1: \NeedsTeXFormat{LaTeX2e}
2: \documentclass[12pt,a4paper]{article}
3:
4:
5: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6: % Definitions for the article comin %
7: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8:
9: %%%%%%%%%%%%%%
10: % Packages %
11: %%%%%%%%%%%%%%
12:
13: \usepackage{a4,enumerate}
14: \usepackage{amsthm}
15: \usepackage{amssymb}
16: \usepackage{amsmath}
17: \usepackage{latexsym}
18: \usepackage{pstricks,pst-coil,pst-plot,pst-node}
19: \usepackage{graphicx}
20:
21: %%%%%%%%%%%%%%%%%%%%%%%%
22: % Arbeitseinstellungen %
23: %%%%%%%%%%%%%%%%%%%%%%%%
24: %\usepackage{showkeys}
25: %\long\def\regie#1{------{\bf #1}------}
26: %\long\def\regie#1{}
27: %\newcommand{\mnote}[1]{\marginpar{\tiny\em #1}}
28: %\def\mnote#1{}
29: \long\def\bild#1{#1}
30: %\long\def\bild#1{\ \newline{\bf !!!!!!Hier erscheint ein Bild !!!!}\newline}
31: \def\url#1{{\tt #1}}
32:
33: %%%%%%%%%%%%%%%%%%%%%%%%%%
34: % Globale Formatierungen %
35: %%%%%%%%%%%%%%%%%%%%%%%%%%
36:
37: \parindent0cm
38: \parskip=.5\baselineskip
39:
40:
41: %%%%%%%%%%%%%%%%%
42: %Seiten-Format %
43: %%%%%%%%%%%%%%%%%
44:
45: \oddsidemargin0.15cm
46: \evensidemargin0.73cm
47: \topmargin0.6cm
48: \headsep.9cm
49: \textwidth15cm
50: \textheight42\baselineskip
51:
52:
53: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
54: % Temporaer verwendete Groessen %
55: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
56:
57: \newdimen\templaenge
58:
59: %%%%%%%%%%%%%%%%
60: % Fonts %
61: %%%%%%%%%%%%%%%%
62:
63: %%% Doppelbalkenfonts
64: \DeclareMathAlphabet{\doba}{U}{msb}{m}{n}
65: \gdef\mC{\doba{C}}
66: \gdef\mH{\doba{H}}
67: \gdef\mN{\doba{N}}
68: \gdef\mP{\doba{P}}
69: \gdef\mQ{\doba{Q}}
70: \gdef\mR{\doba{R}}
71: \gdef\mS{\doba{S}}
72: \gdef\mT{\doba{T}}
73: \gdef\mZ{\doba{Z}}
74:
75: %%% Latex2e-Variante von \qed
76: \def\qed{{\leavevmode\unskip\nobreak\hfil\penalty 50\hskip 1em%
77: \hbox{}\nobreak\hfil\lower 1pt\hbox{$\Box$\kern-.5pt}\parfillskip 0pt
78: \finalhyphendemerits 0\par\bigbreak}}
79: \def\qedmath#1{\setbox0\hbox{$\displaystyle #1$}\templaenge=\textwidth\advance\templaenge by -\wd0%
80: \setbox1\hbox{$\Box$}\advance\templaenge by -2\wd1%
81: $$#1\hbox to0pt{\kern.5\templaenge$\Box$\kern-.5pt\hss}$$\par\bigbreak}
82:
83:
84: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
85: % Buchstabenk"urzel %
86: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
87:
88: %%% griechische Buchstaben
89: \def\al{{\alpha}}
90: \def\be{{\beta}}
91: \def\de{{\delta}}
92: \def\De{{\Delta}}
93: \def\om{{\omega}}
94: \def\Om{{\Omega}}
95: \def\la{{\lambda}}
96: \def\ka{{\kappa}}
97: \def\io{{\iota}}
98: \def\si{{\sigma}}
99: \def\Si{{\Sigma}}
100: \def\ga{{\gamma}}
101: \def\ep{{\varepsilon}}
102: \def\Ga{{\Gamma}}
103: \def\Tau{\mathcal{T}}
104: \def\th{{\vartheta}}
105: \def\Th{{\Theta}}
106: \def\ph{{\varphi}}
107: \def\phi{{\varphi}}
108: \def\Ph{{\Phi}}
109: \def\rh{{\rho}}
110: \def\ze{{\zeta}}
111:
112: %%% beinahe griechische Buchstaben
113: \def\na{{\nabla}}
114: \def\pa{{\partial}}
115: \def\el{{\ell}}
116:
117: %%% Kalligraphische Buchstaben
118: \def\cB{\mathcal{B}}
119: \def\cC{\mathcal{C}}
120: \def\cH{\mathcal{H}}
121: \def\cK{\mathcal{K}}
122: \def\cL{\mathcal{L}}
123: \def\cM{\mathcal{M}}
124: \def\cN{\mathcal{N}}
125: \def\cQ{\mathcal{Q}}
126: \def\cR{\mathcal{R}}
127: \def\cS{\mathcal{S}}
128: % \def\cT{\mathcal{T}} %Nicht benutzen: \Tau
129: \def\cW{\mathcal{W}}
130: \def\cV{\mathcal{V}}
131: \def\cZ{\mathcal{Z}}
132:
133: %%% Zeichen und Symbole
134: \def\ohne{\smallsetminus}
135: \def\ti{\tilde}
136: \def\witi{\widetilde}
137: \def\wihat{\widehat}
138: \def\ol{\overline}
139: \def\gdw{\Longleftrightarrow}
140: \def\embed{\hookrightarrow}
141: \def\bs{\backslash}
142: \def\lan{\langle}
143: \def\ran{\rangle}
144: \def\LA{\Leftarrow}
145: \def\RA{\Rightarrow}
146: \def\LRA{\Leftrightarrow}
147: \def\trl{$\Leftrightarrow$}
148: \def\entsp{{\widehat =}}
149: \def\schlange{{\widetilde{\hspace{0.5cm}}}}
150: \def\Dflat{D^{\rm eucl}}
151:
152: \def\torus{T^2}
153: \def\gflat{g}
154: %\def\Re{{\mathop{\rm Re\;}}}
155: %\def\Im{{\mathop{\rm Im\;}}}
156: \def\Ree{{\mathop{\rm Re\;}}}
157: \def\Imm{{\mathop{\rm Im\;}}}
158:
159:
160: %%%% \mR-Varianten
161: \def\Rplus{\mR^+}
162: \def\Roplus{\mR^+_0}
163: \def\Rminus{\mR^-}
164: \def\Rominus{\mR^-_0}
165:
166: %Text-Makros
167: \def\ie{i.\thinspace e.\ \ignorespaces}
168: \def\Wlog{W.\thinspace l.\ \ignorespaces o.\thinspace g.\ \ignorespaces}
169: \def\nummerarray#1#2{\par\noindent\setbox0\hbox{\rm (#1)}\setbox1\hbox{$#2$}\unhcopy0%
170: \dimen0=.5\textwidth \advance\dimen0 by -\wd0 \advance\dimen0 by -.5\wd1 \kern\dimen0 \unhcopy1}
171:
172:
173: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
174: % Grosse Ausdr"ucke %
175: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
176:
177: %%% Mathematical operators
178: \def\ker{\mathop{{\rm ker}}}
179: \def\image{\mathop{{\rm image}}}
180: \def\grad{{\mathop{{\rm grad}}}}
181: \def\Span{\mathop{{\rm span}}}
182: \def\spec{\mathop{{\rm spec}}}
183: \def\supp{\mathop{{\rm supp}}}
184: \def\dim{\mathop{{\rm dim}}}
185: \def\dist{\mathop{{\rm dist}}}
186: \def\vol{{\mathop{{\rm vol}}}}
187: \def\dvol{{\mathop{{\rm dvol}}}}
188: \def\area{{\mathop{{\rm area}}}}
189: \def\diam{{\mathop{{\rm diam}}}}
190: \def\injrad{{\mathop{{\rm injrad}}}}
191: \def\min{\mathop{{\rm min}}}
192: \def\max{\mathop{{\rm max}}}
193: \def\osc{\mathop{{\rm osc}}}
194: \def\mod{\mathop{{\rm mod}}}
195: \def\can{\mathop{{\rm can}}}
196: \def\Cl{\mathop{{\mC \rm l}}}
197: \def\ClR{\mathop{{\rm Cl}_{\mR}}}
198: \def\id{\mathop{{\rm id}}}
199: \def\Id{\mathop{{\rm Id}}}
200: \def\End{{\mathop{{\rm End}}}}
201: \def\GL{{\mathop{{\rm GL}}}}
202: \def\SO{\mathop{{\rm SO}}}
203: \def\U{{\rm U}}
204: \def\SU{{\rm SU}}
205: \def\Sp{{\rm Sp}}
206: \def\Spin{\mathop{{\rm Spin}}}
207: \def\CSO{{\mathop{{\rm CSO}}}}
208: \def\CSpin{{\mathop{{\rm CSpin}}}}
209: \def\PCSO{{P_{\rm CSO}}}
210: \def\PCSpin{{P_{\rm CSpin}}}
211: \def\PSpin{P_{\mbox{\rm \scriptsize Spin}}}
212: \let\psp\PSpin
213: \def\PSO{{P_{\mbox{\rm \scriptsize SO}}}}
214: \let\pso\PSO
215: \def\pson{{P_{\mbox{\rm \scriptsize SO(n)}}}}
216: \def\Hom{{\mathop{{\rm Hom}}}}
217: \def\scal{{\mathop{{\rm scal}}}}
218: \def\Ric{{\mathop{{\rm Ric}}}}
219: \def\im{{\mathop{{\rm im}}}}
220: \def\imD{{\im_{C^\infty} D_{g_0}}}
221: \def\imDg{{\im_{C^\infty} D_g}}
222: \def\Isom{{\mathop{{\rm Isom}}}}
223:
224: %%% Mathematische Ausdr"ucke
225: \def\gnull{g}
226: \def\gf{g_f}
227: \def\Df{D_f}
228: \def\Dnull{D_1}
229: \def\Laf{\De_f}
230: \def\dvolf{\dvol_f}
231: \def\const{{{\mathop{\it const}}}}
232:
233: \def\geucl{{g_{\mathop{\rm eucl}}}}
234: \def\gcan{{g_{\mathop{\rm can}}}}
235: \def\gstand{g_0}
236: \def\surfnc{{M_1}}
237:
238: \def\platz{{\,\mathord{.}\,}}
239: \def\leersp{\lan\mathord{.}\mathord{,}\mathord{.}\ran}
240: \def\leerSP{(\mathord{.}\mathord{,}\mathord{.})}
241: %\def\CP{\mC{\mathrm P}}
242: %\def\RP{\mR{\mathrm P}}
243: \def\res#1#2{{#1}\lower .11ex\hbox{$|$}\lower .644ex\hbox{$\scriptstyle #2$}}
244: \def\stelle#1#2{\left. {#1}\right|_{#2}}
245: %\def\length#1#2{{\mbox{\rm length}_{#2}({#1})}}
246: %\def\Length#1#2{{\mbox{\rm length}_{#2}\Bigl({#1}\Bigr)}}
247: \let\mo=\mathopen
248: \let\mc=\mathclose
249: \def\GawtM#1#2{{{\ti\Ga}_{#1}^{#2}}}
250: \def\GaN#1#2{{{\Ga}_{#1}^{#2}}}
251: \def\nov{\mathbf{n}}
252: \def\zweiff{{\rm I\!I}}
253: \def\nob{{\mbox{\rmfamily\bfseries N}}}
254: \def\mkv{\mathbf{H}}
255: \def\tr{{\mathop{\rm tr}}}
256: \def\image{{\mathop{\rm image}}}
257: \def\lammin{{\lambda_{\rm min}^+}}
258:
259: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
260: % Environments und Strukturierelemente %
261: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
262:
263: \long\def\komment#1{}
264: %\long\def\longver#1{#1}
265: \long\def\longver#1{}
266:
267: \def\proof#1{{\par\medbreak\noindent {\bf Proof\setbox0\hbox{#1}%
268: \ifdim\wd0=0pt .\else\ \ignorespaces #1.\fi}\enspace}}
269: \def\iop#1{{\par\medbreak\noindent {\bf Idea of proof\setbox0\hbox{#1}%
270: \ifdim\wd0=0pt .\else\ \ignorespaces #1.\fi}\enspace}}
271: %\def\example{{\noindent {\bf Example. }}}
272: \def\examples{{\noindent {\bf Examples. }\par\kern-\baselineskip}}
273: %\def\exampleno#1{{\noindent {\bf Example {#1}. }}}
274:
275:
276: \newtheoremstyle{remarks}{3pt}{3pt}{}{}{\bfseries}{}{ }{}
277: \newtheoremstyle{appendix}{3pt}{3pt}{}{}{\bfseries}{}{ }{}
278:
279: \makeatletter
280: \@addtoreset{equation}{section}
281: \renewcommand{\theequation}{\the\c@section.\the\c@equation}
282: \makeatother
283: %\newtheorem{theorem}{\sc Theorem}[section]
284: \newtheorem{theorem}[equation]{\bf T{\footnotesize HEOREM}}
285: \newtheorem{proposition}[equation]{\bf P{\footnotesize ROPOSITION}}
286: \newtheorem{lemma}[equation]{\bf L{\footnotesize EMMA}}
287: \newtheorem{corollary}[equation]{\bf C{\footnotesize OROLLARY}}
288: \newtheorem{conjecture}[equation]{\bf C{\footnotesize ONJECTURE}}
289: \newtheorem*{cmcprinciple}{\bf Principle for construction of cmc-surfaces}
290: \newtheorem*{maintheorem}{\bf M{\footnotesize AIN} T{\footnotesize HEOREM}}
291: \newtheorem{openproblem}{\bf O{\footnotesize PEN} P{\footnotesize ROBLEM}}
292: \newtheorem{theorema}{\bf T{\footnotesize HEOREM}}[section]
293: \newtheorem{lemmaa}[theorema]{\bf L{\footnotesize EMMA}}
294: \newtheorem{propositiona}[theorema]{\bf P{\footnotesize ROPOSITION}}
295: \theoremstyle{definition}
296: \newtheorem*{remark}{Remark}
297: \newtheorem{definitionnr}[equation]{Definition}
298: \newtheorem*{definition}{Definition}
299: \newtheorem*{example}{Example}
300:
301: \theoremstyle{remarks}
302: %\newtheorem*{remarks}{Remarks}
303:
304: \hyphenation{know-ledge Ha-bi-li-ta-tions-schrift}
305:
306: \def\eref#1{{\rm (\ref{#1})}}
307:
308: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
309: % Extra-Definitionen fuer comin.tex %
310: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
311:
312: \def\pD{{p_D}}
313: \def\pY{{p_Y}}
314: \def\qD{{q_D}}
315: \def\Dir{D}
316: \def\cD{{\cal D}}
317: \def\cF{{\cal F}}
318: \let\<\langle
319: \let\>\rangle
320: \let\La\Lambda
321: \def\inj{\mathop{\rm inj}}
322: \def\Conf{{\rm Conf}}
323: \def\extconf{\ol{[g_0]}}
324: \def\extco#1{\ol{{\cM}(#1)}}
325: \def\RHS{{\rm RHS}}
326:
327:
328:
329: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
330: % %
331: % T h e D o c u m e n t b e g i n s h e r e ! %
332: % %
333: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
334:
335:
336: \begin{document}
337:
338: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
339: \title{The smallest Dirac eigenvalue in a spin-conformal class
340: and cmc-immersions}
341: \author{Bernd Ammann%
342: %\thanks{Partially supported
343: %by The European Contract Human Potential Programme,
344: %Research Training Networks HPRN-CT-2000-00101 and HPRN-CT-1999-00118}
345: }
346: \date{September 2005}
347: %\date{\today}
348: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
349: \maketitle
350:
351: %Kommentar
352: \begin{abstract}
353: Let us fix a conformal class $[g_0]$ and a spin structure $\si$ on
354: a compact manifold $M$. For any $g\in [g_0]$, let $\la^+_1(g)$
355: be the smallest positive eigenvalue of the Dirac operator $D$ on $(M,g,\si)$.
356: In a previous article we have shown that
357: $$\lammin(M,g_0,\si):=\inf_{g\in [g_0]} \la_1^+(g)\vol(M,g)^{1/n}>0.$$
358: In the present article, we enlarge the conformal class by adding
359: certain singular metrics. We will show that
360: if $\lammin(M,g_0,\si)<\lammin(S^n)$,
361: then the infimum is attained on the enlarged conformal class.
362: For proving this, we solve a system of
363: semi-linear partial differential equations involving a nonlinearity
364: with critical exponent:
365: $$D\phi= \la |\phi|^{2/(n-1)}\phi.$$
366: The solution of this problem has many analogies
367: to the solution of the Yamabe problem.
368: However, our reasoning is more involved than in the Yamabe problem
369: as the eigenvalues of the Dirac operator tend to $+\infty$ and $-\infty$.
370:
371: Using the spinorial Weierstra\ss{} representation,
372: the solution of this equation in dimension 2 shows the existence
373: of many periodic constant mean curvature surfaces.
374:
375: %{\bf Preliminary version, please do not distribute!}
376: \end{abstract}
377: {\bf Keywords:} Dirac operator, eigenvalues, conformal geometry,
378: critical Sobolev exponents
379:
380: {\bf Mathematics Classification:} 58J50, 53C27 (Primary),
381: 58C40, 35P15, 35P30, 35B33 (Secondary)
382:
383:
384: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
385: \section{Introduction}
386: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
387:
388: %The problem that we want to study in this article is originally motivated
389: %from the spectral theory of Dirac operators. Because of the conformally
390: %invariant character of the functional and the associated Euler-Lagrange
391: %equations, blowup phenomena similar to those in the Yamabe problem arise.
392: %If the underlying manifold has dimension $2$,
393: %the solutions to the Euler-Lagrange equations
394: %can locally be translated into conformal constant mean curvature immersions
395: %into $\mR^3$. This translation can be used to construct periodic
396: %constant mean curvature surfaces.
397:
398: %In this introduction, we will give a brief
399: %overview over the history, explain the relations to the
400: %Yamabe problem, and sketch the implied techniques; then we
401: %will describe the application to constant mean curvature surfaces.
402:
403: Let $M$ be a compact $n$-dimensional
404: manifold, $n\geq 2$, with a fixed conformal class $[g_0]$ and a fixed
405: spin structure $\si$.
406: Let $g$ be conformal to $g_0$, i.e.\ $ g\in [g_0]$.
407: The classical Dirac operator $D_g$
408: on $(M,g,\si)$ has discrete real spectrum
409: with finite multiplicities. The eigenvalues tend to $+\infty$ and $-\infty$.
410: The dimension of the kernel of $D_g$ is a spin-conformal invariant, i.e.
411: $\dim \ker D_g=\dim \ker D_{g_0}$ for all $g\in [g_0]$.
412: We denote the first (=smallest) positive eigenvalue of $D_g$
413: by $\la_1^+(g)$,
414: and the first (=largest) negative eigenvalue by $\la_1^-(g)$.
415:
416: Finding bounds for this eigenvalue has attracted much interest
417: during the last decades. Among several estimates
418: \cite{lichnerowicz:63,friedrich:80,kirchberg:86,kirchberg:88,
419: kramer.semmelmann.weingart:98,kramer.semmelmann.weingart:99}
420: in terms of a positive scalar curvature bound,
421: let us mention the following estimate due to Friedrich~\cite{friedrich:80}.
422: If the minimum of the scalar curvature
423: of $(M,g)$ is at least $s>0$, then
424: $\left(\la_1^\pm(g)\right)^2\geq {n\over 4(n-1)}s$.
425:
426: An improvement of this inequality that is important in conformal
427: geometry was derived by Hijazi \cite{hijazi:86}
428: in terms of the first eigenvalue
429: $\la_1(L_g)$
430: of the conformal Laplacian $L_g=4\,{n-1\over n-2}+\scal_g$,
431: namely
432: \begin{equation}\label{ineq.hij}
433: \left(\la_1^\pm(g)\right)\geq {n\over 4(n-1)} \la_1(L_g)
434: \end{equation}
435: if $n\geq 3$. On the other hand, if the
436: \emph{Yamabe invariant}
437: $$Y(M,[g_0])=\inf_{g\in [g_0]}{\int_M \scal_g\dvol_g\over
438: \vol(M,g)^{(n-2)/n}}\in (-\infty, n(n-1)]$$
439: is non-negative, one can use the conformal
440: transformation formula
441: for the scalar curvature \cite[Theorem~1.159]{besse:87}, and one obtains
442: that
443: $$ \la_1(L_g)\vol(M,g)^{2/n}\geq Y(M,[g_0]).$$
444: Together with \eref{ineq.hij}, we get
445: \begin{equation}\label{ineq.hij.der}
446: \left|\la_1^\pm(g)\right|\vol(M,g)^{1/n} \geq
447: \sqrt{{n\over 4(n-1)}\, Y(M,[g_0])}.
448: \end{equation}
449: That this inequality even holds in
450: dimension $n=2$, was proved by C.~B\"ar \cite{baer:92b}. In this special case,
451: the Gauss-Bonnet theorem tells us that $Y(M)= 4\pi \chi(M)$, hence we obtain
452: \begin{equation}\label{ineq.baer}
453: \left|\la_1^\pm(g)\right|\area(M,g)^{1/2}\geq 2\sqrt{\pi}
454: \end{equation}
455: if $M$ is diffeomorphic to $S^2$, but we do not get a bound for other surfaces.
456: Equality in \eref{ineq.hij.der} and \eref{ineq.baer} hold for the round
457: spheres.
458:
459:
460: Hence, we have obtained an explicit positive
461: lower bound $\left|\la_1^\pm(g)\right|\vol(M,g)^{1/n}$ that is uniform
462: on the conformal class $[g_0]$.
463: From now on, we will restrict to the first \emph{positive} eigenvalue,
464: in order to simplify the presentation (see the remark below).
465:
466:
467: The \emph{existence} of a positive lower
468: bound for $\la_1^+(g)\vol(M,g)^{1/n}$
469: had already been derived before by J. Lott. His estimate
470: does not require that $Y(M,[g_0])$ is positive, but needs the weaker assumption
471: that the Dirac operator is invertible, i.e.\ has $0$-dimensional kernel.
472: Unfortunately, his bound is not explicit, and for most spin-conformal
473: manifold the determination of the
474: value of the associated infimum
475: \begin{equation}\label{def.lammin}
476: \lammin(M,[g_0],\si)
477: :=\inf_{g\in [g_0]} \la_1^+(g)\vol(M,g)^{1/n}
478: \end{equation}
479: is still a challenging open problem.
480:
481: A key idea of J. Lott's article is to derive a lower
482: bound for the first Dirac eigenvalue in terms of the
483: supremum of a conformally invariant functional, a version of this
484: functional will be explained in Section~\ref{sec.varprin}.
485:
486: In our article \cite{ammann:03}, we started to study this
487: functional in more detail. In particular, we showed that Lott's result
488: extends to the case that the Dirac operator has non-trivial kernel,
489: namely
490: $$\lammin(M,[g_0],\si)>0$$
491: for any compact Riemannian spin manifold $(M,[g_0],\si)$.
492:
493: Furthermore, it was shown that
494: $$\lammin(M,[g_0],\si) \leq \lammin(\mS^n),$$
495: where $\mS^n=(S^n,g_{\rm can})$ denotes the sphere with its standard
496: Riemannian metric of constant sectional curvature $1$.
497: This bound has been proven in \cite{ammann:03} unless
498: $\ker D\not=\{0\}$ and $n=2$. The remaining case $\ker D\not=\{0\}$ and $n=2$
499: was given in \cite{grosjean.humbert:p05}.
500:
501: In the present article, we discuss whether the associated infimum in
502: \eref{def.lammin} is attained. For having a well-behaved
503: minimization problem, it is reasonable to replace the conformal class
504: $[g_0]$ in
505: \eref{def.lammin} by its $L^\infty$-completion $\overline{[g_0]}$.
506: Here, by definition a metric $f^2g_0$ is in $\overline{[g_0]}$ if
507: $f$ is a real-valued $L^\infty$ function. The first positive
508: eigenvalue of the Dirac operator extends naturally to this completion
509: (see Section~\ref{sec.spd}).
510: We show that if we have the strict inequality
511: \begin{equation}\label{eq.strict}
512: \lammin(M,g_0,\si)< \lammin(\mS^n)= {n\over 2}\,\om_n^{1/n},
513: \end{equation}
514: then the infimum is obtained by a generalized metric $g\in \overline{[g_0]}$.
515: This minimizer has the form $g:=|\phi|^{4/(n-1)}g_0$ where
516: $\phi$ is a spinor of regularity $C^2$. The set
517: $\phi^{-1}(0):=\{x\in M\,|\,\phi(x)=0\}$ is
518: called the \emph{nodal set of $\phi$} or the
519: \emph{set of degeneration of $g$}.
520:
521: \begin{theorem}\label{theo.main}
522: Let $M$ be a compact manifold of dimension $n\geq 2$
523: with a fixed conformal
524: class $[g_0]$ and a spin structure $\si$.
525: Assume that \eref{eq.strict}
526: holds. Let $\al:= 2/(n-1)$ if $n\geq 4$, and let $\al\in (0,1)$ if $n\in\{2,3\}$.
527: \begin{enumerate}[\rm (A)]
528: \item Then there is a spinor field
529: $\phi\in C^{2,\al}(\Si M)\cap C^\infty(\Si(M\ohne \phi^{-1}(0)))$
530: on $(M,g_0)$ such that
531: \begin{equation}\label{eq.dirac.nonlin}
532: D_{g_0}\phi = \lammin\, |\phi|^{2/(n-1)}\phi,\qquad \|\phi\|_{2n/(n-1)}=1.
533: \end{equation}
534: \item There is a $g\in \extconf$ such that
535: $$\la^+_1(g)\vol(M,g)^{1/n}=\lammin.$$
536: The metric has the form $g=|\phi|^{4/(n-1)}g_0$ where
537: $\phi$ is a spinor as in (A).
538: \item If $\dim M=2$, then the metric $g$ is smooth
539: and the set of degeneration of $g$, denoted $\cS_g$,
540: is finite.
541: Furthermore
542: $$\#\cS_g< {\rm genus}(M).$$
543: In particular, if $M$ is diffeomorphic to a
544: $2$-torus, then the set of degeneration $\cS_g$ is empty.
545: \end{enumerate}
546: \end{theorem}
547:
548: Inequality \eref{eq.strict} has been proven for several
549: classes of manifolds.
550: It is known that it holds for non-conformally-flat manifolds of dimension
551: $\geq 7$ \cite{ammann.humbert.morel:p03av2}.
552:
553: Inequality~\eref{eq.strict} has also been shown if
554: $M$ is conformally flat, if $D$ is invertible,
555: and if the mass endomorphism is not identically zero
556: (after a possible change of orientation if $\dim M \equiv 3 \mod 4$)
557: \cite{ammann.humbert.morel:p03b}.
558: The mass endomorphism is a section of $\End(\Si M)$
559: defined as the zero order term of the development of the Green function
560: for the Dirac operator at the diagonal with respect to a
561: conformal coordinate map.
562:
563: Furthermore, \eref{eq.strict} is known for many Riemann surfaces (i.e. $n=2$),
564: e.g.\ all rectangular tori have a spin structure such that
565: \eref{eq.strict} holds.
566:
567: The Euler-Lagrange equation~\eref{eq.dirac.nonlin} of the above minimization
568: problem has a particularly nice interpretation in dimension~$n=2$.
569: Locally the equation~\eref{eq.dirac.nonlin}
570: can be translated into a conformal constant mean curvature
571: immersion into $\mR^3$.
572: This translation
573: is a spinorial extension of the Weierstrass representation (see
574: Section~\ref{sec.weierrep}). By pasting together these local surfaces
575: one obtains a ``periodic branched conformal cmc immersion''.
576:
577: More exactly, let $(M,g)$ be a compact Riemann surface together
578: with its universal covering $\pi:\witi M\to M$. A
579: \emph{periodic branched conformal cmc immersion based on $(M,g)$}
580: is by definition a smooth map $F:\witi M\to \mR^3$ together with
581: finitely many points $p_1,\ldots,p_k\in M$, the so-called
582: \emph{branching points}, such
583: that the following properties hold:
584: \begin{enumerate}[{\rm (1)}]
585: \item {\it Periodicity:} There is a homomorphism
586: $h:\pi(M)\to \mR^3$, the \emph{periodicity map}, such that
587: for any $\gamma\in \pi_1(M)$, and $x\in\witi M$ one has
588: $$F(x\cdot \gamma)=F(x)+h(\gamma).$$
589: Here $\cdot$ denotes the action of $\pi_1$ on $\witi M$ via Deck transformation.
590: \item {\it Conformality:} The restriction of $F$ to $\witi M\setminus \pi^{-1}(\{p_1,\ldots,p_k\})$
591: is a conformal immersion.
592: \item {\it Branching points:} We have
593: $dF_q=0$ for any $q\in \pi^{-1}(\{p_1,\ldots,p_k\})$.
594: The order of the first non-vanishing term in the
595: Taylor development of $dF$ in $q$ is called the
596: \emph{branching index} of $F$ at $q$.
597: \item {\it CMC:} The image $F\left(\witi M\setminus \pi^{-1}(\{p_1,\ldots,p_k\}\right)$
598: is an immersed surface
599: with constant mean curvature.
600: \end{enumerate}
601:
602: The principle yields the existence of many periodic branched conformal
603: cmc immersions. The set of periodic branched conformal
604: cmc immersions, $H\neq 0$, is
605: essentially (see Section~\ref{sec.app.cmc} for a precise
606: statement)
607: in bijection with the stationary points of the variational problem
608: associated to our minimization problem.
609: In particular, all minimizers of \eref{def.lammin} give rise
610: to periodic branched conformal
611: cmc immersions. We obtain
612:
613: \begin{cmcprinciple}
614: Assume that the Riemann spin surface $(M,g,\si)$ carries a metric
615: $g$ such that the first positive eigenvalue of the Dirac operator is smaller
616: than $2\sqrt{\pi/\area(M,g)}$. Then there is a periodic branched
617: conformal cmc immersion $F$ based on $(M,g)$.
618: The regular homotopy class of $F$ is determined by the spin-structure $\si$.
619: The indices of all branching points are even,
620: and the sum of these indices
621: is smaller than $2 {\rm genus}(M)$. In particular,
622: if $M$ is a torus, there are no branching points.
623: \end{cmcprinciple}
624:
625: Some examples of branched
626: conformal cmc immersions that may arise by this principle
627: are given in Section~\ref{sec.app.cmc}.
628:
629: The problem that we are discussing in this article has many relations to the
630: Yamabe problem. For a given compact conformal manifold $(M,[g_0])$
631: the Yamabe problem is the problem to find a metric of constant scalar curvature
632: $g$ in $[g_0]$. The problem has been affirmatively solved by Trudinger,
633: Aubin, Schoen and Yau, see \cite{lee.parker:87} for a good overview.
634: At first, due to the
635: conformally invariant character of our problem and the Yamabe problem,
636: certain bounded, but non-compact Sobolev embeddings play an important role.
637: In the solution of both problems it is useful to break the conformal
638: invariance by perturbing a parameter. For this perturbed parameter
639: the embeddings are compact, and standard methods yield the existence
640: of a minimizer. It then has to be checked whether the perturbed minimizers
641: converge to a minimizer of the unperturbed problem, or whether they concentrate
642: in some points. In the Yamabe problem
643: the minimizers converge if
644: \begin{equation}\label{eq.yamab}
645: Y(M,[g_0])<Y(\mS^n),
646: \end{equation}
647: holds whereas in our problem, the minimizers converge if
648: \eref{eq.strict} holds.
649: Secondly, \eref{ineq.hij.der} implies that any manifold satisfying
650: \eref{eq.strict} satisfies \eref{eq.yamab} as well. Hence, proving
651: \eref{eq.strict} for a given spin-conformal manifold $(M,[g_0],\si)$
652: solves the Yamabe problem on this manifold.
653: The third relationship is that some proofs of inequality \eref{eq.strict}
654: in special cases resemble to proofs of inequality,
655: e.g.\ \cite{ammann.humbert.morel:p03av2} resembles to \cite{aubin:76}.
656:
657: The structure of the article is as follows.
658: In Section~\ref{sec.varprin} we reformulate our problem
659: as a variational problem. We will see that it is natural to admit
660: in the infimum \eref{def.lammin} certain singular metrics namely ``metrics''
661: conformal to $g_0$ whose conformal factor might vanish
662: somewhere. Such metrics --- called \emph{generalized metrics} ---
663: are the subject of Section~\ref{sec.spd}. In the following section,
664: we discuss the round sphere. This example is helpful to obtain a deeper
665: understanding for the analytical difficulties. However, it can be skipped
666: if the reader is only interested in the main results of the article.
667: Section~\ref{sec.reg.theo} is devoted to certain regularity issues that will
668: become important in Section~\ref{sec.solution}, where
669: the variational principle is finally solved. We then show in
670: Section~\ref{sec.proofmain} how this implies the main theorem.
671: The singularities of the minimizers are discussed in Section~\ref{sec.degen}.
672: In Section~\ref{sec.weierrep} we recall the spinorial
673: Weierstrass representation. Section~\ref{sec.app.cmc} uses the spinorial
674: Weierstrass representation to derive the application to constant mean
675: curvature surfaces. Several examples are included.
676: In Appendix~\ref{sec.regularity} we summarize (without proofs)
677: some analytical tools.
678: This appendix shall also serve as a reference for fixing the notations for
679: Sobolev spaces and H\"older spaces. In Appendix~\ref{sec.scha.power}
680: we proof a proposition about H\"older spaces, as we could not find a
681: proof in the literature.
682:
683:
684: \begin{remark}[The first negative eigenvalue]
685: The infimum is also attained in \eref{def.lammin}
686: if we replace $\la_1^+$ by $|\la_1^-|$.
687: In the case $n\not\equiv 3\mod 4$ this is obvious:
688: there exists an automorphism
689: of the spinor bundle anticommuting with $D$,
690: hence $\la_1^-(g)=-\la_1^+(g)$.
691: In the case $n\equiv 3\mod 4$, the proof for $\la_1^-$
692: is up to the obvious sign changes completely identical.
693: In almost all references cited in the introduction,
694: all statements for $\la_1^+$ also hold for $|\la_1^-|$ and vice versa.
695: The only exception is \cite{ammann.humbert.morel:p03b}.
696: \end{remark}
697:
698: {\bf Acknowledgements.}
699: I am very much indebted to Christian B\"ar for various support.
700: Thank you also to Emmanuel Humbert for his continuing interest in conformal
701: spin geometry and his deep analytical ideas. Many electronic and personal
702: discussions with Robert Kusner and Karsten Grosse-Brauckmann were also very helpful. I also want to thank B.~Booss-Bavnbek, Oussama Hijazi,
703: Sergiu Moroianu,Victor Nistor, Reiner Sch\"atzle and Guofang Wang
704: for several stimulating discussions related to this article.
705: %, and thanks to
706: %B.~Booss-Bavnbek for pointing us to \cite{booss.marcolli.wang:p02}.
707:
708:
709: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
710: \section{The associated variational principle}\label{sec.varprin}
711: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
712:
713: The goal of this section is to reformulate the problem of minimizing
714: the first Dirac eigenvalue in a conformal class as a variational problem.
715: The choice of a good functional is not very easy, as
716: we would like to find a functional which is both bounded and
717: conformally invariant.
718: %This is more involved than in the solution of the Yamabe problem.
719: The Dirac operator $D$ has a simple behavior under
720: conformal change of the metric, its square $D^2$ transforms in a more
721: complicated way. Hence, it is desirable to use a functional that
722: contains only terms in $D\phi$ and $\phi$ and no term depending
723: on $D^2\phi$. The $\cF_q$-functional defined in \eref{def.F}, $q=2n/(n+1)$,
724: is such a conformally invariant, bounded functional, and it will
725: be turn out that by working with this functional we obtain
726: the desired results.
727:
728: \begin{remark} In case, that the reader of our article
729: is already familiar with the analytical problems of the Yamabe problem,
730: he might find it enlightening to
731: compare this problem to our problem.
732: Many techniques from the resolution of the Yamabe problem
733: can be carried over to our setting. However, several arguments from
734: the resolution of the Yamabe problem fail. Two main problems arise:
735: on the one hand the spectrum of $D$ is neither bounded from
736: below nor from above, on the other hand, we are working with sections
737: of a vector bundle instead of functions, hence the standard maximum principle
738: is not available.
739: Such arguments will have to be replaced by other approaches.
740: \end{remark}
741:
742: %In contrast to the Yamabe operator which is bounded from below,
743: %the Dirac operator is neither bounded from above nor from below.
744: %This is why spectral geometers typically work
745: %with the square $\Dir^2$ of the Dirac operator instead of the Dirac operator
746: %itself.
747: %In our problem, working with $\Dir^2$
748: %is not a good approach, as the conformal
749: %transformation formula for $\Dir^2$ is more complicated than for $\Dir$.
750: %It is difficult to construct a conformally invariant
751: %functional in the style of the Yamabe functional
752: %which contains $\Dir^2$, but not $\Dir$.
753: %A functional that suits to our needs is $\cF_q$ defined in \eref{def.F}.
754:
755: %We will see that the definition of the first positive Dirac eigenvalue
756: %extends naturally to a certain completion of the conformal class. This
757: %completion will provide the natural set in which one should search
758: %for minimizers of $\la^+_1(g)$.
759:
760: At first, we recall some basic definitions and facts from spin geometry.
761: For details we refer to textbooks as for example
762: \cite{lawson.michelsohn:89, roe:88,friedrich:00}
763: or to the beautifully written
764: self-contained introduction \cite{hijazi:01}.
765:
766: Let $M$ be a compact manifold equipped with a Riemannian metric
767: $g_0$ and a spin structure $\si$. Assume that $g$
768: is a metric conformal to $g_0$. One associates to $(M,g,\si)$
769: a natural complex vector bundle over $M$ called the spinor bundle
770: $\Si(M,g,\si)\to M$.
771: Sections of this bundle are called \emph{spinor fields}
772: or simply \emph{spinors}. The bundle carries a hermitian metric, a metric connection,
773: and a Clifford multiplication. These additional structures are used to define
774: the Dirac operator $D_g:\Gamma(\Si(M,g,\si)\to\Gamma(\Si(M,g,\si)$, which
775: is a first order elliptic differential operator.
776: The Dirac operator $D_g$
777: is essentially self-adjoint, and hence it has a self-adjoint
778: extension. As a consequence of standard elliptic theory,
779: the spectrum is real and discrete, and all multiplicities
780: are finite.
781:
782: In the following, the spin structure $\si$ will be fixed (and often suppressed
783: in the notation), whereas the metric $g$ varies inside the conformal class
784: $[g_0]$. Some objects
785: will be noted with an index $g$, which means that they are
786: defined with respect to~$g$ whereas the same object without the index
787: $g$ indicates that it is defined with respect to the fixed background metric
788: $g_0$ (with some exceptions that are apparent from the context).
789: For example, $\dvol_g$ is the volume element associated to $g$
790: and $\dvol=\dvol_{g_0}$ is the one associated to $g_0$.
791: We will frequently use several norms, as for example
792: $\|\phi\|_{L^p}$ and $\|\phi\|_{H_1^q}$
793: that are defined with respect to $g_0$
794: unless otherwise stated. We summarize their definition and some important
795: analytical tools in Appendix~\ref{sec.regularity}.
796:
797: The spectrum of the Dirac operator $D_g$ will be denoted as
798: $$\ldots \leq \la_2^-(g)\leq \la_1^-(g)<0=\ldots=0
799: <\la_1^+(g)\leq\la_2^+(g)\leq \ldots,$$
800: where each eigenvalue appears with its multiplicity. Note that $0$
801: may be an eigenvalue or not. In the former case $D_g$ is not invertible,
802: otherwise it is. Elliptic theory shows
803: $\lim_{k\to\infty}\la_k^+(g) =-\lim_{k\to\infty}\la_{-k}^+(g)=\infty$.
804:
805: The following transformation formula will be of central importance.
806: To our knowledge, the formula was found by Hitchin \cite{hitchin:74}.
807: Another reference written up in a more self-contained manner
808: is \cite{hijazi:01}.
809: %are \cite{hijazi:86,hijazi:01}.
810:
811: \begin{proposition}[Conformal transformation formula for $D$]\label{prop.conf.change}
812: Let $g=f^2g_0$, $f:M\to \mR$ smooth and positive.
813: There is an isomorphism of vector bundles
814: ${F:\Si(M,g_0,\si)\to \Si(M,g,\si)}$
815: which is a fiberwise isometry such that
816: $$\Dir_g(F(\phi))=F\left(f^{-{n+1\over 2}}\Dir_{g_0} f^{{n-1\over 2}}\phi\right).$$
817: \end{proposition}
818:
819: It is convenient to define
820: $$\witi{F}(\phi)=F(f^{-{n-1\over 2}}\phi).$$
821: We will use this isomorphism $\witi{F}$
822: to identify spinors associated
823: to conformal metrics. With this identification the conformal transformation
824: formula reads as
825: $$\Dir_g(\phi)= f^{-1}\Dir_{g_0}(\phi),$$
826: and $|\phi|_g=f^{-{n-1\over 2}}|\phi|_{g_0}$.
827: In particular, with this identification the kernels of $D_g$ and
828: $\Dir_{g_0}$ coincide.
829: It is easy to verify that with this identification,
830: the expression $\int \<\Dir_g \phi,\phi\>_g\dvol_g$ is conformally invariant.
831: The $L^p$-norm $\|\phi\|_{L^p(g)}:=\left(\int |\phi|_g^p\dvol_g\right)^{1/p}$
832: is conformally invariant
833: if and only if $p$ has the value $\pD:=2n/(n-1)$.
834: Similarly, $\|\Dir_g\phi\|_{L^q(g)}$ is conformally invariant if and only if
835: $q$ has the value $\qD:=2n/(n+1)$. Note that $\qD^{-1}+\pD^{-1}=1$.
836:
837: For any $q\in [\qD,2]$ and for any $H_1^q$-spinor $\phi$
838: that is not in the kernel of the
839: Dirac operator we define
840: \begin{equation}\label{def.F}
841: \cF_q^g(\phi)= {\int\<\Dir_g\phi ,\phi\>_g\dvol_g\over \|\Dir_g\phi\|_{L^q(g)}^2}, \qquad
842: \mu_q^g :=\mu_q(M,g,\si):= \sup \cF_q^g(\psi),
843: \end{equation}
844: and $\cF_q:= \cF_q^{g_0}$, $\mu_q:=\mu_q^{g_0}$.
845: The well-definedness of $\cF_q^g$ and some basic properties are given by the
846: following lemma.
847:
848: \begin{lemma}\label{lem.f.elem}
849: Let $q \in [\qD,\infty)$. Let $\phi$ be a spinor field
850: of regularity $H_1^q$, $D\phi\neq 0$.
851: Then,
852: $\cF_q(\phi)$ is well-defined and real. Furthermore
853: $\cF_q:H_1^q\setminus \ker D \to\mR$
854: is Fr\'echet differentiable with derivation given by
855: \begin{equation}\label{eq.abl}
856: d\cF_q(\phi)(\psi) = {2\over \|\Dir\phi\|_{L^q}^2}
857: \,\int\< \phi -\rho_{q,\phi} |\Dir\phi|^{q-2}\Dir\phi,\Dir\psi\>,
858: \end{equation}
859: where $\rho_{q,\phi}=\cF_q(\phi) \|\Dir\phi\|_{L^q}^{2-q}$.
860: The supremum $\mu_q$ is positive and finite.
861: \end{lemma}
862:
863: From the above considerations it is evident that the
864: functional $\cF_q$ is conformally invariant if and only if $q=\qD$.
865: %This functional will take the role which is played by the Yamabe functional
866: %in the Yamabe problem.
867:
868: \proof{}
869: Let $\phi$ be a spinor field of regularity $H_1^q$, $D\phi\neq 0$.
870: Take $p$ with $p^{-1}+q^{-1}=1$, $q\ge {2n\over n+1}$.
871: Because of $\phi\in H_1^q\embed L^p$, we see with H\"older's inequality
872: that $\<D\phi,\phi\>$ is integrable. Thus, the numerator of $\cF_q$ is
873: well-defined, and hence $\cF_q$ is well-defined. The self-adjointness
874: of $\Dir$ implies that $\cF_q(\phi)$ is real.
875: Moreover, because $H_1^q\embed L^p$ is bounded we see that
876: $|\int_M \<D\phi,\phi\>|\leq \|D\phi\|_{L^q}\,\|\phi\|_{L^p}$
877: is bounded from above by a multiple of $\|\phi\|_{H_1^q}^2$.
878: Using Theorem~\ref{theo.global.lp} we obtain
879: $\|\phi\|_{H_1^q}\leq \|D\phi\|_{L^q}+\|\phi\|_{L^q}$,
880: and we see that $\cF_q$ is bounded on
881: $H_1^q\cap (\ker D)^\perp\setminus \{0\}$. An arbitrary spinor field
882: $\phi\in H_1^q$ is written as the sum of $\phi_1\in \ker D$ and a non-zero
883: $\phi_2\perp\ker D$. As $\cF_q(\phi)=\cF_q(\phi_2)$, we see that
884: $\mu_q$ is finite.
885:
886: Because of $\phi\in H_1^q$, we have
887: $|\Dir \phi|^{q-2}\Dir\phi \in L^p$.
888: Hence the right hand side of \eref{eq.abl} defines a continuous
889: functional on $L^p$ and thus on $H_1^q$. We denote the functional
890: by $\psi\mapsto \RHS_\phi(\psi)$.
891: Similarly, one sees that
892: $$\cF_q(\phi+ \psi)-\cF_q(\phi)- \RHS_\phi(\psi)\leq o(\|\psi\|_{H_1^q}),$$
893: hence $\cF_q$ is Fr\'echet differential with derivative
894: $\psi\mapsto \RHS_\phi(\psi)$.
895:
896: If $\phi$ is an eigenspinor to a positive eigenvalue, then
897: $\mu_q\geq \cF_q(\phi)>0$.
898: \qed
899:
900:
901: \begin{proposition}[Properties of $\mu_q^g$]\label{prop.func}
902: The function $\mu_q^g:[\qD,\infty)\to (0,\infty)$ is continuous from the right,
903: and
904: $$\mu_2(M,g,\si)=(\la^+_1(g))^{-1}$$
905: Furthermore, if $\vol(M,g)=1$, then $\mu_q^g$ is non-increasing in $q$.
906: \end{proposition}
907:
908:
909: \proof{}
910: We assume $\vol(M,g)=1$, the statements for $\vol(M,g)\not=1$
911: then follow by rescaling.
912:
913: That $\mu_q^g$ is non-increasing follows easily from the H\"older inequality.
914:
915: In order to show
916: the continuity from the right, let $q\geq \qD$ be given.
917: We take a smooth spinor field $\phi$ such that
918: $\cF_q^g(\phi)\geq \mu_q^g-\ep$.
919: Observe that $$\cF_{q'}^g(\phi)= { \|\Dir_g\phi\|_{L^q(g)}^2 \over \|D_g\phi\|_{L^{q'}(g)}^2}\,\cF_{q}^g(\phi).$$
920: The function $q'\mapsto \|\Dir_g\phi\|_{L^{q'}(g)}$ is continuous, hence
921: if $q'\geq q$ is sufficiently close to~$q$, then
922: $$\mu_{q'}^g\geq \cF_{q'}^g(\psi)\geq \cF_q^g(\psi)-\ep \geq \mu_q^g -2 \ep.$$
923: Because $q\mapsto \mu_q^g$ is non-increasing, the continuity from the right follows.
924:
925: The formula
926: $\mu_2^g=(\la^+_1(g))^{-1}$ follows directly if one writes~$\phi$ as a sum of eigenspinors and evaluates
927: $\cF_2^g(\phi)$.
928: \qed
929:
930: \begin{figure}
931: %
932: % Dies ist die Datei graph.pic.
933: % Dieses Bild zeigt
934: % die Funktion q\mapsto \mu_q
935: %
936: %
937: %
938: \newdimen\axdim
939: \axdim=1pt
940: \newdimen\cudim
941: \cudim=2pt
942: \def\ticklen{.2}
943: \def\abst{.5}
944:
945: \begin{center}
946: \psset{unit=1cm}
947:
948: \begin{pspicture}(-1,-\ticklen)(10.5,6)
949: \psset{linewidth=\axdim}
950: \psaxes[linewidth=\axdim,labels=none,ticks=none]{->}(0,0)(9.5,5)
951: \rput(9.9,0){$q$}
952: \rput(0,5.4){$\mu_q$}
953: \psline(2.4,-\ticklen)(2.4,\ticklen)
954: \psline(3.5,-\ticklen)(3.5,\ticklen)
955: \rput[t](2.4,-\abst){$\qD$}
956: \rput[t](3.5,-\abst){$2$}
957: \psline(-\ticklen,2.8)(\ticklen,2.8)
958: \psline(-\ticklen,3.7)(\ticklen,3.7)
959: \rput[r](-\abst,2.8){$(\la^+_1)^{-1}$}
960: \rput[r](-\abst,3.7){$\mu_\qD$}
961: \psline[linestyle=dotted]{-}(2.4,0)(2.4,3.7)
962: \psline[linestyle=dotted]{-}(3.5,0)(3.5,2.8)
963: \psline[linestyle=dotted]{-}(0,3.7)(2.4,3.7)
964: \psline[linestyle=dotted]{-}(0,2.8)(3.5,2.8)
965: \psset{linewidth=\cudim}
966: \psecurve[showpoints=false](2.2,3.74)(2.4,3.7)(3.5,2.8)(4,2.0)(5.1,1.6)(7.2,1.4)(7.5,1.2)
967: \end{pspicture}
968: \end{center}
969: \caption{$\mu_q$ as a function of $q$}
970: \end{figure}
971:
972: \begin{remark}
973: In Proposition~\ref{prop.sol.subcrit} we will see that the supremum
974: defining $\mu_q^g$
975: is attained for $q>\qD$ by a $C^{2,\al}$-spinor. This implies
976: that the function
977: $$[\qD,\infty)\to (0,\infty],\quad q\mapsto \mu_q^g$$
978: is also continuous from the left.
979: \end{remark}
980:
981:
982: %\begin{corollary} The smallest positive eigenvalue
983: %of the Dirac operator is bounded
984: %from below by $\mu_{\qD}^{-1}$:
985: % $$\la^+_1\geq \mu_{\qD}^{-1}\qquad \lammin\geq \mu_{\qD}^{-1}.$$
986: %\end{corollary}
987:
988: \begin{proposition}\label{prop.lamu}
989: $$\mu_{\qD}(M,g_0,\si)={1\over\lammin(M,g_0,\si)}$$
990: \end{proposition}
991: \proof{}
992: We have already seen that $\mu_{\qD}$ is conformally invariant, \ie for $g_1\in [g_0]$
993: $$\mu_{\qD}(M,g_0,\si)=\mu_{\qD}(M,g_1,\si).$$
994: The previous proposition states that
995: $$\mu_{\qD}(M,g_1,\si)\geq\mu_2(M,g_1,\si)=\left(\la_1^+(g_1)\right)^{-1}$$
996: if $\vol(M,g_1)=1$, and it follows
997: $$\mu_{\qD}(M,g_0,\si)\geq {1\over \lammin(M,g_0,\si)}.$$
998:
999: It remains to show the inverse inequality which amounts to showing
1000: $$\sup_{g\in [g_0]} \la_1^+(g)^{-1}\vol(M,g)^{-1/n}\geq \sup \cF_\qD.$$
1001: For any $\ep>0$ we take a smooth spinor $\phi_\ep$ with
1002: $\cF_\qD(\phi_\ep)\geq \sup \cF_\qD-\ep$ and $\|\Dir\phi_\ep\|_{L^\qD}=1$.
1003: After a small perturbation of $\phi_\ep$ we can assume that
1004: $\Dir\phi_\ep$ has no zeros.
1005: We set
1006: $$g_\ep:=|\Dir\phi_\ep|^{4/(n+1)}g_0.$$
1007: Then we have $\vol(M,g_\ep)=1$, and $|\Dir\phi_\ep|_{g_\ep}$ has constant length $1$.
1008: Hence,
1009: $$\cF_\qD^{(M,g_0,\si)}(\phi_\ep)=\cF_\qD^{(M,g_\ep,\si)}(\phi_\ep)
1010: = \cF_2^{(M,g_\ep,\si)}(\phi_\ep)\leq \mu_2(M,g_\ep,\si)=\la_1^+(g_\ep)^{-1}.$$
1011: This implies the proposition.\qed
1012:
1013: Now, as we have understood the relationship between the supremum of $\cF_\qD$ and the infimum of $\la_1^+\vol^{1/n}$,
1014: we want to establish a relationship of the maximizers of $\cF_\qD$ and the minimizers of $\la_1^+\vol^{1/n}$.
1015: This will require some knowledge about the Euler-Lagrange-equation.
1016:
1017: One easily sees that
1018: $$\cF_q(\alpha \phi+\psi)= \cF_q(\phi)$$
1019: for any $\psi\in \ker\Dir$ and $\alpha\in \mR^*$.
1020: Hence maximizers of $\cF_q$ appear in families and it will be
1021: convenient to choose a good representative for each family of maximizers.
1022:
1023: \begin{lemma}[Euler-Lagrange equations of $\cF_q$.]\label{lem.euler.lagrange}
1024: Let $q\in [\qD,2]$, and choose $p$ with ${q^{-1}+p^{-1}=1}$.
1025: Suppose that $\cF_q$ has a maximizing spinor
1026: $\phi_1\in H_1^q\setminus \ker \Dir$.
1027: Then there is a maximizing spinor $\phi\in H_1^q\setminus \ker \Dir$, $\phi\in \mR^*\phi_1 + \ker D$,
1028: such that
1029: \begin{equation}\label{eq.nonlin.p}
1030: \Dir\phi= \mu_q^{-1}\, |\phi|^{p-2} \phi,\qquad \phi\in H_1^q, \qquad
1031: \|\ph\|_{L^p}=1
1032: \end{equation}
1033: \end{lemma}
1034:
1035: \proof{} We normalize $\phi_1$ such that $\|\Dir \phi_1\|_{L^q}=1$, thus $\rho_{q,\phi_1}=\mu_q$.
1036: As $\phi_1$ is a maximizer, $d\cF_q(\phi_1):H_1^q\setminus \ker \Dir\to \mR$ is identically zero.
1037: Because of \eref{eq.abl} we see that $\tau:=\phi_1-\mu_q|\Dir\phi_1|^{q-2}\Dir\phi_1\in L^p$ is a weak solution
1038: of $D\tau=0$, hence it is smooth and in the kernel of $\Dir$. Then, $\phi_2:=\phi_1-\tau=\mu_q|\Dir\phi_1|^{q-2}\Dir\phi_1$ satisfies $D\phi_1=D\phi_2$ and hence
1039: \begin{equation}\label{eq.nonlin.q}
1040: \phi_2 = \mu_q |D\phi_2|^{q-2}D\phi_2, \qquad
1041: \phi_2\in H_1^q, \qquad\|D\phi_2\|_{L^q}=1.
1042: \end{equation}
1043: Taking norms we obtain
1044: $$|\phi_2|=\mu_q|\Dir\phi_2|^{q-1}=\mu_q|\Dir\phi_2|^{q/p}.$$
1045: $$|D\phi_2|^{q-2}=|D\phi_2|^{-(q-1)(p-2)}=\left(\mu_q^{-1} |\phi_2|\right)^{-(p-2)}$$
1046: Hence $\phi:=\mu_q^{-1}\phi_2$ satisfies \eref{eq.nonlin.p}.
1047: \qed
1048:
1049: \begin{theorem}\label{theo.var.corresp}\ \\[-8mm]
1050: \begin{enumerate}[{\rm (a)}]
1051: \item Let $\psi$ be a maximizing spinor of $\cF_\qD$,
1052: and suppose that $\psi$ is smooth and that $\Dir \psi$ vanishes nowhere.
1053: Then $g:=|\Dir \psi|^{4/(n+1)}g_0$ is a smooth metric minimizing
1054: $\la_1^+\vol^{1/n}$ in the conformal class $[g_0]$.
1055: \item Let $g\in [g_0]$ be a (smooth) metric minimizing $\la_1^+\vol^{1/n}$,
1056: and let $\psi$ be an eigenspinor of $D_g$ to the eigenvalue $\la_1^+(g)$,
1057: then the length of $|\psi|_g$ is constant and $\psi$ maximizes
1058: $\cF_\qD$.
1059: \end{enumerate}
1060: \end{theorem}
1061:
1062: \proof{}
1063: \begin{enumerate}[(a)]
1064: \item Let $\psi$ be a maximizing spinor. According to the previous lemma,
1065: there is an $\al\in \mR^*$ and a $\tau\in\ker \Dir$ such that
1066: $\phi:=\al\psi +\tau$ satisfies~\eref{eq.nonlin.p} qith $q=\qD$ and $p=\pD$.
1067: In particular,
1068: $|\phi|^{\pD-1}=\mu_\qD |\Dir \phi|=\al\mu_\qD|\Dir\psi|$ vanishes nowhere.
1069: The metric $g_\phi:=|\phi|^{4\over n-1}g_0$ satisfies
1070: $\vol(M,g_\phi)=1$ and $|\phi|_{g_\phi}=1$.
1071: Then
1072: $$\left(\la_1^+(g_\phi)\right)^{-1}\geq \cF_2^{g_\phi}(\phi)=
1073: \cF_\qD^{g_\phi}(\phi)=
1074: \cF_\qD^{g_0}(\phi)=
1075: {\mu_\qD^{-1}\int|\phi|^{\pD}\over \mu_\qD^{-2}
1076: \left(\int|\phi|^{(\pD-1)\qD}\right)^{2/\qD}}
1077: =\mu_\qD.$$
1078: As $\mu_\qD^{-1}=\lammin$ we see that $\la_1^+(g_\phi)\leq\lammin$,
1079: hence $g_\phi$ minimizes $\la_1^+\vol^{1/n}$. By a simple rescaling argument
1080: one sees that
1081: $$g:=|\Dir \psi|^{4/(n+1)}g_0=\left({1\over \al}\right)^{4/(n+1)}|\Dir \phi|^{4/(n+1)}g_0=\left({\lammin\over \al}\right)^{4/(n+1)}g_\phi$$
1082: minimizes $\la_1^+\vol^{1/n}$ as well.
1083: \item By rescaling we can assume that $\vol(M,g)=1$. Unless otherwise
1084: indicated all volume measures, norms, scalar products and
1085: Dirac operators in this proof are with respect to $g$.
1086: In order to show that $|\psi|_g$ is constant, we define
1087: $$f_t:={1+t|\psi|_g^2\over \left(\int \left(1+t|\psi|_g^2\right)^n\right)^{1/n}}.$$
1088: One calculates ${d\over dt}|_{t=0}f_t= |\psi|_g^2- \int|\psi|_g^2$.
1089: All metrics $f_t^2 g$ have volume $1$.
1090: Hence, as the infimum of $\la_1^+\vol^{1/n}$ is attained in $g$, we
1091: have $\la_1^+(g)\leq\la_1^+(g_t)$, and hence
1092: $$\mu_2(M,g,\si)=\cF_2^g(\psi)\geq \cF_2^{f_t^2g}(\psi)=
1093: {\int \<D\psi,\psi\>\over \int f_t^{-1}|D\psi|^2}.$$
1094: For $t=0$ equality is attained. Hence
1095: $$\int |D\psi|^2\leq \int f_t^{-1} |D\psi|^2.$$
1096: Using $D\psi=\la_1^+(g) \psi$ and deriving with respect to $t$ yields
1097: $$0=- \left(\la_1^+(g)\right)^2\int \left(|\psi|_g^2- \int|\psi|_g^2\right)|\psi|_g^2.$$
1098: The right hand side is equal to
1099: $-\left(\la_1^+(g)\right)^2\left\||\psi|_g^2- \int|\psi|_g^2\right\|_{L^2}=0$,
1100: hence $|\psi|_g$ is constant.
1101:
1102: This implies that $\cF_q^g(\psi)$ is independent of $q$, thus
1103: $$ \cF_\qD(\psi)=\cF_2^g(\psi)=\left(\la_1^+(g)\right)^{-1}
1104: =\left(\lammin\right)^{-1}.$$
1105: And hence $\cF_\qD$ attains it supremum in $\psi$.
1106: \end{enumerate}
1107: \qed
1108:
1109: \begin{remark}
1110: Later on, we will see that maximizers of $\cF_\qD$ that vanish nowhere
1111: are always smooth.
1112: \end{remark}
1113:
1114:
1115: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1116: \section{Generalized metrics}\label{sec.spd}
1117: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1118:
1119: Unfortunately, we cannot exclude that maximizers of $\cF_q$ vanish
1120: somewhere. Maximizers with zeros correspond
1121: to metrics with certain singularities, more precisely to metrics
1122: whose conformal factor has zeros.
1123: These metrics are the main object of this section.
1124: We will summarize some facts about the size of the zero set of maximizers
1125: in Section~\ref{sec.degen}.
1126:
1127: Roughly speaking, generalized metrics are metrics of the form $f^2g_0$
1128: where $f\in L^\infty, f\geq 0$. However, for technical and formal
1129: reasons it is better to use the following definition.
1130:
1131: \begin{definition}
1132: Let $g_0$ be a smooth metric on a compact manifold $M$.
1133: A \emph{generalized metric} is a tuple $(f,g)$ where $g\in [g_0]$ and
1134: $f\in L^\infty$, $f\geq 0$. If $h>0$ is smooth, we identify
1135: $(fh,g)$ with $(f,h^2g)$. Furthermore we identify $g\in[g_0]$ with $(1,g)$.
1136: Generalized metrics having a representative of the form $(f,g_0)$ are called
1137: \emph{conformal} to $g_0$, and those having a representative $(1,g)$,
1138: $g\in [g_0]$ are called \emph{regular metrics} --- they correspond to metrics
1139: in the ordinary sense.
1140: The set of all generalized metrics conformal to $g_0$ is called the
1141: \emph{$L^\infty$-completion} $\ol{[g_0]}$ of the conformal class $[g_0]$.
1142: The \emph{volume of $g\in \ol{[g_0]}$} is defined as $\int f^n \dvol_{g_0}$.
1143: For a generalized metric we say that $f^{-1}(0)$ is the \emph{set of
1144: degeneration}.
1145: \end{definition}
1146:
1147: \begin{remark}
1148: The reader should pay some attention to the following technical difficulty:
1149: If $f$ vanishes on an open set, then the
1150: $L^\infty-(2,0)$-tensor $f^2g_0$ does not determine (the equivalence class)
1151: $(f,g_0)$.
1152: \end{remark}
1153:
1154: However, despite of this remark and slightly abusing the notation,
1155: we will write $f^2g_0$ instead of $(f,g_0)$.
1156: Formally $f^2g_0$ is a generalized metric in the above sense,
1157: not an $L^\infty-(2,0)$-tensor.
1158:
1159: Let us now assume that $M$ carries a fixed spin structure $\si$.
1160: For any generalized metric $g=f^2g_0$
1161: we want to define a spinor bundle and a Dirac operator
1162: on $(M,g,\si)$ in such a way that the results of the previous section
1163: carry over to this generalization. In particular, the functional $\cF_q$
1164: has to be defined and has to be conformally invariant for $q=\qD$.
1165:
1166: As a vector bundle the spinor bundle $\Si(M,g,\si)$
1167: is defined to be $\Si(M,g_0,\si)$, and due
1168: to our identification of spinors for different metrics in a fixed conformal
1169: class this construction does only depend on the conformal class $[g_0]$
1170: and not on the metric $g_0$ itself.
1171: For any $x\in M$ and any spinor $\phi$ in the fiber of $\Si(M,g,\si)$ over $x$
1172: we define the pointwise norm
1173: $$|\phi|_g:= \left\{
1174: \begin{matrix}
1175: f(x)^{-{n-1\over 2}}|\phi|_{g_0}\hfill & \mbox{if $f(x)\neq 0$}\hfill\cr
1176: \infty\hfill& \mbox{if $f(x)=0$ and $\phi(x)\neq 0$}\hfill\cr
1177: 0\hfill& \mbox{if $f(x)=0$ and $\phi(x)= 0$}\hfill
1178: \end{matrix}
1179: \right.
1180: $$
1181: Again, this norm depends only on the conformal class. For smooth sections
1182: $\phi$ of $\Si(M,g,\si)$ such that $|\phi|_g<\infty$ almost everywhere
1183: we define the $H_1^2(M,g,\si)$-norm as
1184: $$\left(\int f^{-1}|\Dir_{g_0}\phi|_{g_0}^2\,\dvol_{g_0}\right)^{1/2}
1185: + \left(\int |\phi|_{g_0}^\pD\,\dvol_{g_0}\right)^{1/\pD}$$
1186: where we used the conventions $0^{-1}r=\infty$ for $r>0$ and $0^{-1}0=0$.
1187: The Sobolev space
1188: $H_1^2(M,g,\si)$ is the associated completion.
1189: \begin{lemma}
1190: There is a natural inclusion
1191: $$H_1^2(M,g,\si)\embed H_1^2(M,g_0,\si)$$
1192: \end{lemma}
1193: \proof{}
1194: Cauchy sequences with respect
1195: to the norm $H_1^2(M,g,\si)$ are also Cauchy sequences with respect to
1196: $H_1^2(M,g_0,\si)$. Thus we obtain a bounded map
1197: $H_1^2(M,g,\si)\to H_1^2(M,g_0,\si)$. In order to prove injectivity
1198: of this map one shows that if $\phi_i$ is a Cauchy sequence in $H_1^2(M,g,\si)$
1199: converging to $0$ with respect to $H_1^2(M,g_0,\si)$, then it converges to $0$
1200: with respect to the $H_1^2(M,g,\si)$ as well.
1201: \qed
1202:
1203: We define the Dirac operator
1204: $\Dir_g:H_1^2(M,g,\si)\to L^2(M,g,\si)$,
1205: $\Dir_g(\phi):=f^{-1}\Dir_{g_0}(\phi)$.
1206: The spinor $D_g (\phi)$ is well-defined almost everywhere, as
1207: $\{x\in M\,|\,D_{g_0}(\phi)(x)\neq 0 \mbox{ and }f(x)=0\}$ has measure zero.
1208: It is easy to verify that all these definitions only depend
1209: on the conformal class of $g_0$ and not on $g_0$ itself.
1210: Furthermore, we can reformulate the definition of the above Sobolev space as
1211: \begin{eqnarray*}
1212: H_1^2(M,g,\si) &=& \Bigl\{\phi \in \Gamma(\Si(M,g,\si))\;
1213: \Big|\;{\textstyle\int} |\Dir_{g}\phi|_{g}^2\,\dvol_{g}<\infty\Bigr\}.
1214: \end{eqnarray*}
1215:
1216: We now extend the definitions $\cF_2^g$ and $\la_1^+(g)$
1217: to the $L^\infty$-completion of the conformal class.
1218: \begin{definition}
1219: For $g=f^2g_0\in \ol{[g_0]}$ and any $\phi\in H_1^2(M,g,\si)\setminus\ker\Dir$
1220: we define
1221: $$\cF_2^g(\phi):={\int \<\Dir_{g_0}\phi,\phi\>_{g_0}\,\dvol_{g_0}
1222: \over \int f^{-1} |\Dir_{g_0}\phi|_{g_0}^2\,\dvol_{g_0}}.$$
1223: \end{definition}
1224: Because of
1225: \begin{eqnarray*}
1226: \int f^{-1} |\Dir_{g_0}\phi|_{g_0}^2\,\dvol_{g_0}
1227: &\geq& {1\over \|f\|_{L^\infty}}\, \int|\Dir_{g_0}\phi|_{g_0}^2\,\dvol_{g_0}\\
1228: \cF_2^g(\phi)&\leq & \|f\|_{L^\infty}\,\cF_2^{g_0}(\phi)
1229: \end{eqnarray*}
1230: the functional is well-defined and bounded on $H_1^2(M,g,\si)\setminus \ker D$.
1231:
1232: It is also not hard to see that the supremum is attained. In fact,
1233: let $(\phi_i)$
1234: be a sequence of spinors in $H_1^2(M,g,\si)\setminus \ker D$ with
1235: $\cF_2^g(\phi_i)\to \sup \cF_2^g$,
1236: normalized such that $\int f^{-1} |\Dir\phi_i|^2=1$.
1237: Then a subsequence $(\phi_{i_k})$ converges weakly in $H_1^2(M,g,\si)$,
1238: weakly in $H_1^2(M,g_0,\si)$ and strongly in $L^2(M,g_0,\si)$ towards
1239: a $\phi_\infty\in H_1^2(M,g,\si)\setminus \ker D$.
1240: Hence, $\lim \cF_2^g(\phi_{i_k})\leq \cF_2^g(\phi_\infty)$. Thus,
1241: the supremum is attained in $\phi_\infty$. In analogy to
1242: Lemma~\ref{lem.f.elem} one gets for any smooth
1243: test spinor $\psi$
1244: $$\int \<f^{-1}\Dir \phi - (\cF_2^g(\phi))^{-1}\,\phi,\Dir \psi\>,$$
1245: which implies
1246: % in analogy to Lemma~\ref{lem.euler.lagrange}
1247: $\tau:=\Dir_g\phi- (\cF_2^g(\phi))^{-1}\,\phi\in \ker D$ and
1248: finally for $\witi\phi_1:=\phi_\infty+\cF_2^g(\phi)\tau$
1249: $$\Dir_g\witi\phi_1=(\cF_2^g(\witi\phi_1))^{-1}\,\witi\phi_1.$$
1250:
1251: %Note that we have the embeddings
1252: % $$H_1^2(M,g,\si)\embed H_1^2(M,g_0,\si)\embed H_{1/2}^2(M,g_0,\si).$$
1253: %Hence, $\mu_2(M,g,\si):=\sup \cF_2^g<\infty$.
1254: \begin{definition}
1255: The \emph{first positive Dirac eigenvalue}
1256: of $(M,g,\si)$, $g=f^2g_0\in \ol{[g_0]}$
1257: is
1258: $$\la_1^+(g):=\left(\sup \left\{\cF_2^g(\phi)\,|\,\phi \in H_1^2(M,g,\si)\setminus\ker\Dir\right\}\right)^{-1}.$$
1259: A non-trivial spinor with
1260: $$\Dir_g\phi=\la_1^+(g)\phi$$
1261: is an \emph{eigenspinor to the eigenvalue $\la_1^+(g)$}.
1262: \end{definition}
1263:
1264: As we have already seen, this definition coincides with the
1265: definition of the first positive Dirac eigenvalue and a corresponding
1266: eigenspinor if $g$ is regular.
1267:
1268: Most of the statements of the previous section still hold in a
1269: modified version for generalized metrics, the proofs are nearly identical.
1270: For example one can extend Proposition~\ref{prop.lamu} to
1271: \begin{proposition}\label{prop.bar.nonbar.gen}
1272: $$\inf_{g\in [g_0])}\la^+_1(g)\vol(M,g)^{1/n}
1273: =\inf_{g\in \ol{[g_0]}}\la^+_1(g)\vol(M,g)^{1/n} = \mu_{\qD}(M,g,\si)^{-1}.$$
1274: \end{proposition}
1275: The equation $\inf_{g\in [g_0]}\la^+_1(g)\vol(M,g)^{1/n}=\mu_{\qD}(M,g,\si)^{-1}$
1276: is exactly the statement of Proposition~\ref{prop.lamu} and the equation
1277: $\inf_{g\in \ol{[g_0]}}\la^+_1(g)\vol(M,g)^{1/n} = \mu_{\qD}(M,g,\si)^{-1}$
1278: can be proven with exactly the same proof.
1279:
1280: Similarly, we obtain an analogue of Theorem~\ref{theo.var.corresp}.
1281:
1282: \begin{theorem}\label{theo.var.corresp.gen}
1283: Let $(M,g_0,\si)$ be a compact Riemannian spin manifold.\\[-7mm]
1284: \begin{enumerate}[{\rm (a)}]
1285: \item Let $\psi$ be a maximizing spinor of $\cF_\qD$
1286: (with regularity $H_1^\infty$).
1287: Then $g:=|\Dir \psi|^{4/(n+1)}g_0$ is a generalized metric minimizing
1288: $\la_1^+\vol^{1/n}$ in $\ol{[g_0]}$, the $L^\infty$-completion of
1289: the conformal class $[g_0]$.
1290: \item Let $g=f^2g_0\in \ol{[g_0]}$ be a generalized
1291: metric minimizing $\la_1^+\vol^{1/n}$,
1292: and let $\psi$ be an eigenspinor of $D_g$ to the eigenvalue $\la_1^+(g)$,
1293: then the length of $|\psi|_g$ is constant on $M\setminus f^{-1}(0)$ and
1294: $\psi$ maximizes $\cF_\qD$.
1295: \end{enumerate}
1296: \end{theorem}
1297:
1298: The proof of this theorem is essentially the same as the proof
1299: of Theorem~\ref{theo.var.corresp}.
1300: Later, we will see that any maximizing spinor has regularity
1301: $C^{2,\al}$, hence $g=|D\psi|^{4/(n+1)}g_0$ is always a generalized metric.
1302:
1303:
1304:
1305:
1306:
1307: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1308: \section{The case of the sphere}\label{sec.sphere}
1309: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1310:
1311:
1312: The sphere $\mS^n=(S^n,\can)$ with the round sphere
1313: is an important example. On the one hand the invariant $\lammin$
1314: of the round sphere is contained
1315: in many equations and inequalities. On the other hand the
1316: round sphere has a large conformal group,
1317: and hence studying the minimizers on the sphere helps to understand
1318: the analytical difficulties. In particular, we will see
1319: why the conclusion in Theorem~\ref{theo.anydim} does not hold if
1320: $\mu_\qD=\mu_\qD^{\mS^n}$. (Recall $\qD=2n/(n+1)$.)
1321:
1322: The invariant $\lammin(\mS^n)$ is not hard to calculate. Recall
1323: that the Yamabe invariant $Y(\mS^n)$ is given by
1324: $$Y(\mS^n):=\inf_{g\in [\can]}{\int \scal_g\,\dvol_g\over \vol(M,g)^{(n-2)/n}}.$$
1325: It is attained for $g=\can$, and hence $Y(\mS^n)= n (n-1) \, \om_n^{2/n}$,
1326: where $\om_n:=\vol(\mS^n)$.
1327: The Hijazi inequality (\ref{ineq.hij},\ref{ineq.hij.der}) tells us, that
1328: \begin{equation}\label{ineq.sphere.one}
1329: \lammin(\mS^n)^2\geq {n\over 4 (n-1)}\,Y(\mS^n)= {n^2\over 4}\, \om_n^{2/n}.
1330: \end{equation}
1331: Recall that the sphere of constant sectional curvature $1$ carries
1332: a Killing spinor $\psi$ to the constant $-1/2$, i.e. it satisfies
1333: $$\na_X \psi= -(1/2) X\cdot \psi.$$
1334: Note that this condition implies
1335: that the length of $\psi$ is constant. Because of
1336: $D\psi= (n/2) \psi$ we obtain
1337: \begin{equation}\label{ineq.sphere.two}
1338: \lammin(\mS^n)\leq \la_1^+(\mS^n)\vol(\mS^n)^{1/n}\leq {n\over 2}\, \om_n^{1/n}.
1339: \end{equation}
1340: %Together with Proposition~\ref{prop.lamu} we obtain
1341: And then
1342: $$\lammin(\mS^n)
1343: = {n\over 2}\, \om_n^{1/n}.$$
1344: The above Killing spinor satisfies
1345: $$\left(\cF_{\qD}(\psi)\right)^{-1}
1346: = {n\over 2}\, \om_n^{1/n} \left(\sup \cF_{\qD}\right)^{-1},$$
1347: hence $\cF_{\qD}$ attains its supremum in such Killing spinors.
1348:
1349: Let $A:\mS^n\to \mS^n$ be an orientation preserving
1350: conformal diffeomorphism. Then the pullback of any spinor
1351: $\phi$ with respect to $A$ is a section
1352: of $A^*(\Si(S^n,\can))=\Si(S^n,A^*\can)$,
1353: and as before we identify $\Si(S^n,A^*\can)\cong \Si(S^n,\can)$ by using
1354: the map
1355: $\witi F$ described after Proposition~\ref{prop.conf.change}.
1356: If $\phi$ is a solution of \eref{eq.nonlin.p}, then this pullback,
1357: denoted by $A^*\phi$, is a solution of this equation as well.
1358: Furthermore if $\phi$ maximizes $\cF_{\qD}$, then $A^*\phi$ is also a maximizer.
1359: As a consequence, all conformal images
1360: of Killing spinors to the constant $-1/2$ are maximizers of $\cF_{\qD}$.
1361: The following proposition shows that there
1362: are no other maximizers on $\mS^n$.
1363:
1364: %We will not use in the proof of our main results. However, it will enter into
1365: %Remark~\ref{rem.sphereconc}.
1366:
1367: \begin{proposition}
1368: %On the standard sphere we have
1369: % $$\sup \cF_{\qD}={2\over n}\,\om_n^{-1/n},$$
1370: %and if the supremum is attained in $\psi$,
1371: If $\psi$ is a spinor of regularity $C^2$ that
1372: attains the supremum of $\cF_{\qD}$,
1373: then there is a Killing spinor $\phi$ to the Killing constant $-1/2$
1374: and an orientation preserving conformal diffeomorphism $A:\mS^n\to \mS^n$
1375: with $A^*\phi=\psi$.
1376: \end{proposition}
1377:
1378: Later on we will see that any maximizer of regularity $H_1^q$
1379: with $q>\qD$ is even $C^2$. Hence the statement also holds under this
1380: weaker assumption.
1381:
1382: \begin{lemma}\label{lem.scal.bound}
1383: Let $(M,g,\si)$ be an arbitrary Riemannian spin manifold
1384: (not necessarily complete or compact). Assume that there is a spinor $\psi$
1385: of constant length $1$ and with $D\psi=\lambda \psi$. Then
1386: $$\scal = 4 {n-1\over n}\,\lambda^2\,-4 |\witi \nabla \psi|^2,$$
1387: where $\witi\nabla_X \psi:=\na_X\psi + {\la\over n} X\cdot \psi$ denotes
1388: the \emph{Friedrich connection} on spinors.
1389: \end{lemma}
1390:
1391: \proof{of the lemma}
1392: The Friedrich connection is a metric connection, hence
1393: $$0={1\over 2}\,d^*d\<\psi,\psi\> = {\rm Re} \<\witi\na^* \witi\na \psi,\psi\>
1394: - \<\witi\na\psi,\witi\na\psi\>.$$
1395: The twisted version of the Schr\"odinger-Lichnerowicz formula
1396: yields
1397: $$\left(D-{\lambda\over n}\right)^2=\witi\na^*\witi\na + {\scal\over 4}- {(n-1) \lambda^2 \over n^2}.$$
1398: Hence
1399: $$\left(\la-{\la\over n}\right)^2= \<\witi\na^*\witi\na \psi,\psi\>+ {\scal\over 4}- {(n-1) \lambda^2 \over n^2}.$$
1400: We obtain
1401: $$ {n-1\over n} \lambda^2
1402: = \<\witi\na\psi,\witi\na\psi\> + {\scal\over 4}.$$
1403: \qed
1404: \proof{of the proposition}
1405: %As we have already seen above, Killing spinors to
1406: %the Killing constant $-1/2$ are maximizers of $\cF_\qD$ and its images
1407: %under conformal maps.
1408:
1409: %As $\cF_\qD$ is conformally
1410: %invariant, the supremum is also attained by the image of such a Killing spinor
1411: %under an orientation preserving
1412: %conformal map $\mS^n\to \mS^n$.
1413:
1414: We have to show that the supremum is not attained by any other spinor.
1415: For proving this, assume that $\psi$ is a maximizer,
1416: $\|\psi\|_{L^\qD}=1$.
1417: As the Dirac operator on $\mS^n$ has kernel $\{0\}$, the spinor
1418: satisfies the Euler-Lagrange
1419: equation \eref{eq.nonlin.p}.
1420:
1421: On the open subset $S^n\setminus \psi^{-1}(0)$ we define the metric
1422: $$g_1:= |\psi|^{4\over n-1}\can.$$
1423: In this metric \eref{eq.nonlin.p} transforms into a solution of
1424: $$D_{g_1}\psi=\lammin \psi\qquad |\psi|_{g_1}\equiv 1.$$
1425:
1426: On the one hand, one calculates
1427: \begin{eqnarray*}
1428: \scal_{g_1}&=&
1429: 4 {n-1\over n-2}\,|\psi|^{-{n+2\over n-1}}\Delta_{\can} |\psi|^{n-2\over n-1}
1430: + \scal_{\can}|\psi|^{-{4\over n-1}}\\
1431: \end{eqnarray*}
1432: and integration yields
1433: \begin{eqnarray}%\label{ineq.preyama}
1434: \int_{S^n\setminus \psi^{-1}(0)}\scal_{g_1}\,\dvol_{g_1}
1435: &=& \int_{M\setminus \psi^{-1}(0)} 4 {n-1\over n-2}\,|\psi|^{{n-2\over n-1}}\Delta_{\can} |\psi|^{n-2\over n-1}
1436: + n(n-1)|\psi|^{2{n-2\over n-1}}\,\dvol_{\can}\nonumber\\
1437: &=& \int_{S^n}
1438: 4 {n-1\over n-2}\,\left|d|\psi|^{n-2\over n-1}\right|_{\can}^2
1439: + n(n-1)|\psi|^{2{n-2\over n-1}}\,\dvol_{\can},\label{ineq.preyamb}
1440: \end{eqnarray}
1441: where the last term arises by partial integration. In order to make this step
1442: precise one has to exhaust $M\setminus \psi^{-1}(0)$ by smooth manifolds
1443: with boundary and partially integrate over these exhausting manifolds.
1444: The boundary terms vanish in the limit, as $\psi\to 0$ on the boundaries and $d|\psi|$ is bounded.
1445:
1446: It is a standard fact from the resolution of the Yamabe problem (see e.g.\
1447: \cite{lee.parker:87})
1448: that any $H_1^2$ function $f$ satisfies
1449: \begin{eqnarray}\label{ineq.yamafunc}
1450: {\int_{S^n}
1451: 4 {n-1\over n-2}\,\left|df\right|^2
1452: + n(n-1)f^2\,\dvol_{\can}\over \left(\int_{S^n} f^{2n\over n-2}\,
1453: \dvol_{\can}\right)^{n-2\over n}} &\geq& Y(\mS^n)=n(n-1)\om_n^{2/n}.
1454: \end{eqnarray}
1455: Setting $f:=|\psi|^{n-2\over n-1}$, we obtain $\int_{S^n} f^{2n\over n-2}=1$,
1456: and hence the right side of \eref{ineq.preyamb} is bounded from below
1457: by $n(n-1)\om_n^{2/n}$.
1458:
1459: One the other hand, the previous lemma provides
1460: $$\scal_{g_1}\leq 4 {n-1\over n}\left(\lammin\right)^2=n(n-1)\om_n^{2/n},$$
1461: and as $\vol(S^n\setminus \psi^{-1}(0),g_1)=1$, we see that we must have
1462: equality in all inequalities involved, in particular in~\eref{ineq.yamafunc}.
1463: An application of the maximum principle
1464: \cite{lee.parker:87} yields that $f$ does not vanish, and hence $\psi$
1465: has no zeros. Furthermore, $g_1$ is a metric of constant scalar curvature
1466: conformal to $\mS^n$, and such a metric is necessarily of the form
1467: $g_1:=A^*\can$ for an orientation preserving conformal diffeomorphism
1468: $A:\mS^n\to \mS^n$. With respect to $g_1$ we obtain
1469: $\witi \na \psi=0$, hence $\psi$ is a Killing spinor on $(S^n,g_1)$.
1470: This implies that $\phi:=(A^{-1})^*\psi$ is a Killing spinor on $\mS^n$.
1471: \qed
1472:
1473:
1474: %According to standard results in the Yamabe problem, any metric in the form
1475: %$g=f^2 g_0$, $f$ continuous on $M$ and smooth on $U':=M\setminus f^{-1}(0)$
1476: %satisfies
1477: % $$Y(M,[g_0])\leq{\int_{U'} \scal_g\dvol_g\over \vol(U',g)^{n-2\over n}}.$$
1478: %Equality in this inequality
1479: %implies that $f^{-1}(0)=\emptyset$, and that $g$ is a Yamabe metric.
1480: %Applied to $M=S^n$, $f:= |\psi|^{2\over n+1}$ we obtain
1481: % $$n(n-1)\om_m^{2/n}=Y(S^n,\gcan)\leq 4 {n-1\over n}
1482: % \left(\lammin\right)^2\,\int_U\dvol_{g_1} = (n-1)n\, \om_n^{2/n}.$$
1483: %Hence, we have equality, which implies that $g_1$ is a Yamabe metric of
1484: %volume $1$ on $S^n$. This implies that there is an orientation preserving
1485: %conformal map $h:S^n\to S^n$ such that
1486: %$g_1:=\om_n^{-2/n} h^* \gcan$ and that $\psi$ is the pullback under $h$
1487: %of a Killing spinor on $(S^n,\gcan)$.
1488: %\qed
1489:
1490:
1491: \begin{remark}
1492: There are solutions to \eref{eq.nonlin.p} that do not maximize the functional.
1493: An easy construction of such a solution is as follows.
1494: The map $A:\mS^2=\mC\cup\{\infty\}\to \mS^2=\mC\cup\{\infty\}$,
1495: $z\mapsto z^k$, $k\in\mN\setminus\{0\}$
1496: is conformal with branching points $0$ and $\infty$. If $\phi$ is a solution
1497: of $D\phi=c|\phi|^2\phi$, then
1498: %with the help of the identification
1499: %of spinors for conformally related metrics following
1500: %Proposition~\ref{prop.conf.change}, one can pullback
1501: %the solution to a solution
1502: the pullback $\psi:=A^*\phi$ is a solution of
1503: $D\psi=c|\psi|^2\psi$ on $\mS^2\setminus\{0,\infty\}$.
1504: Here $\psi$ is a section of the pull-backed spinor bundle, which is defined
1505: using the pull-backed spin structure. The pull-backed spin structure
1506: on $\mS^2\setminus\{0,\infty\}$ coincides with the standard spin structure
1507: iff $k$ is odd. If $k$ is odd, one can show that the extension of $\psi$ by
1508: setting $\psi(0)=0$ and $\psi(\infty)=0$ is a solution of
1509: $D\psi=c|\psi|^2\psi$ on $\mS^2$.
1510: However, $\int|\psi|^4= k \int|\phi|^4$. This implies
1511: $\cF(\psi)=k^{-1/3}\cF(\phi)$. Hence, if $\phi$ is a Killing spinor,
1512: and $k\geq 3$, $k$ odd, then $\psi$ is a non-maximizing solution.
1513: \end{remark}
1514:
1515:
1516: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1517: \section{Regularity theorems for the Euler-Lagrange equations}\label{sec.reg.theo}
1518: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1519:
1520: In this section we want to study solutions of the Euler-Lagrange equations.
1521: The first subsection ``Removal of singularities'' will be used in
1522: the following section to
1523: extend a solution of $\mR^n=\mS^n\setminus\{\mbox{South Pole}\}$
1524: to a solution on $\mS^n$. The second subsection states that solutions
1525: of the Euler-Lagrange equations are $C^{2,p-2}$.
1526:
1527: \subsection{Removal of singularities}
1528:
1529: \begin{theorem}[Removal of singularities theorem]\label{theo.sing.rem}
1530: Let $p\in[{n\over n-1},\infty)$.
1531: Let $(U,g)$ be a (not necessarily complete) Riemannian manifold
1532: equipped with a spin structure, let $x\in U$.
1533: Assume that $\phi\in L^p(\Si(U\setminus\{x\}),g)$
1534: satisfies weakly on $U\setminus\{x\}$ the equation
1535: \begin{equation}\label{eq.singsol}
1536: D\phi = \la |\phi|^{p-2}\phi.
1537: \end{equation}
1538: Then this equation even holds weakly on $U$. In particular, the distribution
1539: $D\phi$ does not have singular support in $x$ and is contained in
1540: $L^q$.
1541: \end{theorem}
1542:
1543: \proof{}
1544: Let $\psi$ be a smooth spinor compactly supported in $U$.
1545: We have to show
1546: \begin{eqnarray}\label{eq.singsol.show}
1547: \int_U\<\phi,\Dir \psi\>= \la \int_U \< |\phi|^{p-2}\phi,\psi\>.
1548: \end{eqnarray}
1549:
1550: For any small $\ep>0$ we choose a smooth
1551: cut-off function $\eta_\ep:U\to[0,1]$
1552: with $\eta_\ep\equiv 1$ on $B_\ep(x)$, with $|\na \eta_\ep|\leq 2/\ep$
1553: and with support in $B_{2\ep}(x)$.
1554: We rewrite the left hand side as
1555: \begin{eqnarray}
1556: \int_U\<\phi,\Dir \psi\>
1557: &=&\int_U\Bigl\<\phi,\Dir \Bigl( (1-\eta_\ep) \psi + \eta_\ep \psi\Bigr)\Bigr\>\\
1558: &=&\int_U\Bigl\<\phi,\Dir \Bigl( (1-\eta_\ep)\Bigr) \psi\Bigr\>
1559: + \int_U\<\phi,\eta_\ep \Dir \psi\> + \int_U\<\phi,\na \eta_\ep \cdot \psi\>\nonumber
1560: \end{eqnarray}
1561: As $\phi$ is a weak solution of \eref{eq.singsol}
1562: on $U\setminus\{x\}$, the first term
1563: equals to
1564: $$\la \int \< |\phi|^{p-2}\phi,(1-\eta_\ep)\psi\>,$$
1565: and for $\ep\to 0$ it tends to the right hand side of~\eref{eq.singsol.show}.
1566:
1567: Let $q$ be related to $p$ via $1/q+1/p=1$.
1568: The absolute value of the second term is bounded by
1569: $$\|\phi\|_{L^p(B_{2\ep})}\,\|D\psi\|_{L^q(B_{2\ep})}$$
1570: which tends to $0$ for $\ep\to 0$.
1571:
1572: Finally, the absolute value of the third term is bounded by
1573: $${2\over \ep}\,\|\phi\|_{L^p(B_{2\ep})}\,\|\psi\|_{L^q(B_{2\ep})}\leq C\,\|\phi\|_{L^p(B_{2\ep})}\,
1574: \ep^{{n\over q}-1}.$$
1575: Our condition $p\geq {n\over n-1}$ yields $q\leq n$, and hence the third term
1576: also tends to $0$ for $\ep\to 0$.
1577: \qed
1578:
1579:
1580: \subsection{Regularity}
1581:
1582: \begin{theorem}[$C^{1,\alpha}$-regularity theorem]\label{theo.reg}
1583: Suppose that $\phi\in H_1^q$, $q\in [\qD,2]$
1584: is a solution of equations~\eref{eq.nonlin.p}.
1585: Suppose that there is an $r>\pD$ such that $\|\phi\|_{L^r}<\infty$.
1586: Then $\phi$ is $C^{1,\alpha}$ for any $\alpha\in (0,1)$.\\
1587: Furthermore, we obtain a uniform bound of the $C^{1,\alpha}$-norm in the
1588: following sense.
1589: Let us choose $k,K>0$ such that $\|\phi\|_{L^r}<k$ and $\mu_q\geq K$.
1590: Then for any $\al\in (0,1)$ there is a constant
1591: $C$ depending only on $(M,g,\si)$, $p$, $r$, $K$, $k$ and $\al$ with
1592: $$\|\phi\|_{C^{1,\al}}\leq C.$$
1593: \end{theorem}
1594:
1595: \begin{remark}The following
1596: example shows that the theorem cannot hold without the $L^r$-bound.
1597: Let $M=\mS^n$ and $p=\pD$.
1598: Let $\psi$ be a Killing spinor to the Killing constant $-1/2$.
1599: Suppose $\|\psi\|_{L^\pD}=1$.
1600: Let $A:\mS^n\to \mS^n$ be a M\"obius transformation,
1601: such that the differential in $x\in \mS^n$
1602: satisfies $(dA)_x=2 \Id$.
1603: Then $\psi_i=d(A^i)\psi$ is a family of solutions of \eref{eq.nonlin.p},
1604: maximizing $\cF_\qD$.
1605: However, for any $r>\pD$ one can show that $\|\psi_i\|_{L^r}\to \infty$
1606: for $i\to \infty$.
1607: Hence, the $L^r$-bound is necessary for the theorem to hold.
1608: \end{remark}
1609:
1610: \begin{remark}
1611: The theorem will be applied in several versions. At first, we will apply it when
1612: $p<\pD$. In this case the Sobolev embedding already provides the required
1613: $L^r$-bound on $\phi$. However, the $r$ given by the Sobolev embedding depends on $p$. It will
1614: be of central importance to obtain a bound that is uniform for $p\to \pD$.
1615: After having proved Theorem~\ref{theo.anydim}
1616: the uniformity statement in the above theorem will be used to
1617: obtain a bound that is uniform for $p\to \pD$.
1618: Finally, the regularity theorem will be applied in the case $p=\pD$. In this case an additional $L^r$-bound
1619: is required as well.
1620: \end{remark}
1621:
1622: \proof{}
1623: The proof uses the following ``bootstrap argument''.
1624: At first, we assume $r<n\,{n+1\over n-1}$.
1625: As $\phi$ is $L^r$, the right hand side of \eref{eq.nonlin.p}, i.e.\ $|\phi|^{p-2}\phi$,
1626: is $L^{r/(p-1)}\embed L^s$ with $s:={r/(\pD-1)}= r {n-1\over n+1}< n$.
1627: We apply the Global $L^p$-estimates~\ref{theo.global.lp} and get $\phi\in H_1^s$.
1628: Using the Sobolev embedding I, Theorem~\ref{theo.sobo} (a)
1629: one obtains $\phi\in L^{r'}$ with $r'={ns\over n-s}={rn {n-1\over n+1}\over n-r{n-1\over n+1}}$.
1630: Using $r>\pD=2n/n-1$ one sees that $r'>r$, hence we have obtained stronger regularity for $\phi$.
1631: We iterate this argument and get $L^{\ti r}$-bounds for arbitrarily
1632: large~${\ti r}$. For any
1633: $\ti r>n\,{n+1\over n-1}$, we obtain $\phi\in H_1^{\ti s}$ with $\ti s:={\ti r/(\pD-1)}>n$. We apply
1634: the Sobolev embedding theorem~II, Theorem \ref{theo.sobo} (c) and obtain $\phi\in C^{0,\al}$ for any $\al\in (0,1)$.
1635: Hence $|\phi|^{p-2}\phi$ is $C^{0,\al}$ as well, and applying Schauder estimates \ref{theo.schauder}
1636: we get $\phi\in C^{1,\al}$ for arbitrary $\al$.
1637:
1638: The uniformity of the upper bound is clear from the construction.
1639: \qed
1640:
1641: The bootstrap can be continued and we obtain better regularity.
1642:
1643: \begin{proposition}[Improved regularity]\label{prop.imp}
1644: We assume the assumptions of the previous theorem.
1645: \begin{enumerate}[{\rm (1)}]
1646: \item Let $U:=M\setminus\phi^{-1}(0)$. Then $\phi|_U\in C^{\infty}(U)$.
1647: \item If $p > 2$,
1648: then $\phi\in C^{2,\alpha}$ for any $\alpha\in (0,1)\cap(0, p-2]$.
1649: Furthermore,
1650: $$\|\phi\|_{C^{2,\al}}\leq C,$$
1651: where $C$ depends only on $(M,g,\si)$, $p$, $r$, $K$, $k$ and $\al$.
1652: \item If $n=2$ and $p=\pD=4$, then $\phi\in C^\infty$.
1653: Furthermore,
1654: $$\|\phi\|_{C^m}\leq C,$$
1655: where $C$ depends only on $(M,g,\si)$, $p$, $r$, $K$, $k$ and $m$.
1656: \end{enumerate}
1657: \end{proposition}
1658:
1659: \proof{} (1) On $U$ we can continue the
1660: bootstrap argument and apply inductively the Schauder estimates
1661: Theorem~\ref{theo.schauder}.
1662: We conclude that $\phi$ is smooth on $U$. We obtain~(1).
1663: However, uniform bounds will be difficult to obtain.
1664:
1665: (2) The case $p=2$ is trivial.
1666: Let $p>2$. We know that $\phi$ is $C^{1,\al}$ for any $\alpha$.
1667: Hence, using Appendix~\ref{sec.scha.power} one sees that
1668: $|\phi|^{p-2}\phi$ is $C^{1,\al}$. The Schauder estimates imply that $\phi$
1669: is $C^{2,\al}$.
1670:
1671: (3) Similarly, if $n=2$ and $p=4$,
1672: then $\phi\mapsto |\phi|^{p-2}\phi$ is also smooth in $0$. Hence the bootstrap
1673: can go on with higher order Schauder
1674: estimates, and we inductively get $\phi\in C^m(M)$ for any $m$.
1675: The construction obviously provides uniform bounds. We obtain~(3).
1676: \qed
1677:
1678:
1679:
1680: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1681: \section{Solution of the variational principle}\label{sec.solution}
1682: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1683:
1684: \begin{proposition}\label{prop.sol.subcrit}
1685: Let $q$ and $p$ be related via $1/p+1/q=1$. For $q\in(\qD,2)$
1686: the supremum $\mu_q$ of $\cF_q$ is attained by a spinor field
1687: $\phi\in C^{2,\al}$, $\alpha\in (0,1)\cap(0, p-2]$,
1688: $1/p+1/q=1$. The spinor $\phi$ can be chosen such that $\phi$
1689: is a solution of \eref{eq.nonlin.p}.
1690: \end{proposition}
1691:
1692: \proof{}
1693: Let $\phi_i$ be a maximizing sequence for $\cF_q$, \ie
1694: $\cF_q(\phi_i)\to \mu_q$. We may assume $\|\Dir \phi_i\|_{L^q}=1$, and that
1695: $\phi_i$ is orthogonal to $\ker \Dir$.
1696: After taking a subsequence there is a $\phi_\infty\in H_1^q$
1697: such that $\phi_i$ converges weakly to $\phi_\infty$ in $H_1^q$.
1698: Thus $\|\Dir\phi_\infty\|_{L^q}\leq \liminf \|\Dir\phi_i\|_{L^q}=1$.
1699: The compactness of the embedding $H_1^q\hookrightarrow L^p$ provides
1700: a subsequence, that converges strongly to
1701: $\phi_\infty$ in $L^p$.
1702: This implies
1703: $$\int\<D\phi_i,\phi_i\>=
1704: \underbrace{\int\<D\phi_i,\phi_i-\phi_\infty\>}_{\leq\|D\phi_i\|_{L^q}\|\phi_i-\phi_\infty\|_{L^p}}
1705: +\underbrace{\int\<D\phi_i,\phi_\infty\>}_{\to\int\<D\phi_\infty,\phi_\infty\>}\to \int\<D\phi_\infty,\phi_\infty\>.
1706: $$
1707: %This implies
1708: %$\int\<\Dir \phi_i,\phi_i\>\to\int\<\Dir \phi,\phi\>$.
1709: Hence,
1710: $$\mu_q= \lim \cF_q(\phi_i)= \lim
1711: {\int\<\Dir \phi_i,\phi_i\>\over \|\Dir\phi_i\|_{L^q}^2}\leq
1712: {\int\<\Dir \phi_\infty,\phi_\infty\>\over \|\Dir\phi_\infty\|_{L^q}^2}=
1713: \cF_q(\phi_\infty)\leq \mu_q.$$
1714: As a consequence, we have equality in all inequalities, in particular
1715: $\|\Dir\phi_\infty\|_{L^q}=1$.
1716: According to Lemma~\ref{lem.euler.lagrange} one can
1717: find $\al\in \mR^*$ and $\tau\in \ker \Dir$ such that
1718: $\phi=\al\phi_\infty+\tau$
1719: solves \eref{eq.nonlin.p}.
1720: Proposition~\ref{prop.imp} (Improved Regularity)
1721: tells us that $\phi$ is $C^{2,\al}$.
1722: \qed
1723:
1724:
1725: \begin{theorem}[Uniform $C^0$-estimate]\label{theo.anydim}
1726: Let $\phi$ be a solution of \eref{eq.nonlin.p} with $p\in [2,\pD)$
1727: and $\mu_q \geq \mu_{\qD}^{\mS^n}+\ep$, $\ep>0$.
1728: Then there is a constant $C=C(M,g,\si,\ep)$ such that
1729: $$\|\phi\|_{C^0} < C.$$
1730: \end{theorem}
1731:
1732: \begin{remark}\label{rem.sphereconc}
1733: The conclusion of this theorem does not hold any longer, if we drop
1734: the condition $\mu_q>\mu_\qD^{\mS^n}$. In fact, if $q=\qD$ and
1735: $(M,g,\si)=\mS^n$, then there is the following counterexample:
1736: Let $A$ be a non-isometric, but conformal map $\mS^n\to\mS^n$
1737: fixing North and South pole. Let $\phi$ be a Killing spinor. As seen in the
1738: previous section $\phi$ is a maximizer of $\cF_\qD$ and a solution
1739: of \eref{eq.nonlin.p}. The images $\phi_k:=(A^k)_*\phi$ under the conformal
1740: maps $A^k$ are also maximizers of $\cF_\qD$ and solutions
1741: of \eref{eq.nonlin.p}. However, as easily seen, $\|\phi_k\|_{C^0}$
1742: tends to $\infty$ for $k\to \infty$.
1743: \end{remark}
1744:
1745:
1746: \proof{of the theorem}
1747: Assume that such a constant does not exist. Then we find a sequence
1748: of solutions $\phi_k$ of
1749: \begin{equation}\label{eq.nonlin.pk}
1750: \Dir\phi_k= \mu_{q_k}^{-1}\, |\phi_k|^{p_k-2} \phi_k,\qquad \phi_k\in H_1^{q_k}, \qquad
1751: \|\phi_k\|_{L^{p_k}}=1
1752: \end{equation}
1753: $1/q_k = 1 - 1/p_k$, $\mu_{q_k} \geq \mu_{\qD}^{\mS^n}+\ep$ and
1754: \begin{eqnarray}\label{form.div}
1755: \|\phi_k\|_{C^0}\to \infty.
1756: \end{eqnarray}
1757:
1758: Let us assume for a moment that $p_\infty:=\liminf p_k < \pD$.
1759: In this case, we can choose a subsequence with $p_k\to p_\infty$.
1760: We have
1761: $1=\|\phi_k\|_{L^{p_k}}^{p_k}= \Bigl\|\,|\phi_k|^{p_k-2}\phi_k\Bigr\|_{L^{q_k}}^{q_k}={\mu_{q_k}}^{q_k}\|D\phi_k\|_{L^{q_k}}^{q_k}$.
1762: We conclude that
1763: $\phi_k$ is bounded in $H_1^{\ti q}$ for a $\ti q > \qD$,
1764: and hence in $L^r$ for an $r>\pD$.
1765: Then the regularity theorem (Theorem~\ref{theo.reg}) says that
1766: $\|\phi_k\|_{C^0}$ is bounded, in contradiction to \eref{form.div}.
1767: Hence, the case $p_\infty>\pD$ can not occur, i.e.\ $\lim p_k = \pD$.
1768:
1769: There is a sequence of points $s_k\in M$ with
1770: $$m_k:=|\phi_k(s_k)|=\max\left\{|\phi_k(x)|\,\big|\,x\in M \right\}\to \infty.$$
1771: %Since $M$ is compact, we can assume that $s_k$ converges to $p\in M$.
1772: The idea is to blow up suitably the metric such that we obtain
1773: in the limit a solution on Euclidean $\mR^n$.
1774:
1775: We define
1776: \begin{eqnarray*}
1777: \ti g_k&:=&(m_k)^{2(p_k-2)}g\\
1778: \ti \phi_k&:=& (m_k)^{(p_k-2)\frac{n-1}{2}-1} \phi_k.
1779: \end{eqnarray*}
1780: One easily verifies
1781: \begin{equation}\label{eq.normed}
1782: |\ti \phi_k(s_k)|_{\ti g_k}=1,
1783: \end{equation}
1784: and obviously \eref{eq.nonlin.p} transforms into
1785: \begin{equation}\label{eq.preser}
1786: D_{\ti g_k}\ti \phi_k= \frac{1}{\mu_{q_k}}|\ti\phi_k|_{\ti g_k}^{p_k-2}\ti\phi_k.
1787: \end{equation}
1788: We calculate
1789: \begin{eqnarray*}
1790: \left\|\ti\phi_k\right\|_{L^{p_k}(\ti g_k)}^{p_k}
1791: &= &\mu_{q_k}\int\<D_{\ti g_k}\ti\phi_k,\ti\phi_k\>_{\ti g_k}\,\dvol_{\ti g_k}\\
1792: &=& \mu_{q_k}\int\<D_g\ti\phi_k,\ti\phi_k\>\,\dvol_g\\
1793: &=&\left(m_k\right)^{2 \left((p_k-2)\frac{n-1}{2}-1\right)}
1794: \underbrace{\left\|\phi_k\right\|_{L^{p_k}(g)}^{p_k}}_{=1}.
1795: \end{eqnarray*}
1796: We can assume that $m_k\geq 1$.
1797: As $p_k\leq\pD$ implies $(p_k-2)\frac{n-1}{2}-1\leq 0$, we obtain
1798: \begin{equation}\label{ineq.unif.lp}
1799: \left\|\ti\phi_k\right\|_{L^{p_k}(\ti g_k)}\leq 1.
1800: \end{equation}
1801: The injectivity radius of $(M,\ti g_k)$ tends to infinity, i.e.\
1802: for any $R>0$ there is a $k_0=k_0(R)\in\mN$
1803: such that for all $k\geq k_0$ the exponential map
1804: $\exp^{\ti g_k}_{s_k}:T_{s_k}M\to M$
1805: with respect to $\ti g_k$ and based in $s_k$ is a diffeomorphism
1806: on the ball of radius $R$ around $0$.
1807: Now, we identify $(T_{s_k}M,\ti g_k)$ with $(\mR^n,\geucl)$.
1808: Then $\overline g_k:=(\exp^{\ti g_k}_{s_k})^*{\ti g_k}$
1809: is Riemannian metric on $B_R(0)$ that
1810: coincides with $\geucl$ in $0$.
1811: In the limit $k\to \infty$ the metrics $\overline g_k$ converge
1812: to $\geucl$ in the $C^\infty$-topology.
1813:
1814: As already said in previous sections,
1815: the construction of the spinor bundle, its scalar product
1816: and its connection depends on the Riemannian metric (and the spin structure).
1817: In order to work out the blowup construction one has to define
1818: a pull-back of the spinors
1819: $\witi\phi_k\in \Gamma(\Si(M,g_k,\si))$ via normal coordinates and then obtain
1820: a spinor in $\Gamma(\Si(B_R(0),\geucl))$.
1821: In the literature two such pull-backs are used, a pullback
1822: construction carried out in \cite{ammann.humbert.morel:p03av2} and inspired by
1823: \cite{bourguignon.gauduchon:92}, or via radial parallel transport
1824: (see e.g.\ the solution of the index problem in \cite{roe:88}
1825: using Getzler rescaling). Both pullbacks can be used here.
1826: The pullback in \cite{ammann.humbert.morel:p03av2} has better
1827: approximation properties, and is an important tool if one wants to study
1828: fine asymptotics of the blowup. However, the pullback via
1829: radial parallel transport is technically simpler to introduce,
1830: hence it will be used here.
1831:
1832: For $R>0$ let $\Si_0(B_R(0),\geucl)$ (resp. $\Si_{s_k}(M,\overline g_k,\sigma)$)
1833: be the fiber of $\Si(B_R(0),\geucl)$ over $0$
1834: (resp. $\Si(M,\overline g_k,\sigma)$ over $s_k$).
1835: Let us define the radial vector
1836: field $X=r{\partial r}=\sum x^i\partial{x^i}$ on $B_R(0)\subset \mR^n$. Its
1837: length is the distance from $0$.
1838: For sufficiently large $k$, the exponential map of $(M,\overline g_k)$ based
1839: in $s_k$, denoted by $\exp^{\ti g_k}_{s_k}$ is a diffeomorphism from
1840: $B_R(0)$ onto its image.
1841: One chooses a (complex) linear isometry
1842: $\Si_0(B_R(0),\geucl)\to\Si_{s_k}(M,\overline g_k,\sigma)$.
1843: This map extends uniquely to a fiber preserving map
1844: $A:\Si(B_R(0),\geucl)\to\Si(M,\overline g_k,\sigma)$, such that
1845: $$\begin{matrix}
1846: %\matrix{
1847: \Si(B_R(0),\geucl)& \stackrel{A}{\longrightarrow} & \Si(M,\overline g_k,\sigma)\cr
1848: \downarrow & & \downarrow\cr
1849: B_R(0) & \stackrel{\exp^{\ti g_k}_{s_k}}{\longrightarrow} & M
1850: %}
1851: \end{matrix}
1852: $$
1853: commutes and such that
1854: $$A(\nabla_X \phi)= \nabla_{(\exp^{\ti g_k}_{s_k})_*(X)}A\phi.$$
1855: In a neighborhood of $s_k$, this condition can be equivalently
1856: characterized by saying that
1857: $\phi\mapsto A\circ \phi \circ (\exp^{\ti g_k}_{s_k})^{-1}$ maps
1858: parallel sections of $\Si(B_R(0),\geucl)$ to radially parallel
1859: sections of $\Si(B_R(s_k),\overline g_k,\si)$.
1860: One easily sees that $A$ is a fiberwise isometry, but the connection
1861: on the spinor bundle is not preserved. However, this will not matter, as
1862: in the limit $k\to \infty$, $R$ fixed, the connections converge in the
1863: $C^\infty$-topology.
1864:
1865: Let $k\geq k_0(R)$. On $(B_R(0),\geucl)$ we define the spinor
1866: $$\overline \phi_k:=A^{-1}\circ \witi\phi_k\circ \exp^{\ti g_k}_{s_k}$$
1867: and the operator $D_k:\Gamma(\Si(B_R(0),\geucl))\to\Gamma(\Si(B_R(0),\geucl))$,
1868: $$D_k:=A^{-1}\circ D_{\overline g_k}\circ A,$$
1869: where
1870: $D_{\ol g_k}$ is the Dirac operator on $(B_R(0),\overline g_k)$.
1871: We obtain
1872: $$D_k\overline{\phi}_k= {1\over \mu_{q_k}}\,|\overline{\phi}_k|^{p_k-2}\overline{\phi}_k$$
1873: Note that %with respect to $\overline g_k$
1874: $$\|\overline\phi_k\|_{C^0(B_R(0))}\leq |\overline\phi_k(0)|=1.$$
1875: Hence, we may apply the interior $L^p$- and the Schauder-estimates
1876: \ref{theo.int.lp} and \ref{theo.schauder}
1877: to conclude that
1878: $$\|\overline{\phi}_k\|_{C^{1,\al}(B_{R/2}(0))}\leq C(R),$$
1879: with constants $C(R)$ and $k(R)$.
1880: The constant $C(R)$ does not depend on $k$ if $\overline g_k$ is sufficiently
1881: $C^\infty$-close to $\geucl$. In particular, the $C^\infty$-convergence
1882: $\overline g_k\to \geucl$ says that $C(R)$ does not depend on
1883: $k$ for $k\geq k_1(R)$.
1884: %The constant $C(R)$ can be chosen such that it
1885: %depends only on $R$, $n$, $\al$ and the
1886: %$L^\infty$-norms of the Riemann tensor and its first derivative
1887: %in a sufficiently small neighborhood of $p$. However, the choice of
1888: %$k(R)$ also depends on the sequence $m_k$.
1889: % Note that the interior Schauder estimates do depend on the diameter
1890: % of the domain of definition.
1891:
1892: Compare $D_k$ with the Dirac operator $\Dflat$ on Euclidean $\mR^n$.
1893: (For similar and more explicit calculations the reader might consider
1894: e.g.\ \cite{pfaefflediss} for pullback via radial parallel transport,
1895: and \cite{ammann.humbert.morel:p03av2} for the other pullback method).
1896: We have
1897: $$\|(\Dflat -D_k)\overline\phi_k\|_{C^{0,\al}(B_{R/2}(0),\geucl)}\leq \tau_{k,R} \|\overline\phi_k\|_{C^{1,\al}(B_{R/2}(0),\overline g_k)},$$
1898: with $\lim_{k\to \infty}\tau_{k,R}=0$.
1899: The convergence is not uniform in $R$,
1900: but this will not matter in the following.
1901:
1902: We choose a sequence of radii $R_m\to \infty$.
1903: For each $R_m$, the Arcela-Ascoli theorem (Theorem~\ref{theo.arc.asc})
1904: allows us to choose a
1905: subsequence of $(\overline\phi_k)$ converging in
1906: $C^1(B_{R_m}(0),\geucl)$.
1907: After passing to a diagonal sequence, we see that there is a spinor
1908: $\overline{\phi}_\infty$ on $\mR^n$, such that
1909: $\res{\overline{\phi}_k}{B_{R}(0)}$
1910: converges to
1911: $\res{\overline{\phi}_\infty}{B_{R}(0)}\in C^1(B_R(0),\geucl)$
1912: for all $R>0$.
1913:
1914: Then $\overline\phi_\infty$ is a solution of
1915: $$\Dflat \overline\phi_\infty = {1\over\mu_\qD}\, |\overline\phi_\infty|^{\pD-2}
1916: \overline\phi_\infty$$
1917: on $\mR^n$.
1918:
1919: The estimate~\eref{ineq.unif.lp} says that
1920: $\|\witi\phi_k\|_{L^{p_k}(B_R(s_k),\witi g_k)}\leq 1$. Hence,
1921: for any $\ep>0$ and $R>0$ there is $k_2=k_2(R,\ep)$ such that
1922: $$\|\overline\phi_k\|_{L^{p_k}(B_R(0))}\leq 1 + \ep$$
1923: for all $k\geq k_2$.
1924: Because of the $C^1$-convergence $\overline\phi_k\to \overline\phi_\infty$,
1925: Fatou's lemma yields
1926: $$\|\overline\phi_\infty\|_{L^{\pD}(B_R(0))}\leq 1$$
1927: for any $R\in (0,\infty)$, and finally for $R=\infty$.
1928:
1929:
1930: We identify $\overline\phi_\infty$ via stereographic projection with an
1931: $L^\pD$-spinor $\widehat\phi_\infty$ on $\mS^n\setminus\{\mbox{South pole}\}$
1932: with the identification provided by the application $\witi F$ directly after
1933: Proposition~\ref{prop.conf.change}. We obtain
1934: \begin{eqnarray}\label{eq.Sn}
1935: D^{\mS^n} \widehat\phi_\infty = {1\over \mu_\qD}\, |\widehat\phi_\infty|^{\pD-2}
1936: \widehat\phi_\infty
1937: \end{eqnarray}
1938: and $\|\widehat \phi_\infty\|_{L^\pD}\leq 1$.
1939: The removal of singularities theorem, i.e.\ Theorem~\ref{theo.sing.rem},
1940: says that \eref{eq.Sn} holds on the whole sphere $\mS^n$.
1941:
1942:
1943: \begin{eqnarray*}
1944: \int_{\mS^n}\<D\widehat\phi_\infty,\widehat\phi_\infty\> &=&
1945: \mu_\qD^{-1}\,\|\widehat\phi_\infty\|_{L^\pD(\mS^n)}^\pD,\\
1946: \|D\widehat\phi_\infty\|_{L^\qD(\mS^n)}&=&
1947: \mu_\qD^{-1}\,\Bigl\|\,|\widehat\phi_\infty|^{\pD-1}\widehat\phi_\infty\Bigr\|_{L^\qD(\mS^n)}
1948: = \mu_\qD^{-1}\,\|\widehat\phi_\infty\|_{L^\pD(\mS^n)}^{\pD-1},\\
1949: \mu_\qD^{\mS^n}\geq \cF^{\mS^n}_\qD(\widehat\phi_\infty)&=&
1950: \mu_\qD\|\widehat\phi_\infty\|_{L^\pD(\mS^n)}^{2-\pD}\geq \mu_\qD
1951: \end{eqnarray*}
1952: which is apparently a contradiction to our assumption
1953: $\mu_\qD\geq\mu_\qD^{\mS^n}+\ep$.
1954: \qed
1955:
1956:
1957: \begin{proposition}
1958: If there is a $p_0<\pD$ and an $r>\pD$ such that for all
1959: $t\in (p_0,\pD)$ there is a solution $\phi_t$ of equation~\eref{eq.nonlin.p}
1960: with $p=t$, $1/q+1/p=1$ and such that $\|\phi_t\|_{L^{r}}$ is bounded
1961: by a constant $C$ independent from $t$,
1962: then there is a sequence $t_i\to \pD$ such that $\phi_{t_i}$ converges
1963: in the $C^1$-topology
1964: to a solution of equation~\eref{eq.nonlin.p} with $p=\pD$.
1965: \end{proposition}
1966:
1967: \proof{}
1968: For $p$ sufficiently close to $\pD$, we know because of
1969: Proposition~\ref{prop.func}
1970: that $\mu_t$ is bounded from below by a positive constant. Thus,
1971: we can apply the regularity theorem (Theorem~\ref{theo.reg}) which tells us
1972: that $(\phi_t)$ is uniformly bounded
1973: in $C^{1,\al}$.
1974: Hence, for a sequence $(t_i)$ with $t_i<\pD$,
1975: converging to $\pD$,
1976: the spinor fields $\phi_{t_i}$ converge in the $C^1$-topology to a
1977: $C^1$-spinor field $\phi_{\pD}$ which is a solution of
1978: equation~\eref{eq.nonlin.p} with $p=\pD$. \qed
1979:
1980:
1981:
1982:
1983:
1984: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1985: \section{Proof of the main theorem}\label{sec.proofmain}
1986: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1987:
1988: In this section we want to prove the main result of this publication, namely
1989: Theorem~\ref{theo.main}.
1990:
1991: \proof{of Theorem~\ref{theo.main}}
1992: Proposition~\ref{prop.sol.subcrit} tells us that
1993: for any $q\in (\qD,2)$, $\qD:=2n/(n+1)$, the functional $\cF_q$ is attained by a maximizer
1994: denoted $\phi_q$ satisfying
1995: %\begin{equation}%\label{eq.nonlin.p}
1996: $$
1997: \Dir\phi_q= \mu_q^{-1}\, |\phi_q|^{p-2} \phi_q,\qquad \phi_q\in H_1^q, \qquad
1998: \|\phi_q\|_{L^p}=1
1999: $$
2000: %\end{equation}
2001: where $p$ and $q$ are related via $p^{-1}+q^{-1}=1$.
2002: We have assumed that $\lammin(M,g,\si)< \lammin(\mS^n)$
2003: which is equivalent to $\mu_\qD>\mu_\qD^{\mS^n}$.
2004: As the function $q\mapsto \mu_q$ is continuous from the right
2005: (Proposition~\ref{prop.func}), we see that there is an $\ep>0$ such that
2006: $\mu_q>\mu_\qD^{\mS^n}+\ep$ for $q$ close to $\qD$.
2007: For such $q$, Theorem~\ref{theo.anydim} implies that $\phi_q$ are uniformly bounded in the $C^0$-norm,
2008: and then we can use Theorem~\ref{theo.reg} to conclude that these $\phi_q$ are even uniformly bounded
2009: in $C^{1,\al}$. The Theorem of Arcela-Ascoli
2010: (Theorem~\ref{theo.arc.asc})
2011: implies that there is there is a sequence $q_i\to \qD$ such
2012: that $\phi_{q_i}$ converges in the $C^1$-norm
2013: to a solution $\phi$ of
2014: %\begin{equation}%\label{eq.nonlin.p}
2015: $$
2016: \Dir\phi= \lammin\, |\phi|^{\pD-2} \phi,\qquad \phi\in C^1, \qquad
2017: \|\phi\|_{L^\pD}=1
2018: $$
2019: %\end{equation}
2020: Theorem~\ref{theo.reg} and Proposition~\ref{prop.imp} then show that $\phi$ has the desired regularity,
2021: and statement (A) is proven.
2022:
2023: We will now show that statement (B) follows from
2024: statement (A).
2025: If we have a solution as in (A), then we set $g_1:=f^{2/(n-1)}g_0$ with
2026: $f=\<\phi,\phi\>$. Note that $\vol(M,g_1)=\int |\phi|^{2n/(n-1)}=1$.
2027:
2028: The transformation formula for the Dirac operator under conformal changes
2029: (Proposition~\ref{prop.conf.change}) implies that there is a
2030: spinor $\phi_1$ on $(M,g_1,\si)$ such that
2031: %Then $\phi_1:=\phi/|\phi|$ is a bounded solution of
2032: $$D_{g_1}\phi_1=\la \phi_1,\qquad |\phi_1|_{g_1}\equiv 1.$$
2033: Then obviously, $\la^+_1(g_1) = \lammin$ and (B) follows.
2034:
2035: Statement (C) of Theorem~\ref{theo.main} will follow from
2036: Proposition~\ref{prop.zero.est},
2037: proven in the next section.
2038:
2039: %With Proposition~\ref{prop.lamu} we see that Theorem~\ref{theo.first}
2040: %is equivalent to Theorem~\ref{theo.main} (B).
2041:
2042: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2043: \section{The nodal set}\label{sec.degen}
2044: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2045:
2046: In this section we want to study that the zero set of solutions
2047: of the Euler-Lagrange equation
2048: \begin{equation}\label{eq.nonlin.p.nn}
2049: D\phi=c|\phi|^{p-2}\phi.
2050: \end{equation}
2051: Note that very often the zero set of a function $\phi$
2052: solving an equation of the above type is denoted as the
2053: \emph{nodal set of $\phi$}.
2054:
2055: The following theorem is due to C. B\"ar \cite{baer:97} for smooth
2056: $P$.
2057:
2058: \begin{theorem}[Nodal sets for Dirac Operators]
2059: Let $(U,g)$ be a Riemannian manifold and let $\phi$ be a solution of
2060: $$D\phi= P \cdot\phi$$
2061: where $P$ is a smooth function.
2062: Then the nodal set of $\phi$ has Hausdorff dimension at most $n-2$.
2063: \end{theorem}
2064:
2065: Unfortunately, \eref{eq.nonlin.p.nn} has not the desired form
2066: as $P=c|\phi|^{p-2}$ is not smooth for non-integer $p$, and
2067: B\"ar's proof does not extend to such a $P$.
2068: Nevertheless, we conjecture that the theorem also holds for
2069: $P=c|\phi|^{p-2}$, $p>2$.
2070: \begin{conjecture}
2071: The nodal set of any solution of
2072: \eref{eq.nonlin.p.nn} has Hausdorff dimension
2073: at most $n-2$.
2074: \end{conjecture}
2075:
2076: %Using another method, we prove the following theorem for arbitrary
2077: %locally bounded $p$ in \cite{ammann:nodal}.
2078: %
2079: %
2080: %The theorem tells us that the nodal set of
2081: %solutions of equation~\eref{eq.nonlin.p.nn} has Hausdorff dimension
2082: %$\leq n-2$.
2083:
2084: If $n=2$ and $p=\pD=4$, then
2085: we have better regularity.
2086: In this case solutions of equation~\eref{eq.nonlin.p.nn}
2087: and the corresponding $P=c|\phi|^2$ are smooth.
2088: Hence, using \cite[Main Theorem]{baer:97},
2089: one sees that in this case
2090: the nodal set of a solution is a discrete subset. The following proposition
2091: controls its cardinality.
2092:
2093:
2094: \begin{proposition}\label{prop.zero.est}
2095: On a compact spin surface $(M,g,\si)$ of genus $\gamma$
2096: let $\phi$ be a solution of equation
2097: $$D\phi= \lambda |\phi|^2\phi,\quad \|\phi\|_{L^4}=1.$$
2098: Then the number of zeros of $\phi$ is at most
2099: $\gamma-1 + {\lambda^2\over 4\pi}.$
2100: \end{proposition}
2101:
2102: In particular, this implies part (C) of Theorem~\ref{theo.main}.
2103: \proof{}
2104: We set $g_1:= |\phi|^4 g$. Then outside the zero set we know by
2105: Lemma~\ref{lem.scal.bound} that
2106: the Gauss curvature of $g_1$ is at most $\lambda^2$. Furthermore
2107: $\vol(M,g_1)=1$.
2108: Let $\phi(p)=0$. The integral
2109: of the geodesic curvature with respect to $g_1$
2110: over small simply closed loop around $p$ is close to $-2 (2j_p+1) \pi$,
2111: where $j_p$ is the order of the first non-vanishing term in the Taylor
2112: expansion of $\phi$ in $p$.
2113: We remove small open disks
2114: around the zeros of $\phi$ from $M$, and we obtain a surface with boundary
2115: $M'$.
2116: With the Gauss-Bonnet theorem we obtain
2117: $$2\pi\, \chi(M')=\int_{M'} K_{g_1} + \int_{\pa M'} k_{g_1}
2118: \leq \lambda^2-\sum (2j_p+1) 2\pi.$$
2119: And hence $2\pi\, (2-2\gamma)=2\pi\,\chi(M)\leq\lambda^2- 4\pi \sum j_p$,
2120: which implies the proposition.
2121: \qed
2122:
2123:
2124: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2125: \section{The spinorial Weierstrass representation}\label{sec.weierrep}
2126: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2127:
2128: The aim of this section is to recall the spinorial
2129: Weierstrass representation.
2130:
2131: Weierstrass published a representation of minimal surfaces
2132: in $\mR^3$ in terms of holomorphic functions
2133: \cite{weierstrass:66}.
2134: His article deals
2135: only with local questions, everything is described in a fixed
2136: conformal chart of the surface. From a modern (chart free)
2137: point of view, it is clear that
2138: these holomorphic sections should be interpreted as a section of
2139: the spinor bundle, and ``holomorphy'' translates into a ``harmonicity'',
2140: i.e.\ the surface is minimal iff the corresponding spinor $\phi$ satisfies
2141: $D\phi=0$.
2142: %However, this modern language was not available in the 19th century,
2143: %and the local statement already admitted strong applications.
2144:
2145: During the 20th century several attempts were undertaken to
2146: globalize the Weierstrass representation and to adapt it to arbitrary surfaces.
2147: Unfortunately, most approaches replaced Weierstrass' original
2148: approach by a formulation in terms of
2149: a holomorphic $1$-form and a holomorphic function.
2150: The corresponding formula were quite involved,
2151: and hence not very suitable for applications.
2152:
2153: An amazing breakthrough was achieved by work of
2154: D.~Sullivan, R.~Kusner, and N.~Schmitt around 1990,
2155: and independently by U.~Abresch.
2156: In early 1989, Dennis Sullivan put together some unpublished notes explaining
2157: the spinorial character of the Weierstrass representation for minimal
2158: surfaces. In spring 1989, Robert Kusner realized that the spinor
2159: formalism is not limited to minimal surfaces, but extends to conformal
2160: immersions of arbitrary surfaces, as described below. These techniques were
2161: presented in Sullivan's CUNY seminar in 1992, and
2162: spread around among the experts rapidly.
2163: Kusner's results found their continuation
2164: in the PhD thesis of Nick Schmitt \cite{schmitt:diss}.
2165: %, a PhD student of Kusner, the thesis was
2166: %published in 1993.
2167: Schmitt found many interesting applications of the spinorial Weierstrass.
2168: The results of Kusner and Schmitt led to the publications \cite{schmitt:diss}
2169: and \cite{kusner.schmitt:p96}.
2170:
2171: Independently, Abresch developed a spinorial Weierstrass for constant mean
2172: curvature surfaces.
2173: %, that can be understood as the special case $H=\const$
2174: %in Kusner's representation.
2175: Unfortunately,
2176: the resulting document of Abresch, some handwritten lecture notes from a
2177: conference in Luminy, were never published.
2178:
2179: We also want to mention an earlier result of Pinkall \cite{pinkall:85a}.
2180: In the special class of oriented surfaces, Pinkall's result
2181: establishes a bijection
2182: between regular homotopy classes of immersions of
2183: oriented surfaces $M$ and $\mZ_2$-valued quadratic forms on $H^1(M,\mZ_2)$.
2184: These quadratic forms are
2185: in bijection with the spin structures obtained in
2186: the spinorial Weierstrass representation.
2187:
2188: %In the following years many generalizations of the spinorial Weierstrass
2189: %representation were developed. We are aware of \cite{.....}, but we certainly
2190: %missed some important papers.
2191:
2192: \komment{
2193: The idea to use spinors for this representation seems natural
2194: when one reads Weierstrass' original article
2195: \cite{weierstrass:66}, but it is very amazing that
2196: the representation takes such a simple formula in terms of spinors.
2197: Unfortunately, the center of
2198: mathematicians' interest moved away from Weierstrass' original
2199: approach to a formulation in terms of
2200: a holomorphic $1$-form and a holomorphic function, which
2201: dominated the literature until the late 1980s.
2202:
2203: Then, R. Kusner and D. Sullivan found out
2204: that the spinor formulation gives a simple equation for
2205: non-minimal surfaces, and these ideas led to the preprint
2206: \cite{kusner.schmitt:p96}.
2207: Although this preprint was not
2208: published until now, this manuscript is the first document,
2209: where the spinorial Weierstrass representation was explained.
2210:
2211: It also has to be mentioned here that
2212: independently form Kusner, Sullivan
2213: and Schmitt, Abresch \cite{abresch:89} had already
2214: worked with another, less simple version of the spinorial Weierstrass
2215: representation before, but unfortunately,
2216: the resulting document, some handwritten lecture notes from a conference in
2217: Luminy, seem to be unavailable to the public.
2218: }
2219:
2220: More recent literature concerning this
2221: representation can be found for example in
2222: \cite{baer:98,friedrich:98,ammann:diss} and in articles by Pinkall, Taimanov,
2223: M.U. Schmidt, Morel, Voss and their collaborators. However, this list
2224: is far from being exhaustive.
2225:
2226: In our exposition we roughly follow \cite{kusner.schmitt:p96,friedrich:98}.
2227: The setting for the spinorial Weierstrass representation is as follows.
2228: Let $M$ be a compact Riemann surface of genus $\gamma$.
2229: The vector bundles
2230: $\Lambda^{1,0}T^* M$ and $\Lambda^{0,1}T^* M$ are defined
2231: as the complex linear part and the complex anti-linear part of
2232: $T^*M\otimes_\mR \mC$. The compositions
2233: $$I^{1,0}:T^*M\to T^*M\otimes_\mR \mC\to \Lambda^{1,0}M$$
2234: $$I^{0,1}:T^*M\to T^*M\otimes_\mR \mC\to \Lambda^{0,1}M$$
2235: of the complexification and the projection on $\Lambda^{1,0} M$
2236: resp.\ $\Lambda^{0,1} M$
2237: define vector space homomorphisms $T^*M\cong \Lambda^{1,0}M$ and
2238: $T^*M\cong \Lambda^{0,1}M$. These map $I^{1,0}$and $I^{0,1}$
2239: preserve the natural
2240: connections. However, one should pay attention
2241: to the fact, that these maps do not preserve lengths, but
2242: $2\Ree g(I^{1,0}(\al),I^{1,0}(\be))=2\Ree g(I^{0,1}(\al),I^{0,1}(\be))=g(\al,\be)$.
2243: The maps $2(I^{1,0})^{-1}$ resp.\ $2(I^{0,1})^{-1}$ is denoted
2244: as the \emph{real part}, namely $\Ree (\al)=2(I^{1,0})^{-1}(\al)$ for
2245: $\al\in\Ga(\Lambda^{1,0}M)$ and the same notation is used in the $(0,1)$ case.
2246: \emph{Complex conjugation} maps $\Lambda^{0,1}M$ to $\Lambda^{1,0}M$
2247: and vice versa.
2248:
2249: As the second Stiefel-Whitney-class of $M$ is the $\mod 2$ reduction
2250: of the Euler-class of $TM\to M$, one sees that
2251: $M$ is spin. However, the space of spin structures on $M$ is not unique:
2252: it is an affine space
2253: for the group $H^1(M,\mZ_2)=(\mZ_2)^{2\gamma}$, where $\gamma$ denotes
2254: the genus of $M$. Hence, there are $4^\gamma$ spin structures on $M$.
2255:
2256: If a spin structure is fixed, then the associated vector bundle with respect
2257: to the standard representation of $S^1$ on $\mC$ is a complex
2258: line bundle $\Si^+M$ satisfying $\Si^+M \otimes \Si^+M\cong \Lambda^{0,1}M$.
2259: If we equip $\Si^+M$ with the natural hermitian metric, we can choose this map such that
2260: the hermitian metric (res.\ the connection) on $\Lambda^{0,1}M$ is
2261: the tensor product metric (resp.\ tensor product connection).
2262:
2263: We define $\Si^-M$ to be $\Si^+M$ with the conjugated complex structure.
2264: In particular,
2265: there is a natural anti-linear conjugation map $\Si^+M\to \Si^-M$, the
2266: hermitian product defines a complex bilinear metric contraction
2267: $\Si^-M\otimes \Si^+M\to \mC$, and $\Si^-M\otimes \Si^-M\cong \La^{1,0}M$.
2268: In particular,
2269: $\Si^-M\otimes \Lambda^{0,1}M=\Si^-M\otimes \Si^+M \otimes \Si^+M= \Si^+M$, and hence
2270: the Dolbeault operator is a map $\overline\pa:\Gamma(\Si^-M)\to \Gamma(\Si^+M)$,
2271: and similarly $\overline\pa^*=-\partial: \Gamma(\Si^+M)\to \Gamma(\Si^-M)$.
2272:
2273: We define $c^{1,0}$ and $c^{0,1}$ as the compositions
2274: $$TM\stackrel{b}{\to}T^*M\stackrel{I^{1,0}}\to \Lambda^{1,0}M=\Si^-M\otimes \Si^-M=\Hom_\mC(\Si^+M,\Si^-M),$$
2275: $$TM\stackrel{b}{\to}T^*M\stackrel{I^{0,1}}\to \Lambda^{0,1}M=\Si^+M\otimes \Si^+M=\Hom_\mC(\Si^-M,\Si^+M).$$
2276: Composing $c^{1,0}$ with complex conjugation yields $c^{1,0}$ and vice versa.
2277: %and by inserting the complex conjugation $\Lambda^{0,1}M\to \Lambda^{1,0}M$, one obtains $c^{1,0}$
2278: %as the composition
2279: % $$TM\stackrel{b}{\to}T^*M\stackrel{I}\to\Lambda^{0,1}M\to \Lambda^{1,0}M=
2280: % \Si^-M\otimes \Si^-M=\Hom_\mC(\Si^+M,\Si^-M).$$
2281: One calculates
2282: $$c^{1,0}(X) c^{0,1}(Y)+c^{1,0}(Y) c^{0,1}(X)=g(\overline{I(X)},I(Y))+ g(\overline{I(Y)},I(X))=g(X,Y).$$
2283: As a consequence, the map
2284: $$TM\to\End\left(\Si^+M\oplus \Si^-M\right), \qquad X\mapsto
2285: \sqrt{2} \begin{pmatrix}
2286: 0 &c^{0,1}(X)\\-c^{1,0}(X) & 0
2287: \end{pmatrix}$$
2288: satisfies the Clifford relations.
2289: \longver{
2290: $$c^{0,1}(X)={1\over 2}\,(X^b- i J(X)^b)$$
2291: $$c^{1,0}(X)={1\over 2}\,(X^b+ i J(X)^b)$$
2292: Assume $e_1=\pa_x$, $e_2=J(e_1)=\pa_y$.
2293: We calculate on $\Si^-M$
2294: $$e_1\pa_{e_1}= {1\over \sqrt{2}}(dx-i\,dy){\pa\over \pa x}$$
2295: $$e_2\pa_{e_2}= {1\over \sqrt{2}}(dy+i\,dx){\pa\over \pa y}= {1\over \sqrt{2}}i(dx-i\,dy){\pa\over \pa y}$$
2296: $$D = \sqrt{2} \,{dx- i\, dy\over 2}\left({\pa\over \pa x}+ i\,{\pa\over \pa y}\right)=
2297: \sqrt{2} \,\pa_{\ol z} \,d{\ol z}=\sqrt{2}\,\ol{\pa}$$
2298: }
2299:
2300: One sees, that
2301: the sum $\Si M:=\Sigma^+M\oplus \Sigma^-M$ can be identified with the standard spinor bundle on $M$ in
2302: such a way that the above map is the Clifford multiplication, and such that
2303: $\Si^+M$ resp.\ $\Si^-M$ are the positive resp.\ negative half-spinors.
2304:
2305: The Dirac operator can be written in this notation as
2306: $$D= \sqrt{2}
2307: \begin{pmatrix}
2308: 0 &\overline\partial \\-\partial & 0
2309: \end{pmatrix}:
2310: \Gamma\Big(\Sigma^+\oplus \Sigma^-\Big)\to
2311: \Gamma\Big(\Sigma^+\oplus \Sigma^-\Big).$$
2312:
2313: Now let us assume that
2314: $\phi=(\phi_+,\phi_-)$ is a solution of~$D\phi=H|\phi|^2\phi$, where
2315: $H$ is a real-valued function on $M$.
2316: This means
2317: $$-\sqrt{2}\pa \phi_+ = H |\phi|^2\phi_-$$
2318: $$\sqrt{2}\pa \ol{\phi_-} = H |\phi|^2\ol{\phi_+}$$
2319: We define
2320:
2321:
2322: $$ \alpha := %\mathrm{Re}
2323: {\sqrt{2}}\,
2324: \begin{pmatrix}
2325: \phi_+\otimes\phi_+ \,+\,\overline{\phi_-}\otimes \overline{\phi_-}\cr
2326: i \phi_+\otimes \phi_+\,-i\,\overline{\phi_-}\otimes \overline{\phi_-}\cr
2327: 2 i\phi_+\otimes \overline{\phi_-}
2328: \end{pmatrix}\in\Gamma(\Lambda^{0,1}\otimes_\mR \mR^3).
2329: $$
2330: \longver{
2331: $$\pa \al= 2 i H |\phi|^2
2332: \begin{pmatrix}
2333: -2\Imm (\phi_+\otimes \phi_-) \cr
2334: -2\Ree (\phi_+\otimes \phi_-) \cr
2335: |\phi|^2
2336: \end{pmatrix}\in\Gamma(i\mR^3).$$
2337: }
2338:
2339: Let $\witi M$ denote the universal covering of $M$, and $\pi_1(M)$ the group
2340: of Deck transformations.
2341:
2342: As $\pa\al=d\al$ is imaginary, we can find a function $F:\witi M\to \mR^3$,
2343: such that $dF=\Ree \alpha$, and there is a homomorphism
2344: ${V:\pi_1(M)\to \mR^3}$, such that
2345: $$F(p\cdot\gamma)= F(p) + V(\gamma)\quad \forall p\in \witi M,
2346: \;\gamma \in \pi_1(M)$$
2347:
2348: One calculates that
2349: $F$ is a conformal map with possible branching points,
2350: $$|dF|=|\Ree \alpha|={1\over \sqrt{2}}\,|\alpha|=|\phi|^2,$$
2351: and that $F(M)$ has mean curvature $H$.
2352:
2353: Hence, the map $F$ satisfies Properties (1) to (3) from the introduction,
2354: i.e. it is a \emph{periodic branched
2355: conformal immersion $F$ based on $(M,g)$} with mean curvature $H$.
2356:
2357: In any zero of of the spinor $\phi$, the map $F$ has a branching point.
2358: If $F$ vanishes of order $k$, then $\al$ vanishes of order $2k$.
2359: Hence, all branching points of $F$ are necessarily of even order.
2360:
2361: Summarizing the above statement, we obtain for any solution of
2362: $D\phi=H\,|\phi|^2\phi$
2363: a periodic branched conformal immersion of $\witi M$ into $\mR^3$
2364: which is uniquely
2365: determined up to translation. If $\phi$ solves $D\phi=H\,|\phi|^2\phi$,
2366: then $-\phi$ as well, and the corresponding $F$ is the same. Hence, we obtain
2367: a well-defined map
2368: $$\left\{\vbox{\hbox{solutions of}\vtop{\hbox{$D\phi=H\;|\phi|^2\phi$}%
2369: \hbox{on $M$}}}\right\}/{\pm 1}
2370: \quad\longrightarrow\quad
2371: %$$
2372: %$$\stackrel{1:1}{\longleftrightarrow}$$
2373: %$$
2374: \left\{\vbox{\hbox{conformal periodic $H$-immersions}\vtop{\hbox{
2375: $\witi M\to\mR^3$ %($S^3$, $H^3$)
2376: with branching}
2377: \hbox{points of even order}}}\right\}/\vtop{\mbox{trans-}\\\mbox{lations}}$$
2378: and one can show that this map is even a bijection.
2379:
2380: The inverse of this map is given by restricting a parallel spinor
2381: on $\mR^3$ to $F(M)$ and
2382: by performing a conformal change \cite{baer:98}.
2383:
2384: If $H$ is constant, then
2385: there is also another version of the spinorial Weierstrass
2386: representation, where the target
2387: space is $S^3$ instead of $\mR^3$. We view $S^3$ as $\SU(2)$
2388: with a bi-invariant metric of constant curvature $1$,
2389: the multiplication in $\SU(2)$ is denoted with
2390: $\bullet$. The periodicity condition (1) has to be replaced
2391: by
2392: \begin{enumerate}[{\rm (1')}]
2393: \item {\it Left periodicity:}
2394: There is a homomorphism
2395: $h:\pi(M)\to \SU(2)$, the \emph{periodicity map}, such that
2396: for any $\gamma\in \pi_1(M)$, and $x\in\witi M$ one has
2397: $$F(x\cdot \gamma)=h(\gamma)\bullet F(x).$$
2398: Here $\cdot$ denotes the action of $\pi_1$ on $\witi M$ via Deck transformation.\end{enumerate}
2399:
2400: One obtains a bijection
2401: $$\left\{\vbox{\hbox{solutions of}\vtop{\hbox{$D\phi=c\;|\phi|^2\phi$}%
2402: \hbox{on $M$}}}\right\}/{\pm 1}
2403: \quad\longrightarrow\quad
2404: %$$
2405: %$$\stackrel{1:1}{\longleftrightarrow}$$
2406: %$$
2407: \left\{\vbox{\hbox{conformal left periodic $H$-im-}\vtop{\hbox{
2408: mersions $\witi M\to\SU(2)$ %($S^3$, $H^3$)
2409: with}
2410: \hbox{branching points of even order}}}\right\}/\vtop{\mbox{Left mul-}\\\mbox{tiplication}}$$
2411: where $c=\sqrt{H^2+1}$.
2412: For details see
2413: \cite{voss:dipl,morel:02,ammann:habil}.
2414:
2415: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2416: \section{Applications to constant mean curvature surfaces}\label{sec.app.cmc}
2417: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2418:
2419: If the dimension of $M$ is $2$, then \eref{eq.dirac.nonlin} reads as
2420: \begin{equation}\label{eq.nonlin.2}
2421: \Dir\phi=\lammin\, |\phi|^{2} \phi,\qquad \phi\in H_1^{4/3}, \qquad
2422: \|\phi\|_{L^4}=1
2423: \end{equation}
2424: and according to the regularity theory $\phi$ is even smooth.
2425:
2426: The spinorial Weierstrass representation explained in the previous section
2427: tells us, that such a solution can be used to construct
2428: certain immersions with constant mean curvature.
2429:
2430: Combining the previous results we obtain the following application
2431: that is a stronger version of the
2432: ``Principle for construction of cmc-surfaces''
2433: mentioned in the introduction.
2434:
2435: \begin{proposition}
2436: Assume that the Riemann spin surface $(M,g,\si)$ satisfies
2437: \begin{equation}\label{ineq.surface}
2438: \lammin(M,[g],\si)<2\pi.
2439: \end{equation}
2440: Then there is a periodic branched
2441: conformal cmc immersion $F:\witi M\to \mR^3$ based on $(M,g)$.
2442: The mean curvature is equal to $\lammin(M,[g],\si)$ and the area of a
2443: fundamental domain is $1$.
2444: The regular homotopy class of $F$ is determined by the spin-structure $\si$.
2445: The indices of all branching points are even,
2446: and the sum of these indices
2447: is smaller than $2 {\rm genus}(M)$. In particular,
2448: if $M$ is a torus, there are no branching points.
2449: \end{proposition}
2450:
2451:
2452: The proof is a direct consequence of
2453: Theorem~\ref{theo.main} and the previous section.
2454:
2455: There are many examples of stationary points of the functional.
2456: However, it is still open whether they are the maximizers or not.
2457: Note, that by changing the orientation of a surfaces,
2458: a maximizers $\psi$ of $\cF_\qD$
2459: on $M$ turns into a minimizer $\psi'$ of $\cF_\qD$ on the surface with
2460: reversed orientation $M'$, with $\cF^{M'}_\qD(\psi')=-\cF^{M}_\qD(\psi)$.
2461: Let us study some examples.
2462:
2463: \begin{examples}
2464: \begin{enumerate}[{\rm (a)}]
2465: \item
2466: Let $(M,g)$ be a $2$-dimensional torus. Via a conformal change
2467: we can achieve that $g$ is flat, i.e.\ $M=\mR^2/\Gamma$, equipped
2468: with the Euclidean metric.
2469: We assume that the lattice $\Gamma$ is generated by
2470: $\begin{pmatrix}1\cr 0\end{pmatrix}$ and
2471: $\begin{pmatrix}x\cr y\end{pmatrix}$, with $y>0$.
2472: The spinor bundle of a flat manifold is flat as well, hence the holonomy
2473: is a map $\Gamma\to \SU(\Si_p M)$. Indeed, the image of this map is contained
2474: in $\{\pm \Id\}$.
2475: We obtain a homomorphism $\chi:\Gamma\to \{\pm \Id\}$.
2476: This homomorphisms
2477: characterizes the spin structure $\si$ in the sense that two spin
2478: structures on $(M,g)$ are isomorphic iff the homomorphisms $\chi$
2479: coincide, and to each such homomorphism there is a spin structure.
2480: The case $\chi\equiv+\Id$ corresponds to the so-called \emph{trivial}
2481: spin structure $\si_{\mathrm{tr}}$, the other cases correspond to
2482: \emph{non-trivial}
2483: spin structures.
2484: \footnote{Note that this notation is a bit misleading, as it is only the
2485: \emph{trivial} spin structure on $M$ that defines a \emph{non-trivial}
2486: element in the bordism class.}
2487:
2488: At first, we deal with the case $\si=\si_{\mathrm{tr}}$.
2489: In this case, after a possible rotation and a possible homothety,
2490: we can achieve
2491: $$|x|\leq {1\over 2},\quad
2492: y^2+x^2\geq 1,\quad y> 0.$$
2493: On $(M,g,\si_{\mathrm{tr}})$ the kernel of $\Dir$ has complex dimension $2$,
2494: and consists of parallel spinors. If one carries out the constructions
2495: from the last sections for a parallel spinor, then one obtains
2496: an affine conformal map $F$. Such an $F$ is trivially
2497: a periodic conformal immersion $F$ based on $(M,g)$
2498: with \emph{vanishing} mean curvature.
2499: However, if $y>\pi$ then one easily sees that
2500: $\lammin(M,g,\si_{\mathrm{tr}})<2\pi$. The proposition
2501: yields the existence of a periodic conformal immersion $F$
2502: based on $(M,g)$ with constant mean curvature $\lammin(M,g,\si_{\mathrm{tr}})$,
2503: and the area of a fundamental domain is $1$.
2504: However, the proposition only provides the existence of the solution, but
2505: we cannot characterize the maximizer.
2506: In our example, a family of solutions can be explicitly written, namely
2507: it is the family of parameterized cylinders
2508: \begin{eqnarray*}
2509: F:\mR^2&\to &\mR^3\\
2510: \begin{pmatrix} a\cr b\end{pmatrix}& \mapsto &
2511: P\begin{pmatrix}{\sqrt{y}\over 4\pi}\cos{4\pi b\over y}\cr{\sqrt{y}\over 4\pi}\sin{4\pi b\over y}\cr {a\over \sqrt{y}} \end{pmatrix}+ X_0
2512: \end{eqnarray*}
2513: for any $P\in O(3)$, $X_0\in \mR^3$. We conjecture that these solutions are
2514: exactly those that correspond to the maximizers and the minimizers of
2515: $\cF_\qD$, when we normalize such that all spinors have $L^4$-norm $1$.
2516:
2517: In the case $\si\not=\si_{\mathrm{tr}}$
2518: we can achieve
2519: that
2520: $$\chi\begin{pmatrix}1\cr 0\end{pmatrix}=\Id\qquad \chi\begin{pmatrix}1\cr 0\end{pmatrix}=-\Id,$$
2521: $$|x|\leq {1\over 2},\quad
2522: y^2+\left(|x|-{1\over 2}\right)^2\geq {1\over 4},\quad y> 0.$$
2523: The Dirac operator is always invertible.
2524:
2525: One easily sees $\lammin(M,g,\si)\leq {\pi\over \sqrt{y}}$.
2526: Hence, the proposition yields solutions for $y>{4\over \pi}$.
2527: Once again, solutions can be explicitly written, namely
2528: the parameterized cylinder
2529: \begin{eqnarray*}
2530: F:\mR^2&\to &\mR^3\\
2531: \begin{pmatrix} a\cr b\end{pmatrix}& \mapsto &
2532: P\begin{pmatrix}{\sqrt{y}\over 2\pi}\cos{2\pi b\over y}\cr{\sqrt{y}\over 2\pi}\sin{2\pi b\over y}\cr {a\over \sqrt{y}} \end{pmatrix}+ X_0
2533: \end{eqnarray*}
2534: for any $P\in O(3)$, $X_0\in \mR^3$.
2535: However, in some cases, e.g.\ if $x=0$ and $4/\pi<y<1$,
2536: these solutions no longer correspond to maximizers and minimizers,
2537: but to saddle points of the functional. We conjecture
2538: that in the case $x=0$, $y<1$ the maximizers and minimizers
2539: correspond to the unduloid immersions (see Figure~\ref{fig.undu}).
2540: An unduloid is a surface of revolution of constant mean curvature.
2541: \begin{figure}
2542: \begin{center}
2543: \bild{
2544: \includegraphics[scale=1]{unduloid1.eps}
2545: %\includegraphics[scale=.5]{unduloid2.eps}
2546: }
2547: \end{center}
2548: \caption{An unduloid in $\mR^3$, visualized by Nick Schmitt.}\label{fig.undu}
2549: \end{figure}
2550: \item If $M$ has genus $2$, then as in the case of the torus,
2551: the dimension of the kernel is independent of the metric, however it depends
2552: on the spin structure.
2553: If $\si$ is a spin structure such that
2554: $(M,\si)$ is spin-cobordant $0$, then the Dirac operator
2555: is invertible for any metric. Again, as in the torus case, one can find
2556: for any $\ep>0$ a conformal classes
2557: ${g}$ on $M$ with
2558: $\lammin(M,[g],\si)<\ep$. \cite{ammann.humbert:gluing}
2559: \item If the genus is larger than $2$, then the kernel of the Dirac operator
2560: on a Riemannian spin manifold~$(M,\si)$
2561: depends on the metric. For example if $M$ is
2562: a surface of genus $3$ equipped with the spin structure $\si$ and the conformal
2563: structure ${g_0}$ associated to the periodic conformal immersion with
2564: vanishing mean curvature indicated in Figure~\ref{fig.kgb.min}.
2565: \footnote{N. Schmitt's illustrations in this paragraph are available
2566: on his website \url{http://www.gang.umass.edu/gallery/cmc/cmcgallery0302.html},
2567: K. Grosse-Brauckmann's illustrations are available on
2568: \url{http://www.math.uni-bonn.de/people/kgb/Research/folie\_{}iwp.gif}}
2569: This immersion induces a harmonic spinor on $(M,g,\si)$. However, as
2570: $(M,\si)$ is spin-cobordant $0$, there is a perturbation $[g_t]$
2571: of the conformal structure such that the Dirac operator on $(M,g_t,\si)$
2572: has a trivial kernel for small $t\neq 0$ \cite{maier:97}. In this case
2573: $$\lim_{t\to 0\atop t\neq 0}\lammin(M,[g_t],\si)=0,$$
2574: hence there exist
2575: solutions of~\eref{eq.nonlin.2}. Such a solution is visualized in
2576: Figure~\ref{fig.kgb.cmc}.
2577: \begin{figure}
2578: \begin{center}
2579: \bild{\includegraphics[scale=.5]{kgb_min.ps}}
2580: \end{center}
2581: \caption{A periodic branched conformal minimal surface%
2582: % based on a Riemann surface of genus $4$
2583: , visualized by K.~Grosse-Brauckmann}\label{fig.kgb.min}
2584: \end{figure}
2585:
2586: \begin{figure}
2587: \begin{center}
2588: \bild{\includegraphics[scale=.5]{kgb_cmc.ps}}
2589: \end{center}
2590: \caption{A periodic branched conformal cmc surface%
2591: % based on a Riemann surface of genus $4$
2592: , visualized by K.~Grosse-Brauckmann}\label{fig.kgb.cmc}
2593: \end{figure}
2594: \item Constant mean curvature immersions of $T^2$ into $\mR^3$,
2595: in particular Wente tori and twisty tori (Figure~\ref{fig.twisty})
2596: also correspond to
2597: stationary points of $\cF_\qD$. However, as they are not embedded
2598: \cite{li.yau:82} tells us that $\int H^2 \geq 8\pi$.
2599: On the other hand maximizers and minimizers of $\cF_\qD$ satisfy
2600: $\int H^2= \lammin^2\leq 4\pi$, hence these tori do \emph{not}
2601: correspond to maximizers or minimizers.
2602: \begin{figure}
2603: \begin{center}
2604: \bild{
2605: \includegraphics[scale=1]{twisty-torus4.eps}
2606: }\end{center}
2607: \caption{A twisty torus, a cmc immersed torus in $\mR^3$,
2608: visualized by Nick Schmitt}\label{fig.twisty}
2609: \end{figure}
2610: \end{enumerate}
2611: \end{examples}
2612:
2613: Similar propositions also hold for immersions into $S^3$ and
2614: into hyperbolic space $H^3$, see \cite{ammann:habil}. We will only
2615: specify the case of $S^3=\SU(2)$.
2616:
2617: \begin{proposition}
2618: Assume that the Riemann spin surface $(M,g,\si)$ satisfies
2619: \begin{equation}
2620: \lammin(M,[g],\si)<2\pi.
2621: \end{equation}
2622: Let $a\in(0,\lammin(M,[g],\si)$ be given.
2623: Then there is a left periodic branched
2624: conformal cmc immersion $F:\witi M\to \SU(2)$ based on $(M,g)$.
2625: The mean curvature is equal to $H=\sqrt{(\lammin(M,[g],\si)/a)^2-1}$,
2626: and the area of a
2627: fundamental domain is $a^2$.
2628: The regular homotopy class of $F$ is determined by the spin-structure $\si$.
2629: The indices of all branching points are even,
2630: and the sum of these indices
2631: is smaller than $2 {\rm genus}(M)$. In particular,
2632: if $M$ is a torus, there are no branching points.
2633: \end{proposition}
2634:
2635: An example where the image of the periodicity map has a finite image in
2636: $\SU(2)$ is given in Figure~\ref{fig.undu.s}.
2637:
2638: \begin{figure}
2639: \begin{center}
2640: \bild{
2641: \includegraphics[scale=.3]{torus_undu_sphere.eps}
2642: }\end{center}
2643: \caption{An unduloid in $S^3$, visualized by Nick Schmitt.}\label{fig.undu.s}
2644: \end{figure}
2645:
2646:
2647: \begin{remark}
2648: If $(M,g)$ is an analytic Riemannian manifold of dimension $3$, then there
2649: is an analogue of the spinorial Weierstrass representation. This implies
2650: that solutions $\phi$ of \eref{eq.nonlin.p}
2651: can be geometrically interpreted as
2652: a conformal cmc embedding of $M\setminus \phi^{-1}(0)$
2653: into a (non-complete) $4$-dimensional Riemannian manifold $(N,h)$.
2654: This manifold $(N,h)$ carries a parallel spinor whose restriction
2655: to $(M,g)$ is again the solution of~\eref{eq.nonlin.p}. The manifold
2656: $(N,h)$ depends on $\phi$ and is unique up to restriction to subsets
2657: and coverings. A more detailed exposition of this $3$-dimensional version
2658: is work in progress.
2659: \end{remark}
2660:
2661:
2662: \begin{appendix}
2663:
2664: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2665: \section{Elliptic regularity}\label{sec.regularity}
2666: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2667:
2668:
2669: In this section we want to collect some facts about elliptic regularity
2670: for Dirac operators. The proofs of these statements are
2671: analogous to proofs of the corresponding statements for the Laplace operators
2672: as done e.g.\ in \cite{gilbarg.trudinger:77} and \cite{adams:75}.
2673: Details on how to prove most statements of this section
2674: are given also in \cite{ammann:habil}. Either a proof is provided there or
2675: it is sketched how the statements are reduced to standard theorems.
2676:
2677: Let $(N,g)$ be a Riemannian manifold, possibly with boundary,
2678: with a spin structure~$\si$. The interior of $N$ is denoted as $N_0$.
2679: The spinor bundle $\Si(N,g,\si)$ is a complex vector bundle
2680: carrying a natural connection and a natural hermitian metric.
2681:
2682: We now define the Sobolev norms of spinors fields.
2683:
2684: \begin{definition}[Sobolev spaces]
2685: For any spinor $\psi$ smooth on $N$,
2686: and $q\in (1,\infty)$, $k\in \mN\cup\{0\}$,
2687: we define the $H^q_k$-norm of $\psi$ as
2688: $$\|\psi\|_{H^q_k(N,g,\si)}
2689: = \sum_{l=0}^k\,\Biggl(\int_N\Bigl|\underbrace{\na \ldots \na}_{l-\mbox{times}}\psi\Bigr|^q\,\dvol_g\Biggr)^{1/q}.$$
2690: When the domain of integration is clear, we simply write
2691: $\|\psi\|_{H^q_k}$ instead of $\|\psi\|_{H^q_k(N,g,\si)}$.
2692: The closure of the space of
2693: smooth spinors with respect to this norms is denoted as $H^q_k(\Si(N,g,\si))$.
2694: If the underlying structure of Riemannian spin manifold
2695: is clear from the context, we often write shortly
2696: $H^q_k$, or in other situations where we want to emphasize the metric
2697: we write $H^q_k(g)$. We will also write
2698: $L^q$ for $H^q_0$.
2699: \end{definition}
2700:
2701: \begin{theorema}[Interior $L^p$ estimates]\label{theo.int.lp}
2702: Let $(N,g,\si)$ be a compact Riemannian spin
2703: manifold, possibly \emph{with boundary}. Let $K$ be a compact
2704: subset of $N_0$.
2705: Let $\psi$ be in $H_k^q(\Si(N,g,\si))$ and let $\phi\in H_1^1(\Si(N,g,\si))$
2706: be a weak solution (i.e.\ in the sense of distributions) of
2707: $$D\phi = \psi$$
2708: on $N_0$.
2709: Then $\phi|_K\in H_{k+1}^q(\Si(K,g,\si))$ and
2710: $$\|\phi\|_{H_{k+1}^q(\Si(K,g,\si))}\leq C\cdot
2711: \left(\|\psi\|_{H_k^q(\Si(N,g,\si))} + \|\phi\|_{L^q(\Si(N,g,\si))}\right),$$
2712: where $C=C(N,g,\si,K,k,q)$, i.e.\ $C$ depends only on $N,g,\si,K,k$ and $q$.
2713:
2714: Furthermore,
2715: if a sequence of metrics $(g_i)_{i\in \mN}$
2716: converges in the $C^\infty$-topology
2717: to a Riemannian metric $g$, then the constants $C$ can be chosen
2718: such that
2719: $$\sup_{i\in \mN} C(N,g_i,\si,K,k)<\infty.$$
2720: \end{theorema}
2721: %\cite[Theorem~7.63]{adams:75}
2722:
2723: In order to prove the theorem, it is sufficient to prove
2724: it in the case that $N$ and $K$ are small concentric geodesic balls, the radius
2725: being controlled in terms of curvature bounds, and bounds on derivatives
2726: of the curvature. As explained in \cite{ammann:habil}
2727: this can be done in a way analogous to the corresponding statement
2728: for the Laplacian in \cite{gilbarg.trudinger:77}.
2729: The general statement then can be reduced to the special
2730: case by covering (the general) $K$ by finitely many small open balls and an
2731: associated partition of unity.
2732:
2733: In the special case that the boundary of $N$ is the empty set,
2734: and that $K=N=N_0$,
2735: then there is a stronger version, that will be used as well.
2736:
2737: \begin{theorema}[Global $L^p$ estimates]\label{theo.global.lp}
2738: Let $(N,g,\si)$ be a compact Riemannian spin manifold
2739: \emph{without boundary} and $\psi\in H_k^q(\Si(N,g,\si))$.
2740: Then any weak solution $\phi\in H_1^1(\Si(N,g,\si))$ of
2741: $$D\phi = \psi$$
2742: satisfies $\phi\in H_{k+1}^q(\Si(N,g,\si))$, and there is a constant $C=C(N,g,\si)$
2743: such that
2744: $$\|\phi\|_{H_{k+1}^q}\leq C
2745: \left(\|\psi\|_{H_{k}^q}+\|\pi_{\mathop{\rm ker}(D)}(\phi)\|_{L^q}\right)$$
2746: where $\pi_{\mathop{\rm ker}(D)}$ is the $L^2$-orthogonal projection
2747: to the kernel of $D$.
2748: \end{theorema}
2749:
2750: The proof of this theorem is not difficult if one uses some facts
2751: about pseudodifferential operators, explained for example in
2752: \cite{taylor:81}. We abbreviate $\pi:=\pi_{\mathop{\rm ker}(D)}$.
2753: The spectrum of the elliptic operator
2754: $D+\pi$ is bounded away from $0$,
2755: hence it is invertible and its inverse is a pseudodifferential operator
2756: of degree $-1$. Pseudodifferential operators of degree $-1$
2757: are continuous from $L^q$ to $H_1^q$. Hence,
2758: $$\|\phi\|_{H_1^q}=\|(D+\pi)^{-1}(D+\pi)\phi\|_{H_1^q}\leq C\|(D+\pi)\phi\|_{L^q}\leq C\left(\|D\phi\|_{L^q}+\|\pi(\phi)\|_{L^q}\right).$$
2759: This is the statement of the theorem in the case $k=0$, and the statement
2760: for $k>0$ follows for example by using Theorem~\ref{theo.int.lp}.
2761:
2762: However, a uniformity statement for converging $g_i\to g$ as in
2763: Theorem~\ref{theo.int.lp} does not hold. If
2764: $\dim \ker D_g >\limsup_i\dim \ker D_{g_i}$, then one easily shows that
2765: there are eigenspinors $\phi_i$ for $D_{g_i}$ to eigenvalues
2766: $\la_i\to 0$, $\la_i\neq 0$ with
2767: $${\|D\phi_i\|_{L^q(g_i)}+\|\pi_{g_i}(\phi_i)\|_{L^q(g_i)}\over \|\phi_i\|_{H_1^q(g_i)}}=\la_i,$$
2768: which would contradict a uniform version of Theorem~\ref{theo.global.lp}.
2769:
2770: We also needs some facts about H\"older norms.
2771:
2772: \begin{definition}[H\"older spaces]
2773: Let $\al\in (0,1]$.
2774: On $C^\infty(\Si(N,g,\si))$ we define the
2775: \emph{H\"older norms}
2776: \begin{eqnarray*}
2777: \|\phi\|_{C^{0,\al}} & := & \mbox{h\"ol}_\al(\phi)\\
2778: \|\phi\|_{C^{1,\al}} & := & \|\phi\|_{C^0} + \mbox{h\"ol}_\al(\nabla \phi)\\
2779: \mbox{h\"ol}_\al(Q)& := & \sup \Bigl\{{|Q(x)-P_\ga Q(y)|\over
2780: d(x,y)^\al}\;|\; x,y\in M, x\neq y, P_\ga \mbox{ is the }\\
2781: && \mbox{parallel transport along a shortest geodesic $\ga$ from $x$ to $y$}.\Bigr\}
2782: \end{eqnarray*}
2783: If the shortest geodesic is not unique, then we use the convention that
2784: the supremum runs over all possible choices of shortest geodesics.
2785:
2786: The completions of $C^\infty(\Si(N,g,\si))$ with respect to the
2787: H\"older norms $C^{0,\al}$ and $C^{1,\al}$ define the
2788: \emph{H\"older spaces}
2789: $C^{0,\al}=C^{0,\al}(\Si(N,g,\si))$ and
2790: $C^{1,\al}=C^{1,\al}(\Si(N,g,\si))$.
2791: \end{definition}
2792:
2793: H\"older norms are important as they admit Schauder estimates.
2794:
2795: \begin{theorema}[Schauder estimates]\label{theo.schauder}
2796: Let $(N,g,\si)$ be a compact Riemannian spin manifold, possibly
2797: \emph{with boundary}, and let $K$ be a compact subset of $N_0$.
2798: Suppose $\psi\in C^{k,\al}(\Si(N,g,\si))$, $k\in \mN\cup\{0\}$.
2799: Then for any weak solution $\phi\in L^1 (\Si(N,g,\si))$ of
2800: $$D\phi = \psi$$
2801: we have $\phi|_K\in C^{k+1,\al}(\Si(K,g,\si))$ and
2802: $$\|\phi\|_{C^{k+1,\al}(\Si(K,g,\si))}\leq
2803: C \cdot (\|\psi\|_{C^{k,\al}(\Si(N,g,\si))}
2804: +\|\phi\|_{C^0(\Si(N,g,\si)})$$
2805: where $C=C(N,g,\si,K,\al)$.
2806:
2807: Furthermore,
2808: if a sequence of metrics $(g_i)_{i\in \mN}$
2809: converges in the $C^\infty$-topology
2810: to a Riemannian metric $g$, then the constants $C$ can be chosen
2811: such that
2812: $$\sup_{i\in \mN} C(N,g_i,\si,K,k)<\infty.$$
2813: \end{theorema}
2814:
2815: As before, one important special case is that $N$ has empty boundary
2816: and $K=N=N_0$.
2817:
2818: The proof can be done in a way analogous
2819: to the proof of the corresponding statements for the Laplacian in
2820: \cite{gilbarg.trudinger:77}. Again it is sufficient to prove it for
2821: small concentric geodesic balls, and to glue them together.
2822: We omit the details. For a proof for concentric geodesic
2823: balls is provided for example in \cite{ammann:habil}.
2824: % where a version
2825: %with a more general uniformity statement is proven.
2826:
2827: In local charts one easily reduces the following theorem to the standard
2828: Arcela-Ascoli theorem.
2829:
2830: \begin{theorema}[Arcela-Ascoli] \label{theo.arc.asc}
2831: Let $(N,g,\si)$ be a Riemannian spin
2832: manifold, possibly \emph{with boundary}.
2833: For $m\in \mN\cup \{0\}$, the inclusion
2834: $C^{m,\al}(\Si(N,g,\si))\to C^m(\Si(N,g,\si))$ is compact, i.e.\
2835: a bounded sequence in $C^{m,\al}(\Si(N,g,\si))$ has a subsequence
2836: convergent in $C^m(\Si(N,g,\si))$.
2837: \end{theorema}
2838:
2839:
2840: Sobolev and H\"older spaces are related by a several embedding theorems,
2841: some of them are compact, others are only bounded. We summarize the embeddings
2842: that are needed in the article.
2843: %We will only need the Sobolev embeddings and the Rellich-Kondrakov theorem
2844: %in the global version, i.e.\ for compact manifolds \emph{without boundary}.
2845:
2846: \begin{theorema}[Embedding theorems]\label{theo.sobo}\
2847: Let $k,s\in \mN\cup\{0\}$, $k\geq s$ and $q,r\in (1,\infty)$.
2848: Let $(M,g,\si)$ be a compact Riemannian spin manifold \emph{without boundary}.
2849: All spaces of functions are defined on sections of $\Si(M,g,\si)$.
2850: \begin{enumerate}[{\rm (a)}]
2851: \item {\rm (Sobolev embedding theorem I).} If
2852: %\begin{equation}\label{eq.sob.a}
2853: $$\phantom{\mathrm{(A.1)}}\qquad\qquad\qquad\qquad {1\over r} - {s\over n}\geq {1\over q} - {k\over n},\qquad\qquad\qquad\qquad \mathrm{(A.1)}$$
2854: %\end{equation}
2855: then $H_k^q$ is continuously embedded into $H_s^r$.
2856: \item {\rm (Rellich-Kondrakov theorem).}
2857: If strict inequality holds in {\rm (A.1)} %\eref{eq.sob.a}
2858: and if $k>s$, then the inclusion
2859: $H_k^q\embed H_s^r$ is a
2860: compact map.
2861: \item {\rm (Sobolev embedding theorem II).}
2862: Suppose $0<\al <1$, $m\in\{0,1\}$ and
2863: $${1\over q}\leq {k-m-\al \over n}.$$
2864: Then $H_k^q$ is continuously embedded into $C^{m,\al}$.
2865: \end{enumerate}
2866: \end{theorema}
2867:
2868:
2869: \komment{
2870: +++++++++++++++++++++
2871:
2872: Let $(M,g)$ be a Riemannian manifold with a spin structure $\si$.
2873:
2874: \begin{definition}[Sobolev spaces]
2875: For any smooth spinor $\psi$, and $q\in (1,\infty)$, $k\in \mN\cup\{0\}$,
2876: we define the $H^q_k$-norm of $\psi$ as
2877: $$\|\psi\|_{H^q_k}= \|\underbrace{\na \ldots \na}_{k-\mbox{times}}\psi\|_{L^q}.$$
2878: \end{definition}
2879:
2880: Obviously, the $H_q^k$-norms for different connections are equivalent.
2881:
2882: If $M$ is compact, an alternative way to introduce
2883: Sobolev norms on spinors is by setting
2884: $$\big\|\psi\big\|_{\witi H^q_k}=\Big\|\,|D|^k\psi\Big\|_{L^q} + \|\pi\psi\|_1,$$ where $\pi$ is the $L^2$-orthogonal projection to the kernel of $D$,
2885: where $\|\,\cdot\,\|_1$ is an arbitrary norm on the kernel, and
2886: where $|D|^k$ should be understood in the spectral sense, i.e.\
2887: if $\phi$ is an eigenspinor of $D$ to the eigenvalue $\la$,
2888: then $|D|^k\phi= |\la|^k$. Such powers of differential operators
2889: are well-understood because $M$ is compact (see \cite{seeley:67,taylor:81}).
2890: One consequence of the properties of such operators is that the norms
2891: $\ti H^q_k$ are equivalent to the $H^q_k$-norms.
2892: The definition of the $\witi H^q_k$-norms
2893: has the advantage that it extends to arbitrary $k\in \mR$.
2894: \komment{
2895: The definition of Sobolev-norms with $k\not\in \mZ$
2896: on non-compact manifolds can be easily extended, but we omit their definition
2897: as they will not be needed.
2898: }
2899:
2900: We define the Sobolev space $H^q_k(\Si M)=H^q_k$ as
2901: the completion of the smooth spinors with respect to this norm.
2902:
2903:
2904:
2905: \begin{definition}[H\"older spaces]
2906: For $\al\in (0,1]$, the \emph{H\"older-spaces}
2907: $C^{0,\al}(\Si M)$ and $C^{1,\al}(\Si M)$ are defined to be the completions of
2908: $C^\infty(\Si M)$ with respect to the
2909: \emph{H\"older norms}
2910: \begin{eqnarray*}
2911: \|\phi\|_{C^{0,\al}} & := & \mbox{h\"ol}_\al(\phi)\\
2912: \|\phi\|_{C^{1,\al}} & := & \|\phi\|_{\rm sup} + \mbox{h\"ol}_\al(\nabla \phi)\\
2913: \mbox{h\"ol}_\al(Q)& := & \sup \Bigl\{{|Q(x)-P_\ga Q(y)|\over
2914: d(x,y)^\al}\;|\; x,y\in M, x\neq y, P_\ga \mbox{ is the }\\
2915: && \mbox{parallel transport along a shortest geodesic $\ga$ from $x$ to $y$}.\Bigr\}
2916: \end{eqnarray*}
2917: \end{definition}
2918:
2919: %
2920: %In the following we assume for simplicity that either $s\in \mZ$ or $r=2$.
2921: %
2922: From elliptic theory, we know the following statements \cite{ammann:habil}.
2923:
2924: \begin{theorema}[Sobolev embedding theorem]\label{theo.sob}\
2925: Let $k,s\in \mR$, $k\geq s$ and $q,r\in (1,\infty)$.
2926: \begin{enumerate}[{\rm (a)}]
2927: \item If
2928: %\begin{equation}\label{eq.sob.a}
2929: $$\phantom{\mathrm{(A.1)}}\qquad\qquad\qquad\qquad {1\over r} - {s\over n}\geq {1\over q} - {k\over n},\qquad\qquad\qquad\qquad \mathrm{(A.1)}$$
2930: %\end{equation}
2931: then $H_k^q(\Si M)$ is continuously embedded into $H_s^r(\Si M)$.
2932: \item {\rm (Rellich-Kondrakov theorem).}
2933: If strict inequality holds in (A.1)%\eref{eq.sob.a}
2934: , then the inclusion
2935: $H_k^q(\Si M)\embed H_s^r(\Si M)$ is a
2936: compact map.
2937: \item Suppose $0<\al <1$, $m\in\{0,1\}$ and
2938: $${1\over q}\leq {k-m-\al \over n}.$$
2939: Then $H_k^q(\Si M)$ is continuously embedded into $C^{m,\al}(\Si M)$.
2940: \item (Arcela-Ascoli) for $m\in \mN\cup \{0\}$, the inclusion
2941: $C^{m,\al}\to C^m$ is compact, i.e.\
2942: a bounded sequence in $C^{m,\al}$ has a subsequence convergent in $C^m$.
2943: \end{enumerate}
2944: \end{theorema}
2945:
2946: \begin{theorema}[Interior $L^p$ estimates]\label{theo.int.lp}
2947: Let $\Omega$ be open in $\mR^n$, equipped with a Riemannian metric and
2948: the spin structure $\si$ induced from $\mR^n$, and
2949: $K\subset \Omega$ compact.
2950: We assume
2951: $g(v,v)\geq \zeta g_{\rm eucl}(v,v)$ for all $v\in T\Omega$ and
2952: $g\in C^{k+1}$ with $\|g\|_{C^{k+1}}\leq Z$. Let $D$ be the Dirac operator on
2953: $(\Omega,g,\si)$.
2954: Let $\psi$ be in $H_k^q(\Si \Omega)$ and let $\phi$ be a weak solution of
2955: $$D\phi = \psi$$
2956: on $\Omega$.
2957: Then $\phi\in H_{k+1}^q(\Si K)$ and
2958: $$\|\phi\|_{H_{k+1}^q(\Si K)}\leq C\cdot
2959: \left(\|\psi\|_{H_k^q(\Si \Omega)} + \|\phi\|_{L^q(\Si \Omega)}\right),$$
2960: where $C=C(\zeta,Z,\Omega,K)$.
2961: \end{theorema}
2962: %\cite[Theorem~7.63]{adams:75}
2963:
2964:
2965: \begin{theorema}[Interior Schauder estimates]\label{theo.int.schauder}
2966: Let $\Omega$ be open in $\mR^n$, equipped with a Riemannian metric and
2967: the spin structure induced from $\mR^n$, and $K\subset \Omega$ compact.
2968: Let $\psi$ be a $C^{0,\al}$-spinor on $\Omega$.
2969: We assume
2970: $g(v,v)\geq \zeta g_{\rm eucl}(v,v)$ for all $v\in T\Omega$ and
2971: $\|g\|_{C^{k+1,\al}(\Omega)}\leq Z$.
2972: Then for any $C^1$-solution $\phi$ of
2973: $$D\phi = \psi$$
2974: we have $\phi\in C^{k+1,\al}(K)$ and
2975: $$\|\phi\|_{C^{k+1,\al}(K)}\leq C \cdot (\|\psi\|_{C^{k,\al}(\Omega)}+\|\phi\|_{C^0(\Omega)})$$
2976: where $C$ only depends on $n$, $\al$, $\diam (\Omega)$,
2977: $\dist(K,\partial \Omega)$, $\zeta$ and $Z$.
2978: \end{theorema}
2979:
2980: By gluing together the local versions via charts, we obtain
2981: global $L^p$ and Schauder estimates.
2982:
2983: \begin{theorema}[Global $L^p$ estimates]\label{theo.glob.lp}
2984: Let $(M,g,\si)$ be a compact Riemannian spin manifold
2985: and $\psi\in H_k^q(\Si M)$.
2986: Then any weak solution of
2987: $$D\phi = \psi$$
2988: satisfies $\phi\in H_{k+1}^q(\Si M)$, and there is a constant $C=C(M,g,\si)$
2989: such that
2990: $$\|\phi\|_{H_{k+1}^q}\leq C \left(\|\psi\|_{H_{k}^q}+\|\phi\|_{L^q}\right).$$
2991: \end{theorema}
2992: \begin{theorema}
2993: Let $(M,g,\si)$ be a compact Riemannian spin manifold, and
2994: let $\phi\in L^1(\Si M)$ with $D\phi\in H_k^q(\Si M)$.
2995: Then $\phi\in H_{k+1}^q(\Si M)$,
2996: and there is a constant $C=C(M,g,\si)$ with
2997: $$\|\phi\|_{H_{k+1}^q}\leq C
2998: \left(\|D\phi\|_{H_{k}^q}+\|\pi_{\mathop{\rm ker}(D)}(\phi)\|_{L^q}\right)$$
2999: where $\pi_{\mathop{\rm ker}(D)}$ is the $L^2$-orthogonal projection
3000: to the kernel of $D$.
3001: \end{theorema}
3002:
3003: \begin{theorema}[Global Schauder estimates]\label{theo.glob.schauder}
3004: Let $(M,g,\si)$ be a compact Riemannian spin
3005: manifold and $\psi\in C^{k,\al}(\Si M)$.
3006: Then any solution of
3007: $$D\phi = \psi,$$
3008: satisfies $\phi\in C^{k+1,\al}(\Si M)$,
3009: and there is a constant $C=C(M,g,\si)$ such that
3010: $$\|\phi\|_{C^{k+1,\al}}\leq C \left(\|\psi\|_{C^{k,\al}}+\|\phi\|_{C^{0}}\right).$$
3011: \end{theorema}
3012: }
3013:
3014: \komment{
3015: Similar norms and completions
3016: can be defined on $C^\infty(M)$ using the Laplacian
3017: on functions. The optimal constant in the embedding
3018: $H_1^2\embed L^{p_Y}$, $p_Y= 2n/(n+2)$ has been determined by Hebey.
3019:
3020:
3021: We recall an untechnical version of the theorem, which gives
3022: sufficient control for our application:
3023: %
3024: %We recall the following theorem which is a part of Hebey's research on the
3025: %optimal constant problem for Sobolev embeddings.
3026: %We set $p_Y=2n/(n-2)$, the critical exponent for the conformal Laplacian.
3027:
3028: \begin{theorema}[{\cite[Theorem 4.12]{hebey:96}}]\label{theo.hebey}
3029: Let $(M,g)$ be a complete Riemannian $n$-manifold such that the Riemann
3030: curvature tensor fulfills $|R|\leq \La_1$ and $|\na R|\leq \La_2$ for some
3031: $\La_1, \La_2>0$ and $\inj(M,g)\geq i$ for some $i>0$. Let $\pY=2n/(n-2)$
3032: Then there exists a positive constant $B=B(n,\La_1,\La_2,i)$
3033: depending only on $n$, $\La_1$, $\La_2$ and $i$, such that for any $u\in H_1^2(M)$,
3034: $$\|u\|^2_{L^{\pY}}\leq {4\over n (n-2)\, \om_n^{2/n}}\, \|du\|_{L^2}^2 +
3035: B \,\|u\|^2_{L^2}.$$
3036: \end{theorema}
3037: }
3038:
3039: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3040: \section{Some facts about H\"older spaces}\label{sec.scha.power}
3041: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3042:
3043: In this section we want to include some proofs of probably
3044: well-known statements about H\"older spaces.
3045:
3046: \begin{lemmaa}
3047: Let $V$ be a Euclidean vector space, $\al\in (0,1)$, $m\in \mN$.
3048: Then the map $V\to \bigotimes_m V$,
3049: $$T:x\mapsto |x|^{\al-m}\underbrace{x\otimes\cdots\otimes x}_{\mbox{$m$ times}}$$
3050: is $C^\al$.
3051: \end{lemmaa}
3052:
3053: \proof{}
3054: Let $x,y\in V$. At first, we suppose that
3055: $\|x-y\|<\de$ and $\|x\|\geq \de$.
3056: Then we calculate
3057: %\begin{equation}\label{est.hoeeins}
3058: $$
3059: \|x\|\,\Big\|{x\over \|x\|}-{y\over \|y\|}\Big\|=\Big\|x-{\|x\|\over \|y\|}y\Big\|\leq \|x-y\|+\Big|1- {\|x\|\over \|y\|}\Big|\,\|y\|\leq 2\de
3060: $$
3061: %\end{equation}
3062: Hence
3063: %\begin{equation}\label{est.hoezwei}
3064: $$
3065: \Big\|{x\over \|x\|}\otimes \cdots \otimes {x\over \|x\|}
3066: - {y\over \|y\|}\otimes \cdots \otimes {y\over \|y\|}\Big\|\leq {2\de m\over \|x\|}
3067: $$
3068: %\end{equation}
3069: This implies
3070: $$|T(x)-T(y)|\leq
3071: \Big|\|x\|^\al -\|y\|^\al\Big|+ \|x\|^\al \,{2\de m\over \|x\|}\leq \|x-y\|^\al+ \|x\|^{\al-1}2\de m\leq (2m+1)\de^\al.$$
3072:
3073: Now suppose that
3074: $\|x-y\|<\de$ and $\|x\|< \de$. Then
3075: $$|T(x)-T(y)|\leq \|x\|^\al + \|y\|^\al\leq 3\de^\al.$$
3076: Hence, we obtain $|T(x)-T(y)|\leq (2m+1)\|x-y\|$ for all $x,y\in V$, and hence $T$ is $C^\al$.
3077: \qed
3078:
3079:
3080: \begin{propositiona}
3081: If $V$ is a vector bundle over a compact Riemannian manifold,
3082: and if $\phi$ is a
3083: $C^{1,\al}$-section of $V$, $\al\in (0,1)$ then
3084: $\psi=|\phi|^\beta\phi$, $\be>0$ is a $C^{1,\gamma}$-section
3085: for $\gamma:=\min\{\al,\be\}$.
3086: \end{propositiona}
3087:
3088: \proof{}
3089: The section $\phi$ is obviously Lipschitz, hence $|\phi|^\be$ is $C^\be$.
3090: We have to show that
3091: \begin{eqnarray*}
3092: \na\psi&=&|\phi|^\be\na\phi + \beta\<\na \phi,\phi\>|\phi|^{\be-2}\phi
3093: \end{eqnarray*}
3094: is $C^\ga$. The first summand is a product of $C^\be$ and $C^\al$,
3095: hence $C^\gamma$. According to the previous lemma,
3096: $|\phi|^{\be-2}\phi\otimes \phi$ is $C^\be$, hence the second summand
3097: is a product of $C^\al$ and $C^\be$, hence also $C^\ga$.
3098: \qed
3099:
3100:
3101: \end{appendix}
3102:
3103:
3104:
3105:
3106:
3107: %\bibliographystyle{amsalpha}
3108: %\bibliographystyle{amsbernd}
3109: %\bibliography{literatur}
3110:
3111: \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
3112: \providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR }
3113: % \MRhref is called by the amsart/book/proc definition of \MR.
3114: \providecommand{\MRhref}[2]{%
3115: \href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2}
3116: }
3117: \providecommand{\href}[2]{#2}
3118: \begin{thebibliography}{AHM03b}
3119:
3120: \bibitem[Ada75]{adams:75}
3121: R.~Adams, \emph{{S}obolev {S}paces}, Academic Press, 1975.
3122:
3123: \bibitem[AH05]{ammann.humbert:gluing}
3124: B.~Ammann and E.~Humbert, \emph{A gluing formula for the spinorial
3125: $\tau$-invariant (preliminary title)}, Work in progress, 2005.
3126:
3127: \bibitem[AHM03a]{ammann.humbert.morel:p03b}
3128: B.~Ammann, E.~Humbert, and B.~Morel, \emph{Mass endomorphism and spinorial
3129: {Y}amabe type problems}, to appear in {C}omm. {A}nal. {G}eom., 2003.
3130:
3131: \bibitem[AHM03b]{ammann.humbert.morel:p03av2}
3132: \bysame, \emph{A spinorial analogue of {A}ubin's inequality}, {A}r{X}iv
3133: math.{DG}/0308107, 2003.
3134:
3135: \bibitem[Amm98]{ammann:diss}
3136: B.~Ammann, \emph{{S}pin-{S}trukturen und das {S}pektrum des
3137: {D}irac-{O}perators}, Ph.D. thesis, University of {F}reiburg, {G}ermany,
3138: 1998, Shaker-Verlag Aachen 1998, ISBN 3-8265-4282-7.
3139:
3140: \bibitem[Amm03a]{ammann:03}
3141: \bysame, \emph{A spin-conformal lower bound of the first positive {D}irac
3142: eigenvalue}, {D}iff. {G}eom. {A}ppl. \textbf{18} (2003), 21--32.
3143:
3144: \bibitem[Amm03b]{ammann:habil}
3145: \bysame, \emph{A variational problem in conformal spin geometry},
3146: Habilitationsschrift, Universit\"at Hamburg, 2003.
3147:
3148: \bibitem[Aub76]{aubin:76}
3149: T.~Aubin, \emph{{{\'E}quations diff{\'e}rentielles non lin{\'e}aires et
3150: probl{\`e}me de Yamabe concernant la courbure scalaire.}}, J. Math. Pur.
3151: Appl., IX. Ser. \textbf{55} (1976), 269--296.
3152:
3153: \bibitem[B{\"a}r92]{baer:92b}
3154: C.~B{\"a}r, \emph{Lower eigenvalue estimates for {D}irac operators}, Math. Ann.
3155: \textbf{293} (1992), 39--46.
3156:
3157: \bibitem[B{\"a}r97]{baer:97}
3158: \bysame, \emph{On nodal sets for {D}irac and {L}aplace operators}, Comm. Math.
3159: Phys. \textbf{188} (1997), 709--721.
3160:
3161: \bibitem[B{\"a}r98]{baer:98}
3162: \bysame, \emph{Extrinsic bounds for eigenvalues of the {D}irac operator}, Ann.
3163: Global Anal. Geom. \textbf{16} (1998), 573--596.
3164:
3165: \bibitem[Bes87]{besse:87}
3166: A.~L. Besse, \emph{Einstein manifolds}, Ergebnisse der Mathematik und ihrer
3167: Grenzgebiete, 3.~Folge, no.~10, Springer-Verlag, 1987.
3168:
3169: \bibitem[BG92]{bourguignon.gauduchon:92}
3170: J.-P. Bourguignon and P.~Gauduchon, \emph{Spineurs, op{\'e}rateurs de {D}irac
3171: et variations de m{\'e}triques}, Comm. Math. Phys. \textbf{144} (1992),
3172: 581--599.
3173:
3174: \bibitem[Fri80]{friedrich:80}
3175: T.~Friedrich, \emph{Der erste {E}igenwert des {D}irac-{O}perators einer
3176: kompakten {R}iemannschen {M}annigfaltigkeit nicht-negativer {K}r{\"u}mmung},
3177: Math. Nach. \textbf{97} (1980), 117--146.
3178:
3179: \bibitem[Fri98]{friedrich:98}
3180: \bysame, \emph{{On the spinor representation of surfaces in Euclidean
3181: $3$-space.}}, J. Geom. Phys. \textbf{28} (1998), 143--157.
3182:
3183: \bibitem[Fri00]{friedrich:00}
3184: \bysame, \emph{{D}irac {O}perators in {R}iemannian {G}eometry}, {G}raduate
3185: {S}tudies in {M}athematics 25, AMS, Providence, Rhode Island, 2000.
3186:
3187: \bibitem[GH05]{grosjean.humbert:p05}
3188: F.~Grosjean and E.~Humbert, \emph{The first eigenvalue of {D}irac and {L}aplace
3189: operators on surfaces (preliminary title)}, preprint in preparation, 2005.
3190:
3191: \bibitem[GT77]{gilbarg.trudinger:77}
3192: D.~Gilbarg and N.~Trudinger, \emph{Elliptic partial differential equations of
3193: second order}, Grundlehren der mathematischen Wissenschaften, no. 224,
3194: Springer-Verlag, 1977.
3195:
3196: \bibitem[Hij86]{hijazi:86}
3197: O.~Hijazi, \emph{A conformal lower bound for the smallest eigenvalue of the
3198: {D}irac operator and {K}illing spinors}, Comm. Math. Phys. \textbf{104}
3199: (1986), 151--162.
3200:
3201: \bibitem[Hij01]{hijazi:01}
3202: \bysame, \emph{{Spectral properties of the {D}irac operator and geometrical
3203: structures.}}, {Ocampo, Hernan (ed.) et al., Geometric methods for quantum
3204: field theory. Proceedings of the summer school, Villa de Leyva, Colombia,
3205: July 12-30, 1999. Singapore: World Scientific. 116-169 }, 2001.
3206:
3207: \bibitem[Hit74]{hitchin:74}
3208: N.~Hitchin, \emph{Harmonic spinors}, Adv. Math. \textbf{14} (1974), 1--55.
3209:
3210: \bibitem[Kir86]{kirchberg:86}
3211: K.-D. Kirchberg, \emph{An estimation for the first eigenvalue of the {D}irac
3212: operator on closed {K}\"ahler manifolds of positive scalar curvature}, Ann.
3213: Global Analysis and Geometry \textbf{4} (1986), 291--325.
3214:
3215: \bibitem[Kir88]{kirchberg:88}
3216: \bysame, \emph{Compact six-dimensional {K}{\"a}hler spin manifolds of positive
3217: scalar curvature with the smallest possible first eigenvalue of the {D}irac
3218: operator}, Math.\ Ann. \textbf{282} (1988), 157--176.
3219:
3220: \bibitem[KS96]{kusner.schmitt:p96}
3221: R.~Kusner and N.~Schmitt, \emph{The spinor representation of surfaces in
3222: space}, preprint, http://www.arxiv.org/abs/dg-ga/9610005, 1996.
3223:
3224: \bibitem[KSW98]{kramer.semmelmann.weingart:98}
3225: W.~Kramer, U.~Semmelmann, and G.~Weingart, \emph{The first eigenvalue of the
3226: {D}irac operator on quaternionic {K}\"ahler manifolds}, Comm. Math. Phys.
3227: \textbf{199} (1998), no.~2, 327--349.
3228:
3229: \bibitem[KSW99]{kramer.semmelmann.weingart:99}
3230: \bysame, \emph{Eigenvalue estimates for the {D}irac operator on quaternionic
3231: {K}\"ahler manifolds}, Math. Z. \textbf{230} (1999), no.~4, 727--751.
3232:
3233: \bibitem[Lic63]{lichnerowicz:63}
3234: A.~Lichnerowicz, \emph{Spineurs harmoniques}, C. R. Acad. Sci. Paris
3235: \textbf{257} (1963), 7--9.
3236:
3237: \bibitem[LM89]{lawson.michelsohn:89}
3238: H.-B. Lawson and M.-L. Michelsohn, \emph{Spin geometry}, Princeton University
3239: Press, Princeton, 1989.
3240:
3241: \bibitem[LP87]{lee.parker:87}
3242: J.~M. Lee and T.~H. Parker, \emph{{The Yamabe problem.}}, Bull. Am. Math. Soc.,
3243: New Ser. \textbf{17} (1987), 37--91.
3244:
3245: \bibitem[LY82]{li.yau:82}
3246: P.~Li and S.-T. Yau, \emph{A new conformal invariant and its applications to
3247: the {W}illmore {C}onjecture and the first eigenvalue of compact surfaces},
3248: Invent. Math. \textbf{69} (1982), 269--291.
3249:
3250: \bibitem[Mai97]{maier:97}
3251: S.~Maier, \emph{Generic metrics and connections on spin- and
3252: spin-$\,^c$-manifolds}, Comm. Math. Phys. \textbf{188} (1997), 407--437.
3253:
3254: \bibitem[Mor02]{morel:02}
3255: B.~Morel, \emph{Surfaces in {$S^3$} and {$H^3$} via spinors}, preprint, 2002,
3256: math.{DG}/0204090.
3257:
3258: \bibitem[Pf{\"a}02]{pfaefflediss}
3259: F.~Pf{\"a}ffle, \emph{{E}igenwertkonvergenz f{\"u}r {D}irac-{O}peratoren},
3260: Ph.D. thesis, University of {H}amburg, {G}ermany, 2002, Shaker Verlag Aachen
3261: 2003, ISBN 3-8322-1294-9.
3262:
3263: \bibitem[Pin85]{pinkall:85a}
3264: U.~Pinkall, \emph{Regular homotopy classes of immersed surfaces}, Topology
3265: \textbf{24} (1985), 421--434.
3266:
3267: \bibitem[Roe88]{roe:88}
3268: J.~Roe, \emph{Elliptic operators, topology and asymptotic methods}, Pitman
3269: Research Notes in Mathematics Series, no. 179, Longman, 1988.
3270:
3271: \bibitem[Sch93]{schmitt:diss}
3272: N.~Schmitt, \emph{Minimal surface with planar embedded ends}, {P}h.{D}.
3273: dissertation, University of Amherst, 1993.
3274:
3275: \bibitem[Tay81]{taylor:81}
3276: M.~E. Taylor, \emph{Pseudodifferential operators}, Princeton University Press,
3277: Princeton, N.J., 1981.
3278:
3279: \bibitem[Vos99]{voss:dipl}
3280: L.~Voss, \emph{Eigenwerte des {D}irac-{O}perators auf {H}yperfl{\"a}chen},
3281: {D}iplomarbeit, {H}umboldt {U}niversit\"at zu {B}erlin, 1999.
3282:
3283: \bibitem[Wei66]{weierstrass:66}
3284: K.~Weierstrass, \emph{{U}ntersuchungen {\"u}ber die {F}l{\"a}chen, deren
3285: mittlere {K}r{\"u}mmung {\"u}berall gleich {N}ull ist}, available in
3286: {W}eierstrass, {M}athematische {W}erke. {III}, 39--52, 1866.
3287:
3288: \end{thebibliography}
3289:
3290: \pagebreak[2]
3291:
3292: \vspace{1cm}
3293: Author's address:
3294: \nopagebreak
3295: \vspace{5mm}\\
3296: \parskip0ex
3297: \vtop{
3298: \hsize=7cm\noindent
3299: \obeylines
3300: Bernd Ammann
3301: Institut \'Elie Cartan BP 239
3302: Universit\'e Henri Poincar\'e, Nancy 1
3303: 54506 Vandoeuvre-l\`es-Nancy Cedex
3304: France
3305: }
3306:
3307:
3308: \vspace{0.5cm}
3309:
3310: E-Mail:
3311: {\tt bernd.ammann@gmx.de}
3312:
3313:
3314: WWW:
3315: {\tt http://www.berndammann.de/uni}
3316:
3317: \end{document}
3318: