
6303: $$
6304: $$
6305: \ti f_{\nu,k}(t, \theta)
6306: \deff \msmall{1\over \eta_\nu} f_{\nu,k+k_\nu}(t+k_\nu, \theta),
6307: \qquad
6308: \ti J_{\nu, k} \deff J_{\nu, k+k_\nu},
6309: \qquad
6310: \ti v^\iota_{\nu,k} \deff
6311: \msmall{1\over \eta_\nu} v^\iota_{\nu,k+k_\nu},\; \iota=+,0,-,
6312: $$
6313: In other words, we shift all the picture from $Z_{k_\nu}$ to $Z_0$ and
6314: rescale the maps $u_\nu$, the vectors $v^\iota_{\nu,k}$, $\iota=+,0,-$, and so on
6315: in a way as to make the left hand side of \eqqref(5.2.12) equal to $1$.
6316:
6317: It follows from \eqqref(5.2.12) that $A^{1+s}_{\nu, k+k_\nu} \le C \nu^{-\half}
6318: \eta_\nu =o(\eta_\nu)$ for any fixed $k$. Consequently, $\norm{\ti f_{\nu, k}}
6319: _{L^{1,2} (Z_k)} \lrar 0$ for any fixed $k$ and $\nu\lrar \infty$. This implies
6320: that the norms $\norm{\ti w_{\nu, k} - \ti v_{\nu, k}}_{L^{1,2}(Z_k)}$ remain
6321: uniformly bounded in $\nu$ for any fixed $k$.
6322:
6323: Represent every $\ti w_{\nu, k}$ as the Laurent series
6324: \begin{equation}\eqqno(5.2.14a)
6325: \textstyle
6326: \ti w_{\nu, k}(t, \theta)= \sum_{m=-\infty}^{+\infty}
6327: e^{m(-t + J_{\nu, k}\theta)} w^m_{\nu, k}
6328: \end{equation}
6329: and denote by $\ti w'_{\nu, k}$ the sum of terms with $m=0, \pm1$, \ie
6330: $$
6331: \ti w'_{\nu, k}(t, \theta)\deff
6332: e^{t - J_{\nu, k}\theta} w^{-1}_{\nu, k} + w^0_{\nu, k} +
6333: e^{-t + J_{\nu, k}\theta} w^1_{\nu, k}.
6334: $$
6335: It follows from the construction of $\ti v_{\nu,k}$ that
6336: \begin{equation}\eqqno(5.2.15)
6337: \norm{\ti w'_{\nu, k} - \ti v_{\nu,k} }_{L^{1,2}(Z_k)}=
6338: O(\norm{\ti f_{\nu, k}}_{L^{1,2}(Z_k)}) \lrar 0
6339: \end{equation}
6340: for any fixed $k$. Indeed, $\ti v_{\nu,k}$, considered as a function in $\theta$,
6341: is a linear combination of a constant and the trigonometric functions $\cos\theta$
6342: and $\sin\theta$. So it is orthogonal to the remaining terms $e^{m(-t + J_{\nu, k}
6343: \theta)} w^m_{\nu, k}$, $|m|\ge2$. Thus $\ti w'_{\nu, k}$ is the best
6344: approximation of $\ti w_{\nu, k}$ by such linear combinations, whereas the
6345: difference $\ti w'_{\nu, k} - \ti v_{\nu,k}$ appears as the best approximation of
6346: $\ti f_{\nu, k}$.
6347:
6348: Since $\ti J_{\nu, 0}$ is bounded uniformly in $\nu$, there exists a subsequence,
6349: still indexed by $\nu$, which converges to a linear complex structure $\ti J$.
6350: It follows from the definition of $\ti J_{\nu, k}$ and the estimate on the
6351: diameter of $u_\nu(Z_k)$ for $k\approx k_\nu$ that for any fixed $k$
6352: the structures $\ti J_{\nu, k}$ also converge to $\ti J$.
6353:
6354: Now we show that, after going to a subsequence, $\ti u_{\nu, 0} - \ti v_{\nu, 0}$
6355: converges weakly in the $L^{1,2}(Z(-2,1))$-topology to a $\ti J$-holomorphic
6356: function, and that this convergence is strong in the $L^{1,2}(Z_0)$-topology. The
6357: inequality \eqqref(5.2.12) together with the choice of $\eta_\nu$ and the
6358: construction of $\ti u_{\nu, k}$ gives boundedness of the norms $\norm{\ti u_{\nu,
6359: 0} - \ti v_{\nu, 0} }_{L^{1,2}(Z(-2,1))}$ uniform in $\nu$. So the weak
6360: convergence follows. From \eqqref(5.2.14) and $A^{1+s}_{\nu,k} =o(\eta_\nu)$ we
6361: obtain the vanishing $\norm{\dbar_{\ti J _{\nu, 0}} \ti u_{\nu, 0} }_{L^{1,2}
6362: (Z(-2,1))} \lrar0$. The estimate \eqqref(5.2.15) and $\norm{\ti f_{\nu, k}}
6363: _{L^{1,2} (Z_k)} \lrar 0$ yield $\norm{\dbar_{\ti J _{\nu, 0}} \ti v_{\nu, 0} }
6364: _{L^{1,2} (Z(-2,1))} \lrar0$. Now the desired strong $L^{1,2}(Z_0)$-convergence
6365: follows from elliptic regularity of $\ti J _{\nu, 0}$. In the same way for $k=
6366: \pm1$ we obtain the weak $L^{1,2}$-convergence of $\ti u_{\nu, k} -\ti v_{\nu, k}$
6367: in $Z_k$.
6368:
6369: For $k=0,\pm1$, let $\ti u_k \deff \lim \ti u_{\nu, k} -\ti v_{\nu, k}$ be the
6370: limit functions obtained above. Since $\norm{\ti f_{\nu, k}} _{L^{1,2} (Z_k)}
6371: \lrar 0$, these are $\ti J$-holomorphic functions in $Z_k$, $\ti J = \lim
6372: \ti J_{\nu,k}$.
6373:
6374: Observe that the functions $\ti v_{\nu,\pm1} -\ti v_{\nu,0}$ are linear
6375: combinations of constants and the function $e^{\pm t}\cos\theta$, $e^{\pm t}\sin
6376: \theta$ which are uniformly bounded in the $L^{1,2}(Z_0)$-norm. Consequently,
6377: after taking a subsequence, we also obtain the strong $L^{1,2}$-convergence
6378: in $Z_0$. This implies the strong $L^{1,2}$-convergence in $Z(-2,1)$. Finally, from
6379: \eqqref(5.2.15) we conclude that the Laurent series for each $\ti u_k$ does not
6380: contain terms of degree $m=0,\pm1$, \ie a constant term and a multiple of
6381: $e^{\pm(-t + \ti J \theta)}$. This, in turn, implies that, first, the $\ti u_k(t,
6382: \theta)$ are restrictions to $Z_k$ of the same $\ti J$-holomorphic function
6383: $\ti u$, and second, $\lim \ti v_{\nu,-1} -\ti v_{\nu,0} = \lim \ti v_{\nu,1} -
6384: \ti v_{\nu,0} =0$.
6385:
6386: Substituting into \eqqref(5.2.12), we see that $\ti u$ satisfies the inequality
6387: \begin{equation}\eqqno(5.2.16)
6388: \norm{\ti u}^2_{L^{1,2}(Z_0)}
6389: \ge \gamma_s \cdot\left(
6390: \norm{\ti u}^2_{L^{1,2}(Z_{-1})} +
6391: \norm{\ti u}^2_{L^{1,2}(Z_1)} \right).
6392: \end{equation}
6393: On the other hand, the absence of the terms of degree $m=0,\pm1$ in the Laurent
6394: decomposition of type \eqqref(5.2.14a) for $\ti u$ implies the inequality
6395: \begin{equation}\eqqno(5.2.17)
6396: \norm{\ti u}^2_{L^{1,2}(Z_0)}
6397: \le \gamma_2 \cdot\left(
6398: \norm{\ti u}^2_{L^{1,2}(Z_{-1})} +
6399: \norm{\ti u}^2_{L^{1,2}(Z_1)} \right).
6400: \end{equation}
6401: with $\gamma_2 \deff {1\over \cosh(4)}$. This inequality is easily obtained for
6402: mutually orthogonal terms $\ti u^m e^{m(-t+\ti J \theta)}$. However, since $s<1$,
6403: $\gamma_2 < \gamma_s= {1\over \cosh(2 +2s)}$, which is a contradiction.
6404:
6405: This implies the validity of \eqqref(5.2.11) for all $k= k_0,\ldots, l-k_0$
6406: with $k_0$ independent of $J$, $l$, and $u$.
6407:
6408: \smallskip\noindent {\sl Step 2}.
6409: We now turn back to the proof of the theorem. To show that \eqqref(5.2.11)
6410: implies \eqqref(5.2.10), we set for $k=0, \ldots, l$
6411: \begin{multline}
6412: A'_k \deff k\cdot \msmall{\cosh(2+2s) \over \sinh(2+2s)} \cdot C_1 \cdot
6413: e^{-2(1+s)k} \norm{du}^2_{L^2(Z(0,2))}+
6414: \\
6415: + (l-k)\cdot \msmall{\cosh(2+2s) \over \sinh(2+2s)} \cdot C_1 \cdot
6416: e^{-2(1+s)(l-k)} \norm{du}^2_{L^2(Z(l-2,l))}.\
6417: \end{multline}
6418: Then $A'_k$ satisfies the equality
6419: $$
6420: A'_k = \msmall{\gamma_s \over 2} \cdot (A'_{k-1} + A'_{k+1})
6421: + C_1 \cdot \bigl(e^{-2(1+s)k} \norm{du}^2_{L^2(Z(0,2))} +
6422: e^{-2(1+s)(l-k)}\norm{du}^2_{L^2(Z(l-2,l))}\bigr).
6423: $$
6424: Consequently,
6425: \begin{multline}\eqqno(5.2.18)
6426: \norm{u -v_k}^2_{L^{1,2}(Z_k)} - A'_k +
6427: \norm{v_{k-1} -v_k}_{L^{1,2}(Z_k)}^2 +
6428: \norm{v_{k+1} -v_k}_{L^{1,2}(Z_k)}^2 \le
6429: \\
6430: \le \gamma_s \cdot\left(
6431: \norm{u -v_{k-1}}^2_{L^{1,2}(Z_{k-1})} - A'_{k-1} +
6432: \norm{u -v_{k+1}}^2_{L^{1,2}(Z_{k+1})} - A'_{k+1})\right).\quad
6433: \end{multline}
6434: As in the proof of \lemma{lem5.2.2}, \eqqref(5.2.18) implies the estimate
6435: \begin{multline}\eqqno(5.2.19)
6436: \norm{u -v_k}^2_{L^{1,2}(Z_k)} - A'_k \le
6437: \\
6438: \le C_2 \cdot \bigl(e^{-2(1+s)k} \norm{du}^2_{L^2(Z(0,2))} +
6439: e^{-2(1+s)(l-k)}\norm{du}^2_{L^2(Z(l-2,l))}\bigr)
6440: \end{multline}
6441: for all $k= k_0,\ldots, l-k_0$ with $k_0$ independent of $J$, $l$, and $u$.
6442: Substitution the definition of $A'_k$ yields
6443: \begin{multline}\eqqno(5.2.20)
6444: \norm{u -v_k}^2_{L^{1,2}(Z_k)} +
6445: \norm{v_{k-1} -v_k}_{L^{1,2}(Z_k)}^2 +
6446: \norm{v_{k+1} -v_k}_{L^{1,2}(Z_k)}^2 \le
6447: \\
6448: \le C_3 \cdot \bigl(k\cdot e^{-2(1+s)k} \norm{du}^2_{L^2(Z(0,2))} +
6449: (l-k) \cdot e^{-2(1+s)(l-k)}\norm{du}^2_{L^2(Z(l-2,l))}\bigr)
6450: \end{multline}
6451:
6452:
6453: \smallskip\noindent {\sl Step 3}. For concrete $J$, $l$, and $u$ as in the
6454: hypotheses of the theorem, find $k^*$ for which the right hand side of
6455: \eqqref(5.2.20) takes its minimum. Set $J\st \deff J_{k^*} = J(x_{k^*})$,
6456: and $v^\iota \deff v^\iota_{k^*}$, $\iota=-,0,+$, and $v(t,\theta) \deff v_{k^*}(t,
6457: \theta)$. In view of \eqqref(5.2.20), for the proof of the theorem it is
6458: sufficient to estimate $\norm{v_k - v}_{L^{1,2}(Z_k)}$.
6459:
6460: We do this by descending recursion starting from $k=k^*$. Assume that we have
6461: shown that
6462: \begin{equation}\eqqno(5.2.21)
6463: \norm{v_k - v}_{L^{1,2}(Z_k)} \le
6464: C_4 \, k^{\half}\, e^{-(1+s)k}\,\norm{du}_{L^2(Z(0,2))}
6465: \end{equation}
6466: for all $k=k_1+1,\ldots, k^*$ with the constant $C_4$ to be chosen below.
6467: By our choice of $k^*$, for $k=1,\ldots, k_1$ we obtain from
6468: \eqqref(5.2.20)
6469: $$
6470: \norm{v_k - v_{k+1}}_{L^{1,2}(Z_k)} \le
6471: 2\,C_3\, k^{\half}\, e^{-(1+s)k}\,\norm{du}_{L^2(Z(0,2))}.
6472: $$
6473: Observe that for any function $w(t,\theta)$ of the form
6474: $$
6475: w(t,\theta)= w^0
6476: + (e^t w^+_c + e^{-t} w^-_c) \cos(\theta)
6477: + ( e^t w^+_s + e^{-t} w^-_s) \sin(\theta)
6478: $$
6479: with constant vectors $w^0, w^\pm_c, w^\pm_s \in \rr^{2n}$---so are all our
6480: differences $v_k - v_{k'}$---we have the estimate
6481: $$
6482: \norm{w}_{L^{1,2}(Z_k)} \le e \cdot \norm{w}_{L^{1,2}(Z_{k+1})}.
6483: $$
6484: Applying this, we obtain
6485: \begin{align*}
6486: \norm{v_{k_1} - v}_{L^{1,2}(Z_{k_1})}
6487: &\le \norm{v_{k_1} - v_{k_1+1}}_{L^{1,2}(Z_{k_1})} +
6488: \norm{v_{k_1+1} -v }_{L^{1,2}(Z_{k_1})}
6489: \\
6490: \le& 2\,C_3\, k_1^{\half}\, e^{-(1+s)k_1} \,\norm{du}_{L^2(Z(0,2))}+
6491: e \cdot \norm{v_{k_1+1} -v }_{L^{1,2}(Z_{k_1+1})}
6492: \\
6493: \le& 2\,C_3\, k_1^{\half}\, e^{-(1+s)k_1} \,\norm{du}_{L^2(Z(0,2))}+
6494: C_4\, (k_1+1)^{\half}\, e^{1-(1+s)(k_1+1)}\,\norm{du}_{L^2(Z(0,2))}
6495: \\
6496: =&2\,C_3\, k_1^{\half}\, e^{-(1+s)k_1} \,\norm{du}_{L^2(Z(0,2))}+
6497: e^{-s} \,C_4\, (k_1+1)^{\half}\, e^{-(1+s)k_1} \,\norm{du}_{L^2(Z(0,2))}.
6498: \end{align*}
6499: Assume additionally that $e^{-s/2}(k_1+1)^{\half} \le k_1^{\half}$. Then
6500: setting $C_4 \deff {2C_3 \over 1- e^{-s/2}}$ we can conclude that \eqqref(5.2.21)
6501: also holds for $k=k_1$. Since the condition $e^{-s/2}(k_1+1)^{\half}
6502: \le \cdot k_1^{\half}$ is equivalent to $k_1 \ge { 1 \over e^s-1 }$, our recursive
6503: construction implies \eqqref(5.2.21) for all $k\in \left[{ 1 \over e^s-1 }, k^*
6504: \right]$. For the remaining $k\in \left[1, { 1 \over e^s-1 }\right]$ the
6505: estimate \eqqref(5.2.21) follows from \lemma{lem5.2.2}.
6506:
6507: Making a similar recursive construction for $k=k^*, \ldots, l$
6508: we obtain the estimate
6509: \begin{equation}\eqqno(5.2.22)
6510: \norm{v_k - v}_{L^{1,2}(Z_k)} \le
6511: C_4 \, (l-k)^{\half}\, e^{-(1+s)(l-k)}\,\norm{du}_{L^2(Z(l-2,l))}
6512: \end{equation}
6513: for all $k=k^*,\ldots, l-1$ with the the same constant $C_4$. Now \eqqref(5.2.21),
6514: \eqqref(5.2.22), and \eqqref(5.2.20) imply the desired estimate
6515: \eqqref(5.2.10). \qed
6516:
6517:
6518: \newsubsection[5.3]{Deformation of a node and gluing}
6519: The cycle topology on $\barm$, introduced in \refsubsection{5.2c},
6520: has the nice property that $\pr_{\!\!\scrj}: \barm \to \scrj$ is continuous and
6521: {\sl proper}. The last property is follows from the Gromov compactness for
6522: closed curves. However, it is desirable to have a better understanding of the
6523: topological structure of $\barm$. Recall that in \refsubsection{5.2c} we obtained
6524: a natural stratification of $\barm$ in which the strata are distinguished by a
6525: topological type of curves. Moreover, every stratum $\scrm_\bftau$ has a natural
6526: structure of a $C^\ell$-smooth Banach manifold such that the restricted
6527: projection $\pr_{\!\!\scrj}: \scrm_\bftau \to \scrj$ is Fredholm.
6528: So to understand of the topology of $\barm$
6529: means to describe how different strata are attached to each other.
6530: The most important problem is to describe deformations of the standard node.
6531: Let us formulate the question as follows:
6532:
6533: \state Gluing problem. {\sl Let $J_0\in \scrj$ be an almost complex structure and
6534: $u_0: \scra_0 \to X$ a $J_0$-holomorphic map. Describe possible $J$-holomorphic
6535: maps $u: Z(0,l) \to X$ with $J\in \scrj$ sufficiently close to $J_0$ and $l\gg0$
6536: which are sufficiently close to $u_0$ \wrt the Gromov topology}. In other words,
6537: we try to reverse the bubbling and construct a single map $u$ of a long cylinder
6538: $Z(0,l)$ by gluing together the components $u'_0, u''_0: \Delta \to X$ of the map
6539: $u_0$.
6540:
6541: \smallskip
6542: Moreover, one would like to have a smooth structure on the set of such deformation,
6543: so that the transversality techniques could be applied. This means that one seeks
6544: a family of deformations of a given $u_0: \scra_0 \to (X, J_0)$ depending
6545: smoothly on the parameter.
6546:
6547: \smallskip
6548: As the main result of this paragraph we give a satisfactory solution to
6549: the {\sl Gluing problem}. Let us start with introducing some notation.
6550:
6551: \newdefi{def5.3.1} For a fixed sufficiently small $\epsi>0$, let
6552: \begin{equation}
6553: \scra \deff \{\, (z^+, z^-) \in \Delta^2\;:\;
6554: |z^+| \cdot |z^-| < \epsi \,\}
6555: \end{equation}
6556: with the projection
6557: \begin{equation}
6558: \pr_\scra : \scra \to \Delta(\epsi),
6559: \quad \pr_\scra(z^+, z^+) = \lambda(z^+, z^+) \deff z^+ \cdot z^- .
6560: \end{equation}
6561: Further, for $\lambda\in \Delta(\epsi)$ define the analytic sets
6562: \begin{equation}
6563: \scra_\lambda \deff \{\, (z^+, z^-) \in \Delta^2\;:\;
6564: z^+ \cdot z^- = \lambda\,\} = \pr_\scra \inv(\lambda).
6565: \end{equation}
6566: For $\lambda=0$ this is the standard node and for $\lambda \not=0$ a cylinder of
6567: conformal radius $R= \log{1\over|\lambda|}$. Define
6568: $$
6569: \scra^\pm_\lambda \deff
6570: \{(z^+,z^-) \in \scra_\lambda: |\lambda| \le z^\pm <1 \}.
6571: $$
6572: Then $\scra^\pm_\lambda$ are subannuli for $\lambda \not=0$ and $\scra^\pm_0$ are
6573: discs $\Delta^\pm$, the irreducible components of $\scra_0$. In any case, $\scra
6574: _\lambda = \scra^+_\lambda \cup \scra^-_\lambda$.
6575:
6576: To describe a Hermitian metric on a complex manifold $X$, it is sufficient to
6577: indicate only the corresponding K\"ahler form $\omega$. In this case we shall say
6578: that $\omega$ {\sl induces} a metric on $X$ or even that $\omega$ {\sl is} a
6579: metric on $X$. The author begs the reader's pardon for such informality in the
6580: terminology. The advantage of such notation is that the restriction of a metric on
6581: a complex
6582: submanifold is given by the restriction of the corresponding K\"ahler form.
6583: In this notation, the standard metric on the disc $\Delta$ with the coordinate
6584: $z$ is given by the form ${\isl \over 2}dz \wedge d\bar z$.
6585:
6586: We equip $\scra_\lambda$ with the Riemannian metric induced from $\Delta^2$. This
6587: gives the standard metric ${\isl \over 2}dz^\pm \wedge d\bar z^\pm$ on each
6588: component $\Delta^\pm$ of $\scra_0$ and the {\sl hyperbola metric}
6589: \begin{equation}\eqqno(5.3.0.1)
6590: {\isl \over 2}\Bigl(1 + {|\lambda|^2
6591: \over |z^+|^4}\Bigr) dz^+ \wedge d\bar z^+ =
6592: {\isl \over 2}\Bigl(1 + {|\lambda|^2
6593: \over |z^-|^4}\Bigr) dz^- \wedge d\bar z^-
6594: \end{equation}
6595: on $\scra_\lambda$ with $\lambda \not=0$.
6596:
6597: Set
6598: \begin{equation*}
6599: \check\scra \deff \{\, (z^+, z^-) \in \scra\;:\;
6600: z^+ \cdot z^- \not=0 \,\} =
6601: \sqcup_{\lambda \in \check\Delta(\epsi)} \scra_\lambda
6602: \end{equation*}
6603: and
6604: \begin{equation*}
6605: V^\pm \deff \{ 1-\epsi < |z^\pm | < 1 \}, \qquad V \deff V^+ \sqcup V^-.
6606: \end{equation*}
6607: For a given $\lambda$ we have the canonical imbedding $V \to \scra_\lambda$,
6608: defined by the coordinate functions $z^\pm$ on $\scra_\lambda$ and on $V^\pm$.
6609: This imbedding defines the restriction map
6610: \begin{equation*}
6611: u \in L^{1,p}(\scra_\lambda, X) \mapsto u\ogran_V \in L^{1,p}( V, X)
6612: \qquad
6613: u \mapsto (u(z^+)\ogran_{V^+}, \; u(z^-)\ogran_{V^-}).
6614: \end{equation*}
6615: \end{defi}
6616:
6617: \newdefi{def5.3.2} For a nodal curve $C$ with smooth boundary $\d C=
6618: \sqcup_i \gamma_i$, $\gamma_i \cong S^1$, let $\scrp(C)$ be the set of
6619: {\sl stable pseudoholomorphic maps between $C$ and $X$},
6620: \begin{equation}
6621: \scrp(C) \deff \{\, (u, J) \in L^{1,p}(C, X) \times \scrj
6622: \;:\; \dbar_J u=0,\; \text{ $u$ is stable}\;\}.
6623: \end{equation}
6624: Equip $\scrp(C)$ with the topology induced from $L^{1,p}(C, X) \times \scrj$.
6625: In particular, $\scrp(V)$ consists of triples $(u^+, u^-, J)$, where
6626: $u^\pm: V^\pm \to X$ is $L^{1,p}$-smooth $J$-holomorphic map.
6627: Denote by $\scrp^*(C)$ the subset of $(u,J) \in \scrp(C)$ for which
6628: $u$ is {\sl non-multiple on the union of compact components of $C$}.
6629: Further, we define
6630: \begin{equation}
6631: \scrp(\scra) \deff \sqcup_{\lambda \in \Delta(\epsi)}
6632: \scrp(\scra_\lambda) ,
6633: \qquad
6634: \scrp(\check\scra) \deff \sqcup_{\lambda \in \check\Delta(\epsi)}
6635: \scrp(\scra_\lambda)
6636: \end{equation}
6637: and equip this spaces with the topology induced by the Gromov convergence in the
6638: interior of $\scra_\lambda$ and $L^{1,p}$-convergence near boundary. This means
6639: that $(u_n, J_n, \lambda_n)$ converges to $(u_\infty, J_\infty, \lambda_\infty)$
6640: if $(J_n, \lambda_n)$ converges to $(J_\infty, \lambda_\infty)$ in $\scrj \times
6641: \Delta (\epsi)$, the restrictions $u_n\ogran_V$ converges to $u_\infty \ogran_V$
6642: \wrt $L^{1,p}$-norm, and $u_n$ converges to $u_\infty$ in the sense of {\sl
6643: Definition 5.2.7}. Elements of $\scrp(\scra)$ will be denoted by $(u, J,
6644: \lambda)$. As usual, $\pr_{\!\!\scrj}$ stands for the natural projections from
6645: $\scrp(C)$ or $\scrp(\scra)$ to $\scrj$.
6646: \end{defi}
6647:
6648:
6649: \newthm{thm5.3.1} The natural map $\pr_V: \scrp(\scra) \to \scrp(V)
6650: \times \Delta(\epsi)$, defined by
6651: \begin{equation}
6652: \pr_V(u, J, \lambda) \deff (u(z^+)\ogran_{V^+}, u(z^-)\ogran_{V^-}, J;
6653: \lambda)
6654: \end{equation}
6655: is an imbedding of a {\sl topological} Banach submanifold.
6656:
6657: Moreover, for every $(u_0,J_0) \in \scrp(\scra_0)$ there exists a neighborhood
6658: $\scru \subset \scrp(\scra_0)$ of $(u_0,J_0)$, an $\epsi'>0$, and a map
6659: $\Phi: \scru \times \Delta(\epsi') \to \scrp(\scra)$ such that
6660: \begin{itemize}
6661: \item $\Phi$ is a homeomorphism onto the image;
6662: \item for every $\lambda \in \Delta(\epsi')$ the restricted map $\Phi_\lambda
6663: \deff \Phi\ogran_{\scru \times \{\lambda\}}: \scru \to \scrp(V) \times \Delta(
6664: \epsi)$takes values in $\scrp(\scra_\lambda) \subset \scrp(\scra)$ and is a
6665: $C^1$-diffeomorphism;
6666: \item the family of maps $\Psi_\lambda \deff \pr_V \scirc \Phi_\lambda: \scru \to
6667: \scrp(V)$ depends continuously on $\lambda \in \Delta(\epsi')$ \wrt the
6668: $C^1$-topology.
6669: \end{itemize}
6670: \end{thm}
6671:
6672: In other words, the theorem statess that $\scrp(\scra) =\sqcup \scrp(\scra_\lambda)$
6673: is a continuous family of $C^1$-submanifolds. Before proving this result, we state
6674: and prove a corollary which provides a technique which allows one to smooth nodal
6675: points on pseudoholomorphic curves.
6676:
6677: \newthm{thm5.3.1a} Let $C^*$ be a closed connected nodal curve parameterized by
6678: a real surface $S$, $J^*\in \scrj$, $u^*: C^* \to X$ a $J^*$-holomorphic map, and
6679: $(C^*,u^*,J^*) \in \barm^{Gr}=\barm^{Gr}(S, X, [C^*])$ the corresponding
6680: element of the Gromov compactification of the total moduli space. Assume
6681: that the map $u^*:C^* \to X$ is non-multiple.
6682:
6683: Then there exist $(C', u', J') \in \scrm(S, X, [C^*])$ arbitrarily close to
6684: $(C^*,u^*,J^*)$ \wrt the Gromov topology such that $C'$ is a smooth curve.
6685: \end{thm}
6686:
6687: The notation used in this theorem was introduced in {\sl Definitions
6688: \ref{def5.2.6}} and {\sl\ref{def5.2c.2}}. Note that the condition of
6689: non-multiplicity of $u^*:C^* \to X$ is equivalent to the absence of ghost and
6690: multiple components.
6691:
6692: \proof Let $\{ z^*_1, \ldots z^*_k\}$ be the set of nodal points of $C^*$.
6693: For every $z^*_i$ fix a neighborhood $V_i$ isomorphic to the standard node.
6694: We may also assume that the sets $V_i$ are pairwise disjoint.
6695: Let $u^*_i$ denote the restriction of $u^*$ to $V_i$. Applying \refthm{thm5.3.1}
6696: we can perturb $V_i\cong \scra_0$ to an annulus $V'_i = \scra_{\lambda_i}$ and
6697: $u^*_i: V_i \to X$ to a $J^*$-holomorphic map $u'_i: V'_i \to X$. If these
6698: perturbations $(V'_i, u'_i)$ are made small enough, then we can adjust the
6699: structure $J^*$ and the map $u^*$ on the remaining part of the curve $C^*$ in a way
6700: yielding the desired $(C', u', J') \in \scrm(S, X, [C^*])$. \qed
6701:
6702:
6703: Modifying the proof of \refthm{thm5.3.1a} one can also obtain
6704:
6705: \newprop{prop5.3.1b} Let $C^*$ be a closed connected nodal curve parameterized by
6706: a real surface $S$, $J^*\in \scrj$, $u^*: C^* \to X$ a $J^*$-holomorphic map, and
6707: $(C^*,u^*,J^*) \in \barm^{Gr}=\barm^{Gr}(S, X, [C^*])$ the corresponding
6708: element of the Gromov compactification of the total moduli space. Assume
6709: that the map $u^*:C^* \to X$ is non-multiple.
6710:
6711: Then in a neighborhood of $(C^*,u^*,J^*)$ the space $\barm^{Gr}$ is a topological
6712: Banach manifold and the natural projection $\pr^{Gr}:\barm^{Gr} \to \barm$
6713: a homeomorphism.
6714: \end{prop}
6715:
6716:
6717: Since the result of \propo{prop5.3.1b} is not needed for the
6718: purposes of this paper, we leave it without a proof.
6719:
6720:
6721: \medskip
6722: The proof of \refthm{thm5.3.1} is divided in the subsequent lemmas. The first two
6723: are simple but useful technical results.
6724:
6725: \newlemma{lem5.3.2} Let $C$ be a nodal curve without closed compact components,
6726: and $E$ a holomorphic vector bundle over $C$. Then any operator $D: L^{1,p}(C,
6727: E) \to L^p_{(0,1)}(C, E)$ of the form $D= \dbar_E + R$ with $R \in L^p$ is
6728: surjective and its kernel $\sfh^0_D(C, E)$ admits a closed complement.
6729: \end{lem}
6730:
6731: \proof Imbed $C$ into a compact nodal curve $\wt C$ and extend $E$ to
6732: a holomorphic vector bundle $\wt E$ over $\wt C$. Without loss of generality
6733: we may assume that the Chern numbers $\la c_1(\wt E), \wt C_i\ra$ are
6734: sufficiently large for each component $\wt C_i$ of $\wt C$. Now extend
6735: $R \in L^p(C, \homr(E, E\otimes \Lambda^{(0,1)}))$ to $\wt R \in L^p(\wt C,
6736: \homr(\wt E, \wt E\otimes \Lambda^{(0,1)}))$ and set $\wt D \deff \dbar
6737: _{\wt E} + \wt R$. Adjusting $\wt R$ on the complement $\wt C \bs C$ we may
6738: assume that $\wt D: L^{1,p}(\wt C, \wt E) \to L^p_{(0,1)}(\wt C, \wt E)$
6739: is surjective. The sufficient condition for existence of such an adjustment
6740: is provided by the condition on the Chern numbers of $\wt E$. Since any $\eta
6741: \in L^p_{(0,1)}(C, E)$ extends to $\wt \eta \in L^p_{(0,1)}(\wt C, \wt E)$,
6742: the surjectivity of $\wt D$ implies the surjectivity of $D$.
6743:
6744: The existence of a closed complement to the kernel of $D$ is shown in
6745: \cite{Iv-Sh-2} in the case when $D= \dbar$. This proof applies also in our case
6746: with only minor adjustments.
6747: %with necessary minor corrections. \
6748: \qed
6749:
6750: \state Remark. The existence of a closed complement to the kernel of
6751: $D$ allows us
6752: %ensures possibility
6753: to apply the implicit function theorem.
6754:
6755: \smallskip
6756: \newlemma{lem5.3.3}
6757: \sli The space $L^{1,p}(C, X)$ is a smooth Banach
6758: manifold with tangent space
6759: \begin{equation}
6760: T_uL^{1,p}(C, X)=L^{1,p}(C, E_u).
6761: \end{equation}
6762:
6763: \slii The space $\scrp^*(C)$ is a $C^\ell$-smooth
6764: submanifold of $L^{1,p}(C, X) \times \scrj$ with tangent space
6765: \begin{equation}
6766: T_{(u,J)}\scrp^*(C)= \{ (v, \dot J)\in L^{1,p}(C, E_u) \times T_J\scrj\,:\,
6767: D_{u,J}v + \dot J \scirc du \scirc J_C =0 \}.
6768: \end{equation}
6769:
6770: \sliii $\pr_V: \scrp(\scra_0) \to \scrp(V)$ and $\pr_V: \scrp(\check\scra)
6771: \to \scrp(V) \times \check\Delta(\epsi)$ are $C^\ell$-smooth imbeddings on
6772: Banach submanifolds.
6773: \end{lem}
6774:
6775: The definitions of the spaces which are involved here are given in
6776: \refdefi{def5.2.3a} and \refdefi{def5.3.2}.
6777:
6778: \proof Let $C= \cup C_i$ be the decomposition of $C$ into components
6779: and $\{(z'_a, z''_a)\}$ the set of pairs of points on the normalization
6780: $\hat C$ corresponding to the nodal points.
6781:
6782: \sli The space $L^{1,p}(C, X)$ is a subset of a smooth Banach
6783: manifold $\prod_i L^{1,p}(C_i, X)$ defined by equations $u(z'_a) =u(z''_a)$.
6784: One checks the transversality condition and computes the tangent space.
6785:
6786: \slii One can use the same arguments as in part \slip.
6787:
6788: \sliii First we note that \lemma{lem1.2.4} \slii implies the following unique
6789: continuation property: Any
6790: $J$-holomorphic map $u$, defined on an open set $U$ of a nodal curve $C$ admits
6791: at most one $J$-holomorphic extension to an irreducible component $C'$ of $C$
6792: provided $C'\cap U \not =\emptyset$. This shows that the restriction maps $F_0:
6793: \scrp(\scra_0) \to \scrp(V)$ and $\check F: \scrp(\check\scra) \to \scrp(V)\times
6794: \check\Delta(\epsi)$ are set-theoretically injective. Note that we have
6795: introduced new notation, $F_0$ and $\check F$, for the restrictions of
6796: the map $\pr_V$ to the corresponding definition domains.
6797:
6798: Further, $F_0: \scrp(\scra_0) \to \scrp(V)$ is obviously $C^\ell$-smooth
6799: and the differential $dF_0: T_u\scrp(\scra_0) \to T_u\scrp(V)$ is simply
6800: the restriction map $dF_0: (v, \dot J) \mapsto (v\ogran_V, \dot J)$.
6801: \lemma{lem5.3.2} shows that the restriction map
6802: \begin{equation}
6803: \{\, v \in L^{1,p}(\scra_0, E_u)\;:\; D_{u,J}v=0\, \}\mapsto
6804: \{\, v \in L^{1,p}(V, E_u)\;:\; D_{u,J}v=0\, \}
6805: \eqqno(5.3.2)
6806: \end{equation}
6807: is a closed imbedding and splits, \ie admits a closed complement.
6808:
6809: \smallskip
6810: The claim about $\check F: \scrp(\check\scra) \to \scrp(V) \times
6811: \Delta(\epsi)$ is proven in a similar way. Details are left to the reader.
6812: \qed
6813:
6814: \smallskip
6815: A crucial point in the proof of \refthm{thm5.3.1} is to find an apriori estimate
6816: for the operator $D_{u,J,\lambda}$ which is uniform as $\lambda\lrar0$. Because of
6817: local nature of the estimate it is sufficient to work with the ball $B\subset
6818: \rr^{2n} \cong \cc^n$ equipped with the standard complex structure $J\st$. We
6819: start with introducing a chart for the space $L^{1,p}(\scra, \cc^n) \deff \sqcup
6820: _{\lambda \in \Delta(\epsi)} L^{1,p}(\scra_\lambda, \cc^n)$.
6821:
6822: \newdefi{def5.3.3} For a nodal complex curve $C$ and a complex manifold $X$ let
6823: $\scrh(C, X)$ denote the space of holomorphic maps $f: C \to X$ which are
6824: $L^{1,p}$-smooth up to the boundary $\d C$. In the case $X=\cc$ we abbreviate
6825: the notation to $\scrh(C)$.
6826: \end{defi}
6827:
6828: \newlemma{lem5.3.3a}%
6829: \footnote{\.The results presented in the lemma have been obtained jointly with
6830: S.~Ivashkovich.} There exist families of homomorphisms $T_\lambda: L^p_{(0,1)}
6831: (\scra_\lambda{,} \cc)
6832: \to L^{1,p}(\scra_\lambda, \cc)$ and isomorphisms $\sfl_\lambda: \scrh(\scra_0)
6833: \to \scrh(\scra_\lambda)$ and $Q_\lambda: L^p_{(0,1)} (\scra_0, \cc) \to
6834: L^p_{(0,1)} (\scra_\lambda, \allowbreak
6835: \cc)$ with the following properties:
6836:
6837: \sli the homomorphisms $T_\lambda$ are right inverses of $\dbar: L^{1,p}(\scra
6838: _\lambda, \cc) \to L^p_{(0,1)} (\scra_\lambda, \cc)$;
6839:
6840: \slii the norms of $T_\lambda$, $\sfl_\lambda$, and $Q_\lambda$, as well as
6841: $\sfl_\lambda\inv$ and $Q_\lambda\inv$, are uniformly bounded;
6842:
6843: \sliii the homomorphisms $T_\lambda$, $\sfl_\lambda$, and $Q_\lambda$ depend
6844: smoothly on $\lambda\not=0$.
6845: \end{lem}
6846:
6847: \proof
6848: By the definition, the hyperbola metric ${\isl \over 2}\left(1 + {|\lambda|^2 \over
6849: |z^+|^4} \right) dz^+ \wedge d\bar z^+$ on $\scra_\lambda$ is the sum of the
6850: standard flat metrics ${\isl \over 2}dz^+ \wedge d\bar z^+$ and ${\isl \over 2}dz^-
6851: \wedge d\bar z^-$. Note also that the function $\left(1 + {|\lambda|^2 \over |z^+|
6852: ^4}\right)$, restricted to the subannulus $\scra^+_\lambda= \{ |\lambda|^\half <
6853: |z^+| <1\} \subset \scra_\lambda$, takes values in the interval $[1, 2]$. This
6854: implies that in every subannulus $\scra^\pm_\lambda$ the metric ${\isl \over 2}
6855: \left(1 + {|\lambda|^2 \over |z|^4}\right) dz^+ \wedge d\bar z^+$ is equivalent to
6856: the disc metric ${\isl \over 2}dz^\pm \wedge d\bar z^\pm$. In particular, the norm
6857: $\norm{v}_{L^{1,p}(\scra_\lambda)}$ is equivalent to the norm
6858: \begin{equation}
6859: \left(\int_{|\lambda|^ < |z^+|^2 <1} (|v| + |dv|)^p\;
6860: \msmall{\isl \over 2}dz^+ \wedge d\bar z^+
6861: \,+\,
6862: \int_{|\lambda|^ < |z^-|^2 <1} (|v| + |dv|)^p\;
6863: \msmall{\isl \over 2}dz^- \wedge d\bar z^-
6864: \right)^{1\over p}.
6865: \end{equation}
6866:
6867: We start with construction of $T_\lambda$. For the discs $\Delta^\pm$ with the
6868: coordinates $z^\pm$ respectively we define $\ti T^\pm: L^p_{(0,1)} (\Delta^\pm,
6869: \cc) \to L^{1,p}(\Delta^\pm, \cc)$ to be the Cauchy-Green operators, \ie
6870: $$
6871: \ti T^+(\phi^+)(z^+) \deff
6872: \msmall{1\over 2\pi \isl} \int_{\zeta \in \Delta}
6873: \msmall{d\zeta \wedge \phi^+(\zeta) \over \zeta -z^+}
6874: $$
6875: and similarly for $\ti T^-$. Then we set
6876: $$
6877: T^\pm(\phi^\pm)(z^\pm) \deff \ti T^\pm(\phi^\pm)(z^\pm) -
6878: \ti T^\pm(\phi^\pm)(0)
6879: $$
6880: So $T^\pm$ are normalizations of $\ti T^\pm$ respectively to the condition
6881: $T^\pm(\phi^\pm)(0) =0$. For a form $\phi \in L^p_{(0,1)} (\scra_\lambda, \cc)$ we
6882: denote by $\phi^\pm (z^\pm)$ its restriction to $\scra^\pm \subset \Delta^\pm$
6883: extended by $0$ to the whole discs $\Delta^\pm$, and set
6884: $$
6885: T_\lambda(\phi) \deff T^+(\phi^+)(z^+) + T^-(\phi^-)(z^-)
6886: $$
6887: For the special case $\lambda=0$ this construction should be modified follows:
6888: Every form $\phi \in L^p_{(0,1)} (\scra_0, \cc)$ has two components $\phi^\pm
6889: (z^\pm)$ corresponding to the decomposition
6890: $$
6891: L^p_{(0,1)} (\scra_0, \cc)= L^p_{(0,1)} (\Delta^+, \cc) \oplus
6892: L^p_{(0,1)} (\Delta^-, \cc),
6893: $$
6894: and the operator $T_0$ transforms $\phi^\pm$ into functions $f^\pm(z^\pm)\deff
6895: T^\pm_0(\phi^\pm)(z^\pm)$ which satisfy $f^+(0)= f^-(0) =0$. Then $f\deff (f^+,f^-)
6896: \in L^{1,p}(\scra_0, \cc)$. The desired properties of $T_\lambda$ can be seen
6897: in a straightforward way.
6898:
6899: \medskip
6900: Defining the operators $\sfl_\lambda$, we recall the identification
6901: $$
6902: \scrh(\scra_0) = \{ (g^+(z^+), g^-(z^-)) \in
6903: \scrh(\Delta^+) \oplus \scrh(\Delta^-) : g^+(0) =g^-(0) \}.
6904: $$
6905: We set $\sfl_0 = \id: \scrh(\scra_0) \to \scrh(\scra_0)$ and
6906: $$
6907: \sfl_\lambda: g=(g^+(z^+), g^-(z^-)) \in \scrh(\scra_0) \mapsto
6908: g^+(z^+) + g^-(z^-) - g(0)
6909: $$
6910: for $\lambda \not=0$. The desired properties of $\sfl_\lambda$ are obvious as well.
6911:
6912: Note also that for $\lambda \not=0$ the inverse operator $\sfl_\lambda\inv$
6913: essentially gives the Laurent decomposition of functions $g \in \scrh(\scra
6914: _\lambda)$.
6915:
6916:
6917: \medskip
6918: The definition of $Q_\lambda$ is more subtle. For $\lambda \not=0 \in
6919: \Delta(\epsi)$ we set
6920: \begin{equation}\eqqno(5.3.a1)
6921: \rho_\lambda(r) \deff
6922: \msmall{ r^2 - {|\lambda|^2 \over r^2} \over 1-|\lambda|^2}\cdot
6923: \end{equation}
6924: Then every $\rho_\lambda$ induces a diffeomorphism of the interval $[|\lambda|,1]$
6925: onto $[-1,1]$, such that $[|\lambda|, |\lambda|^\half]$ and $[|\lambda|^\half,1]$
6926: are mapped onto the intervals $[-1,0]$ and $[0,1]$ respectively. The inverse map
6927: is given by
6928: \begin{equation}\eqqno(5.3.a2)
6929: r= R_\lambda(\rho) = \sqrt{
6930: \msmall{ \rho(1-|\lambda|^2) + \sqrt{ \rho^2(1-|\lambda|^2)^2 +4\,|\lambda|^2}
6931: \over2}} \cdot
6932: \end{equation}
6933: For $\lambda \not=0\in \Delta(\epsi)$ we define the maps $\sigma^\pm_\lambda:
6934: Z(-1,1) = [-1,1] \times S^1 \to \scra_\lambda$ which are given in the coordinates
6935: $z^\pm= r^\pm \, e^{\isl\theta^\pm}$ by relations $r^\pm = R_\lambda(\rho)$ and
6936: $\theta^\pm =\theta$ respectively, so that
6937: $$
6938: \sigma^\pm_\lambda: (\rho, \theta)\in Z(-1,1) \mapsto
6939: z^\pm= R_\lambda(\rho) e^{\isl\theta} \in \scra_\lambda.
6940: $$
6941: Now we can explain the reason for the choice of \eqqref(5.3.a1), which at first
6942: glance probably seems rather wild. %at first look seems probably rather wild.
6943: Our choice is made to insure that the pull-back
6944: by $\sigma_\lambda$ of the natural volume form of $\scra_\lambda$ is a constant
6945: multiple of the standard volume form on $Z(-1,1)$, \ie
6946: \begin{equation}\eqqno(5.3.a3)
6947: \sigma_\lambda^*\left( \left(1+ {|\lambda|^2 \over (r^+)^4} \right)
6948: \msmall{\isl\over2} dz^+ \wedge d\bar z^+ \right)
6949: = (1-|\lambda|^2) d\rho \wedge d\theta .
6950: \end{equation}
6951: Moreover, $1-|\lambda|^2$ remains uniformly bounded as $\lambda$ varies
6952: in $\Delta(\epsi)$, so that the volume forms in the right hand side of
6953: \eqqref(5.3.a3) are equivalent.
6954:
6955: The behavior of $\sigma^\pm_\lambda$ by $\lambda$ close to $0$, which is rather
6956: delicate, can be described as follows. Denote $Z^+ \deff Z(0,1)$ and $Z^- \deff
6957: Z(-1,0)$. Then for $\lambda \lrar 0$ we have convergence of $\sigma^+_\lambda$ on
6958: $Z^+$ to a map $\sigma^+_0: Z^+ \to \scra^+_0 =\Delta^+$, and resp.\ convergence
6959: of $\sigma^-_\lambda$ on $Z^-$ to $\sigma^-_0: Z^- \to \scra^-_0$, which are given
6960: by
6961: \begin{align*}
6962: \sigma^+_0: &(\rho, \theta) \mapsto z^+=\sqrt\rho \, e^{\isl\theta},
6963: \\
6964: \sigma^-_0: &(\rho, \theta) \mapsto z^-=\sqrt\rho \, e^{\isl\theta}.
6965: \end{align*}
6966: Observe also that we obtain the map $R_0(\rho) = \sqrt\rho$ in the limit of
6967: \eqqref(5.3.a2) as $\lambda\lrar0$. The convergence $\sigma^\pm_\lambda \lrar
6968: \sigma^\pm_0$ is in the $C^\infty$-sense in the interiors of $Z^\pm$, and in
6969: the $C^0$-topology up to boundary of $Z^\pm$.
6970:
6971: On the other hand, there is no convergence of $\sigma^\pm_\lambda$ on $Z^\mp$.
6972: The topological reason for the absence of the convergence of $\sigma^\pm_\lambda
6973: \ogran_{Z^\mp}$ is that, making with $\lambda$ a small bypass around $0$, we
6974: perform a {\sl Dehn twist} with $\scra_\lambda$.\footnote{\.The author is
6975: indebted to Bernd Siebert for this remark.}
6976:
6977: Now, we define $Q_\lambda$ representing the components $\phi^\pm$ of every $\phi\in
6978: L^p_{(0,1)} (\scra_\lambda, \cc) = L^p_{(0,1)} (\Delta^+, \cc) \oplus L^p_{(0,1)}
6979: (\Delta^-, \cc)$ in the form $\phi^\pm(z^\pm) = f^\pm(r^\pm, \theta^\pm) d\bar
6980: z^\pm$ with $L^p$-functions $f^\pm(r^\pm, \allowbreak
6981: \theta^\pm)$ and setting
6982: \par\noindent
6983: $
6984: Q_\lambda (\phi) \deff \begin{cases}
6985: f^+\Bigl(\! R_\lambda\bigl( (r^+)^2\bigr), \theta^+\Bigr)d\bar z^+
6986: & \!\!\!\!\text{at the point $z^+= r^+ e^{\isl\theta^+}$ with
6987: $r^+ \in [|\lambda|^\half, 1]$};
6988: \\
6989: f^-\Bigl(\! R_\lambda\big( (r^-)^2 \bigr), \theta^-\Bigr)d\bar z^-
6990: & \!\!\!\!\text{at the point $z^-= r^- e^{\isl\theta^-}$ with
6991: $r^- \in [|\lambda|^\half, 1]$}.
6992: \end{cases}$ \lineeqqno(5.3.a4)
6993:
6994: \smallskip
6995: Let us explain the meaning of the construction of $Q_\lambda$ given in
6996: \eqqref(5.3.a4). The first point is that we essentially transform every form
6997: $\phi \in L^p_{(0,1)}(\scra_\lambda, \cc)$ into a function $f\in L^p(\scra
6998: _\lambda, \cc)$. This is done by representing $\phi$ in the form $\phi(z^+) = f^+
6999: (z^+) d\bar z^+$ on $\scra^+_\lambda$ and in the form $\phi(z^-) = f^-(z^-)
7000: d\bar z^-$ on $\scra^-_\lambda$. Observe that we use different coordinates $z^\pm$
7001: on the different parts $\scra^\pm_\lambda$ of $\scra_\lambda$. Independently of
7002: this, we obtain a well-defined $L^p$-function $f$ on $\scra_\lambda$, and hence a
7003: family of well-defined maps $F_\lambda: L^p_{(0,1)}(\scra_\lambda, \cc) \to L^p(
7004: \scra_\lambda, \cc)$. It is not difficult to see that the $F_\lambda$ are complex
7005: linear isomorphisms and that the operator norms $\norm{F_\lambda}_{\sf{op}}$
7006: and $\norm{F_\lambda\inv}_{\sf{op}}$ are bounded uniformly in $\lambda$.
7007:
7008: The second point is the observation that for the definition of a space $L^p(Y)$
7009: only the involved measure $\mu$ on $Y$ is essential. In particular, if a measurable
7010: map $g: (Y_1, \mu_1) \to (Y_2, \mu_2)$ induces an equivalence of measures, \ie
7011: $g^*\mu_2 = e^\rho(y) \mu_1$ for some bounded $\rho \in L^\infty(Y_1, \mu_1)$,
7012: then the induced map $g^*: L^p(Y_2, \mu_2) \to L^p(Y_1, \mu_1)$ is an isomorphism
7013: of Banach spaces. Thus our construction exploits the fact that the measures
7014: $$
7015: (\sigma^\pm_\lambda)^* \left( \left(1+ {|\lambda|^2 \over (r^\pm)^4} \right)
7016: \msmall{\isl\over2} dz^\pm \wedge d\bar z^\pm \right)
7017: $$
7018: on $Z^\pm$ are equivalent.
7019:
7020: \newcorol{cor5.3.3b} The Banach spaces $L^{1,p}(\scra_0, \cc^n)$ and $L^{1,p}
7021: (\scra_\lambda, \cc^n)$ with $\lambda \not=0$ are isomorphic.
7022: \end{corol}
7023:
7024: \proof One uses $T_\lambda$ to split the exact sequences
7025: $$
7026: 0 \lrar \scrh(\scra_\lambda, \cc^n) \lrar L^{1,p}(\scra_\lambda, \cc^n)
7027: \buildrel \dbar \over \lrar L^p_{(0,1)}(\scra_\lambda, \cc^n)
7028: \lrar 0
7029: $$
7030: for $\lambda=0$ and $\lambda \not=0$. Then one applies $\sfl_\lambda$ and
7031: $Q_\lambda$ to identify $\scrh(\scra_0, \cc^n)$ with $\scrh(\scra_\lambda, \cc^n)$
7032: and respectively $L^p_{(0,1)}(\scra_0, \cc^n)$ with $L^p_{(0,1)}(\scra_\lambda,
7033: \cc^n)$. \qed
7034:
7035: \medskip
7036: Let us explain now the main difficulty in the proof of \refthm{thm5.3.1}. Several
7037: authors (see \eg \cite{Sie}, \cite{Li-T}, \cite{Ru}) have approached the
7038: {\sl Gluing problem} %tried to solve the
7039: by making an appropriate local imbedding $\scrp(\scra) \hook L^{1,p} (\scra,
7040: \cc^n)= \sqcup_{\lambda \in \Delta(\epsi)} L^{1,p}(\scra_\lambda, \cc^n)$
7041: and showing that, roughly speaking, the $\dbar$-equation induces an operator which
7042: defines $\scrp(\scra)$ and has a certain (rather weak) smoothness property.
7043: This property is sufficient for showing that for any fixed $J \in \scrj$ the
7044: compactification $\barr\scrm_J$ has a well-defined ``virtual fundamental class'',
7045: the basic object to define in the theory of Gromov-Witten invariants.
7046:
7047: The difficulty with this approach is that the smooth structure in $L^{1,p}(\scra,
7048: \cc^n)$ given by the isomorphism $L^{1,p}(\scra, \cc^n) \cong L^{1,p}(\scra_0,
7049: \cc^n) \times \Delta(\epsi)$ depends heavily on both the linear and the complex
7050: structures in $\cc^n$. Moreover, one can show that for a generic smooth
7051: diffeomorphism $g: \cc^2 \to \cc^2$ the induced map $u \in L^{1,p}(\scra, \cc^n)
7052: \mapsto u\scirc g \in L^{1,p}(\scra, \cc^n)$ is not even Lipschitzian at generic
7053: $u \in L^{1,p}(\scra_0, \cc^n)$.
7054:
7055: To the contrary, our approach of imbedding $\scrp(\scra)$ into $\scrp(V)\times
7056: \Delta(\epsi)$ by means of tracing the restriction $u\ogran_V$ has the
7057: advantage that the smooth structure in $\scrp(V)$ is natural and canonical.
7058: Here we shall use \lemma{lem5.3.3a} rather in a different way: It
7059: serves for us as an approximative description of the behavior of the Gromov
7060: operator $D_{u,J,\lambda}$ as $\lambda \lrar 0$.
7061:
7062: \newdefi{def5.3.4} Let $X$ be a manifold with a fixed symmetric connection
7063: $\nabla$. Then for $\lambda \in \Delta(\epsi)$, $u\in L^{1,p}(\scra_\lambda, X)$,
7064: and a $C^1$-smooth almost complex structure on $X$ we denote by $J_\lambda$ the
7065: complex structure on $\scra_\lambda$ and by
7066: $$
7067: D_{u,J,\lambda}: L^{1,p}(\scra_\lambda, E_u) \to
7068: L^p_{(0,1)}(\scra_\lambda, E_u)
7069: $$
7070: the operator given by
7071: \begin{equation}\eqqno(5.3.a5)
7072: D_{u,J,\lambda}(v) \deff \nabla v + J \scirc \nabla v \scirc J_\lambda
7073: + \msmall{1\over2} \bigl( \nabla_vJ \scirc du \scirc J_\lambda -
7074: J \scirc \nabla_vJ \scirc du \bigr).
7075: \end{equation}
7076: Observe that in the definition of the space $L^p_{(0,1)}(\scra_\lambda, E_u)$ we
7077: equip $E_u \deff u^*TX$ with the structure $u^*J$. The construction of
7078: $D_{u,J,\lambda}$ is an extension of the definition of the Gromov operator, because
7079: for $(u, J)\in \scrp(\scra_\lambda)$ the definition \eqqref(5.3.a5) coincides
7080: with the original one in \eqqref(1.3.4).
7081: \end{defi}
7082:
7083:
7084: \newlemma{lem5.3.4} Let $B\subset \rr^{2n}$ be the ball and $J^*$ a $C^1$-smooth
7085: almost complex structure in $B$. Then there exist constants $\epsi=\epsi(J)>0$
7086: and $C =C(J) < \infty$ such that for any
7087: almost complex structure $J$ with
7088: $$
7089: \norm{J -J^*}_{C^1(B)} \le \epsi,
7090: $$
7091: any $\lambda \in\Delta(\epsi)$, and any $J$-holomorphic map $u: \scra_\lambda \to
7092: B(\half)$ with
7093: $$
7094: \norm{du}_{L^p(\scra_\lambda)} \le \epsi,
7095: $$
7096:
7097: \smallskip
7098: \sli
7099: one has a uniform estimate
7100: \begin{equation}
7101: \norm{v}_{L^{1,p}(\scra_\lambda)} \le C \cdot
7102: \bigl(\norm{v}_{L^{1,p}(V)} +
7103: \norm{D_{u,J,\lambda}v}_{L^p(\scra_\lambda)} \bigr)
7104: \eqqno(5.3.4)
7105: \end{equation}
7106: for any $v\in L^{1,p}(\scra_\lambda, E_u)$;
7107:
7108: \slii there exists an operator $T_{u,J,\lambda}: L^p_{(0,1)}(\scra_\lambda, E_u)
7109: \to L^{1,p}(\scra_\lambda, E_u)$ with $D_{u,J,\lambda} \scirc T_{u,J,\lambda}= id$;
7110:
7111: \sliii moreover, the family of operators $T_{u,J,\lambda}$ depends continuously
7112: on $(u,J,\lambda)$.
7113: \end{lem}
7114:
7115:
7116: \proof Recall that on each half-annulus $\scra^\pm_\lambda$ the hyperbola metric
7117: \eqqref(5.3.0.1) is equivalent to the (flat) disc metric ${\isl\over2} dz^\pm
7118: \wedge \bar z^\pm$. Thus we can apply the Morrey estimate
7119: $$
7120: \diam(u(\scra_\lambda^\pm)) \le
7121: C \cdot \norm{du}_{L^p(\scra^\pm_\lambda)}
7122: $$
7123: which is uniform in $\lambda$. This gives
7124: \begin{equation}\eqqno(5.3.a6)
7125: \diam(u(\scra_\lambda)) \le C \cdot \norm{du}_{L^p(\scra_\lambda)}
7126: \end{equation}
7127: again uniformly in $\lambda$.
7128:
7129: \smallskip
7130: Consequently, $\diam(u(\scra_\lambda))$ is sufficiently small. Let $J\st$ be a
7131: linear complex structure in $\rr^{2n}$ with coincides with $J$ at some point $x_0=
7132: u(z_0)\in u(\scra_\lambda)$. Then $\norm{J\scirc u -J\st}_{C^0(\scra_\lambda)}$ is
7133: also small enough.
7134:
7135: The canonical trivialization of the tangent bundle $TB$ yields the canonical
7136: trivialization of $E_u = u^*TB$ and the identification $L^{1,p}(\scra_\lambda, E_u)
7137: =L^{1,p}(\scra_\lambda, \cc^n)$. By \refcorol{cor5.3.3b} we obtain a structure of
7138: a continuous Banach bundle on the union $\sqcup_{(u,\lambda)\in L^{1,p}(\scra,B)}
7139: L^{1,p}(\scra_\lambda, E_u)$.
7140:
7141: A similar identification for $L^p_{(0,1)}(\scra_\lambda, E_u)$ requires some
7142: modification, since in the definition of this space one
7143: involves the complex structure $J_u \deff u^*J = J\scirc u$. Therefore we must fix
7144: a {\sl complex} isomorphism $\phi$ of $(TB, J)$ with the trivial complex bundle
7145: $(TB, J\st)$ over $B$. This means that $\phi$ is an $\rr$-linear endomorphism of
7146: $TB$ satisfying $\phi \scirc J = J\st \scirc \phi$. Since $J$ coincides with
7147: $J\st$ at $x_0=u(z_0)$ with $z_0 \in \scra_\lambda$, we may also
7148: assume that $\norm{\phi - \id_{TB}}_{C^0(u(\scra_\lambda))}$ is small enough. Using
7149: $\phi$, we obtain an isomorphism $\phi_*: L^p_{(0,1)}(\scra_\lambda, E_u)\buildrel
7150: \cong \over \lrar L^p_{(0,1)} (\scra_\lambda, \cc^n)$. Moreover, the norms
7151: $\norm{\phi_*}_{\sf{op}}$ and $\norm{\phi_*\inv}_{\sf{op}}$ are bounded uniformly
7152: in $\lambda \in \Delta(\epsi)$. Now the composition $\phi_* \scirc D_{u,J,\lambda}$
7153: is a homomorphism between $L^{1,p}(\scra_\lambda, \cc^n)$ and $L^p_{(0,1)}
7154: (\scra_\lambda, \cc^n)$.
7155:
7156: Using the trivialization $(u^*TB, u^*J\st)$ we define the operator
7157: $$
7158: \dbar: L^{1,p}(\scra_\lambda, \cc^n) \to L^p_{(0,1)} (\scra_\lambda, \cc^n).
7159: $$
7160: By construction, $\norm{ \dbar - \phi_* \scirc D_{u,J,\lambda}}_{\sf{op}}$
7161: is small enough. Consequently, the restriction of $D_{u,J,\lambda}$ to the image
7162: of the operator $T_\lambda$ constructed in \lemma{lem5.3.3a} is an isomorphism.
7163: Now the existence of $T_{u,J,\lambda}$ follows from the ``closed graph theorem''
7164: of Banach.
7165:
7166: Note that the choice of the point $x_0$ and the isomorphism $\phi$ used in the
7167: construction of $T_{u,J,\lambda}$ can be made continuous on $(u,J,\lambda)$.
7168: This means that there exist families $x_0(u,J,\lambda)$ and $\phi(u,J,\lambda)$
7169: which depend continuously on $(u,J,\lambda)$ and have the properties needed above.
7170: For example, one can set $x_0(u,J,\lambda) \deff u(1,\lambda)$, the image of
7171: the point $(1,\lambda) \in \scra_\lambda$. For such a choice of $x_0(u,J,\lambda)$
7172: and $\phi(u,J,\lambda)$ the family $T_{u,J,\lambda}$ also depends
7173: continuously on $(u,J,\lambda)$.
7174:
7175: \smallskip
7176: As a consequence of \sliip, it is sufficient to prove \eqqref(5.3.4)
7177: under the additional condition $D_{u,J,\lambda}v =0$. But then
7178: \begin{multline*}
7179: \norm{v}_{L^{1,p}(\scra_\lambda)} \le C_2 \cdot
7180: \bigl(\norm{v}_{L^{1,p}(V)} +
7181: \norm{\dbar v}_{L^p(\scra_\lambda)} \bigr)=
7182: C_2 \cdot \bigl(\norm{v}_{L^{1,p}(V)} +
7183: \norm{\dbar v- \phi_* \scirc D_{u,J,\lambda}v}_{L^p(\scra_\lambda)} \bigr)
7184: \\
7185: \le
7186: C_2 \cdot \bigl(\norm{v}_{L^{1,p}(V)}
7187: + \norm{\dbar - \phi_* \scirc D_{u,J,\lambda}}_{\sf{op}}
7188: \cdot \norm{v}_{L^{1,p}(\scra_\lambda)}\bigr).
7189: \end{multline*}
7190: This yields \eqqref(5.3.4) provided $C_2 \cdot \norm{\dbar - \phi_* \scirc
7191: D_{u,J,\lambda}}_{\sf{op}} \le \half$.
7192:
7193: \smallskip
7194: Note once more that all the estimates in the proof are uniform in $\lambda \in
7195: \Delta(\epsi)$. \qed
7196:
7197:
7198: \smallskip
7199: \newlemma{lem5.3.7} The union $\sqcup_{\lambda \in \Delta(\epsi)} T\scrp(
7200: \scra_\lambda)$ is a continuous locally trivial Banach bundle over
7201: $\scrp(\scra)$.
7202: \end{lem}
7203:
7204: \proof We use the map $\pr_V$ to identify $\sqcup_{\lambda \in \Delta(\epsi)}
7205: \scrp(\scra_\lambda)$ with its image in $\scrp(V) \times \Delta(\epsi)$. By
7206: \lemma{lem5.3.4} \sliiip, we can we consider $\sqcup_{\lambda \in \Delta(\epsi)}
7207: T\scrp(\scra_\lambda)$ as subbundle of the bundle $T\scrp(V) \times \Delta(\epsi)$
7208: over $\scrp(V) \times \Delta(\epsi)$. This defines the canonical topology on
7209: $\sqcup_{\lambda \in \Delta(\epsi)} T\scrp(\scra_\lambda)$. Moreover,
7210: \lemma{lem5.3.4} \sliii implies the claim at all $(u,J,\lambda)$ with
7211: $\lambda\not=0$. Hence it remains to show the local triviality of the bundle
7212: in question in a neighborhood of a given $(u_0, J_0, 0) \in \scrp(\scra_0)$.
7213:
7214: Consider first the special case when $\norm{du_0}_{L^p(\scra_0)}$ is small
7215: enough. Under this assumption, \lemma{lem5.3.4} provides existence of
7216: a continuous family of splittings $L^{1,p}(\scra_\lambda, E_u) = \ker( D_{u,J,
7217: \lambda}) \oplus \im(T_{u,J,\lambda})$ defined in a neighborhood of $(u_0,J_0,0)$
7218: in $\sqcup_{\lambda \in \Delta(\epsi)} \scrp(\scra_\lambda)$. Moreover, it follows
7219: from the construction of $T_{u,J,\lambda}$ that the map $\dbar:\im(T_{u,J,\lambda})
7220: \to L^p_{(0,1)}(\scra_\lambda, \cc^n)$ is an isomorphism. By \refcorol{cor5.3.3b},
7221: this implies local triviality of $\im(T_{u,J,\lambda})$. In a similar way one
7222: shows that for a continuous family of isomorphisms $\phi_{u,J,\lambda}: (TB, J)
7223: \to (TB, J\st)$ the operators
7224: $$
7225: v \in \ker( D_{u,J, \lambda}) \subset L^{1,p}(\scra_\lambda, E_u)
7226: \mapsto \bigl(\id - T_\lambda \scirc \dbar\bigr)\phi_{u,J,\lambda} v
7227: \in \scrh(\scra_\lambda, \cc^n)\subset L^{1,p}(\scra_\lambda, \cc^n)
7228: $$
7229: are isomorphisms. This gives the local triviality of $\ker( D_{u,J, \lambda})$.
7230:
7231: Now observe that the bundle
7232: \begin{equation}\eqqno(5.3.f1)
7233: \{ (v, \dot J) \in T_{(u,J)}\scrp(\scra_\lambda):
7234: v = -T_{u,J,\lambda}(\dot J \scirc du \scirc J_\lambda) \}
7235: \end{equation}
7236: is a complement to $\ker( D_{u,J, \lambda})$ in $T_{(u,J)}\scrp(\scra_\lambda)$.
7237: Since the bundle \eqqref(5.3.f1) is isomorphic to the lift of $T\!\!\scrj(B)$ onto
7238: $\scrp(\scra_\lambda)$, it is locally trivial over the whole union $\sqcup_\lambda
7239: \scrp(\scra_\lambda)$.
7240:
7241: \smallskip
7242: Turn back to the general case of the lemma. For a given $(u_0, J_0, 0) \in
7243: \scrp(\scra_0)$, find a radius $r>0$ such that for the subnode
7244: $$
7245: \scra'_0 \deff \{ (z^+,z^-) \in \scra_0 : |z^+|<r,\; |z^-|<r\,\}
7246: = \Delta^+(r) \cup \Delta^-(r)
7247: $$
7248: the norm $\norm{du_0}_{L^p(\scra'_0)}$ is small enough. Define
7249: $$
7250: \textstyle
7251: A^+\deff \{ z^+ \in \Delta^+: {r \over2} <|z^+| <1\,\},
7252: \qquad
7253: A^-\deff \{ z^- \in \Delta^-: {r \over2} <|z^-| <1\,\},
7254: $$
7255: $$
7256: \scra'_\lambda \deff \{ (z^+,z^-) \in \scra_\lambda: |z^+|<r,\; |z^-|<r\,\}.
7257: $$
7258: and
7259: $$
7260: V' \deff A^+ \cup A^-,
7261: \qquad
7262: V'' \deff V' \cap \scra'_0.
7263: $$
7264: For $|\lambda|< {r \over2}$ we identify $A^\pm$ with the corresponding subsets in
7265: $\scra_\lambda$ by means of the coordinates $z^\pm$. Then for $|\lambda|< {r \over
7266: 4}$ $A^+$ is disjoint from $A^-$. This gives a family of coverings $\scra_\lambda=
7267: \scra'_\lambda \cup V'$ parameterized by $\lambda$ with $|\lambda|< {r \over4}$.
7268: Moreover, we can consider $V'$ and $V''$ as constant (\ie independent of $\lambda$)
7269: complex curves. Now, for any $(u,J,\lambda) \in \scrp(\scra)$ sufficiently
7270: close to $(u_0,J_0,0)$ we obtain the sequence
7271: \begin{equation}\eqqno(5.3.f2)
7272: 0 \to T_{(u,J,\lambda)} \scrp(\scra_\lambda)
7273: \buildrel \alpha_{u,J,\lambda} \over \lrar
7274: T_{(u,J,\lambda)} \scrp(V')
7275: \oplus T_{(u,J,\lambda)} \scrp(\scra'_\lambda)
7276: \buildrel \beta_{u,J,\lambda} \over \lrar
7277: T_{(u,J,\lambda)} \scrp(V'')
7278: \to 0,
7279: \end{equation}
7280: where we set
7281: $$
7282: \alpha_{u,J,\lambda}(v) \deff
7283: \bigl(v\ogran_{V'}, v\ogran_{\scra'_\lambda} )
7284: \qquad
7285: \beta_{u,J,\lambda}(v, w)
7286: \deff v\ogran_{V''} - w\ogran_{V''}.
7287: $$
7288: We claim that the sequence \eqqref(5.3.f2) is exact and splits.
7289: Since $\alpha_{u,J,\lambda}$ is obviously injective, it is sufficient
7290: to construct a right inverse to each $\beta_{u,J,\lambda}$, \ie a homomorphism
7291: $$
7292: \gamma_{u,J,\lambda}: T_{(u,J,\lambda)} \scrp(V'') \to
7293: T_{(u,J,\lambda)} \scrp(V')
7294: \oplus T_{(u,J,\lambda)} \scrp(\scra'_\lambda)
7295: $$
7296: such that $\beta_{u,J,\lambda} \scirc \gamma_{u,J,\lambda}= \id$. Furthermore,
7297: since $\beta_{u,J,\lambda}$ depends continuously on $(u,J,\lambda) \in \scrp(\scra
7298: )$ close to the $(u_0,J_0,0)$, it is sufficient to construct a right inverse only
7299: for the $\beta_{u_0,J_0,0}$.
7300:
7301: \smallskip
7302: Consider the restriction map between $L^{1,p}(\scra_0)$ and $L^{1,p}(V'')$ induced
7303: by the imbedding $V'' \hook \scra_0$. It is well-known that this map admits a left
7304: inverse, \ie a map $Q:L^{1,p}(V'') \to L^{1,p}(\scra_0)$ such that $Q(v'')\ogran
7305: _{V''} =v''$ for any $v''\in L^{1,p}(V'')$. Let us use the same notation for the
7306: restriction of $Q$ onto $T_{(u_0,J_0,0)}\scrp(V'') \subset L^{1,p}(V'')$. Recall
7307: that in \lemma{lem5.3.4} we have constructed the operator $T_{u_0,J_0,0}: L^p_{
7308: (0,1)}(\scra_0) \to L^{1,p}(\scra_0)$ which is a right inverse to the operator
7309: $D_{u_0,J_0,0}: L^{1,p}(\scra_0) \to L^p_{(0,1)}(\scra_0)$. Denote by $\Chi
7310: _{\scra'_0}: L^p_{(0,1)}(\scra_0, E_{u_0}) \to L^p_{(0,1)}(\scra_0, E_{u_0})$ the
7311: operator given the multiplication on the characteristic function of the subset
7312: $\scra'_0 \subset \scra_0$. Now, we obtain a left inverse $\gamma_{u_0,J_0,0}$ to
7313: the operator $\beta_{u_0,J_0,0}$ setting
7314: $$
7315: \gamma_{u_0,J_0,0} (v'') \deff
7316: \biggl(\!\!
7317: \bigl( T_{u_0,J_0,0} \scirc \Chi_{\scra'_0} \scirc D_{u_0,J_0,0}
7318: \scirc Q(v'') \bigr)\ogran_{V'},
7319: \bigl((T_{u_0,J_0,0} \scirc \Chi_{\scra'_0} \scirc D_{u_0,J_0,0} - \id)
7320: \scirc Q(v'') \bigr)\ogran_{\scra'_0}
7321: \biggr)
7322: $$
7323: for $v'' \in T_{(u_0,J_0,0)} \scrp(V'')$. Let us check that $\gamma_{u_0,J_0,0}$
7324: has the desired properties. For a given $v'' \in T_{(u_0,J_0,0)} \scrp(V'')$,
7325: set $\ti v'' \deff Q(v'')$. Then $D_{u_0,J_0,0}\ti v''$ vanishes on $V''$.
7326: Consequently, $D_{u_0,J_0,0}\ti v''$ is the sum of the forms $\phi, \psi
7327: \in L^p_{(0,1)}(\scra_0, E_{u_0})$ with the supports in $V'\bs V''$ and $\scra'_0
7328: \bs V''$ respectively. Moreover, $\psi = \Chi_{\scra'_0} \scirc D_{u_0,J_0,0}
7329: \ti v''$. From the relation $D_{u_0,J_0,0}\scirc T_{u_0,J_0,0} =\id$ we obtain
7330: $$
7331: D_{u_0,J_0,0} \bigl( T_{u_0,J_0,0} \scirc \Chi_{\scra'_0} \scirc D_{u_0,J_0,0}
7332: \scirc Q(v'') \bigr)\ogran_{V'} =
7333: \bigl( \Chi_{\scra'_0} \scirc D_{u_0,J_0,0}
7334: \scirc Q(v'') \bigr)\ogran_{V'} = 0.
7335: $$
7336: This means that the first component $\bigl( T_{u_0,J_0,0} \scirc \Chi
7337: _{\scra'_0} \scirc D_{u_0,J_0,0} \scirc Q(v'') \bigr)\ogran_{V'}$ of $\gamma
7338: _{u_0,J_0,0} (v'')$ satisfies $D_{u_0,J_0,0}v=0$. Thus the first component
7339: of $\gamma_{u_0,J_0,0}$ takes values in $T_{(u_0,J_0,0)}\scrp(V')$. In the same
7340: way one can show that the second component of $\gamma_{u_0,J_0,0}$ takes values
7341: in $T_{(u_0,J_0,0)}\scrp(\scra'_0)$. Finally, it is obvious that the difference
7342: of the components of $\gamma_{u_0,J_0,0} (v'')$ is $v''$.
7343: The lemma follows.\qed
7344:
7345:
7346: \statep Proof. of \refthm{thm5.3.1}. Take some $(u_0, J_0) \in \scrp(\scra_0)$.
7347: Using \lemma{lem5.3.3} \sliiip, we identify $\scrp(\scra_0)$ with its image
7348: $\pr_V\bigl( \scrp(\scra_0) \bigr) \subset \scrp(V)$ and $T_{(u_0, J_0)}\scrp
7349: (\scra_0)$ with its image $\pr_V\bigl( T_{(u_0, J_0)}\scrp (\scra_0) \bigr)
7350: \subset T_{(u_0, J_0)}\scrp(V)$.
7351:
7352: We claim that there exists a closed complement to $\pr_V\bigl( T_{(u_0, J_0)}
7353: \scrp (\scra_0)\bigr)$
7354: in $T_{(u_0, J_0)}\scrp(V)$. To show this, let us fix a closed nodal complex curve
7355: $C$ and an imbedding $\scra_0 \hook C$. Then there exists the unique open set $\wt
7356: V \subset C$ such that $\wt V \cap \scra_0 =V$ and $\wt V \cup \scra_0 =C$. Extend
7357: the bundle $E_{u_0} = u_0^*\cc^n$ to an $L^{1,p}$-smooth complex bundle $\wt E$
7358: over $C$. Also extend the operator $D_{u_0,J_0,0}: L^{1,p}(\scra_0, E_{u_0}) \to
7359: L^p_{(0,1)} (\scra_0, E_{u_0})$ to an operator $\wt D: L^{1,p}(C, \wt E) \to L^p
7360: _{(0,1)} (C, \wt E)$ of the form $\wt D= \dbar_{\wt E} + \wt R$ where $\dbar_{\wt
7361: E}$ is the Cauchy-Riemann operator corresponding to some holomorphic structure in
7362: $\wt E$ and $\wt R$ is a $\cc$-antilinear $L^p$-integrable homomorphism between
7363: $\wt E$ and $\wt E \otimes \Lambda^{(0,1)}C$, \ie $\wt R \in L^p\bigl(C, \barr
7364: \hom_\cc (\wt E, \wt E \otimes \Lambda^{(0,1)}C) \bigr)$. It is not difficult to
7365: show that the extensions $\wt E$ and $\wt D$ can be made in such a way that the
7366: operator $\wt D$ is an isomorphism. In particular, there exists the inverse
7367: operator $\wt T: L^p_{(0,1)}(C, \wt E) \to L^{1,p}(C, \wt E)$. For an open set
7368: $U \subset C$ define
7369: $$
7370: \scrh(U) \deff \{ v \in L^{1,p}(U, \wt E) : \wt Dv=0\}.
7371: $$
7372: In particular, we have $\scrh(\scra_0)=\pr_V\bigl( T_{(u_0,J_0)}\scrp(\scra_0)
7373: \bigr)$ and
7374: similar identifications for $V$ and $V'$. Define the operator $\beta: \scrh(\wt V)
7375: \oplus \scrh(\scra_0) \to \scrh(V)$ setting $\beta(v,w) \deff v\ogran_V -
7376: w\ogran_V$. Then there exists an operator $\gamma: \scrh(V)\to \scrh(\wt V)
7377: \oplus \scrh(\scra_0)$ which is left inverse to $\beta$. Indeed,
7378: the construction of the operator $\gamma_{u_0,J_0,0}$ made in the proof of
7379: \lemma{lem5.3.7} can be applied with appropriate modifications. In particular,
7380: one uses $\wt T$ instead of $T_{u_0,J_0,0}$.
7381:
7382: Observe that the kernel $\ker(\beta)$ can be identified in a natural way with
7383: the kernel of $\wt D: L^{1,p}(C, \wt E) \to L^p_{(0,1)} (C, \wt E)$. This implies
7384: that $\beta$ is an isomorphism and $\gamma$ its inverse. Consequently, every
7385: $v\in \scrh(V)$ can be uniquely represented in the form $v=v_1 + v_2$ such that
7386: $v_1$ extends to $\ti v_1 \in \scrh(\wt V)$ and $v_1$ extends to $\ti v_2 \in
7387: \scrh(\scra_0)$. The set of all $v_1\in \scrh(V)= T_{u_0,J_0}\scrp(V)$ obtained in
7388: this way forms the desired closed compliment to $\pr_V\bigl( T_{(u_0, J_0)}\scrp
7389: (\scra_0) \bigr)$ in $T_{(u_0, J_0)}\scrp(V)$.
7390:
7391: \medskip
7392: The existence of such a complement implies that there exists a small ball $\scru
7393: \subset \scrp(\scra_0)$ centered at $(u_0, J_0)$ and an open imbedding $F: \scru
7394: \times \scrb \hook \scrp(V)$ with the following properties:
7395: \begin{itemize}
7396: \item $\scrb$ is a small ball in a closed complement of $\pr_V \bigl(T_{(u_0,
7397: J_0)}\scrp(\scra_0) \bigr)$ in $T_{\pr_V(u_0, J_0)} \scrp(V)$;
7398: \item the map $F$ is a $C^1$-diffeomorphism onto image;
7399: \item the restricted map $F\ogran_{\scru \times \{0\}}: \scru \to \pr_V(\scru)
7400: \subset \pr_V(\scrp(\scra_0)) \subset \scrp(V)$ coincides with $\pr_V$.
7401: \end{itemize}
7402: In other words, $\scru \times \scrb$ appears as a chart for $\scrp(V)$ in
7403: which $\pr_V(\scru)= F(\scru \times \{0\})$ is a linear subspace.
7404:
7405: Now consider the natural projection $\pi_\scru: \scru \times \scrb \to \scru$
7406: restricted to $F\inv\bigl(\pr_V(\scrp(\scra_\lambda))\bigr)$. By
7407: \lemma{lem5.3.7} this is a $C^1$-diffeomorphism for all sufficiently small
7408: $\lambda$, provided the ball $\scru$ is small enough. Denote by
7409: $$
7410: \Phi_\lambda: \scru \to \pr_V\inv\bigl(\pr_V(\scrp(\scra_\lambda)) \cap
7411: F(\scru \times \scrb) \bigr) \subset \scrp(\scra_\lambda)
7412: $$
7413: the inverse map. We claim that the family $\Phi_\lambda$, parameterized by $\lambda
7414: \in\Delta(\epsi')$ with a sufficiently small $\epsi'>0$, has the properties
7415: stated in \refthm{thm5.3.1}. In fact, it remains to check only the fact that
7416: the whole map
7417: $$
7418: \Phi: \scru \times \Delta(\epsi') \to \sqcup_{\lambda
7419: \in\Delta(\epsi')} \scrp(\scra_\lambda)
7420: $$
7421: is continuous. However, this follows from the construction of $\Phi$. \qed
7422:
7423:
7424: \newsection[6]{Symplectic isotopy problem in $\cp^2$}
7425:
7426: \newsubsection[6.1]{Symplectic isotopy problem}
7427: %The symplectic isotopy problem comes in question as one of possible applications
7428: %of \refthm{thm4.5.1}. We describe it as follows.
7429: Let $\Sigma, \; \Sigma'$ be two (connected) symplectically imbedded surfaces in a
7430: symplectic 4-fold $(X, \omega)$. Assume that they have the same homology class.
7431: Then they have the same genus, see \lemma{lem1.1.2}. Thus one can ask whether or
7432: not there exists an isotopy $\{ \Sigma_t \}_{t \in [0, 1]}$ from $\Sigma$ to
7433: $\Sigma'$ such that all $\Sigma_t$ are also symplectically imbedded.
7434: This is refered to as the symplectic isotopy problem.
7435:
7436: The example of Fintushel and Stern \cite{Fi-St} shows that there is no hope
7437: %strangles the hope
7438: to obtain a results of this type in the case when $\la c_1(X), [\Sigma] \ra \le 0$.
7439: Namely, they proved that under certain conditions on a symplectic 4-fold $(X,
7440: \omega)$ there exists an infinite
7441: collection of symplectic imbeddings $\Sigma_i \hook X$, such that $\Sigma_i$
7442: represent the same homology class $[C] \in \sfh_2(X, \zz)$ but are pairwise
7443: non-isotopic, even smoothly. Moreover, the class of symplectic 4-folds with these
7444: conditions is sufficiently wide, so that one has enough examples of this type.
7445:
7446: \smallskip
7447: On the other hand, \refthm{thm4.5.1} hints that a satisfactory solution for the
7448: symplectic isotopy problem in the case $\la c_1(X), [C] \ra \ge 1$ is possible.
7449: We state the problem in a more precise form.
7450:
7451: \state Conjecture 1. {\sl (Symplectic isotopy problem). \it Let $(X,\omega)$ be
7452: a compact symplectic $4$-dimensional manifold and $[C]\in \sfh_2(X,\zz)$ a
7453: homology class with $\la c_1(X), [C] \ra \ge1$. Then every two symplectically {\sl
7454: immersed} surfaces $\Sigma$ and $\Sigma'$ in the class $[C]$ are symplectically
7455: isotopic provided they have the same genus $g$ and the only singularities are
7456: {\sl positive} nodal points.}
7457:
7458:
7459: Recall, there exists a complete classification of compact symplectic 4-folds
7460: $X$ which come in question. Namely, {\bf Corollary 1.5} in \cite{McD-Sa-3},
7461: claims
7462:
7463: \newprop{prop6.1.0a}
7464: Let $X$ be a symplectic manifold and $\Sigma
7465: \subset X$ a symplectically imbedded surface which is not an exceptional
7466: sphere. Then $X$ is the blow-up of a rational or ruled manifold.
7467: \end{prop}
7468:
7469: The complete description of possible symplectic structures on such $X$ was done in
7470: \cite{McD-4}, \cite{La-McD}, and \cite{McD-Sa-3}, see also \cite{Li-Liu},
7471: \cite{Liu}.
7472:
7473:
7474: As the main result of this paper we give a positive solution of the symplectic
7475: isotopy problem for imbeddings of low degree in $\cp^2$.
7476:
7477: \newthm{thm6.1.1} Any two symplectically imbedded surfaces $\Sigma,\; \Sigma'
7478: \subset \cp^2$ of the same degree $d\le6$ are symplectically isotopic.
7479: \end{thm}
7480:
7481: The case $d=1$ and $2$ of the theorem has been proven by Gromov in \cite{Gro},
7482: the case $d=3$ by Sikorav \cite{Sk-3}.
7483:
7484: \smallskip
7485: In this connection a result of S.~Finashin about (non-symplectic) isotopy problem
7486: in $\cp^2$ should be mentioned. He proves in \cite{Fin} that for any
7487: even degree $d=2k\ge6$ there exist infinitely
7488: many isotopy classes of imbedded real surfaces in $\cp^2$ having the degree $d$
7489: and the genus $g$ given by the genus formula, \ie $g={(d-1)(d-2) \over2}\cdot\;$
7490: Note that \refthm{thm6.1.1} claims that for $d=6$ only one of these isotopy
7491: classes is realizable by a {\sl symplectic} imbedding.
7492:
7493: \medskip
7494: Let us explain main ideas of the proof of \refthm{thm6.1.1}. First we observe
7495: that existence of a symplectic isotopy $\{ \Sigma_t \}_{t \in [0,1]}$ between
7496: surfaces $\Sigma,\; \Sigma'$ in a symplectic manifold $(X,\omega)$ implies
7497: existence of an ``accompanying'' homotopy $\{ J_t \}_{t \in [0, 1]}$ of tame
7498: almost complex structures, such that the imbeddings $\Sigma_t \hook X$ are
7499: $J_t$-holomorphic. Conversely a homotopy of $\omega$-tame $J_t$-holomorphic
7500: imbeddings is necessarily a symplectic isotopy. So given $\Sigma _0$ and
7501: $\Sigma_1$, the natural thing to do is to outfit them with compatible structures
7502: $J_0$ and $J_1$, take a generic curve $J_t$ and attempt to find appropriate
7503: liftings $\Sigma _t$. We do this using the following theorem of Harris \cite{Ha}
7504: for an intermediate construction.
7505:
7506:
7507: \newprop{prop6.1.2} Any two irreducible nodal algebraic curves $C_0$ and $C_1$ in
7508: $\cp^2$ of the same degree $d$ and the same geometric genus $g$ are {\sl
7509: holomorphically isotopic}, \ie can be connected by an isotopy $\{C_t\}_{t\in [0,
7510: 1]}$ consisting of nodal algebraic curves.
7511: \end{prop}
7512:
7513: By this result, in order to construct the symplectic isotopy, it is enough
7514: to construct a lifting as above for the case where $J_1$ is the standard
7515: integrable structure on $\cp^2$ and $\Sigma_1$ is some smooth algebraic curve.
7516:
7517: \medskip
7518: Obviously, \refthm{thm4.5.1} would imply existence of symplectic isotopy if we
7519: could show that for a generic path $\{J_t \}_{t \in [0,1]}$ the moduli space $\scrm
7520: _{J_t}$ is non-empty. An obstruction to this is the fact that the projection $\pi
7521: _\scrj:\scrm \to \scrj$ is not proper. This means that we must understand the
7522: structure of the total moduli space $\scrm$ ``at infinity''. In
7523: \refsubsection{5.2c} we have constructed a completion $\barm$ of $\scrm$ and
7524: eqquiped it with a natural stratification such that every stratum is a smooth
7525: Banach manifold. In particular, the transversality technique developed in
7526: \refsection{2} can be applied to every such stratum.
7527:
7528: The next idea in the proof of \refthm{thm6.1.1} is to construct a path $\ti\gamma
7529: _t \deff(C_t, J_t) \in \barr\scrm$ which goes piecewisely along some strata and
7530: which can be ``pushed'' into the ``main stratum'' $\scrm$ yielding the desired
7531: isotopy $(\Sigma_t, J_t)$. The main difficulty in realization this idea is to
7532: ensure that pushing $\ti\gamma_t$ into $\scrm$ we still remain in the same
7533: connected component of $\scrm$ so that the symplectic isotopy class is preserved.
7534: This means that we are interested in describing possible connected components of
7535: $\scrm$ in a neighborhood of a given curve $(C^*, J^*) \in \barr\scrm$. Moreover,
7536: the positive solution of a symplectic isotopy problem would follow immediately from
7537: the fact that locally exactly one such component exists. Indeed, it would be then
7538: sufficient to construct {\sl any} path $\ti\gamma_t\deff(C_t, J_t) \in \barr\scrm$
7539: connecting $\Sigma$ and $\Sigma'$. But existence of such a path follows easily
7540: from \refthm{thm4.5.1} in the case $c_1(X)[C]>0$.
7541:
7542: \smallskip
7543: The result of \refthm{thm6.1.1} is obtained via the proof of the local uniqueness
7544: of such a component of $\scrm$ near a given $(C^*, J^*) \in \barr\scrm$ in the
7545: special case when $C^*$ contains no multiple components. The restriction $d\le6$
7546: in the theorem comes from the fact that in this case it is possible to avoid the
7547: appearance of multiple components in $C^*$. We are able to do so by demanding that
7548: the pseudoholomorphic curves $\Sigma_t$ in the isotopy path pass through fixed
7549: generic $3d-1$ points on $X=\cp^2$. Note that the number $3d-1$ is the maximal
7550: possible in \refthm{thm4.5.3}.
7551:
7552:
7553:
7554: \newsubsection[6.2]{Local symplectic isotopy problem} As we have explained
7555: in the previous paragraph, we are interested in the possible symplectic isotopy
7556: classes of pseudoholomorphic curves $C$ in a neighborhood of a given singular curve
7557: $C^*$ with no multiple components. The main difficulty in this case is, of course,
7558: to understand the local behavior of curves $C$ near singular points of $C^*$.
7559: In this way we come to the following question.
7560:
7561: \state The Local Symplectic Isotopy Problem.
7562: {\sl Let $B$ be the unit ball in $\rr^4$ equipped with the standard symplectic
7563: structure $\omega\st$, $J^*$ an $\omega\st$-tame almost complex structure, and
7564: $C^*\subset B$ a connected $J^*$-holomorphic curve in $B$ with a unique isolated
7565: singularity at $0\in B$ and without multiple components. Describe the possible
7566: symplectic isotopy classes of curves $C$ in $B$ which lie sufficiently close to
7567: $C^*$ \wrt the cycle topology and which have prescribed singularities, \eg
7568: prescribed number of nodes and ordinary cusps.}
7569:
7570:
7571: \medskip
7572: We start with a construction of certain symplectic isotopy classes of nodal
7573: pseudoholomorphic curves. For $C^*$ as above, let $C^*= \cup_i C^*_i$ be the
7574: decomposition
7575: into irreducible components. Then there exist $J^*$-holomorphic parameterizations
7576: $u^*_i : S_i \to B$, $u^*_i(S_i) = C^*_i$. Shrinking $C^*_i$, if needed, we may
7577: assume that all $S_i$ are compact and smooth boundaries $\d S_i$, each
7578: consisting of finitely many circles. Note that the images of the boundary
7579: circles are imbedded in $B$ and mutually disjoint. Further, we can also suppose
7580: that $u^*_i$ are $L^{1,p}$-smooth up to boundaries $\d S_i$. Set $S\deff \sqcup
7581: S_i$ and define $u^*: S \to B$ by $u^*\ogran_{S_i} \deff u^*_i$. Denote by $J^*
7582: _S$ the complex structure on $S$ induced by $u^*: S \to B$ from $C^*$.
7583:
7584:
7585: \newlemma{lem6.2.1} \sli The set $\scrp(S,B)_\nod$ of those $(u, J_S, J) \in
7586: \scrp(S,B)$ for which the map $u:S \to B$ is an immersion and the singularities
7587: of the image $C \deff u(S)$ are only nodal points is open and dense in $\scrp(S,B)$
7588: and is connected;
7589:
7590: \slii For $(u', J'_S, J')$ and $(u'', J''_S, J'') \in \scrp(S,B)_\nod$,
7591: sufficiently close to $(u^*, J^*_S, J^*) \in \scrp(S,B)$, the pseudoholomorphic
7592: curves $C' \deff u'(S)$ and $C' \deff u'(S)$ are symplectically isotopic;
7593:
7594: \sliii For a fixed $J_S \in \scrj_S$ and $J\in \scrj(B)$, the subspace of nodal
7595: curves in each of the spaces $\scrp(S; B, J)$, $\scrp(S,J_S; B)$, and $\scrp(S,J_S;
7596: B, J)$ is open and dense in thecorresponding space.
7597: \end{lem}
7598:
7599: \proof By results of \refsection{4}, the complement to $\scrp(S,B)_\nod$ in
7600: $\scrp(S,B)$ is closed and consists of submanifolds of real codimension at least
7601: 2. This shows \sli and implies \slii. Part \sliii is obtained similarly.
7602: \qed
7603:
7604: \newdefi{def6.2.1} In the situation of \lemma{lem6.2.1}, we call $C_\nod=u(S)$ a
7605: {\sl maximal nodal deformation of $C^*$} and the number $\delta$ of nodes on $C
7606: _\nod$ the {\sl nodal number of $C^*$} at the singular point $0\in C^*$. In other
7607: words, a maximal nodal deformation is a nodal pseudoholomorphic curve obtained from
7608: $C^*= u^*(S)$ by a (sufficiently small) generic deformation of the parameterization
7609: map $u^*: S \to X$, $C^*= u^*(S)$.
7610:
7611: Further, a {\sl canonical smoothing of $C^*$} is a $J^*$-holomorphic curve
7612: $C^\dag$ obtained from a maximal nodal nodal deformation $C_\nod$ by smoothing of
7613: all nodes. We use the notion of canonical smoothing for both the construction and
7614: the resulting curve. Further, we shall always assume that a canonical smoothing
7615: $C^\dag$ is sufficiently close to $C^*$ \wrt the cycle topology.
7616: \end{defi}
7617:
7618: It follows immediately from \lemma{lem6.2.1} that the symplectic isotopy class
7619: of a canonical smoothing of $C^*$ is well-defined.
7620:
7621:
7622: \newprop{prop6.2.1a} Any two curves $C^\dag_1$ and $C^\dag_2$ obtained from $C^*$
7623: by the construction of canonical smoothing are symplectically isotopic. Moreover,
7624: such an isotopy can be be carried out sufficiently close to the identity map.
7625: \end{prop}
7626:
7627: \smallskip
7628: Note that the number $\delta(C_\nod)$ of nodes on a maximal nodal deformation
7629: $C_\nod$ of $C^*$ equals the nodal number $\delta(0, C^*)$ of $C^*$ at $0$. Observe
7630: also that one can smooth {\sl some} number of nodes on $C_\nod$ producing further
7631: symplectic isotopy classes. It is easy to show that these new classes are
7632: determined by the set of the nodes on $C_\nod$ which are smoothed. We conjecture
7633: that these are all possible symplectic isotopy classes of nodal curves in a
7634: neighborhood of $C^*$ \wrt the cycle topology.
7635:
7636: \state Conjecture 2. {\sl (Local symplectic isotopy problem for nodal curves).
7637: \it Let $J^*$ be a $C^2$-smooth $\omega\st$-tame almost
7638: complex structure in $B \subset \rr^4$ and $C^* \subset B$ a $J^*$-holomorphic
7639: curve with a unique isolated singular point at $0\in B$ and without multiple
7640: components. Assume that $J$ is an almost complex structure in $B$ which is $C^{0,
7641: \alpha}$-smooth for $\alpha>0$ and sufficiently close to $J^*$ \wrt the $C^{0,
7642: \alpha}$-topology.
7643:
7644: Then any {\it nodal} $J$-holomorphic curve $C$ sufficiently close to $C^*$ \wrt
7645: the cycle topology is symplectically isotopic to a $J^*$-holomorphic curve
7646: obtained from a maximal nodal deformation $C_\nod$ of $C^*$ by smoothing some
7647: number of nodes on $C_\nod$.}
7648:
7649:
7650: \medskip
7651: We give a proof the conjecture for the case of {\sl imbedded} curves. Observe
7652: that here we have only one candidate, namely the canonical smoothing.
7653:
7654:
7655: \newthm{thm6.2.2} In the situation described in {\sl Conjecture 2}, let $C^\dag$
7656: be $J^*$-holomorphic curve obtained by the canonical smoothing of $C^*$.
7657:
7658: Let $J$ be an almost complex structure on $B$ sufficiently close to $J^*$ \wrt the
7659: $C^{0,\alpha}$-topology and $C$ an imbedded $J$-holomorphic $C$ sufficiently close
7660: to $C^*$ \wrt the cycle topology. Then there exist a homotopy $J_t$ which is
7661: $C^0$-sufficiently close to $J^*$ and connects $J^*$ with $J$, and an isotopy
7662: $C_t$ of $J_t$-holomorphic curves which connects $C^\dag$ with $C$ and is
7663: sufficiently close to $C^*$ \wrt the cycle topology.
7664: \end{thm}
7665:
7666: \smallskip
7667: The proof will be given after some preparatory results. We shall always assume
7668: that the hypotheses of the theorem are fulfilled. Denote by $S$ the real surface
7669: parameterizing $C$. In other words $S$ is the curve $C$, considered as
7670: real oriented surface without complex structure.
7671:
7672: \smallskip
7673: Our first observation is that the theorem holds in the case when $C^*$ and the
7674: approximating curve $C$ are holomorphic in the usual sense. The result is
7675: well-known, see \eg \cite{Mil}. Its proof is based on the main advantage of the
7676: holomorphic case: the fact that one can represent a holomorphic curve as the zero
7677: divisor of a holomorphic function.
7678:
7679: \newlemma{lem6.2.2a} Let $f^*$ be a holomorphic function in the ball $B$ in
7680: $\cc^2$ whose zero divisor is a holomorphic curve $C^*$ with a single singular
7681: point at $0 \in B$ and without multiple components. Assume that $f^*$ and $C^*$
7682: are sufficiently smooth also at the boundary $\d B$. Then
7683: \begin{itemize}
7684: \item[\slip] a canonical smoothing $C$ is obtained as the zero divisor of
7685: a sufficiently small perturbation $f$ of $f^*$;
7686: \item[\sliip] for two generic sufficiently small perturbations $f_1$ and $f_2$
7687: of $f^*$ their zero divisors $C_0$ and $C_1$ are non-singular and {\sl
7688: holomorphically isotopic}, \ie can be connected by a homotopy consisting
7689: of holomorphic non-singular curves $C_t$.
7690: \end{itemize}
7691: \end{lem}
7692:
7693:
7694: \smallskip
7695: Denote by $\delta^*$ the nodal number of $C^*$ at $0\in C^*$. We may assume
7696: inductively that the claim of \refthm{thm6.2.2} holds for all curves $C'$ which
7697: satisfy the hypotheses of the theorem but have the nodal number $\delta(C')$ at
7698: $0\in C'$ which is strictly less than $\delta^*$. Further, we assume that $\delta^*
7699: \ge2$, since otherwise $\delta^*=1$ and $0\in C^*$ is a nodal point,
7700: the case covered by \refsubsection{5.3}.
7701:
7702: \medskip
7703: Recall that by the theorem of Micallef and White (see \lemma{lem1.2.1}) in a
7704: neighborhood of $0\in B$ there exists a $C^1$-diffeomorphism $\phi$ of $B\subset
7705: \rr^4$ such that $\phi(0)=0$, $\phi_*(J^*(0))=J\st$, the standard complex structure
7706: in $\rr^4=\cc^2$, and such that $\phi(C^*)$ is a $J\st$-holomorphic curve.
7707: Obviously, we may also assume that $d\phi : T_0B \to T_0B$ is the identity map.
7708: This means that the form $\phi_*\omega\st$ coincides with $\omega\st$ at $0\in B$,
7709: $\phi_* (\omega\st)\ogran_{T_0B}=\omega\st$, and similarly $\phi_*(J^*(0))=J^*(0)
7710: =J\st$. Consequently, $\phi_*(J^*)$ is $\omega\st$-tame in a sufficiently small
7711: ball $B(r)$, $r\ll1$. Let us fix such a radius $r$.
7712:
7713: Moreover, since $C^*\subset B$ is imbedded outside $0$, we can additionally assume
7714: that $\phi$ is {\sl smooth} outside $0\in B$.
7715:
7716: Below, we translate the original situation by means of such $\phi$ and work with
7717: a {\sl holomorphic} curve $\phi(C^*)\cap B(r)$. This leads to the difficulty that
7718: $\phi_*(J^*)$ is apriori only {\sl continuous} at $0\in B(r)$. This requires an
7719: additional control on the behavior of pairs $(C,J) \in \scrp(B)$ approximating
7720: $(C^*, J^*)$.
7721:
7722:
7723:
7724: \newlemma{lem6.2.3} Let $(u_n, J_{S,n}, J_n) \in \scrp(S, B)$ be a sequence such
7725: that $J_n$ converges to $J^*$ in the $C^{0,\alpha}$-topology with $0<\alpha <1$,
7726: and $C_n \deff u_n(S)$ converges to $C^*$ \wrt the cycle topology and \wrt the
7727: $L^{1,p}$-topology near boundary $\d C_n = u_n(\d S)$. Further, let $\phi: B \to
7728: B$ be the diffeomorphism introduced above. Then for all sufficiently big $n$
7729:
7730: \sli $u_n: S \to B$ is an imbedding;
7731:
7732: \slii there exists a sequence $J^*_n$ of $C^\ell$-smooth almost complex structures
7733: in $B$ such that
7734: \begin{itemize}
7735: \item $u_n$ are $(J_{S,n}, J^*_n)$-holomorphic;
7736: \item $\phi_*(J^*_n)$ converges to $J\st$ in the $C^0$-topology in $B(r)$ and
7737: in the $C^{0, \alpha}$-topology outside $0\in B(r)$.
7738: \end{itemize}
7739: \end{lem}
7740:
7741: \proof The first part follows from \lemma{lem1.2.2a}, applied to a smaller
7742: ball $B(\rho)$, $\rho<1$, and curves $C^* \cap B(\rho)$, $C_n \cap B(\rho)$.
7743:
7744: \smallskip
7745: Define $J^\sharp$ as the pull-back of $J\st$ \wrt $\phi$, $J^\sharp \deff \phi^*(
7746: J\st)$. Then the second part is equivalent to the convergence $J^*_n \lrar
7747: J^\sharp$ in the appropriate topology.
7748:
7749: Fix some sufficiently small $\epsi>0$. Since $J^*(0)=J\st=J^\sharp(0)$, there
7750: exists a positive radius $\rho\ll r$ such that $\norm{J^* -J^\sharp}_{C^0(B(\rho))}
7751: < \epsi$. This implies that $\norm{J_n -J^\sharp}_{C^0(B(\rho))} < \epsi$ for all
7752: sufficiently big $n$.
7753:
7754: Now observe that in $B\bs B(\rho)$ we have the $C^{1,\alpha}$ convergence $C_n
7755: \lrar C^*$. In particular, in $B\bs B(\rho)$ we have $C^{0,\alpha}$-convergence of
7756: tangent bundles $TC_n \lrar TC^*$. This implies that for $n \gg1$
7757: we can extend every $J_n$ from $B(\rho)$ to $B(r)$ as a $C^\ell$-smooth structure
7758: $J^*_n$ which is defined {\sl along} $C_n$ and obeys the estimate
7759: \begin{subequations}
7760: \eqqno(6.2.1)
7761: \begin{align}
7762: &\norm{J^*_n -J^\sharp}_{C^0(C_n\cap B(r))} < \epsi,
7763: \\
7764: &\norm{J^*_n -J^\sharp}_{C^{0,\alpha}(C_n \cap ( B(r) \bs B(2\rho) ))}
7765: < \epsi.
7766: \end{align}
7767: \end{subequations}
7768: Finally we extend the constructed $J^*_n$ from $C_n \cup B(\rho)$ to the whole
7769: ball $B$ preserving the estimates \eqqref(6.2.1). \qed
7770:
7771: \state Remark. In fact, below we shall merely make use of the weaker $C^0
7772: $-convergence $\phi_*(J^*_n) \to J\st$. The H\"older $C^{0,\alpha} $-convergence
7773: $J_n \to J^*$ was used only to provide the $C^0$-convergence of
7774: tangent bundles $TC_n \to TC^*$ outside $0\in C^*$. In particular, it would be
7775: sufficient to have only $C^0$-convergence $J_n \to J^*$ in $B$ and the $C^{0,
7776: \alpha}$-convergence outside $0\in B$. On the other hand, in the case when the
7777: convergence $J_n \to J^*$ is better, say in the $C^\ell$-topology with
7778: non-integer $\ell >1$, we could achieve just as as well the $C^\ell$-convergence in
7779: $B(r)$ outside $0$.
7780:
7781: \medskip
7782: \lemma{lem6.2.3} insures that we can reduce the problem to the case when $C^*$ is
7783: holomorphic in the usual sense, \ie \wrt the structure $J\st$. Further, observe
7784: that for the proof of \refthm{thm6.2.2} it is sufficient to show that for any
7785: sequence $(u_n, J_{S,n}, J_n)$ satisfying the hypotheses of \lemma{lem6.2.3} the
7786: curves $C_n \deff u_n(S)$ are symplectically isotopic to $C^\dag$ for $n\gg1$. An
7787: equivalent problem is to show that $\phi(C_n)$ are symplectically isotopic to
7788: $\phi(C^\dag)$ in $B(r)$. Thus we can replace our initial objects by their
7789: $\phi$-images in $B(r)$. For the sake of simplicity we maintain the original
7790: notations for these new objects, \eg $B$ for $B(r)$, $C^*$ and $C_n$ for
7791: respectively $\phi(C^*) \cap B(r)$ and $\phi(C_n) \cap B(r)$, $J^*$ and $J_n$ for
7792: respectively $\phi_*(J^*)\ogran_{B(r)}$ and $\phi_* (J_n)\ogran_{B(r)}$, and so
7793: on. On the other hand, $J\st$ and $\omega\st$ remain the standard structures in
7794: $B$. Observe that now we have the weaker $C^0$-convergence $J_n \lrar J^*$.
7795:
7796: \smallskip
7797: Imbed $B$ in $\cp^2$ in the standard way so that $J\st$ becomes the standard
7798: integrable structure, still denoted by $J\st$. Then we can extend $\omega$ to a
7799: global symplectic form on $\cp^2$ taming $J\st$. We maintain the notation $\omega$
7800: for this extension.
7801:
7802: We claim that $C^*$ also extends to $\cp^2$ as a compact closed pseudoholomorphic
7803: curve. Moreover, we claim that there exists an extension $\ti C^*$ with the
7804: following properties
7805: \begin{itemize}
7806: \item all irreducible components of $\ti C^*$ are rational, \ie
7807: parameterized by the sphere $S^2$;
7808: \item except for the original singularity at $0\in \ti C^*$, all new singularities
7809: are only nodal points.
7810: \end{itemize}
7811:
7812: Indeed, every irreducible component of $C^*\subset B$ is
7813: parameterized by a holomorphic map $u_i=u_i(z): \Delta \to B$ with $u_i(0)=0$. For
7814: every $u_i(z)$ we take the Taylor polynomials $u_i^{(d)}(z)$ of degree $d$ chosen
7815: sufficiently high to satisfy the following conditions:
7816: \begin{itemize}
7817: \item every $u_i^{(d)}(z)$ is non-multiple;
7818: \item the images $u_i^{(d)}(\Delta)$ are pairwise distinct holomorphic discs.
7819: \end{itemize}
7820:
7821: Then every $u_i^{(d)}(z)$ can be considered as an algebraic map $f_i$ from $\cp^1=
7822: S^2$ to $\cp^2$. Making a generic perturbation of $f_i$ outside $B$, we obtain
7823: desired curve $\ti C^* \subset \cp^2$ as the union of the images $\ti f_i(S^2)$ of
7824: the perturbed maps. Observe that $d$ appears as the degree of every component
7825: $\ti C^*_i \deff \ti f_i(S^2)$.
7826:
7827: \newlemma{lem6.2.4} There exist an almost complex structure $\ti J^*$ and points
7828: $x_\alpha$ on $\ti C^*$ satisfying the following conditions:
7829: \begin{itemize}
7830: \item[{\sl (a)}] the points $x_\alpha$ are pairwise distinct, and there
7831: are exactly $3d-1$ of them on every component $\ti C^*_i$;
7832: \item[{\sl (b)}] $\ti J^*$ is $C^\ell$-smooth and $\omega$-tame, $\ti C^*$ is
7833: $\ti J^*$-holomorphic, and $\ti J^*$ coincides with $J^*$ on $B$;
7834: \item[{\sl (c)}] any $\ti J^*$-holomorphic curve $C'$ which
7835: \begin{itemize}
7836: \item passes through the fixed points $x_\alpha$;
7837: \item is sufficiently close to $\ti C^*$ \wrt the cycle topology;
7838: \item has the same number of singular points as $\ti C^*$;
7839: \item[$*$] has a singular point $x'\in C'$ with the nodal number $\delta^*$
7840: at $x'$
7841: \end{itemize}
7842: must coincide with $\ti C^*$.
7843: \end{itemize}
7844: \end{lem}
7845:
7846: The last property asserts that every pseudoholomorphic curve $C'\not=\ti C^*$ with
7847: the properties {\sl(c)} except $(*)$ has simpler singularities than $\ti C^*$. So
7848: the induction assumption can be applied to such a $C'$.
7849:
7850: \proof We use the results of {\sl Sections \ref{sec:2}} and {\sl\ref{sec:4}}.
7851: Fix non-singular points $x_\alpha$ on $\ti C^*$ such that condition {\sl(a)}
7852: is fulfilled. Let $\mbfx_i$ be the $(3d-1)$-tuple of the points lying on the
7853: component $\ti C^*_i$. Denote by $\scrm'$ the space of pairs $(C',J')$,
7854: where $J'\in \scrj$ and $C'$ is $J'$-holomorphic curve $C'$ satisfying properties
7855: {\sl(c)} except $(*)$. Then by the genus formula \eqqref(1.2.1) any such curve
7856: $C'$ has only rational irreducible components $C'_i$, the number of which is the
7857: same as for $\ti C^*$, and the degree of
7858: every component $C'_i$ is $d$. This means that $\scrm'$ is the fiber product
7859: of the spaces $\scrm(S^2, \cp^2, d, \mbfx_i)$ of rational pseudoholomorphic curves
7860: of degree $d$ in $\cp^2$ passing through $\mbfx_i$. The product is taken
7861: over the space $\scrj$ of almost complex structure in $\cp^2$. By the transversality
7862: technique of \refsection{2}, the space $\scrm'$ a Banach manifold. To compute
7863: the Fredholm index of the natural projection $\pi'_\scrj: \scrm' \to \scrj$ observe
7864: that the expected dimension of rational $J$-holomorphic curves in $\cp^2$ of
7865: degree $d$ passing through $3d-1$ fixed distinct points is $0$. This implies that
7866: the index of the projection $\pi'_\scrj: \scrm' \to \scrj$ is also $0$.
7867:
7868: Further, by results of \refsection{4} the condition $(*)$ defines a proper
7869: $C^\ell$-smooth submanifold $\scrm^*$ in $\scrm'$ of finite codimension, say $m$.
7870: Consequently, the index of the corresponding projection $\pi^*_\scrj\deff \pi'
7871: _\scrj\ogran_{\scrm^*}: \scrm^* \to \scrj$ is negative. Using the transversality
7872: technique of \refsection{2} we can construct a $C^\ell$-smooth submanifold $Y
7873: \subset \scrm^*$ of dimension $m$ such that
7874: \begin{itemize}
7875: \item $(\ti C^*, \ti J^*)\in Y$ for some $\ti J^*$ obeying the condition
7876: {\sl(b)} of the lemma;
7877: \item $Y$ is transversal to $\scrm^*$;
7878: \item the restricted projection $\pi'_\scrj\ogran_Y: Y \to \scrj$ is an
7879: imbedding.
7880: \end{itemize}
7881: Then $(\ti C^*, \ti J^*)$ is an isolated point of the intersection $Y \cap
7882: \scrm^*$. But this means that $\ti J^*$ has the desired properties. \qed
7883:
7884:
7885:
7886: \medskip
7887: Below we shall need a property which is a bit sharper than {\sl(c)} in
7888: \lemma{lem6.2.4}. Roughly speaking, it claims that one can recover a
7889: pseudoholomorphic curve $C$ in $B$ knowing its part $\bigl(\barr B \bs B(\half)
7890: \bigr) \cap C$.
7891:
7892: \newdefi{def6.2.2} Denote by $A$ the spherical annulus $\barr B \bs B(\half)$.
7893: It is a closed subset of the closed unit ball $\barr B \subset \cp^2$. For
7894: closed subsets $Y_1, Y_2 \subset \cp^2$ we denote by $\dist_A(Y_1, Y_2)$ the
7895: Hausdorff distance between $Y_1 \cap A$ and $Y_2 \cap A$,
7896: $$
7897: \dist_A(Y_1, Y_2) \deff \dist(Y_1\cap A, Y_2\cap A),
7898: $$
7899: if both $Y_1 \cap A$ and $Y_2 \cap A$ are non-empty. The standard distance
7900: function in $\cp^2$ is used as the base. If exactly one of the set $Y_i \cap A$
7901: is empty, we set $\dist_A(Y_1, Y_2) \deff \diam(\cp^2)$. If $Y_1 \cap A= Y_2
7902: \cap A =\emptyset$, we define $\dist_A(Y_1, Y_2) \deff 0$. We call $\dist_A$
7903: the {\sl $A$-distance}.
7904: \end{defi}
7905:
7906: It is easy to see that $\dist_A$ is only a pseudo-distance function, \ie it is
7907: non-negative, symmetric, and has the triangle inequality property, but does not
7908: distinguish all closed subsets $Y_1 \not= Y_2 \subset \cp^2$ in general. It turns
7909: out that it induces the cycle topology on the set of pseudoholomorphic curves lying
7910: in a sufficiently small $\dist_A$-neighborhood of $\ti C^*$ provided only $C^1
7911: $-smooth almost complex structures $J$ are used. More precise statement is given in
7912:
7913: \newlemma{lem6.2.4a} There exists an $\epsi>0$ with the following property.
7914:
7915: Let $J \in \scrj$ be a $C^1$-smooth almost complex structure which satisfies the
7916: condition $\norm{J-J^*}_{C^0(\cp^2)} \le \epsi$ and $C$ a $J$-holomorphic curve
7917: which is homologous to $\ti C^*$ and satisfies the condition $\dist_A(C, \ti C^*)
7918: \le \epsi$. Then for any sequence $J_n$ of continuous almost complex
7919: structures $J_n$ converging to $J$ in the $C^0$-topology, $\norm{J_n-J}_{C^0(\cp
7920: ^2)} \lrar 0$, and any sequence of $J_n$-holomorphic curves $C_n$ the condition
7921: $\dist_A(C_n, C) \lrar 0$ implies that $C_n$ converges to $C$ in the cycle
7922: topology.
7923: \end{lem}
7924:
7925: \proof Consider a sequence of almost complex structures $J_n$ in $\cp^2$ which
7926: converges to $J^*$ in the $C^0$-topology, and a sequence $C_n$ of closed $J_n
7927: $-holomorphic curves homologous to $C^*$, for which $\lim \dist_A(C_n, \ti C^*)=
7928: 0$. Then $J_n$ are $\omega\st$-tame for all $n\gg1$. Hence we can apply the Gromov
7929: compactness theorem (see \refthm{thm5.2.1}). This means that some subsequence,
7930: still denoted $C_n$, converges to a $J^*$-holomorphic curve $C^+$ \wrt the cycle
7931: topology. The condition $\lim \dist_A(C_n, \ti C^*)= 0$ implies that $\dist_A(C^+,
7932: \ti C^*)=0$, which means that $\ti C^* \cap A = C^+ \cap A$.
7933:
7934: Observe now that by the construction of $\ti C^*$ every irreducible component $\ti
7935: C^*_i$ of $\ti C^*$ meets the interior $\sf{Int}(A)$ of $A$. By the unique
7936: continuation property of pseudoholomorphic curves proven in \lemma{lem1.2.4}
7937: \sliip, every component $\ti C^*_i$ is contained in $C^+$. Thus $\ti C^* \subset
7938: C^+$. Since $C^+$ is homologous to $\ti C^*$, we must have equality $\ti C^*= C^+$.
7939: This means that $C_n$ converges to $\ti C^*$ in the cycle topology. In particular,
7940: for every sufficiently big $n$ every irreducible component of $C_n$ meets the
7941: interior $\sf{Int}(A)$ of $A$.
7942:
7943: The latter property shows that the same argumentation can be used if we replace
7944: $\ti C^*$ by any $C_n$ with $n\gg1$ and the lemma follows. \qed
7945:
7946: \medskip
7947: Now we are ready to complete the
7948:
7949: \statep Proof. of \refthm{thm6.2.2}. It follows from the construction
7950: of the extension $\ti C^*$ that the sequence $(C_n, J_n)$ can be extended
7951: to a sequence $(\ti C_n, \ti J_n)$ such that $\ti J_n$ is a sequence of
7952: $\omega$-tamed almost complex structures in $\cp^2$ converging to $\ti J^*$
7953: and $\ti C_n$ is a sequence of compact (\ie closed) $\ti J_n$-holomorphic curves
7954: converging to $\ti C^*$. Moreover, we may additionally assume that the curves
7955: $\ti C_n$ pass through the marked points $\mbfx$ for all sufficiently big $n$.
7956:
7957: Observe that all $\ti C_n$ are symplectically isotopic. We denote by $\ti S$
7958: the closed oriented real surface parameterizing $\ti C_n$. It can be obtained
7959: from the surface $S$ parameterizing $C_n$ by gluing in discs to fill out the holes
7960: in $S$.
7961:
7962:
7963: Fix a sequence of homotopies $\{\ti J_{n,t}\}_{t\in [0,1]}$ of almost complex
7964: structures with the following properties:
7965: \begin{itemize}
7966: \item all $\ti J_{n,t}$ are $C^\ell$-smooth and depend $C^\ell$-smoothly on $t$;
7967: \item every initial structure $J_{n,0}$ is $\ti J_n$;
7968: \item for some small $\epsi_0>0$ the structures $\ti J_{n,t}$ are integrable in
7969: $B$ for all $t\in [1-\epsi_0, 1]$;
7970: \item as $n$ goes to infinity, the structures $\ti J_{n,t}$ converge to $\ti J^*$
7971: in the $C^0$-topology uniformly in $t\in [0,1]$, \ie
7972: $$
7973: \lim_{n\lrar\infty}\; \mathop{\sup}\limits_{t\in [0,1]}\;
7974: \norm{\ti J_{n,t} - \ti J^*}_{C^0(\cp^2)} =0;
7975: $$
7976: \item the homotopy $\{\ti J_{n,t}\}_{t \in [0,1]}$ is generic for every $n$.
7977: \end{itemize}
7978:
7979: \smallskip
7980: Now let us try to deform continuously every $\ti C_n$ inside a family $\ti J_{n,
7981: t}$-holomorphic curves preserving the isotopy class. Since we want to control also
7982: the local isotopy class we must impose the condition that the curves in the family
7983: lie sufficiently close to $\ti C^*$. Apriori, it can occur that such a curve does
7984: not exist for all $t\in [0,1]$. Nevertheless, we can find the maximal subinterval
7985: where such a family of curves exists. Moreover, we allow that under the deformation
7986: some nodal points appear. Let us formalize this observation.
7987:
7988: \newprop{prop6.2.6} Fix a sufficiently small $\epsi>0$. Then for every $n\gg1$
7989: there exists a $t^+_n \in (0,1]$ which is {\sl maximal} \wrt the following
7990: condition:
7991:
7992: For any $t <t^+_n$ there exists a $\ti J_{n,t}$-holomorphic curve $\ti C
7993: _{n,t}$ such that
7994: \begin{itemize}
7995: \item $\ti C_{n,t}$ passes through the fixed points $\mbfx$ on $\cp^2$;
7996: \item the curve $\ti C'_{n,t}$, obtained from $\ti C_{n,t}$ by smoothing of all
7997: singular points contained in $B$, is symplectically isotopic to $\ti C_n$;
7998: \item $\dist_A(\ti C_{n,t}, C^*) < \epsi$.
7999: \end{itemize}
8000: \end{prop}
8001:
8002: Recall that for $t$ sufficiently close to $1$ the structures $\ti J_{n,t}$ are
8003: integrable in $B$. So if $t^+_n =1$ for some $n$, then for some $t$ close to $1$
8004: we obtain a {\sl holomorphic} curve $C_{n,t} \deff \ti C_{n,t} \cap B$ whose
8005: smoothing is symplectically isotopic to the original curve $C_n$. In this case
8006: \refthm{thm6.2.2} follows from \lemma{lem6.2.2a}. We claim that it is always
8007: the case for $n\gg1$.
8008:
8009: \smallskip
8010: To show this, let us analyze the possible reasons which could cause the strict
8011: inequality $t^+_n<1$ for a given $n\gg1$.
8012: %To do it let us fix some $n\gg1$ and
8013: Consider
8014: an increasing sequence of parameters $t_\nu$ approaching to $t^+_n$. Then there
8015: exists a sequence of $\ti J_{n, t_\nu}$-holomorphic curves $\ti C_{n, t_\nu}
8016: $ with the properties from \propo{prop6.2.6}. In particular, all $\ti C_{n,
8017: t_\nu}$ are homologous to $\ti C^*$. Taking a subsequence, we may assume that
8018: $\ti C_{n, t_\nu}$ converges to a $\ti J_{n, t^+_n}$-holomorphic curve $\ti C^+_n$
8019: in the cycle topology. Note $\dist_A(\ti C^+_n, \ti C^*)\le \epsi$ by our
8020: construction. By \lemma{lem6.2.4a}, $\ti C^+_n$ is sufficiently close to
8021: $\ti C^*$ also \wrt the cycle topology. Consequently, near every nodal point of
8022: $\ti C^*$ there is exactly one nodal point of $\ti C^+_n$.
8023:
8024: Observe that $\ti C^+_n$ has no singular point $x^+_n \in \ti C^+_n$ with the
8025: nodal number $\ge \delta^*$ at $x^+_n$. Indeed, otherwise we can repeat the
8026: argumentation from the proof of \lemma{lem6.2.4} and show that $\ti C^+_n$ must
8027: consist of rational components the number of which is the same as that for
8028: $\ti C^*$. But the expected
8029: dimension of such curves in the space $\scrm'_{\ti J_{n, t^+_n}, \mbfx}$ with a
8030: singular point of this type is negative and less then $-1$. So the existence of
8031: $x^+_n\in \ti C^+_n$ with $\delta(x^+_n, \ti C^+_n) \ge \delta^*$ contradicts the
8032: genericity of the path $\ti J_{n, t}$. Thus all singularities of $\ti C^+_n$ are
8033: simpler than those of $\ti C^*$. By the induction assumption, the curve $\ti C'_n$
8034: obtained as the canonical smoothing of all singular points of $\ti C^+_n$
8035: contained in $B$ is symplectically isotopic to $\ti C_n$.
8036:
8037: Let $u^+_n: S^+_n \to \cp^2$ be a {\sl normal} parameterization of $\ti C^+_n$
8038: (see \refdefi{def5.2c.3}).
8039: Consider the relative moduli space $\scrm_{h_n, \mbfx}(S^+_n, \cp^2)$ of $h_n(t)=
8040: \ti J_{n,t}$-holomorphic curves which are parameterized by $S^+_n$, are in the
8041: homology class $[\ti C^*]$, and pass through the fixed points $\mbfx$. This space
8042: is non-empty since it contains $(\ti C^+_n, t^+_n)$. \refthm{thm4.5.3} provides
8043: that for some interval $t\in [t^+_n, t^{++}_n]$ with $t^{++}_n >t^+_n$ we can
8044: construct a path of $\ti J_{n,t}$-holomorphic curves $\ti C_{n, t}$ which lies in
8045: $\scrm_{h_n, \mbfx}(S^+_n, \cp^2)$ and starts at $\ti C^+_n$. Then the curves
8046: obtained from such $\ti C_{n, t}$ by smoothing of all singular points contained
8047: in $B$ will be symplectically isotopic to $\ti C_n$. Moreover,
8048: if we would additionally have the strict inequality $\dist_A(\ti C^+_n, \ti C^*)<
8049: \epsi$, then $\dist_A(\ti C_{n, t}, \ti C^*)< \epsi$ for some $t\in ]t^+_n,
8050: t^{++}_n[$, and this would contradict the maximality of $t^+_n$.
8051:
8052: \smallskip
8053: Thus we may assume that $\dist_A(\ti C^+_n, \ti C^*)= \epsi$ for every $n$. Then
8054: for every $n\gg1$ we can fix $t^-_n \in [0, t^+_n]$ and a $\ti J_{n,t^-_n}
8055: $-holomorphic curve $\ti C_{n,t^-_n}$ which has properties from \propo{prop6.2.6}
8056: and satisfies the additional condition
8057: $$
8058: \msmall{\epsi\over2} \le \dist_A(\ti C_{n, t^-_n}, \ti C^*) \le \epsi.
8059: $$
8060: Taking a subsequence, we may assume that $\ti C_{n,t^-_n}$ converges to a $\ti J
8061: $-holomorphic curve $\ti C^+$ in the cycle topology. Then
8062: $$
8063: \msmall{\epsi\over2} \le \dist_A(\ti C^+, \ti C^*)
8064: \le \epsi.
8065: $$
8066: By {\sl Lemmas \ref{lem6.2.4}} and {\sl\ref{lem6.2.4a}}, $\ti C^+$ must have
8067: simpler singularities than $\ti C^*$ provided the constant $\epsi$ was chosen
8068: small enough. By the induction assumption, the curve $\ti C'$ obtained by canonical
8069: smoothing of all singular points of $\ti C^+$ lying in $B$ is symplectically
8070: isotopic to every $\ti C_n$, as also to every $\ti C_{n,t^-_n}$. On the other
8071: hand, $\ti J^*$ coincide in $B$ with the standard structure $J\st$. Thus $C'
8072: \deff \ti C' \cap B$ is a canonical smoothing of $C^*$ by \lemma{lem6.2.2a}.
8073:
8074:
8075:
8076: \newsubsection[6.3]{Global symplectic isotopy in $\cp^2$} In this paragraph give
8077:
8078: \nobreak
8079: \statep Proof. of \refthm{thm6.1.1}. We proceed by making appropriate
8080: modifications of the argumentation used in the proof of \refthm{thm6.2.2}. Let
8081: $\Sigma$ be an imbedded surface in $\cp^2$ of degree $d\le 6$, such that $\omega\st
8082: \ogran_{\Sigma}$ is non-degenerate. By \propo{prop6.1.2}, to prove the theorem it
8083: is sufficient to show that $\Sigma$ is symplectically isotopic to a non-singular
8084: algebraic curve of degree $d$.
8085:
8086: \smallskip
8087: Find an $\omega\st$-tame almost complex structure $J_0$ making $\Sigma$ a $J_0
8088: $-holomorphic curve, denoted by $C_0$. Fix $3d-1$ distinct points $\mbfx=(x_1,
8089: \ldots, x_{3d-1})$ on $C_0$. Perturbing $C_0$ and the points, we may assume that
8090: $x_1,\ldots, x_{3d-1}$ are in generic position \wrt the standard structure $J\st$
8091: in the following sense. For any positive degree $d'\le d$ and any closed oriented
8092: surface $S$, not necessary connected, the moduli space $\scrm_{J\st, \mbfx}(S,
8093: \cp^2; d')$ of $J\st$-holomorphic (and hence {\sl algebraic}) curves of degree
8094: $d'$ with normalization $S$ passing through $\mbfx$ is a (possibly empty)
8095: complex space of the expected dimension.
8096:
8097: Fix a generic path $h(t)$ of $\omega\st$-tame almost complex structures $J_t \deff
8098: h(t)$ connecting $J_0$ with $J\st=J_1$. Without loss of generality we may assume
8099: that all $J_t$ are $C^\ell$-smooth and depend $C^\ell$-smoothly on $t$.
8100:
8101: \newprop{prop6.3.1} There exists a $t^+ \in (0,1]$ which is {\sl maximal} \wrt
8102: the following condition:
8103:
8104: For any $t <t^+$ there exists a $J_t$-holomorphic curve $C_t$ such that
8105: \begin{itemize}
8106: \item[\slip] $C_t$ passes through the fixed points $\mbfx$ on $\cp^2$;
8107: \item[\sliip] $C_t$ is non-multiple, but not necessarily irreducible;
8108: \item[\sliiip] the curve $C'_t$, obtained from $C_t$ by smoothing of all singular
8109: points, is symplectically isotopic to $C_0$.
8110: \end{itemize}
8111: \end{prop}
8112:
8113: To prove the theorem, we must show that $t^+=1$ and that there exist a $J\st
8114: $-holomorphic curve $C_1$ with the properties \slip--\slii in the proposition.
8115:
8116: \smallskip
8117: Let $t_n$ be an increasing sequence converging to $t^+$. Fix $J_{t_n}$-holomorphic
8118: curves $C_n$ with these properties. Property \slii implies that the $C_n$ have
8119: the same degree $d$. Going to a subsequence we may assume that they converge
8120: to a $J_{t^+}$-holomorphic curve $C^+$ in the cycle topology.
8121:
8122: We claim that $C^+$ contains no multiple components. To show this, it is sufficient
8123: to consider only the case when $C^+$ has only two components $C^+_1$ and $C^+_2$
8124: with multiplicities $m_1=1$ and $m_2=2$ respectively. Let $d_i$ be the degree of
8125: $C^+_i$, so that $d_1+2d_2=d$. Then the geometric genus $g_i$ of every $C_i$ is at
8126: most $g_i \le{(d_i-1)(d_i-2) \over 2}\cdot\;$ By the genericity of the path $h(t)=
8127: J_t$, each $C_i$ can contain at most $k_i \le 3d_i -1 + g_i \le{d_i(d_i+3) \over
8128: 2}$ of the fixed points $\mbfx$, see \refsubsection{2.4}. Thus $C^+$ can contain
8129: at most $\le{d_1(d_1+3) + d_2(d_2+3) \over 2}$ points. It is easy to show that for
8130: $d\le 6$ this number is strictly less then $3d-1$. For example, in the worst case
8131: with $d=6$, $d_1=4$, and $d_2=1$ we would have on $C^+$ at most ${4\cdot(4+3)
8132: \over 2} + {1\cdot (1+3)\over 2}= 14 +2 =16$ the marked points $\mbfx$ instead
8133: of the necessary $3\cdot 6-1 =17$. Observe, that this argument remains valid also
8134: in the case $t^+=1$ and $J_1=J\st$.
8135:
8136: Now the results of \refsubsection{6.2} show that the curve $C'$ obtained
8137: from $C^+$ by the
8138: canonical smoothing of all singular points is symplectically isotopic to $C_0$.
8139: This implies the theorem in the case $t^+=1$. Indeed, in this case $C^+$ is the
8140: zero divisor of a homogeneous polynomial $F^+$ of degree $d$. Making a generic
8141: perturbation of coefficients of $F^+$ we obtain a polynomial $F'$ whose
8142: zero divisor is an algebraic (and hence $J\st$-holomorphic) curve $C'$ which
8143: is symplectically isotopic to $C_0$.
8144:
8145: In the case $t^+ <1$ we must show that for some $t^{++}>t^+$ there exists a
8146: $J_{t^{++}}$-holomorphic curve $C^{++}$ with the properties given in
8147: \propo{prop6.3.1}. To do this we fix a normal parameterization $u^+: S^+ \to
8148: C^+ \subset \cp^2$ (see \refdefi{def5.2c.3}) and consider the relative moduli
8149: space $\scrm_{h, \mbfx}(S^+, \cp^2, d)$ of $J_t=h(t)$-holomorphic curves which
8150: are parameterized by $S^+$,
8151: pass through $\mbfx$ and have the degree $d$. This space is non-empty because it
8152: contains $C^+$. The results of \refsubsection{4.5} imply that for some interval
8153: $t\in [t^+_n, t^{++}_n]$ with $t^{++}_n >t^+_n$ we can construct a path of
8154: $J_t$-holomorphic curves $C^+_t$ which lies in $\scrm_{h, \mbfx}(S^+, \cp^2, d)$
8155: and starts at $C^+$. By \refsubsection{6.2}, the $C^+_t$ have the the properties
8156: \slip--\sliii from \propo{prop6.3.1}. This contradicts the maximality of $t^+$
8157: and implies the statement of \refthm{thm6.1.1}. \qed
8158:
8159: \medskip
8160: \state Remark. In fact, the real homotopy $C_t$ from $C_0=\Sigma$ to an algebraic
8161: curve $C_1$ has the property described in \refsubsection{6.1}. Namely, after
8162: fixing a generic homotopy $h(t)=J_t$, one tries to construct {\sl any} path $C_t$
8163: of {\sl imbedded} $J_t$-holomorphic curves $C_t$. Such a path exists for some
8164: interval $t\in [0, t')$. The saddle point property proven in
8165: \refsubsection{4.5} removes the main difficulty in the construction of $C_t$:
8166: the presence of local maxima in the corresponding moduli space $\scrm$. This means
8167: that at end of this interval, when $t \lrar t'$, the curves $C_t$ go to infinity in
8168: $\scrm$. By Gromov compactness, going along some sequence $t'_n \lrar
8169: t'$, we approach a $J_{t'}$-holomorphic curve $C'$ lying on some infinity
8170: stratum of $\barr \scrm$ parameterized by a new moduli space $\scrm'$. As we have
8171: shown in the proof, one can avoid the strata $\scrm'$ corresponding to curves with
8172: multiple components. Now we continue to deform $C'$ as a path $C'_t$ along
8173: $\scrm'$, having in mind that the canonical (in fact, any) smoothing of singular
8174: points of $C'_t$ gives curves symplectically isotopic to $C_0$. The new path $C'_t$
8175: continues until we come to the next infinity stratum $\scrm''$ of $\barr \scrm$,
8176: and so on.
8177:
8178: \medskip
8179: We finish the paper with a remark on {\sl Conjecture 2} about the local symplectic
8180: isotopy problem for nodal curves. The proof of this result would follow from the
8181: corresponding result for {\sl holomorphic} curves, which is essentially a local
8182: version of the Severi problem, see \propo{prop6.1.2}. Indeed, the proof of
8183: \refthm{thm6.2.2} could be applied after appropriate modification.
8184:
8185: \smallskip
8186: \state Conjecture 3. {\sl (Local Severi-Harris problem). \it Let $C^*$ be
8187: a holomorphic curve in the ball $B \subset \cc^2$ with a unique isolated singular
8188: point at $0\in B$ and without multiple components.
8189:
8190: Then any {\it nodal} holomorphic curve $C\subset B$ sufficiently close to $C^*$
8191: \wrt the cycle topology is holomorphically isotopic to a holomorphic curve
8192: $C^\dag$ obtained from a maximal nodal deformation $C_\nod$ of $C^*$ by smoothing
8193: some number of nodes on $C_\nod$.}
8194:
8195: \smallskip
8196: In view of the main results of this paper, the validity of {\sl Conjecture 3},
8197: and hence of {\sl Conjecture 2}, seems quite plausible.
8198:
8199:
8200: \vskip1cm plus 1cm
8201:
8202:
8203: \newpage
8204:
8205: %\input iso_bib.tex
8206: \ifx\undefined\bysame
8207: \newcommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\,}
8208: \fi
8209:
8210: \begin{thebibliography}{sh-sh-sh}
8211: %\show\bibitem
8212: %\makeatletter
8213: %\show\@bibitem
8214:
8215:
8216:
8217: \def\entry#1#2#3#4\par{\bibitem[#1]{#1}
8218: {\textsc{#2 }}{\sl{#3}}#4\par\vskip2pt}
8219: %{ref}{author}{title}.
8220:
8221: \entry{Abi}%
8222: {Abikoff, W.}%
8223: {The real analytic theory of Teichm\"uller space.}
8224: Lecture Notes in Mathematics, {\bf820}. Springer, Berlin, 1980. vii+144 pp.,
8225: {\bf Math.\ Rev.:\.}82a:32028.
8226:
8227: \entry{Ar}{Aronszajn, N.:}{A unique continuation theorem for elliptic
8228: differential equation or inequalities of the second order.} J.\ Math.\ Pures
8229: Appl., {\bf36}, 235--339, (1957).
8230:
8231: %\entry{B-K}{Bedford, E., Klingenberg W.:}{On the~envelope of meromorphy of
8232: %a 2-sphere in $\cc^2$.} J.\ Amer.\ Math.\ Soc., {\bf4}, 623-646, (1981).
8233:
8234:
8235: \entry{B-P-V}{Barth W., Peters, C., Van de Ven, A.:}{Compact complex
8236: surfaces.} Springer Verlag, (1984).
8237:
8238: %
8239: \entry{Bi}{Birman, J.:}{Braids, links, and mapping class groups.}
8240: Annals of Math.\ Studies., {\bf82}(1975), Princeton Univ.\ Press and Univ.\
8241: of Tokyo Press., 229 p., (1975).
8242:
8243: \entry{Bn}{Bennequin, D.:}{Entrelacement et \'equation de Pfaff.}
8244: Ast\'erisque {\bf107--108}(1982), 87--161.
8245:
8246:
8247: \entry{Ch-Sp}{Chern, S.-S., Spanier, E.:}{A theorem on orientable surfaces
8248: in four-dimensional space.} Comment.\ Math.\ Helv., {\bf25}(1951), 205--209.
8249:
8250:
8251: %\entry{Ch}{Chirka E.:}{On generalized Hartogs lemma.} To appear in the
8252: %volume dedicated to B.\.V.\.Shabat, PHASIS, Moscow.
8253:
8254: %\entry{Ch-St}{Chirka, E., Stout L.:}{A Kontinuit\"atssatz.} In ``Topics in
8255: %Complex Analysis'', Banach Center Publications, {\b31}, 143--350, (1995).
8256:
8257: \entry{Diaz}%
8258: {Diaz, Steven}%
8259: {Exceptional Weierstrass points and the divisor on moduli space that they
8260: define.} Mem.\ AMS, {\bf56}(1985), No.\.327, iv+69 pp.,
8261: {\bf Math.\ Rev.:\.} 86j:14022
8262:
8263: \entry{Eli}{Eliashberg, Y.:}{Filling by holomorphic discs and its applications.}
8264: London Math.\ Soc.\ Lecture Notes, {\bf151}, Geometry of low dimensional
8265: manifolds, (1991).
8266:
8267: \entry{Fi-St}%
8268: {Fintushel, Ronald; Stern, Ronald J.:}%
8269: {Symplectic surfaces in a fixed homology class.}
8270: J.\ Diff.\ Geom., {\bf52}(1999), No.\.2, 203--222.
8271:
8272: \entry{Fin}%
8273: {Finashin, Sergey:}%
8274: {Knotting of Algebraic Curves in $\cp^2$.}
8275: Preprint, {\tt arXiv:math.GT/9907108}.
8276:
8277: %\entry{F}{Forstneri\v c, F.:}{Complex tangents of real surfaces in complex
8278: %Ssurfacea.} Duke Math.\ J., {\bf67}, 353--376, (1992).
8279:
8280: %\entry{Gra}{Grauert, H.:}{\"Uber Modifikationen und exzeptionelle analytische
8281: %Mengen.} Math.Ann. {\bf146}, 331--368 (1962).
8282:
8283: \entry{Gr-Ha}{Griffiths, P., Harris, J.:}{Principle of algebraic geometry.}
8284: John Wiley \& Sons, N.-Y., (1978).
8285:
8286:
8287: \entry{Gro}{Gromov, M.:}{Pseudo holomorphic curves in symplectic
8288: manifolds.} Invent.\ math. {\bf82}, 307--347 (1985).
8289:
8290: \entry{Ha}{Harris, J.:}{On the Severi problem.} Invent.\ Math.,
8291: {\bf84}(1986), 445-461.
8292:
8293: \entry{Ha-Mo}%
8294: {Harris, J; Morrison, I.:}%
8295: {Moduli of curves.}
8296: Graduate Texts in Mathematics, {\bf187},
8297: Springer-Verlag, 1998, xiv+366 pp., ISBN 0-387-98438-0,
8298: {\bf Math.\ Rev.:}\.99g:14031.
8299:
8300: %\entry{Hv-Po}{Harvey, R., Polking, J.:}{Removable singularities of
8301: %solutions of partial differential equations.} Acta Math. {\bf125}, 209--226
8302: %(1970).
8303:
8304:
8305: \entry{Hrt-W}{Hartman, P., Winter, A.:}{On the~local behavior of solutions
8306: of non-parabolic partial differential equations.} Amer.\ J.\ Math.,
8307: {\bf75}, 449--476, (1953).
8308:
8309: %
8310:
8311: \entry{Hf}{Hofer, H.:}{Pseudoholomorphic curves in symplectizations with
8312: applications to the Weinstein conjecture in dimension three.} Invent.\ Math.,
8313: {\bf114}, 515--563 (1993).
8314: %
8315:
8316: \entry{H-L-S}{Hofer, H., Lizan, V., Sikorav, J.-C.:}{On genericity for
8317: holomorphic curves in four-dimensional almost-complex manifolds.} J.\ of Geom.\
8318: Anal., {\bf7}, 149--159, (1998).
8319:
8320: \entry{Hum}%
8321: {Hummel, Christoph.:}%
8322: {Gromov's compactness theorem for pseudo-holomorphicmorphic curves.}
8323: Progress in Mathematics, {\bf151},
8324: Birkh\"auser Verlag, Basel, 1997, viii+131pp.,
8325: {\bf Math.\ Rev.:}\.98k:58032.
8326:
8327: %\entry{Iv-1}{Ivashkovich, S.:}{Extension of analytic objects by the method
8328: %of Cartan-Thullen.}In Proceedings of Conference on Complex Analysis and Math.
8329: %Phys., 53--61, Krasnojarsk, (1988).
8330:
8331: %\entry{Iv-2}{Ivashkovich, S.:}{The Hartogs-type extension theorem for
8332: %meromorphic maps into compact K\"ahler manifolds.} Invent. Math. {\bf109},
8333: %47--54 (1992).
8334:
8335: \entry{Iv-Sh-1}{Ivashkovich, S., Shevchishin, V.:}{Pseudoholomorphic curves
8336: and envelopes of meromorphy of two-spheres in $\cc\pp^2$.} Preprint, Bochum
8337: (1995); available as e-print {\tt ArXiv:math.CV/9804014}.
8338:
8339: \entry{Iv-Sh-2}{Ivashkovich, S., Shevchishin, V.:}%
8340: {Deformations of noncompact complex curves, and envelopes of meromorphy of
8341: spheres.} (Russian) Mat.\ Sb.\ {\bf189}(1998), No.9,
8342: 23--60; translation in Sb.\. Math.\ {\bf189}(1998), No.9--10, 1335--1359.
8343:
8344:
8345: \entry{Iv-Sh-3}{Ivashkovich, S., Shevchishin, V.:}{Gromov compactness theorem
8346: for stable curves.} Preprint, Bochum, (1999); available at
8347: {\tt ArXiv:math.DG/9903047}, to appear in Int.\ Math.\ Res.\ Notes.
8348:
8349: \entry{Ko-No}{Kobayashi, S., Nomizu, K.:}{Foundations of differential
8350: geometry.} Vol.II, Interscience Publishers, (1969).
8351:
8352: \entry{K}{Kontsevich M.:}{Enumeration of rational curves via torus actions.}
8353: Proc.\ Conf.\ ``The moduli spaces of curves'' on Texel Island, Netherland.
8354: Birkh\"auser Prog.\ Math., {\bf129}(1995), 335--368.
8355:
8356: \entry{K-M}{Kontsevich M., Manin Yu.:}{Gromov-Witten classes, quantum
8357: cohomology, and enumerative geometry.} Comm.\ Math.\ Phys., {\bf164}(1994),
8358: 525--562.
8359:
8360:
8361: \entry{Knu}%
8362: {Knudsen, Finn F.:}%
8363: {The projectivity of the moduli space of stable curves}
8364: {\sl I.:} Math.\ Scand., {\bf39} (1976), No. 1, 19--55;
8365: {\bf Math.\ Rev.:}\.55\#10465; \
8366: {\sl II.:} Math.\ Scand., {\bf52}(1983), No. 2, 161--199.
8367: {\bf Math.\ Rev.:}\.85d:14038a; \
8368: {\sl II.:} Math.\ Scand., {\bf52}(1983), No. 2, 200--212.
8369: {\bf Math.\ Rev.:}\.85d:14038b. \
8370:
8371: \entry{La-McD}{Lalond, F., Mcduff, D.:}{The classification of ruled
8372: symplectic manifolds.} Math.\ Res.\ Lett., {\bf3}, 769--778, (1996).
8373:
8374: \entry{Li-Liu}{Li, T.-J., Liu, A.-K.:}{Symplectic structure on ruled surfaces
8375: and a generalized adjunctionformula.} Math.\ Res.\ Lett., {\bf2}, 453--471,
8376: (1995).
8377:
8378: \entry{Li-T}%
8379: {Li, Jun; Tian, Gang}%
8380: {Virtual moduli cycles and Gromov-Witten invariants of general symplectic
8381: manifolds.}
8382: Topics in symplectic $4$-manifolds, (Irvine, CA, 1996), 47--83,
8383: First Int. Press Lect. Ser., I,
8384: Internat. Press, Cambridge, MA, 1998,
8385: {\bf Math.\ Rev.:}2000d:53137.
8386:
8387:
8388: \entry{Liu}{Liu, A.-K.:}{Some new applications of general wall crossing
8389: formula, Gompf's conjecture and its applications.}
8390: Math.\ Res.\ Lett., {\bf3}, 569--585, (1996).
8391:
8392:
8393: \entry{Lich}{Lichnerovicz, A:}{Global theory of connections and holonomy
8394: groups.} Noordhoff Int. Publishing, Leiden, (1976).
8395:
8396: \entry{Mi-Wh}{Micallef, M., White, B.:}{The structure of branch points in
8397: minimal surfaces and in pseudoholomorphic curves.} Ann.\ Math., {\bf 139},
8398: 35-85 (1994).
8399:
8400: \entry{Mil}{Milnor, J.:}{Singular points of complex hypersurfaces.} Annals of
8401: Mathematics Studies., {\bf61},. Princeton, N.J.: Princeton University Press and the
8402: University of Tokyo Press. 122 p., (1968).
8403:
8404: \entry{McD-1}{McDuff, D.:}{The local behavior of holomorphic curves in
8405: almost complex 4-manifolds.} J.Diff.Geom. {\bf 34}(1991), 143-164.
8406:
8407: \entry{McD-2}{McDuff, D.:}{Examples of symplectic structures.} Invent.\
8408: Math., {\bf 89}(1987), 13--36.
8409:
8410: \entry{McD-3}{McDuff, D.:}{Singularities and positivity of intersections of
8411: $J$-holomorphic curves.} In {\it``Holomorphic curves in symplectic
8412: geometry".} Ed. by M.~Audin and J.~Lafontaine, Birkh\"auser, (1994).
8413:
8414: \entry{McD-4}{McDuff, D.:}{The structure of rational and ruled
8415: symplectic manifolds.} J.\ AMS, {\bf3}(1990), 679--712; Erratum:
8416: J.\ AMS, {\bf5}(1992), 987--988,.
8417:
8418: \entry{McD-5}{McDuff, D.:}{Blow ups and symplectic embeddings in dimension 4.}
8419: Topology, {\bf30}(1991), 409-421.
8420:
8421: \entry{McD-Sa-1}{McDuff, D., Salamon, D.:}{Introduction to symplectic
8422: topology.} Clarendon Press, Oxford, 425+viii p., (1995).
8423:
8424: \entry{McD-Sa-2}{McDuff, D., Salamon, D.:}{Introduction to symplectic
8425: topology.} {\it 2nd edition}, (1998).
8426:
8427: \entry{McD-Sa-3}{McDuff, D., Salamon, D.:}{A survey of symplectic
8428: $4$-manifolds with $b_+ =1$.} Turk.\ J.\ Math. {\bf20}, 47--60, (1996).
8429:
8430: \entry{Mo}{Morrey, C.:}{Multiple integrals in the calculus of variations.}
8431: Springer Verlag, (1966).
8432:
8433: \entry{MFK}%
8434: {Mumford, D., Fogarty, J., Kirwan, F.:}%
8435: {Geometric invariant theory,}
8436: third edition, Springer-Verlag, Berlin, 1994. xiv+292 pp.,
8437: ISBN 3-540-56963-4,
8438: {\bf Math.\ Rev.:}\.95m:14012
8439:
8440:
8441: \entry{N-1}{Nemirovsky S.:}{Stein domains on algebraic surfaces.} Russian
8442: Math.\ Notes, {\bf60}(1996), 295--298 (1996).
8443:
8444: \entry{N-2}{Nemirovsky S.:}{Holomorphic functions and imbedded real surfaces.}
8445: Math.\ Notes, {\bf63}(1998), 527--532.
8446:
8447: \entry{Pa-Wo}{Parker, Th.\ H., Wolfson, J.\ G.:}{Pseudo-holomorphic maps
8448: and bubble trees.} J.~Geom.\ Anal., {\bf3}, 63-98, (1993).
8449:
8450: \entry{Ra}{Ramis J.-P.:}{Sous-ensembles analytiques d'une vari\'et\'e
8451: banachique complex.} Springer, Berlin (1970).
8452:
8453: %\entry{Rm}{Remmert, R.:}{Holomorphe und meromorphe Abbildungen komplexer
8454: %R\"aume.} Math.\ Ann., {\bf133}, 328--370 (1957).
8455:
8456: \entry{Rf}{Rolfsen, D.:}{Knots and links.} 2nd print, Math.\ Lect.\ Series,
8457: {\bf N7}, xiv+439pp., Publish or Perish, (1990).
8458:
8459: %\entry{R}{Rosay J.-P.:}{An example to the question of Chirka.} Preprint.
8460:
8461: \entry{Ru}%
8462: {Ruan, Yongbin}%
8463: {Virtual neighborhoods and pseudo-holomorphic curves.}
8464: Proceedings of 6th G\"okova Geometry-Topology Conference.
8465: Turkish J.\ Math., {\bf23}(1999), 161--231,
8466: {\bf Math.\ Rev.:} CMP 1\.701\.645 (2000:03).
8467:
8468: \entry{S-U}{Sacks J., Uhlenbeck K.:}{Existence of minimal immersions of
8469: two-spheres} Annal.\ Math., {\bf113}(1981), 1--24.
8470:
8471: \entry{Sk-1}{Sikorav, J.-C.:}{Some properties of holomorphic curves in almost
8472: complex manifolds.} In \it ``Holomorphic curves in symplectic geometry .''
8473: \rm Ed. by M.~Audin and J.~Lafontaine, Birkh\"auser, (1994).
8474:
8475: \entry{Sk-2}{Sikorav, J.-C.:}{Singularities of $J$-holomorphic curves.}
8476: Math.\ Z., {\bf226}, 359--373, (1997).
8477:
8478: \entry{Sk-3}{Sikorav, J.-C.:}{The gluing construction for normally generic
8479: $J$-holomorphic curves.} Preprint No.\.UMPA--2000-n$^\circ$264.
8480:
8481: \entry{Sie}%
8482: {Siebert, Bernd}%
8483: {Symplectic Gromov-Witten invariants.}
8484: New trends in algebraic geometry, (Warwick, 1996), 375--424,
8485: London Math. Soc. Lecture Note Ser., {\bf264},
8486: Cambridge Univ. Press, Cambridge, 1999,
8487: {\bf Math.\ Rev.:\.}1 714 832 (2000:02).
8488:
8489:
8490: %\entry{Si-1}{Siu, Y.-T.:}{Extension of meromorphic maps into K\"ahler
8491: %manifolds.} Annals of Math. {\bf102}, 421--462 (1975).
8492:
8493: %\entry{Si-2}{Siu, Y.-T.:}{Every Stein subvariety admits a Stein
8494: %neighborhood.} Invent.math. {\bf38}, 89--100 (1976).
8495:
8496: \entry{Sh}{Shafarevich I.:}{Basic algebraic geometry. Second edition.}
8497: Springer-Verlag.
8498:
8499: %\entry{Shch}{Shcherbina, N.:}{On the~polynomial hull of a two-dimensional
8500: %sphere in $\cc^2$.} Soviet Math.\ Doklady, {\bf49}, 629--632, (1991).
8501:
8502: \entry{Wn}{Weinstein, A.:}{Lectures on symplectic manifolds.} CBMS. Reg.
8503: Conf. Series, {\bf N29}, AMS(1977).
8504:
8505: \entry{Wa}{Walker, R.:}{Algebraic curves.} Springer Verlag, (1978).
8506:
8507: \entry{Ye}%
8508: {Ye, Rugang}%
8509: {Gromov's compactness theorem for pseudo holomorphic curves.}
8510: Trans.\ AMS, {\bf342}(1994), 671--694,
8511: {\bf Math.\ Rev.:}\.94f:58030.
8512:
8513:
8514:
8515: %%%
8516: \end{thebibliography}
8517:
8518: %%% Local Variables:
8519: %%% mode: latex
8520: %%% TeX-master: "iso"
8521: %%% End:
8522:
8523:
8524:
8525: \newpage
8526: \tableofcontents
8527:
8528: %\printindex
8529: \end{document}
8530:
8531: \font\sevensl=cmsl7 %cmsl10 scaled 694
8532: \font\fivesl=cmsl5 %cmsl10 scaled 482
8533:
8534:
8535: % LocalWords: resp eps
8536:
8537: %%% Local Variables:
8538: %%% mode: latex
8539: %%% TeX-master: t
8540: %%% End:
8541: