1: % ----------------------------------------------
2: % Dressing preserving the fundametal group
3: %
4: % Josef Dorfmeister
5: % Martin Kilian
6: %
7: % May 08, 2003
8: % ----------------------------------------------
9:
10: % debug flag
11: \newif\ifdebug
12: \debugfalse
13:
14: % preliminary flag
15: \newif\ifpreliminary
16: \preliminarytrue
17:
18: \documentclass[12pt,twoside]{article}
19: %\usepackage{a4Camb}
20:
21: \usepackage{amsmath}
22: \usepackage{amstext}
23: \usepackage{amsbsy}
24: \usepackage{amssymb}
25: \usepackage{amsfonts}
26: \usepackage{latexsym}
27:
28: \usepackage{graphicx}
29: %\usepackage[draft]{graphicx}
30:
31: % gzipped images
32: \DeclareGraphicsRule{.eps.gz}{eps}{.eps.bb}{`gunzip -c #1}
33:
34: \ifdebug
35: \usepackage[notref,notcite]{showkeys}
36: \fi
37: \makeatletter
38: \@addtoreset{equation}{subsection}
39: \def\theequation{\thesubsection.\arabic{equation}}
40: \makeatother
41:
42: \renewcommand{\thefootnote}{\fnsymbol{footnote}}
43:
44: \vskip6.94444pt
45:
46: \newcommand{\separate}{\medskip\noindent}
47:
48: \def\newsection{ \separate
49: \refstepcounter{subsection}
50: {\large\bf \thesubsection\kern.3em}
51: }
52: \def\mytheorem#1{
53: \separate{\large\bf Theorem#1:\kern.3em}}
54:
55: \def\mylemma#1{
56: \separate{\large\bf Lemma#1:\kern.3em}}
57:
58: \def\mycorollary#1{
59: \separate{\large\bf Corollary#1:\kern.3em}}
60:
61: \def\myprop#1{
62: \separate{\large\bf Proposition#1:\kern.3em}}
63:
64: \def\myremark#1{
65: \separate{\large\bf Remark#1:\kern.3em}}
66:
67: \def\mydefinition#1{
68: \separate{\large\bf Definition#1:\kern.3em}}
69:
70: \def\myexample#1{
71: \separate{\large\bf Example#1:\kern.3em}}
72:
73: \def\Proof{\separate\underline{Proof:}\kern1em}
74:
75: \newcommand{\field}[1]{\mathbb{#1}}
76: \newcommand{\C}{\field{C}}
77: \newcommand{\D}{\field{D}}
78: \newcommand{\R}{\field{R}}
79: \newcommand{\Q}{\field{Q}}
80: \newcommand{\N}{\field{N}}
81: \newcommand{\Z}{\field{Z}}
82: \newcommand{\CPE}{{\mathchoice{\C{\rm P}_1}{\C{\rm P}_1}{\C\!\!
83: \mbox{\rm\tiny P}_1}{\C\!\!\mbox{\rm\tiny P}_1}}}
84: \newcommand{\ad}{{\mathchoice{\mbox{\rm ad}}{\mbox{\rm ad}}{%
85: \mbox{\scriptsize\rm ad}}{\mbox{\scriptsize\rm ad}}}}
86: \newcommand{\unity}{{\setlength{\unitlength}{1em}
87: \begin{picture}(0.75,1)
88: \put(0,0){$1$}
89: \put(0.34,0){\line(0,1){0.65}}
90: \end{picture}
91: }}
92: \newcommand{\X}{\mathbb X}
93: \newcommand{\Xh}{\widehat{\mathbb X}}
94: \newcommand{\T}{\mathbb T}
95: \newcommand{\hash}{\mbox{\tiny{\#}}}
96: \newcommand{\MT}{\mbox{\small{$\widetilde{M}$}}}
97: \newcommand{\s}{\mbox{\boldmath $\sigma$}}
98: \newcommand{\ci}{\mbox{\boldmath $i$}}
99: \newcommand{\DPWLab}{{\tt dpwlab}}
100: \newcommand{\CP}{\C \mathbb{P}^1}
101: \newcommand{\tr}{\mathrm{tr}}
102: \newcommand{\Id}{\mathrm{Id}}
103: \newcommand{\Order}{{\rm O}}
104: \newcommand{\FID}{\mathcal{F}_{\mbox{\tiny{$\Id$}}}
105: (\MT)}
106:
107: \def\QED{\hfill$\Box$}
108: \def\inv{^{-1}}
109: \def\bref#1{(\ref{#1})}
110: \def\tmatrix#1#2#3#4{
111: \left(\begin{array}{cc} #1 & #2 \\ #3 & #4 \end{array}\right)}
112:
113: %MATRICES
114:
115: \def\SL{{\mbox{\bf SL}(2,\C)}}
116: \def\SO{{\mbox{\bf SO}}}
117: \def\GL{{\mbox{\bf GL}(2,\C)}}
118: \def\SU{{\mbox{\bf SU}(2)}}
119:
120: \newcommand{\Utwo}{\mbox{\bf U}(2)}
121: \def\U1{{\mbox{\bf U}(1)}}
122:
123: \def\B{{\mbox{\bf B}}}
124: \def\G{{\mbox{\bf G}}}
125: \def\sl{{\mbox{\bf sl}(2,\C)}}
126: \def\gl{{\mbox{\bf gl}(2,\C)}}
127: \def\su{{\mbox{\bf su}(2)}}
128:
129: \newcommand{\n}{\noindent}
130:
131: %LOOPGROUPS
132:
133: %GL splitting on r-circle
134:
135: \def\LSL{{\Lambda \SL}}
136: \def\L{{\Lambda_r \GL_\sigma}}
137: \def\LP{{\Lambda_r^+ \GL_\sigma}}
138:
139: \def\LU{{\Lambda_r \U_\sigma}}
140: \newcommand{\LB}{\Lambda_{r,\T}^{\mbox{\tiny{$+$}}} \GL _\sigma}
141:
142: % SL loops on two different circles
143: \def\LGC{{\Lambda_r \SL_\sigma}}
144: \def\LGCs{{\Lambda_s \SL_\sigma}}
145:
146: % Borel factor
147: \newcommand{\LGB}{\Lambda_{r,\T}^{\mbox{\tiny{$+$}}} \SL _\sigma}
148:
149: % Birkhoff factors
150: \def\LGP{{\Lambda_r^{\mbox{\tiny{$+$}}} \SL _\sigma}}
151: \def\LGM{{\Lambda_{r}^{\mbox{\tiny{$-$}}} \SL _\sigma}}
152:
153: % SU loops
154: \def\LGU{{\Lambda_r \SU_\sigma}}
155: \def\LSU{{\Lambda \SU_\sigma}}
156: \def\Lhol{{\Lambda_{\mbox{\tiny{$\C^*$}}}\SU_\sigma}}
157:
158: %LOOPALGEBRAS
159: \def\LAC{{\Lambda_r \sl _\sigma}}
160: \def\LAI{{\Lambda_r \sl _\sigma}}
161: \def\LAE{{\Lambda_r \sl _\sigma}}
162: \def\Lsu{{\Lambda \su _\sigma}}
163: \def\LGe{{\Lambda_r \su _\sigma}}
164:
165: \newcommand{\delpot}{\xi^{\mbox{\tiny{\rm{Del}}}}}
166: \newcommand{\cylpot}{\xi_{\mbox{\tiny{\rm{cyl}}}}}
167:
168:
169: %DPW-POTENTIALS
170: \newcommand{\pot}{\Lambda \Omega(M)}
171: \newcommand{\pott}{\Lambda \Omega(\MT)}
172: \newcommand{\Lsl}{\Lambda \vspace{-.3mm}\hspace{-.2mm}^{\sigma}_{\mbox{\tiny{$-1,\infty$}}} \Sl}
173:
174: \newcommand{\gauge}{\mathcal{G}_r (\MT)}
175: \newcommand{\gaugebase}{\mathcal{G}_{r,\T} (\MT)}
176:
177: \def\u{{\mbox{\bf u}}}
178: \def\b{{\mbox{\bf b}}}
179: \def\g{{\mbox{\bf g}}}
180: \def\k{{\mbox{\bf k}}}
181: \def\p{{\mbox{\bf p}}}
182: \def\s{{\mbox{\bf $\sigma$}}}
183:
184: \def\l{{\mbox{\footnotesize \sc l}}}
185:
186:
187:
188: %\input epsf
189:
190: \setlength{\topmargin}{-.6in}
191: \setlength{\textheight}{8.75in}
192: \ifdebug
193: \setlength{\textwidth}{5in}
194: \else
195: \setlength{\textwidth}{6in}
196: \setlength{\oddsidemargin}{3.5ex}
197: \setlength{\evensidemargin}{0 pt}
198: \fi
199: \setlength{\headsep}{.5in}
200: \setlength{\footskip}{.5in}
201: \setlength{\footskip}{.5in}
202: \setlength{\parskip}{1ex}
203: \setlength{\baselineskip}{1.2\baselineskip}
204: \setlength{\parindent}{0in}
205:
206:
207: % uses lower-case Roman numbers in 'enumerate'
208: \renewcommand{\theenumi}{\roman{enumi}}
209: \renewcommand{\labelenumi}{(\theenumi)}
210: \renewcommand{\theenumii}{\roman{enumii}}
211: \renewcommand{\labelenumii}{(\theenumii)}
212: %\renewcommand{p@enumii}{\theenumi--}
213:
214:
215: \begin{document}
216:
217: \begin{center}
218: {\LARGE Dressing preserving the fundamental group}
219:
220: \vskip1cm
221: \begin{minipage}{6cm}
222: \begin{center}
223: Josef~Dorfmeister \\
224: TU M\"{u}nchen\\
225: Zentrum Mathematik \\
226: Boltzmannstr. 3 \\
227: D-85747 Garching, Germany.
228: \end{center}
229: \end{minipage}
230: \begin{minipage}{6cm}
231: \begin{center}
232: Martin~Kilia$\mbox{n}^*$ \\
233: Mathematical Sciences \\
234: University of Bath \\
235: Claverton Down \\
236: Bath, BA2 7AY, UK.
237: \end{center}
238: \end{minipage}
239: \vspace{0.5cm}
240:
241: \end{center}
242: \footnotetext[0]{1991 Mathematics Subject
243: Classification. Primary 53A10; Secondary 53C20.\\
244: $^*$ Supported by CEC Contract
245: HPRN-CT-2000-00101 EDGE and EPSRC GR/S28655/01. \date{\today}}
246:
247: \begin{abstract}
248: In this note we consider the
249: relationship between the dressing action and
250: the holonomy representation in the context of
251: constant mean curvature surfaces.
252: We characterize dressing elements that preserve
253: the topology of a surface and discuss dressing
254: by simple factors as a means of adding bubbles
255: to a class of non finite type cylinders.
256: \end{abstract}
257:
258: \refstepcounter{subsection}
259:
260: \textbf{Introduction.}
261: The equation for a harmonic map from a Riemann
262: surface to a Riemannian symmetric space has a
263: zero-curvature representation, and so
264: corresponds to a loop of flat connections.
265: Uhlenbeck discovered in her study \cite{Uhl} of
266: harmonic maps into a compact Lie group
267: $G$ that such maps correspond to certain holomorphic
268: maps into the based loop group of $G$ and used
269: this to define the \textit{dressing action} of
270: a certain loop group on the space of harmonic maps.
271: Dressing, or the \emph{vesture method} \cite{ZakS2},
272: was first developed in soliton theory to generate
273: new solutions from old by solving a matrix Riemann
274: problem. Unfortunately, the new solution does not
275: automatically inherit properties of the old,
276: such as domain, periods or asymptotics nor is
277: it easy to control such properties when solving
278: a matrix Riemann problem.
279:
280: When the target is the two dimensional
281: round sphere, harmonic maps are precisely
282: the Gauss maps of surfaces with constant mean
283: curvature \cite{RuhV}. In the article \cite{DorPW}
284: a method was presented by which all conformally
285: immersed surfaces with constant mean curvature
286: (CMC) can be obtained.
287: The construction involves solving a meromorphic
288: linear differential system with values in
289: a loop group and then Iwasawa decomposing the
290: solution to obtain the extended unitary frame of
291: the surface. Both these steps make it
292: difficult to keep track of the topology.
293: Variation of the initial condition
294: corresponds to the dressing action and is
295: an integral part of the theory.
296: Not only can dressing be used to close periods, but
297: also to generate new CMC surfaces from old ones
298: without altering the topology. In this work,
299: we look into the relationship
300: between dressing and topology in the framework of
301: meromorphic ODE's and loop group factorizations
302: in the context of CMC surfaces.
303:
304: The question, when a conformal CMC immersion
305: factors through a given surface has been discussed
306: for tori \cite{Bob:tor}, \cite{DorH:per},
307: \cite{PinS}, cylinders \cite{DorH:cyl},
308: \cite{Kil:thesis},
309: \cite{KilMS} and trinoids \cite{BobPS},
310: \cite{DorW:noids}, \cite{KilMS}, \cite{Sch:tri}.
311: It will be interesting to see more results in
312: this direction, with particular emphasis on
313: complete, non-compact surfaces with genus and
314: compact surfaces of genus $g \geq 2$.
315:
316: Sterling \& Wente \cite{SteW} discovered that
317: one can add ``bubbles'' to a round cylinder
318: while preserving the mean curvature by
319: Bianchi-- B\"{a}cklund transformations.
320: In \cite{Kil:thesis} it was shown that these
321: ``bubbletons'' can be produced by dressing
322: from the standard round cylinder by the so called
323: {\em simple factors} of Terng and Uhlenbeck
324: \cite{TerU}. Subsequently, it was shown in
325: \cite{Kob} that the simple bubbletons in
326: \cite{Kil:thesis} coincide with those of Sterling
327: and Wente \cite{SteW}. Recently, A. Mahler
328: \cite{Mah} has shown that indeed
329: Bianchi -- B\"{a}cklund transformations
330: can be achieved by dressing. Further, it was
331: recently shown in \cite{KilSS} that it is
332: also possible to dress CMC surfaces homeomorphic
333: to the n-punctured sphere with suitably chosen
334: simple factors without changing the topology.
335:
336: Every CMC immersion of a Riemann surface $M$ induces
337: a 'monodromy representation'
338: $\chi = \chi (\gamma,\,\lambda )$ of the
339: fundamental group $\pi_1 (M)$ of $M$ with values
340: in a unitary loop group. Moreover, as a function
341: of the loop parameter $\lambda$, the monodromy
342: matrices are holomorphic on $\C^*$.
343: Dressing a given immersion with some arbitrary
344: matrix will in general destroy the topology and
345: the dressed surface will have a trivial
346: fundamental group. Thus the question is:
347: For which dressing matrices will the dressed
348: immersion have the same fundamental group?
349:
350: To explain this in more detail we note that in
351: our method we associate with a given CMC immersion
352: $f :M \rightarrow \R^3$ its associated family
353: $f_\lambda :\MT \to \R^3$, where $\MT$
354: denotes the universal cover of $M$ and
355: $\lambda \in S^1$. This way we obtain the extended
356: frame $F = F(z,\, \bar{z},\, \lambda )$ for
357: $\lambda \in S^1$ and $z \in \MT$.
358: If $\gamma \in \pi_1(M)$ and we also denote the
359: corresponding deck transformation by $\gamma$ and
360: write
361: $\gamma^* F = F(\gamma(z),\,\overline{\gamma(z)},\,\lambda)$
362: then we have $\gamma^*F = \chi\,F\,k$ for some smooth
363: $k:\MT \to \U1$.
364: Since, by assumption, the immersion $f_1 = f$
365: descends to $M$, the monodromy matrices satisfy
366: the two closing conditions
367: $ \chi (\gamma,\, 1) = \pm \Id$ and
368: $\left.\partial_{\lambda} \chi \right|_{\lambda = 1}
369: = 0$ for all $\gamma \in \pi_1(M)$.
370:
371: If $h = h(\lambda )$ is some dressing element
372: and $\hat F$ denotes the dressed extended
373: frame and we write $hF = \hat{F} V_+$ for
374: the dressing equation, then $\gamma^* \hat{F} =
375: h\,\chi\,h^{-1}\,\hat{F}\, V_+\,\gamma^* V_+^{-1}$.
376: Thus $\hat{F}$ has some monodromy
377: $\hat{\chi}$ if and
378: only if $\hat{L} \hat{F} = \hat{F} \hat{W}_+$,
379: where $\hat{W}_+ =
380: V_+ \, \gamma^* V_+ ^{-1}$ and
381: $\hat{L} = ( h \chi h^{-1})^{-1}
382: \hat{\chi}$. As a consequence,
383: $\hat{\chi} = h \chi h^{-1} \hat{L} =
384: h \chi L h^{-1}$, where $L = h^{-1} \hat{L} h$.
385:
386: We prove
387: %in Lemma \ref{th:iso_problem}
388: that
389: a dressing matrix $h$ preserves the fundamental
390: group of an immersion if and only if for
391: every $\gamma \in \pi_1 (M)$ there exists some
392: matrix $L = L(\gamma,\,\lambda)$, such that
393:
394: (i) $LF = F W_+$,\vspace{-1mm}
395:
396: (ii) $h \chi h^{-1}$ is holomorphic for
397: $\lambda \in \C^*$,\vspace{-1mm}
398:
399: (iii) $h \chi L h^{-1}$ is unitary on $S^1$,
400: and \vspace{-1mm}
401:
402: (iv) The closing conditions are satisfied for
403: $h\chi L h^{-1}$.
404:
405: Clearly, the situation simplifies considerably,
406: if the original immersion has an umbilic point,
407: since then $L = \pm \Id$ \cite{DorH:dre}.
408: Let us assume this for now.
409: Considering in this case condition (ii) we see
410: that $\chi$ is holomorphic for
411: $\lambda \in \mathbb{C}^*$ and also
412: $h \chi h^{-1}$ needs to be holomorphic on
413: $\mathbb{C}^*$, in spite of the fact
414: that $h$ may only be defined on some circle
415: of radius $ 0 < r \leq 1$.
416: We prove
417: %in Theorem \ref{th:h-split}
418: that in
419: the situation discussed here we can assume that
420: $h = \mathcal{M}\,\mathcal{C}$, where
421: $[\,\mathcal{C}, \,\chi \,] = 0$ and $\mathcal{M}$
422: is meromorphic on some open dense subset
423: $\mathbb{S} \subset \mathbb{C}^*$,
424: which contains an open annulus about $S^1$.
425: Moreover, $\mathbb{S}$ is solely defined
426: by the eigenvalues of $\chi$.
427: In view of condition (iii) we further show
428: %in Theorem \ref{th:h-split1}
429: that actually
430: $\mathcal{M}$ can be chosen so that it is
431: unitary on $S^1$.
432:
433: This gives altogether a fairly complete
434: description for the case, where an umbilic
435: point exists. Similarly complete is the case,
436: dealt with in Theorem \ref{th:h-split_cyl},
437: where $\chi$ is the monodromy of the standard
438: round cylinder. For the general umbilic free
439: case we have, at this point, only some partial
440: results.
441:
442: Let us briefly outline the contents of this paper.
443: The first chapter sets the scene by defining the
444: relevant loop groups, dressing and gauge actions
445: and recalls the DPW representation \cite{DorPW}
446: of CMC surfaces.
447: In the second chapter we consider how automorphisms
448: affect maps at the various levels of the DPW
449: construction as a prerequisite to understanding
450: monodromy. In chapter 3 we present a factorization
451: as a necessary condition on the dressing matrix
452: to ensure that the dressed surface retains the
453: topology of the original surface.
454: In the fourth chapter we apply our
455: methods to the dressing orbit of the vacuum
456: and give a self contained account of twisted
457: simple factors. We conclude this work by
458: discussing how dressing with simple factors is
459: related to the monodromy and apply these methods
460: to a class of CMC cylinders with umbilics.
461:
462: The second author wishes to thank F Burstall for
463: many useful discussions and the Department of
464: Mathematical Sciences at the University of Bath
465: for its hospitality.
466:
467: \section{Notation and basic results}\label{basics}
468: \message{[basics]}
469: We begin by collecting some well known results
470: on loop groups and the dressing action.
471: We shall use the following notation
472: for diagonal and off--diagonal matrices:
473: %
474: \begin{equation*}
475: \mathrm{diag}[u,\, v] =
476: \bigl( \begin{smallmatrix}
477: u & 0 \\ 0 & v \end{smallmatrix} \bigr),\,
478: \mathrm{off}[u,\, v] =
479: \bigl( \begin{smallmatrix}
480: 0 & u \\ v & 0 \end{smallmatrix} \bigr).
481: \end{equation*}
482: %
483: \newsection{\bf Loops.} \label{loopgroups}
484: For real $r \in (0,1]$, let
485: $C_r = \{ \lambda \in \C : | \lambda | = r \}$
486: and denote the $r$--Loop group of $\SL$ by
487: %
488: \begin{equation*}
489: \Lambda_r \SL = \left\{ g : C_r \to \SL
490: \mbox{ smooth} \right\}.
491: \end{equation*}
492: %
493: We have an involution on maps $C_r \to \gl$
494: defined by
495: %
496: \begin{equation} \label{eq:sigma}
497: \sigma : g(\lambda) \mapsto
498: \sigma_3 g(-\lambda) \sigma_3^{-1},\,
499: \sigma_3 = \mathrm{diag}[1,-1],
500: \end{equation}
501: %
502: and denote the \textit{twisted} $r$--Loop group
503: of $\SL$ by
504: %
505: \begin{equation*}
506: \LGC = \left\{ g \in \Lambda_r \SL:
507: \, \sigma g = g \right\}.
508: \end{equation*}
509: %
510: Analogously, one can define the Lie algebras of
511: these groups, denoted by
512: $\LAC$. To make these loop groups
513: complex Banach Lie groups, we equip them,
514: as in \cite{DorPW},
515: with some $H^s$ topology for $s>1/2$ or some
516: (possibly weighted) Wiener topology.
517: Elements of these twisted loop
518: groups are matrices whose off--diagonal entries
519: are odd functions,
520: while the diagonal entries are even functions of
521: the parameter $\lambda$.
522: All entries are in the Banach algebra
523: ${\cal A}_r$ of
524: $H^s$--smooth functions or of finite
525: (possibly weighted) Wiener norm on $C_r$.
526: Furthermore, we will use the following subgroups
527: of $\LGC$: Let
528: $I_r = \{ \lambda \in \C : |\lambda|<r \}$
529: and denote
530: %
531: \begin{equation*}
532: \LGP = \left\{ g \in \LGC : g
533: \text{ extends analytically to } I_r \right\}.
534: \end{equation*}
535: %
536: Let $A_r = \{ \lambda \in \C :
537: r<|\lambda|<1/r \}$, and by abuse of
538: notation we denote
539: %
540: \begin{equation*}
541: \LGU = \left\{ g \in \LGC : g
542: \text{ extends analytically to $A_r$
543: and}\left. g \right|_{S^1} \in \SU \right\}.
544: \end{equation*}
545: %
546: Let $g:A_r \to \SL_\sigma$ be analytic.
547: Then $g \in \LGU$ if and only if
548: $\varrho \, g = g$ for
549: %
550: \begin{equation} \label{eq:varrho}
551: (\varrho g)(\lambda) :=
552: \overline{g(1/\bar{\lambda})}^{t-1}.
553: \end{equation}
554: %
555: or alternatively, $g \in \LGU$ if and only if
556: %
557: \begin{equation} \label{eq:unitarity_def}
558: (g^*)(\lambda) :=
559: \overline{g(1/\bar{\lambda})}^t =
560: g(\lambda)^{-1}.
561: \end{equation}
562: %
563: For $r=1$ we will always omit the subscript '$r$'.
564: All the groups above are Banach Lie subgroups
565: of $\LGC$. The corresponding Banach Lie
566: sub algebras of $\LAC$ are defined analogously.
567:
568:
569: %%%%%%%%%%%
570:
571: \newsection{\bf Iwasawa decomposition.}
572: Of fundamental importance in our investigation
573: is a certain Loop group factorization.
574: We modify a result from \cite{McI} to obtain:
575:
576: \separate Multiplication
577: $\LGU \times \LGP \to \LGC$
578: is a real analytic surjection.
579: An associated splitting $g = FB$
580: of $g \in \LGC$ with $F \in \LGU$ and $B \in \LGP$
581: will be called an Iwasawa decomposition of $g$.
582: Note that
583: %
584: \begin{equation} \label{eq:intersect}
585: \LGU \cap \LGP = \U1.
586: \end{equation}
587: %
588: If we demand that $B(\lambda=0)$ has positive real
589: diagonal entries, then the multiplication map
590: above is a real analytic diffeomorphism onto.
591: We will call this decomposition the "unique" Iwasawa
592: decomposition.
593:
594: %%%%%%%%%%%
595:
596: \newsection{\bf Untwisted loops.}
597: We shall be primarily working with the twisted
598: loop groups and algebras. On occasion it is
599: advantageous to switch to the untwisted setting
600: via the isomorphism between untwisted and twisted
601: $r$--loops:
602: let $X_u \in \Lambda_r \SL$ be an
603: untwisted loop.
604: Then the corresponding twisted loop is given by
605: $X_t(\lambda) = D\,X_u(\lambda^2)\,D^{-1}$
606: where $D=\mathrm{diag}[\,\sqrt{\lambda},\,
607: 1/\sqrt{\lambda}\,]$. In matrix notation, this
608: isomorphism is given by
609: %
610: \begin{equation} \label{eq:twist}
611: \begin{pmatrix} a(\lambda) & b(\lambda) \\
612: c(\lambda) & d(\lambda) \end{pmatrix} \cong
613: \begin{pmatrix} a(\lambda^2) &
614: \lambda\, b(\lambda^2) \\
615: \lambda^{-1} c(\lambda^2) &
616: d(\lambda^2)\end{pmatrix}.
617: \end{equation}
618: %
619: One technical issue in switching between untwisted
620: and twisted loops is that under this twisting
621: isomorphism, an untwisted positive loop is not
622: automatically positive when twisted, since the
623: untwisted loop at $\lambda = 0$ need not be
624: upper triangular. To circumvent this issue,
625: we must first perform a Gram-Schmidt
626: orthonormalization at $\lambda = 0$
627: before twisting the loop.
628:
629: More precisely, consider an untwisted loop
630: $\Phi_u \in \Lambda_r\SL$ with an $r$-Iwasawa
631: decomposition $\Phi_u = F_u\,B_u$
632: with untwisted $F_u \in \Lambda_r \SU$ and
633: $B_u \in \Lambda_r^+ \SL$. Now consider the
634: twisted loop $\Phi_t(\lambda) = D\,\Phi_u(\lambda^2)\,D^{-1}$
635: obtained from $\Phi_u$ via the
636: isomorphism \eqref{eq:twist} and let
637: $\Phi_t = F_t\,B_t$ be an $r$-Iwasawa decomposition
638: with twisted $F_t \in \LGU$ and $B_t \in \LGP$.
639: If $B_u$ has an expansion
640: $B_u = B_0 + \lambda B_1 + \ldots$,
641: then as $B_0 \in \SL$, we may write
642: $B_0 = Q\,R$ via Gram-Schmidt orthonormalization
643: with $Q \in \SU$ and $R \in \SL$ upper triangular.
644: Consequently, $Q^{-1}B_u \in \Lambda_r^+\SL$ and
645: $\left.Q^{-1}B_u \right|_{\lambda=0} = R$ and thus
646: $D\,Q^{-1}B_u(\lambda^2)\,D^{-1} \in \LGP$.
647: Furthermore, we then have that
648: %
649: \begin{equation}
650: \Phi_t(\lambda) = \left( D\,F_u(\lambda^2)
651: \,Q\,D^{-1} \right) \left( D\,Q^{-1}
652: B_u(\lambda^2) \,D^{-1} \right)
653: \end{equation}
654: %
655: with $D\,F_u(\lambda^2)\,Q\,D^{-1} \in \LGU$ and
656: $D\,Q^{-1}B_u(\lambda^2)\,D^{-1} \in \LGP$ is an
657: $r$-Iwasawa decomposition. It is in this sense that
658: we may carry an Iwasawa decomposition in the
659: untwisted setting over to the twisted setting.
660:
661: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
662:
663: \newsection{\bf DPW method.} \label{sec:DPW}
664: Let $\Omega(M)$ denote the holomorphic
665: $1$--forms on a Riemann surface $M$ and define
666: the
667: %
668: \begin{equation*}
669: \pot = \Omega(M) \otimes \left\{
670: \xi : \C^* \to \sl
671: \mbox{ holomorphic}: \xi(\lambda) =
672: \sum_{j \geq -1}
673: \xi_j \lambda^j,\,\s \xi = \xi \right\}.
674: \end{equation*}
675: %
676: CMC surfaces come in $S^1$ families,
677: the \textit {associated family}.
678: The DPW representation \cite{DorPW} constructs
679: all conformal CMC immersions of the universal
680: cover $\MT$ in the following three steps:
681: Let $\xi \in \pott$, $\tilde{z}_0 \in
682: \MT$ and $\Phi_0 \in \LGC$ for some $r \in (0,1]$.
683: To avoid totally umbilical surfaces, we assume
684: $\det \xi_{-1} \not \equiv 0$.
685:
686: \noindent 1. Solve the initial value problem
687: %
688: \begin{equation}\label{eq:IVP}
689: d\Phi = \Phi \xi,\, \Phi(\tilde{z}_0)= \Phi_0
690: \end{equation}
691: %
692: to obtain a unique \textit{holomorphic frame}
693: $\Phi : \MT \to \LGC$.
694:
695: \noindent 2. Iwasawa decompose the map
696: $\Phi : \MT \to \LGC$ point-wise on $\MT$
697: %
698: \begin{equation} \label{eq:Iwasawa}
699: \Phi(z,\,\lambda) =
700: F(z,\,\bar{z},\,\lambda) \,
701: B(z,\,\bar{z},\,\lambda)
702: \end{equation}
703: %
704: We will always assume
705: that the factors $F$ and $B$ in \eqref{eq:Iwasawa}
706: are real analytic in $z$.
707: The map $F : \MT \to \LGU$ will be called
708: \textit{unitary frame}. If we use the unique
709: Iwasawa decomposition, then the factors $F$ and $B$
710: are automatically real analytic.
711:
712:
713: \noindent 3. Let $H \in \R^*$ and
714: $\partial_\lambda =
715: \tfrac{\partial}{\partial \lambda}$.
716: Plug $F$ into the Sym-Bobenko formula
717: %
718: \begin{equation} \label{eq:Sym}
719: f_\lambda = - \tfrac{1}{2H}\left(
720: i\lambda \tfrac{\partial F}{\partial\lambda}
721: \,F^{-1} + \tfrac{i}{2} F \sigma_3 F^{-1} \right)
722: \end{equation}
723: %
724: to obtain (possibly branched) conformal
725: CMC immersions $f_\lambda :\MT \to \Lsu$ with mean
726: curvature $H$, that is, for each
727: $\lambda_0 \in S^1$ we have a conformal CMC
728: immersion $f_{\lambda_0} :\MT \to \su \cong \R^3$.
729: Note, $f_\lambda (z)$ is branched at $z_0$ if
730: and only if for $\xi_{-1} = \mathrm{off}[a,\,b]$ we
731: have $a(z_0)=0$.
732:
733: \newsection{\bf Frames.} We call $\xi \in \pott$ a
734: \textit{holomorphic potential}.
735: If $\Phi_0 = \phi_u \phi_+$ is an Iwasawa
736: decomposition, then it is not hard to see that
737: the immersions obtained from
738: $(\xi,\Phi_0,\tilde{z}_0)$ and
739: $(\xi,\phi_+,\tilde{z}_0)$
740: differ by a $\lambda$-dependent rigid motion.
741: Thus the choice of initial
742: condition may be restricted to $\LGP$.
743: Further, if $\Phi_0 \in \LGP$
744: then $F(\tilde{z}_0) \in \U1$ and
745: $\Phi = F F(\tilde{z}_0)^{-1} F(\tilde{z}_0)B$.
746: Hence we may assume without loss of generality that
747: %
748: \begin{equation} \label{eq:F_initial}
749: F(\tilde{z}_0) \equiv \Id \mbox{ for all }
750: \lambda \in S^1.
751: \end{equation}
752: %
753: Let $\sl = \mathfrak{k} \oplus \mathfrak{p}$
754: be the Cartan decomposition induced by the
755: involution $\s$ defined in \eqref{eq:sigma}.
756: The specific form of $\xi$ ensures that the
757: Maurer--Cartan form of $F$ acquires the form
758: %
759: \begin{equation} \label{eq:MC}
760: F^{-1}dF =
761: \lambda^{-1} \alpha_{\mathfrak{p}}^{\prime} +
762: \alpha_{\mathfrak{k}} +
763: \lambda \, \alpha_{\mathfrak{p}}^{\prime\prime}
764: \end{equation}
765: %
766: where $\alpha_{\mathfrak{p}}^\prime$ and
767: $\alpha_{\mathfrak{p}}^{\prime\prime}$ are
768: $(1,\,0)$ respectively $(0,1)$-forms on $\MT$
769: with values in $\mathfrak{p}$ and
770: $\alpha_{\mathfrak{k}}$ takes values in
771: $\mathfrak{k}$.
772: For each $p \in \MT$,
773: $\lambda \mapsto F(p,\lambda)$ is
774: holomorphic on $\C^*$. Maps
775: %
776: \begin{equation}
777: F:\MT \to \Lhol :=
778: \bigcap_{r \in (0,1]}\LGU
779: \end{equation}
780: %
781: with the property \eqref{eq:MC} are called
782: \textit{extended unitary frames}
783: and denoted by $\mathcal{F}(\MT)$.
784: We call the extended unitary frames
785: for which \eqref{eq:F_initial} holds the
786: \textit{normalized extended unitary frames}
787: and denote these by
788: %
789: \begin{equation}
790: \FID = \left\{
791: F \in \mathcal{F}(\MT) :
792: F(\tilde{z}_0) = \Id \mbox{ for some }
793: \tilde{z}_0 \in \MT \right\}.
794: \end{equation}
795: %
796:
797: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
798:
799: \newsection{\bf Gauge.} \label{Gauge}
800: The DPW representation $(\xi,\Phi_0,\tilde{z}_0)
801: \mapsto \mathcal{F}(\MT)$ is a surjective map
802: \cite{DorPW}. Injectivity fails since the
803: \textit{Gauge group}
804: %
805: \begin{equation}
806: \gauge = \{ G : \MT
807: \to \LGP \mbox{ holomorphic } \},
808: \end{equation}
809: %
810: acts by right multiplication on the fibers of
811: this map. On the level of the potential, this
812: \textit{gauge action} is given by
813: %
814: \begin{equation}\label{eq:gauge}
815: \xi.G = G^{-1} \xi \ G + G^{-1} dG.
816: \end{equation}
817: %
818: A computation shows that if $\Phi$ solves
819: \eqref{eq:IVP} with triple
820: $(\xi,\Phi_0 ,\tilde{z}_0 )$ and
821: $G \in \gauge$ then the triples
822: $(\xi,\Phi_0 ,\tilde{z}_0 )$ and
823: $(\xi.G,\Phi_0 G(\tilde{z}_0),\tilde{z}_0 )$
824: induce the same CMC immersions, assuming $H \neq 0$.
825:
826: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
827:
828: \newsection{\bf Dressing.} \label{Dressing}
829: For $r \in (0,1]$, $h \in \LGC$ and
830: $F_0 \in \mathcal{F}(\MT)$ let
831: %
832: \begin{equation}\label{eq:dress}
833: hF_0 = F \, B
834: \end{equation}
835: %
836: be a point-wise Iwasawa decomposition of $hF_0$ in
837: $\LGC$, where every factor is analytic in $z$. Then
838: $F \in \mathcal{F}(\MT)$.
839:
840: Note that $F:\MT \to \LGU$ and
841: $B:\MT \to \LGP$ are not uniquely determined,
842: since for any smooth $U:\MT \to \U1$, we again
843: have an Iwasawa decomposition given by
844: $hF_0 = FU \,U^{-1}B$. We shall write
845: %
846: \begin{equation}
847: F \in [h\#F_0]
848: \end{equation}
849: %
850: to signify that $F:\MT \to \LGU$ satisfies
851: \eqref{eq:dress}. We call $[h\#F_0]$ the
852: \textit{dressing class} of $F_0$ under $h$
853: and say that $F \in [h\#F_0]$ was obtained
854: by \textit{dressing} $F_0 \in \mathcal{F}(\MT)$
855: by $h \in \LGC$.
856:
857: To preserve the base point condition
858: \eqref{eq:F_initial}, one needs to restrict
859: to dressing with elements $h \in \LGP$.
860:
861: \mylemma{}\label{th:initial}
862: Let $h \in \LGC$ and $F_0 \in \FID$ be a normalized
863: extended unitary frame. Then there exists a
864: normalized extended unitary frame in $[h\#F_0]$ if
865: and only if $h \in \LGP$.
866:
867: \Proof A straightforward computation,
868: see Proposition 2.9 of \cite{BurP:dre},
869: shows that the Maurer--Cartan
870: form of any element in $[h\#F_0]$
871: is again of the form \eqref{eq:MC}.
872: The issue is that for
873: $F_0 \in \FID$ we have
874: $F_0(\tilde{z}_0) = \Id$, while for
875: $F \in [h\#F_0]$ we have apriori only
876: $F(\tilde{z}_0) \in \U1$.
877:
878: If $hF_0 = FB$ is an Iwasawa decomposition,
879: then $F_0(\tilde{z}_0) = F(\tilde{z}_0) = \Id$,
880: implies $h = B(\tilde{z}_0) \in \LGP$.
881:
882: Conversely, if $hF_0 = FB$ is an
883: Iwasawa decomposition, then so is
884: %
885: \begin{equation}
886: hF_0 = F F(\tilde{z}_0)^{-1} F(\tilde{z}_0)B.
887: \end{equation}
888: %
889: Hence $F F(\tilde{z}_0)^{-1}\in [h\#F_0]$ and
890: satisfies \eqref{eq:F_initial}. Thus
891: $F F(\tilde{z}_0)^{-1} \in \FID$. \QED
892:
893: We thus have a left action of $\LGP$
894: on $\FID$ and shall write
895: $F = h \# F_0$ to signify that $F \in [h\#F_0]$
896: with $F(\tilde{z}_0) = \Id$. Evaluating $hF_0 =
897: h\#F_0 \, B$ at $\tilde{z}_0$ also gives
898: $h=B(\tilde{z}_0)$.
899: %
900: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
901:
902: \newsection{\bf Isotropies.}
903: We denote the isotropy group of a map
904: $F: \MT \to \LGU$ under dressing by
905: %
906: \begin{equation}
907: \mathrm{Iso}_r(F) = \left\{ h \in \LGC :
908: F \in [h\# F] \right\}.
909: \end{equation}
910: %
911: For a holomorphic frame $\Phi:\MT \to \LGC$
912: we define the isotropy under dressing by
913: %
914: \begin{equation}
915: \mathrm{Iso}_r(\Phi) = \{ g \in \LGC :
916: g \, \Phi = \Phi \, G \mbox{ for some } G
917: \in \mathcal{G}_r(\MT) \}.
918: \end{equation}
919: %
920: If $g \in \mathrm{Iso}_r(\Phi)$ with
921: $g\Phi = \Phi G$ and $\Phi = F\,B$
922: is an Iwasawa decomposition of $\Phi$,
923: then $gF = FBGB^{-1}$ with $BGB^{-1} \in \LGP$.
924: Hence $g \in \mathrm{Iso}_r(F)$. Conversely, if
925: $h \in \mathrm{Iso}_r(F)$, and $hF=FH$ is an Iwasawa
926: decomposition, then $h\Phi = \Phi B^{-1}HB$
927: and thus
928: %
929: \begin{equation} \label{eq:isotropies}
930: \mathrm{Iso}_r(\Phi) = \mathrm{Iso}_r(F).
931: \end{equation}
932: %
933: Further, using \eqref{eq:intersect} it is
934: straightforward to verify that
935: %
936: \begin{equation} \label{eq:iso_intersect}
937: \mathrm{Iso}_r(F) \cap \LGU = \U1 \mbox{ for }
938: F \in \FID.
939: \end{equation}
940: %
941: Umbilic points occur naturally on compact
942: CMC surfaces of genus $g \geq 2$ as well
943: as on complete, open CMC surfaces with more than
944: two ends. In this context, we quote
945:
946: \mytheorem{ \cite{DorH:dre}}
947: \label{th:isotropy}
948: If the associated family of
949: $F \in \mathcal{F}(\MT)$ has umbilics,
950: then $\mathrm{Iso}_r(F) = \{ \pm \Id \}$ for all
951: $r \in (0,\,1]$. \QED
952:
953: \section{Symmetries}
954: %
955: The notion of symmetry for CMC immersions has been
956: discussed in the articles
957: \cite{DorH:per,DorH:sym2}.
958: It turns out that symmetry can be defined on
959: various levels in the context of the DPW
960: representation.
961: The problem of dealing with symmetries in the
962: DPW representation stem mostly from the fact
963: that symmetries on the potential level are,
964: by and large, defined as coming from symmetries
965: on the immersion level and have not yet been
966: defined completely intrinsically on the potential
967: level. Since symmetries associated with
968: automorphisms $Aut(\MT)$ of the universal
969: cover are central to this article, we recall
970: the basic facts, retaining the notation of
971: section \ref{sec:DPW}.
972:
973: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
974: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
975:
976: \newsection{\bf Definitions.} \label{symmetries}
977: Let $f : \MT \to \Lsu$ be a CMC $H \neq 0$ immersion
978: generated by a triple $( \xi, \Phi_0, \tilde{z}_0 )$
979: with $\Phi$ its holomorphic frame and $F$ its
980: unitary frame. In the rest of
981: this section we will always assume
982: $\gamma \in Aut(\MT)$ and for
983: a map G with domain $\MT$ we shall write
984: $\gamma^* G = G \circ \gamma$.
985:
986: \n {\bf{Immersion level}}:
987: A triple
988: $(\gamma,\,\X,\,T)$ with
989: $\X \in \LGU$ and $T \in \LGe$
990: is called a symmetry of $f$ if and only if
991: %
992: \begin{equation}\label{eq:sym_immersion}
993: \gamma^* f = \X \, f \, \X^{-1}
994: + T.
995: \end{equation}
996: %
997:
998: \n {\bf{Unitary frame level}}:
999: A pair $(\gamma,\,\X)$ with
1000: $\X \in \LGU$
1001: is called a symmetry of $F$ if and
1002: only if there exists a smooth map
1003: $k \in C^\infty(\MT,\U1)$ such that
1004: %
1005: \begin{equation}\label{eq:sym_frame}
1006: \gamma^* F = \X \,F \,k.
1007: \end{equation}
1008: %
1009:
1010: \n {\bf{Holomorphic frame level}}:
1011: A pair $(\gamma,\,\X)$ with
1012: $\X \in \LGU$
1013: is a symmetry of $\Phi$ if and only if
1014: there is a map $H \in \gauge$ such that
1015: %
1016: \begin{equation}\label{eq:sym_curve}
1017: \gamma^* \Phi = \X \,\Phi \,H.
1018: \end{equation}
1019: %
1020:
1021: \n {\bf{Potential level}}:
1022: A pair $(\gamma,\,G)$ with
1023: $G \in \gauge$ is called a
1024: symmetry of $\xi$ if and only if
1025: %
1026: \begin{equation}\label{eq:sym_potential}
1027: \gamma^* \xi = \xi.G \,.
1028: \end{equation}
1029: %
1030: where $\xi.G$ denotes the gauge
1031: transformation of $\xi$ by $G$ as defined in
1032: \eqref{eq:gauge}.
1033: á
1034:
1035: \mytheorem{}\label{th:symmetries}
1036: Let $f : \MT \to \Lsu$ be a CMC $H\neq 0$ immersion
1037: generated by the triple
1038: $(\xi,\Phi_0,\tilde{z}_0 )$ with $\Phi$ and
1039: $F$ the holomorphic respectively unitary frame.
1040: %
1041: \begin{enumerate}
1042: \item If $(\gamma,\,\X_1,\,T_1)$ and
1043: $(\gamma,\,\X_2,\,T_2)$ are symmetries
1044: of $f$, then $\X_2 = \pm \X_1$ and
1045: $T_1 = T_2$.
1046: \item If $(\gamma,\,\X_1)$ and $(\gamma,\,\X_2)$
1047: are symmetries of $F$ then $\X_2 = \pm \X_1$.
1048: \item If $(\gamma,\,\X_1)$ and $(\gamma,\,\X_2)$
1049: are symmetries of $\Phi$ then $\X_2 = \pm \X_1$.
1050: \end{enumerate}
1051: %
1052: \Proof (i) If $(\gamma,\,\X_1,\,T_1)$ and
1053: $(\gamma,\,\X_2,\,T_2)$ are symmetries, then
1054: $\gamma^* f = \X_1 f \X_1^{-1} + T_1 =
1055: \X_2 f \X_2^{-1} + T_2$ implies
1056: $\X \, f \, \X^{-1} + T = f$ with
1057: $\X = \X_2^{-1}\X_1$ and
1058: $T = \X_2^{-1}(T_1-T_2)\X_2$. Since a CMC
1059: $H \neq 0$ surface in $\R^3$ is never contained
1060: in an affine plane,
1061: $\X \, f \, \X^{-1} + T = f$ implies
1062: $\X = \pm \Id$ and $T=0$.
1063:
1064: (ii) If $(\gamma,\,\X_1)$ and $(\gamma,\,\X_2)$
1065: are symmetries of $F$, then
1066: $\gamma^* F = \X_1 F k_1 = \X_2 F k_2$.
1067: Plugging $\X_1 F k_1$ and $\X_2 F k_2$ into
1068: the Sym--Bobenko formula and equating yields
1069: %
1070: \begin{equation}
1071: \X_1 f \X_1^{-1} - \tfrac{i\lambda}{2H}
1072: (\partial_{\lambda} \X_1 ) \X_1^{-1} =
1073: \X_2 f \X_2^{-1} - \tfrac{i\lambda}{2H}
1074: (\partial_{\lambda} \X_2 ) \X_2^{-1}.
1075: \end{equation}
1076: %
1077: Now $\X_2^{-1} \X_1 \,f \, \X_1^{-1} \X_2 +
1078: \tfrac{i\lambda}{2H} \X_2^{-1}
1079: \left( (\partial_{\lambda} \X_2) \X_2^{-1} -
1080: (\partial_{\lambda} \X_1 ) \X_1^{-1} \right) \X_2
1081: = f$ and part (i) implies $\X_2 = \pm \X_1$.
1082:
1083: (iii) If $(\gamma,\,\X_1)$ and
1084: $(\gamma,\,\X_2)$ are symmetries of $\Phi$ then
1085: $\gamma^* \Phi = \X_1 \Phi H = \X_2 \Phi G$
1086: implies $\X_2^{-1} \X_1 \Phi = \Phi G H^{-1}$
1087: and shows that
1088: $\X_2^{-1} \X_1 \in \mathrm{Iso}(\Phi)$.
1089: Let $\Phi = F B$ be an Iwasawa decomposition of
1090: $\Phi$. Then
1091: $\X_2^{-1} \X_1 \in \mathrm{Iso}(F)$, since
1092: $\mathrm{Iso}(\Phi) = \mathrm{Iso}(F)$ by
1093: \eqref{eq:isotropies}.
1094: Hence $\X_2^{-1} \X_1 = \pm \Id$
1095: by \eqref{eq:iso_intersect}. \QED
1096:
1097: For symmetries $(\gamma,\,\X_\gamma)$ and
1098: $(\mu,\,\X_\mu)$ of $F \in \mathcal{F}(\MT)$ and
1099: corresponding maps
1100: $k_\gamma,\,k_\mu$,
1101: we have $\mu^*\,\gamma^*\,F =
1102: \X_\gamma \, \X_\mu \,F\,k_\mu
1103: \,\mu^* k_\gamma$ and if we assume that
1104: $((\gamma \circ \mu),\,\X_{\gamma\mu})$
1105: is a symmetry, then part (ii) of Theorem
1106: \ref{th:symmetries} implies $\X_{\gamma \mu} =
1107: \pm \X_\gamma \, \X_\mu$ and consequently
1108: %
1109: \begin{equation}
1110: k_{\gamma \mu} = \pm k_\mu \, \mu^* k_\gamma.
1111: \end{equation}
1112: %
1113:
1114: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1115:
1116: \newsection{\bf Symmetries of $\Phi$.}
1117: We briefly investigate how symmetries
1118: of a holomorphic potential descend to symmetries
1119: of the corresponding holomorphic frame and note a
1120: simple consequence in case the potential is
1121: invariant. In view of these results and our
1122: subsequent inquiry into the relationship with the
1123: dressing action, we allow symmetries on the level
1124: of holomorphic frames of the form $(\gamma, X)$
1125: with $X \in \LGC$.
1126:
1127: \myprop{} Let $\gamma \in Aut(\MT)$ and
1128: $(\gamma,\,G)$ be a symmetry of
1129: $\xi \in \pott$. Let $\Phi:\MT \to \LGC$
1130: be a solution to $d\Phi = \Phi \xi$.
1131: Then there exists $X \in \LGC$
1132: such that $\gamma^* \Phi = X \Phi G$.
1133:
1134: \Proof Since $\gamma^* \Phi$ and
1135: $\Phi G$ both solve the differential equation
1136: $dY = Y\gamma^*\xi$, there exists
1137: $X \in \LGC$ with $\gamma^*\Phi = X \Phi G$.
1138: More explicitly, if $\Phi$ solves
1139: $d\Phi = \Phi \xi,
1140: \, \Phi(\tilde{z}_0) = \Phi_0$, then
1141: $X = \Phi(\gamma(\tilde{z}_0))
1142: G(\tilde{z}_0)^{-1} \Phi_0^{-1}$.
1143: \QED
1144:
1145: We shall again call a pair $(\gamma,\,X)$ with
1146: $\gamma \in Aut(\MT)$ and $X \in \LGC$ a
1147: symmetry of a map $\Phi:\MT \to \LGC$
1148: if there exists a $G \in \gauge$ such that
1149: $\gamma^* \Phi = X\,\Phi \,G$.
1150:
1151: \mycorollary{} \label{th:monophi}
1152: Let $\gamma \in Aut(\MT)$ and
1153: $\xi\in \pott$ with $\gamma^* \xi = \xi$.
1154: Let $\Phi$ be a solution to $d\Phi = \Phi \xi$.
1155: Then there exists $X \in \LGC$ such that
1156: $\gamma^* \Phi = X \Phi$. \QED
1157:
1158:
1159: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1160:
1161: \newsection{\bf Symmetries of $F$.}
1162: In view of the previous section,
1163: we now characterize how symmetries
1164: of the holomorphic frame descend to symmetries
1165: of the corresponding extended unitary frame
1166: and show that for open Riemann surfaces, the
1167: co-cycle factor in equation \eqref{eq:sym_frame}
1168: can be gauged away.
1169:
1170: \mylemma{}\label{th:symF}
1171: Let $(\gamma,\,X)$ be a symmetry
1172: of $\Phi:\MT \to \LGC$. Let
1173: $\Phi = F\,B$ be an Iwasawa decomposition
1174: of $\Phi$. Then there exists an $\X \in \LGU$
1175: such that $(\gamma,\X)$ is a symmetry of
1176: $F$ if and only if there exists an element
1177: $L \in \mathrm{Iso}_r(\Phi)$ with $LX^{-1} \in \LGU$.
1178:
1179: \Proof Let $G \in \gauge$ such that
1180: $\gamma^* \Phi = X \Phi G$. If $(\gamma,\X)$
1181: with $\X \in \LGU$ is a symmetry of $F$, then by
1182: \eqref{eq:sym_frame} there
1183: exists a differentiable map $k:\MT \to \U1$
1184: such that $\gamma^* F = \X F k$. In combination
1185: with $F = \Phi B^{-1}$, after rearranging,
1186: we obtain
1187: %
1188: \begin{equation}
1189: \X^{-1} X \, \Phi =
1190: \Phi B^{-1}k\gamma^*B \, G^{-1}.
1191: \end{equation}
1192: %
1193: Then $L:= \X^{-1} X \in \mathrm{Iso}_r(\Phi)$ as
1194: $B^{-1}k\gamma^*B \, G^{-1} \in \gauge$ and
1195: $L \, X^{-1} = \X^{-1} \in \LGU$.
1196:
1197: Conversely, let $L \in \mathrm{Iso}_r(\Phi)$ and
1198: $L \, X^{-1} \in \LGU$. Then there exists
1199: an element $H \in \gauge$ with
1200: $L\, \Phi = \Phi H$. By \eqref{eq:isotropies},
1201: $L \in \mathrm{Iso}_r(F)$ and $LF=FV$ for
1202: $V = BHB^{-1} \in \gauge$. Then
1203: $\gamma^* \Phi = X \Phi G$ yields
1204: %
1205: \begin{equation}
1206: \gamma^* F = X \,F\, B\, G\, \gamma^* B^{-1} =
1207: X\, L^{-1}F\, V\, B \,G\, \gamma^*B^{-1}.
1208: \end{equation}
1209: %
1210: Define $\X:=X\,L^{-1}$ and
1211: $k:=VBG\gamma^*B^{-1}$.
1212: On the one hand, $k \in \gauge$
1213: while on the other hand
1214: $k= F^{-1}\X^{-1}\gamma^*F$
1215: takes values in $\LGU$, since
1216: $\X \in \LGU$ by assumption. From equation
1217: \eqref{eq:intersect} we conclude that
1218: $k:\MT \to \U1$. \QED
1219: %
1220:
1221: \mytheorem{}
1222: Let $M$ be an open Riemann surface
1223: with Fuchsian group $\Gamma$ and let
1224: $f_\lambda: \MT \to \R^3$ be an associated family
1225: of CMC $H \neq 0$ immersions.
1226: Then there exists an extended unitary frame
1227: $F \in \mathcal{F}(\MT)$ for $f_\lambda$ such
1228: that for every $\gamma \in \Gamma$ there exists
1229: $\X \in \Lhol$ such that $\gamma^*F = \X \,F$.
1230:
1231: \Proof First we apply \cite{DorH:sym2},
1232: Theorem 2.3 and infer that there
1233: exists some extended frame
1234: $\tilde{F} \in \mathcal{F}(\MT)$ for $f_\lambda$
1235: such that for every $\gamma \in \Gamma$ we have
1236: $\gamma^* \tilde F =
1237: \X \,\tilde{F} \,k$, with $\X \in \Lhol$ and
1238: $k:\MT \to \U1$ as in \eqref{eq:sym_frame}.
1239: Writing $\tilde{F} = \Phi \,B$ with
1240: $\Phi:\MT \to \LGC$ holomorphic and
1241: $B \in \gauge$, we obtain
1242: $\gamma^* \Phi = \X\,\Phi\,W$
1243: with $W \in \gauge$. It is easy to see that
1244: $W$ satisfies the co-cycle condition. Therefore,
1245: since the open Riemann surface $\MT$ is Stein,
1246: by \cite{Bun}, $W$ is a co-boundary and whence
1247: $W = W\,\gamma^*W^{-1}$. Hence
1248: $\tilde{\Phi} = \Phi \,W$ satisfies
1249: $\gamma^* \tilde{\Phi} = \X \, \tilde{\Phi}$.
1250: Iwasawa splitting $\tilde{\Phi} = F \hat{B}$ with
1251: $\hat{B}$ normalized such that the $\lambda^0$
1252: coefficient in $\hat{B}$ has positive real entries
1253: we obtain $\gamma^* F = \X\,F$. \QED
1254:
1255: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1256:
1257: \newsection{{\bf Symmetries \& Dressing.}
1258: \label{sec:sym}
1259: Next, we characterize how symmetries between two
1260: dressing equivalent unitary frames are related.
1261:
1262: \mytheorem{} \label{th:sym1}
1263: Let $F_0 \in \mathcal{F}(\MT)$ have symmetry
1264: $(\gamma,\X_0)$ and $h \in \LGC$. Then
1265: $F \in [h \# F_0]$ has symmetry $(\gamma,\X)$
1266: if and only if there exists a
1267: $L \in \mathrm{Iso}_r(F)$ such that
1268: $\X = h \X_0 h^{-1} L^{-1}$.
1269: Furthermore, if both $F_0,\,F \in \FID$, then
1270: $L \in \LGP$.
1271:
1272: \Proof
1273: %
1274: Writing $hF_0 = F B$ for an Iwasawa decomposition
1275: and using $\gamma^* F_0 = \X_0 F_0 k_0$, we obtain
1276: $h \X_0 h^{-1} F =
1277: \gamma^* F \gamma^* B k_0^{-1} B^{-1}$.
1278: If $(\gamma,\X)$ is a symmetry of $F$ then
1279: $\gamma^* F = \X F k$, whence
1280: %
1281: \begin{equation} \label{eq:iso_con}
1282: \X^{-1} h \X_0 h^{-1} F =
1283: F k \gamma^*B k_0^{-1} B^{-1}.
1284: \end{equation}
1285: %
1286: Set $L :=\X^{-1} h \X_0 h^{-1}$ and
1287: $H := k \gamma^*B k_0^{-1} B^{-1}$. Clearly
1288: $L \in \LGC$ and $H \in \gauge$, so equation
1289: \eqref{eq:iso_con} reads $LF=FH$ and is an
1290: Iwasawa decomposition of $LF$. Hence
1291: $L \in \mathrm{Iso}_r(F)$, and by construction
1292: $\X = h \X_0 h^{-1} L^{-1}$.
1293:
1294: Conversely, assume there exists a map
1295: $L \in \mathrm{Iso}_r(F)$
1296: such that $L F = F H$ for some $H \in \gauge$
1297: and $h \X_0 h^{-1} L^{-1} \in \Lhol$.
1298: Defining $\X = h \X_0 h^{-1} L^{-1}$ and
1299: $k = H B k_0 \gamma^* B ^{-1}$, a computation
1300: yields $\gamma^* F = \X F k$.
1301: A priori, $k:\MT \to \LGP$, but as
1302: $\gamma^* F,\, \X$ and $F$ take values
1303: in $\Lhol$ so does $k$.
1304: Hence $k:\MT \to \U1$ and consequently
1305: $(\gamma,\X)$ is a symmetry of $F$.
1306: \QED
1307:
1308: \mycorollary{}
1309: If in addition to the assumptions of Theorem
1310: \ref{th:sym1}, $\mathrm{Iso}_r(F) = \{ \pm \Id \}$,
1311: then $(\gamma,\X)$ is a symmetry of
1312: $F$ if and only if $\X = \pm h\, \X_0 h^{-1}$.
1313:
1314: \Proof
1315: The assumption $\mathrm{Iso}_r(F) =
1316: \{ \pm \Id \}$ implies that for the maps $L,\,H$ in the
1317: proof of Theorem \ref{th:sym1} we have
1318: $L = \Id$ and $H \equiv \Id$, yielding the claim.
1319: \QED
1320:
1321: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1322:
1323: \newsection{\bf Groups of symmetries.}
1324: Extending the results of section \ref{sec:sym}
1325: to groups $\Gamma \subset Aut(\MT)$ of symmetries
1326: one obtains:
1327:
1328: \mytheorem{} \label{th:mainresult}
1329: Let $F_0 \in \mathcal{F}_r(\MT)$ admit symmetries
1330: $(\gamma,\X_0(\gamma))$ for all
1331: $\gamma \in \Gamma$ and let $h \in \LGC$.
1332: Then $F \in [h \# F_0]$
1333: has symmetries $(\gamma,\X(\gamma))$ for all
1334: $\gamma \in \Gamma$ if and only if there exist
1335: maps $L(\gamma) \in \mathrm{Iso}_r(F)$ such that
1336: $\X(\gamma)=h\X_0(\gamma) h^{-1} L(\gamma)^{-1}$.
1337: Moreover, if $\mathrm{Iso}_r(F) = \{ \pm \Id \}$,
1338: then $F$ admits symmetries $(\gamma,\X(\gamma))$
1339: for all $\gamma \in \Gamma$ if and only if
1340: $\X(\gamma) = h \X_0(\gamma) h^{-1}$ for all
1341: $\gamma \in \Gamma$. \QED
1342:
1343: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1344:
1345: \newsection{\bf Symmetries \& Isotropy}
1346: Let $(\gamma,\X)$ be a symmetry of
1347: $F \in \mathcal{F}(\MT)$ and
1348: $\gamma^* F = \X F k$.
1349: If $L \in \mathrm{Iso}_r(F)$ and $LF = FB$ for an
1350: Iwasawa decomposition, then
1351: $\X^{-1}L \X F = F k\gamma^*Bk^{-1}$.
1352: As $(k\gamma^*Bk^{-1})$ is $\LGP$--valued,
1353: $\X^{-1} L \X \in \mathrm{Iso}_r(F)$
1354: and every symmetry
1355: $(\gamma,\X)$ induces an inner automorphism
1356: $S:\mathrm{Iso}_r(F) \to \mathrm{Iso}_r(F),\,
1357: L \mapsto \X^{-1} L \X$.
1358:
1359: Now consider $F_0 \in \FID,\,
1360: h \in \LGP$ and let $F = h \# F_0$. Assume that
1361: $(\gamma,\,\X_0)$ and $(\gamma,\,\X)$ are
1362: symmetries of $F_0$ respectively $F$.
1363: If further $h \in \mathrm{Iso}_r(F)$, then we
1364: obtain a map $S : \mathrm{Iso}_r(F) \to
1365: \mathrm{Iso}_r(F),\,h \mapsto
1366: \X^{-1} h \X_0 h^{-1}$ with the property
1367: $S(h_1 h_2) = S(h_1) h_1 S(h_2) h_1^{-1}$.
1368:
1369: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1370: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1371:
1372: \section{Invariant potentials \& Monodromy}
1373: %
1374: We start our investigation of the relationship
1375: between monodromy and dressing by defining the
1376: notion of monodromy on the level of
1377: holomorphic and unitary frames in the DPW framework.
1378:
1379: \mydefinition{} \label{def:monodromies}
1380: Let $M$ be a connected Riemann surface
1381: with universal cover $\MT$ and $\Delta$
1382: the group of deck transformations.
1383: Let $\xi \in \pott$ and $\Phi: \MT \to \LGC$ be a
1384: solution of $d \Phi = \Phi \xi$ and
1385: $\tau \in \Delta$. A loop
1386: $\varrho(\tau) \in \LGC$ such that
1387: $(\tau,\,\varrho(\tau))$ is a symmetry of
1388: $\Phi$ will be called a \textit{monodromy} of
1389: $\Phi$ with respect to $\tau$.
1390: Let $\Phi = F\,B$ be an Iwasawa decomposition.
1391: A loop $\chi(\tau) \in \LGU$ such that
1392: $(\tau,\,\chi(\tau))$ is a symmetry of
1393: $F$ is called a \textit{monodromy} of $F$
1394: with respect to $\tau$.
1395:
1396: \myremark{} A monodromy of a normalized extended
1397: unitary frame $F \in \FID$ takes values in $\Lhol$.
1398: %
1399: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1400: %
1401:
1402: \newsection{\bf Existence of monodromies.}
1403: Apriori, it is not clear that
1404: for some $\tau \in \Delta$ there exist monodromies
1405: $\varrho(\tau)$ and
1406: $\chi(\tau)$ as in the definitions above.
1407: In case the
1408: corresponding potentials are invariant under
1409: $\Delta$ we can evoke Corollary \ref{th:monophi}
1410: and obtain the following sufficient conditions
1411: for the existence of monodromies.
1412:
1413: \myprop{} (i) Let $\xi \in \pott$ and
1414: $\Phi$ be a solution of $d\Phi = \Phi\,\xi$ with
1415: initial condition $\Phi_0 \in \LGC$.
1416: Let $\tau \in \Delta$ with $\tau^*\xi = \xi$.
1417: Then there exists a monodromy
1418: $\varrho(\tau) \in \LGC$ such that
1419: $\tau^* \Phi = \varrho(\tau) \,\Phi$.
1420:
1421: (ii) Let $F:\MT \to \LGU$ and $\alpha = F^{-1}dF$.
1422: If $\tau^*\alpha = \alpha$ for $\tau \in \Delta$,
1423: then there exists a monodromy $\chi(\tau) \in \LGU$
1424: such that $\tau^* F = \chi(\tau) \,F$. \QED
1425:
1426: It is shown in \cite{DorH:cyl} that CMC
1427: immersions of open Riemann surfaces $M$
1428: can always be generated by $\Delta$--invariant
1429: potentials $\xi \in \pott$.
1430: If $\Phi = F B$ is the Iwasawa
1431: decomposition of a holomorphic frame
1432: $\Phi : \MT \to \LGC$, then we shall need
1433: to study the monodromy of $F$ in order to control
1434: the periodicity of the
1435: resulting immersion \eqref{eq:Sym}.
1436: Even if $F$ is obtained from a potential that
1437: is invariant under $\Delta$, we are a priori
1438: not assured that there exist loops
1439: $\chi(\tau) \in \LGU$ and maps
1440: $k(\tau):\MT \to \U1$ such that
1441: equation \eqref{eq:sym_frame} holds for all
1442: $\tau \in \Delta$ or even more strongly,
1443: that for $\alpha = F^{-1}dF$ we have
1444: $\tau^* \alpha = \alpha$ for all
1445: $\tau \in \Delta$. In analogy to Theorem
1446: \ref{th:sym1}, this issue is characterized
1447: by the following
1448:
1449: \mylemma{} \label{th:iso_problem}
1450: Let $(\xi, \Phi_0, \tilde{z}_0 )$ generate the
1451: holomorphic frame $\Phi:\MT \to \LGC$ and
1452: extended unitary frame $F \in \mathcal{F}(\MT)$.
1453: Let $\tau \in \Delta$ and $\varrho(\tau)$
1454: be a monodromy of $\Phi$.
1455: Then the following are equivalent.
1456: %
1457: \begin{enumerate}
1458: \item There exists a monodromy
1459: $\chi(\tau) \in \LGU$ of $F$ with
1460: respect to $\tau$.
1461: %
1462: \item There exists an element
1463: $L(\tau) \in \mathrm{Iso}_r(\Phi)$
1464: with $L(\tau)\varrho^{-1}(\tau) \in \LGU$.
1465: \end{enumerate}
1466:
1467: \Proof We first show that (i) implies (ii).
1468: Let $\chi(\tau) \in \LGU$
1469: and $k(\tau):\MT \to \U1$ such that
1470: $\tau^* F = \chi(\tau) F\, k(\tau)$.
1471: Using $F = \Phi B^{-1}$ and
1472: $\tau^* \Phi = \varrho(\tau) \Phi$ and
1473: rearranging, we obtain
1474: $\chi(\tau)^{-1}\varrho(\tau) \, \Phi =
1475: \Phi\, B^{-1} k(\tau) \,\tau^* B$.
1476: Then $L(\tau) := \chi(\tau)^{-1}\varrho(\tau)
1477: \in \mathrm{Iso}_r(\Phi)$ and
1478: $L(\tau)\varrho(\tau)^{-1} \in \LGU$.
1479:
1480: Conversely, let $L(\tau) \in \mathrm{Iso}_r(\Phi)$
1481: and assume $L(\tau)\varrho(\tau)^{-1} \in \LGU$.
1482: Then there exists an element
1483: $G(\tau) \in \gauge$ with
1484: $L(\tau) \, \Phi = \Phi \, G(\tau)$.
1485: Multiplying this last equation on the right by
1486: $\tau^* \Phi^{-1}$ and rearranging gives
1487: %
1488: \begin{equation} \label{eq:iso_phi2}
1489: F^{-1}L(\tau) \, \varrho(\tau)^{-1} \,\tau^*F
1490: = B \, G(\tau) \, \tau^*B^{-1}.
1491: \end{equation}
1492: %
1493: The left respectively right hand side of
1494: \eqref{eq:iso_phi2} takes values in $\LGU$
1495: respectively $\LGP$.
1496: Hence, by \eqref{eq:intersect},
1497: both sides are $\lambda$--independent and
1498: $\U1$--valued. Set $k(\tau):= F^{-1}L(\tau)\,
1499: \varrho(\tau)^{-1}\tau^*F$
1500: and $\chi(\tau) := \varrho(\tau)\,L(\tau)^{-1}$.
1501: Then $\tau^* F = \chi(\tau)\, F\, k(\tau)$ and
1502: $\chi(\tau) \in \LGU$.
1503: \QED
1504:
1505: \mycorollary{} \label{th:chi=rho}
1506: Let $(\xi, \Phi_0, \tilde{z}_0 )$ generate the
1507: holomorphic frame $\Phi:\MT \to \LGC$ and
1508: extended unitary frame $F \in \mathcal{F}(\MT)$.
1509: Let $\tau \in \Delta$ and $\varrho(\tau)$
1510: be a monodromy of $\Phi$
1511: and $\chi(\tau)$ a monodromy of $F$ with
1512: respect to $\tau \in \Delta$.
1513: If $\mathrm{Iso}_r(F) = \{ \pm \Id \}$,
1514: then $\chi(\tau) = \pm \varrho(\tau)$. \QED
1515:
1516: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1517:
1518: \newsection{\bf A factorization Theorem.}
1519: In this section we work exclusively with
1520: untwisted loops.Consider an analytic map $H:\C^* \to \SL$,
1521: for which $\tr H$ is not independent of $\lambda$.
1522: Its eigenvalues are
1523: %
1524: \begin{equation}
1525: \mu_{\pm} = \frac{1}{2} \left( \tr H \pm
1526: \sqrt{(\tr H)^2 - 4}\right).
1527: \end{equation}
1528: %
1529: We will need to have $\mu_{\pm}$ holomorphic on
1530: some open and dense subset $\mathbb S$ of $\C^*$.
1531: To this end we slightly generalize the procedure
1532: used in \cite{DorH:cyl}.
1533:
1534: \myprop{}
1535: Let $g:\C^* \to \C,\,g \not\equiv 0$ be holomorphic.
1536: Then there exists some connected, open,
1537: and dense subset $\mathbb S \subset \C^*$
1538: such that there exists a well defined square
1539: root function $\sqrt{g}$ of $g$ on $\mathbb S$.
1540:
1541: \Proof
1542: Since $g(\lambda)$ does not vanish
1543: identically it has at most finitely
1544: many roots on $S^1$. Moving slightly inside,
1545: if necessary, we can choose any $0 < r \leq 1$
1546: such that $g$ does not vanish on $C_r$.
1547: We form a (scalar) Birkhoff splitting
1548: of $g$ on $C_r$: $g = g_- \lambda^N g_+$.
1549: It is easy to verify that, since $g$ is
1550: holomorphic on $\C^*$, also $g_-$ and
1551: $g_+$ are holomorphic on $\C^*$.
1552: Now we introduce some cuts: Consider
1553: the roots of $g_+$. They are all contained
1554: in the complement of $I_r$.
1555: We cut the Riemann sphere from every odd
1556: ordered root of $g_+$ to infinity.
1557: Similarly we cut the Riemann sphere from
1558: each odd ordered root of $g_-$ to the origin.
1559: Thus $g_-$ and $g_+$ have well defined
1560: root functions on the cut domain.
1561: If $N$ is odd, then we also need to
1562: introduce one further cut from the origin to the
1563: point at infinity. This resulting domain will
1564: be denoted by $\mathbb S$.
1565: \QED
1566:
1567: Applying this to the function
1568: $g(\lambda) = {\tr H}^2 -4$ we obtain
1569:
1570: \mycorollary{} \label{th:mu}
1571: The eigenvalue functions $\mu_{\pm}(\lambda)$
1572: of a loop $H \in \Lambda_{\C^*}\SL$ are analytic
1573: on a connected, open and dense set
1574: $\mathbb S \subset \C^*$, obtained by making
1575: radial branch cuts from odd ordered roots $\alpha$
1576: of ${\tr H}^2 -4$ to the origin
1577: respectively the point at infinity, depending
1578: on whether $|\alpha| <r$ respectively
1579: $|\alpha| >r $. In addition there is
1580: possibly a cut from $0$ to $\infty$.
1581: \QED
1582:
1583: For our approach it will be convenient
1584: to diagonalize $H$
1585: %
1586: \begin{equation} \label{eq:diag}
1587: Y\, H \, Y^{-1}= \mathrm{diag}[\mu_+,\,\mu_-]
1588: \end{equation}
1589: %
1590: If the matrix entries $H_{1,2}$ and $H_{2,1}$
1591: of $H$ do not vanish identically,
1592: then one can choose as diagonalizing matrix $Y$
1593: ( see e.g. (3.5.18) in \cite{DorH:cyl})
1594: %
1595: \begin{equation}
1596: Y = \begin{pmatrix}
1597: 1 & -\frac{H_{12}}{u-iv-H_{11}}\\
1598: -\frac{H_{21}}{2iv}& \frac{u + iv-H_{22}}{2iv}
1599: \end{pmatrix},
1600: \end{equation}
1601: %
1602: where $H_{ij}$ denotes the $(i,j)-$coefficient
1603: of $H$ and $ 2u = \tr H$ and $ v^2 = u^2 -1$.
1604: If one of the off-diagonal coefficients vanishes
1605: identically, then we choose the corresponding
1606: matrix $Y$ listed in \cite{DorH:cyl}. In this
1607: case the diagonal entries are $\mu_{\pm}$. Note
1608: that this works in our setting.(Not only in the
1609: somewhat more special setting of \cite{DorH:cyl}.)
1610:
1611: From the specific form of the matrices
1612: $Y$ mentioned above we infer that $Y$
1613: is meromorphic on $\mathbb S $, provided
1614: the trace $\tr H$ of $H$ is not independent
1615: of $\lambda$.
1616:
1617: The next result is crucial for our
1618: characterization of dressing matrices
1619: preserving the fundamental group. As before,
1620: we denote an open annulus about $S^1$ by
1621: $A_r = \{ \lambda \in \C : r<|\lambda|<1/r \}$ and
1622: by $C_r = \{ \lambda \in \C : |\lambda| = r \}$
1623: a circle of radius $r$. With the above notations
1624: we have the following
1625:
1626: \mytheorem{} \label{th:h-split}
1627: Let $h \in \Lambda_r^+\SL$ for some $0<r\leq1$ and
1628: $H \in \Lambda_{\C^*}\SL$ such that the trace
1629: of $H$ is not constant.
1630: Let $\mathbb S$ be as in Corollary \ref{th:mu}.
1631: If $h\, H \,h^{-1}$ is meromorphic on
1632: $\mathbb S \cap A_s$ for some
1633: $0 < s < r$, then $h$ can be factored, on a possibly
1634: segmented circle $C_l \cap \mathbb S$ for
1635: $s\leq l \leq r$, into
1636: %
1637: \begin{equation}\label{eq:h-split}
1638: h = \mathcal{M}\, \mathcal{C},
1639: \end{equation}
1640: %
1641: where $\mathcal{C}$ is $\SL$--valued
1642: on $\mathbb S\cap C_l$ and
1643: $[\, \mathcal{C},\, H \,] = 0$ there. Moreover,
1644: $\mathcal{M}$ is $\SL$--valued and meromorphic
1645: on the connected open set $\mathbb S \cap A_s$.
1646:
1647: \Proof From the above discussion,
1648: there exists some $Y$
1649: such that \eqref{eq:diag}
1650: holds. Let us write
1651: %
1652: \begin{equation}
1653: h Y^{-1} = \begin{pmatrix}
1654: a & b \\
1655: c & d \end{pmatrix}.
1656: \end{equation}
1657: %
1658: Computing $h\,H \,h^{-1}$ and evaluating
1659: the diagonal entries gives that both
1660: %
1661: \begin{align}
1662: &ad\mu_- -bc\mu_+ , \label{eq:r22}\\
1663: &ad\mu_+ - bc\mu_- \label{eq:r11}
1664: \end{align}
1665: %
1666: are meromorphic on $\mathbb S \cap A_s$.
1667: Multiplying \eqref{eq:r22} by $\mu_+$ and
1668: \eqref{eq:r11} by $\mu_-$ and subtracting
1669: the resulting equations,
1670: implies that $bc(\mu_+^2 - \mu_-^2)$
1671: is meromorphic on $\mathbb S \cap A_s$.
1672: Multiplying \eqref{eq:r22} by $\mu_-$ and
1673: \eqref{eq:r11} by $\mu_+$ and subtracting the
1674: resulting equations, implies that
1675: $ad(\mu_-^2 - \mu_+^2)$ is
1676: meromorphic on $\mathbb S \cap A_s$. Assuming
1677: $\mu_-^2 - \mu_+^2 = 0$ would imply
1678: $\mu_{\pm}^4 = 1$ and contradict the assumption
1679: that $\tr H$ is not independent of $\lambda$.
1680: Since $\mu_{\pm}$ are
1681: holomorphic on $\mathbb S \cap A_s$ both $ad$
1682: and $bc$ are meromorphic on $\mathbb S \cap A_s$.
1683: Similarly, evaluating the off--diagonal terms of
1684: $h\,H \,h^{-1}$ we obtain that both $ab$
1685: and $cd$ are meromorphic on $\mathbb S \cap A_s$.
1686:
1687: If $d \neq 0$ on some circle
1688: $C_l$, for $s \leq l \leq r$, then we may write
1689: $h\,Y^{-1}=\widehat{h}\,\mathrm{diag}[\,1/d,\,d\,]$
1690: with
1691: %
1692: \begin{equation} \label{eq:h-split_case1}
1693: \widehat{h} := \begin{pmatrix}
1694: ad & \tfrac{b}{d} \\ cd & 1 \end{pmatrix}
1695: \mbox{ meromorphic on $\mathbb S\cap A_s$.}
1696: \end{equation}
1697: %
1698: We define $\mathcal{C} :=
1699: Y^{-1} \,\mathrm{diag}[\,1/d, d\,]\,Y$ and
1700: $\mathcal{M} := \widehat{h}\, Y$.
1701: Then $h = \mathcal{M}\, \mathcal{C}$
1702: with $\mathcal{C} = \mathcal{C}(\lambda)$
1703: defined on $C_l \cap \mathbb S$ and
1704: $[\,\mathcal{C},\,H \,] = 0$ there. Moreover,
1705: $\mathcal{M}$ is meromorphic on
1706: $\mathbb S \cap A_s$.
1707:
1708: If $d \equiv 0$ on an arc $C_l \cap \mathbb S$ for
1709: some $s \leq l \leq r$, then both
1710: $cd \equiv 0$ on this arc and consequently
1711: $cd \equiv 0$ in $\mathbb S \cap I_r$,
1712: and $ad \equiv 0$ on this arc and consequently
1713: $ad \equiv 0$ in $\mathbb S \cap I_r$.
1714: Thus $d \equiv 0$ on $C_l \cap \mathbb S$
1715: implies that both $b,\,c \neq 0$ on
1716: $\mathbb S \cap I_r$ and we may write
1717: $h\,Y^{-1} = \widetilde{h}\,
1718: \mathrm{diag}[\,c,\,1/c\,]$
1719: with
1720: %
1721: \begin{equation*} \label{eq:h-split_case2}
1722: \widetilde{h} = \begin{pmatrix}
1723: a/c & bc \\ 1 & cd
1724: \end{pmatrix}
1725: \mbox{ meromorphic on $\mathbb S\cap A_s$.}
1726: \end{equation*}
1727: %
1728: Set
1729: $\mathcal{C} = Y^{-1} \mathrm{diag}
1730: [\, c, 1/c \,] Y$ and
1731: $\mathcal{M} = \widetilde{h} Y$. Then
1732: $h = \mathcal{M}\, \mathcal{C}$ with
1733: $[\, \mathcal{C},\,\chi \,] =0$
1734: and $\mathcal{M}$ meromorphic on
1735: $\mathbb S\cap A_s$. \QED
1736:
1737: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1738:
1739: Given an open dense subset $\mathbb S \subset \C^*$,
1740: obtained by making branch cuts from the points
1741: $\{ \alpha_j \} \subset \C^*$ to the origin
1742: respectively $\infty$, depending on whether
1743: $|\alpha_j|<r$ respectively $|\alpha_j| > r$, let
1744: $\hat{\mathbb S}$ be the set obtained by making
1745: branch cuts according to this rule from the points
1746: $\{ 1/\bar{\alpha}_j \}$. Then the set
1747: ${\mathbb S}^* = \mathbb S \cap \hat{\mathbb S}$ is
1748: invariant under $\lambda \mapsto 1/\bar{\lambda}$.
1749: Note though that if there are at least two distinct
1750: $\alpha_j \in S^1$, then $\mathbb S^*$ is not
1751: connected.
1752:
1753: We now improve on these results by exploiting the
1754: reality condition of loops in $\Lambda_{\C^*} \SU$.
1755:
1756: \mylemma{} \label{th:S}
1757: The eigenvalue functions $\mu_{\pm}$ of a loop
1758: $H \in \Lambda_{\C^*}\SU$
1759: are of the form $\mu_{\pm} = u \pm iv$,
1760: where $2u = \tr H$ and
1761: $v^2 =1 - u^2 = \lambda^N \cdot \hat{v}^2$ with
1762: $N \in 2\Z$ and $\hat{v}$ holomorphic on
1763: $\mathbb S$. In particular,
1764: $S^1 \subset \mathbb S = {\mathbb S}^*$.
1765:
1766: \Proof A loop $H \in \Lambda_{\C^*}\SU$
1767: is of the form
1768: %
1769: \begin{equation*}
1770: H(\lambda) = \begin{pmatrix}
1771: x(\lambda) & y(\lambda) \\
1772: -\overline{y(1/\bar{\lambda})} &
1773: \overline{x(1/\bar{\lambda})} \end{pmatrix}
1774: \end{equation*}
1775: %
1776: with analytic functions $x,\,y:\C^* \to \C$. Hence
1777: $\left. u \right|_{S^1} \in [-1,\,1]$.
1778: Consequently, the function $v^2$ is real and
1779: non-negative on $S^1$ and thus has only even
1780: roots on $S^1$. If
1781: $\{ \lambda_1,\,\ldots\,,\lambda_m \}$ are the
1782: roots of $v^2$ on $S^1$, we may write
1783: %
1784: \begin{equation*}
1785: v^2 = \left( \prod_{j=1}^m
1786: (\lambda - \lambda_j)^{2K_j} \right) \, s
1787: \end{equation*}
1788: %
1789: where $s:\C^* \to \C$ is analytic and
1790: $\left.s\right|_{S^1} \neq 0$. We write
1791: $s = c_0\,s_- \, \lambda^N \, s_+$
1792: for the scalar Birkhoff decomposition of
1793: $s$ on $S^1$, with $c_0 \in \C^*$ and $N \in \Z$,
1794: such that $s_-(\infty)=s_+(0)=1$.
1795: With these normalizations, the reality condition
1796: $u^* = u$ implies $s_+^* = s_-$ as well as
1797: %
1798: \begin{equation*}
1799: N = - \sum_{j=1}^m K_j \mbox{ and }
1800: \overline{c_0} = c_0 \, \prod_{j=1}^m
1801: \lambda_j^{2K_j}.
1802: \end{equation*}
1803: %
1804: Whence also $s_+^*(\infty)=1$ and
1805: $v^2 = \lambda^N \hat{v}^2$ with
1806: %
1807: \begin{equation}
1808: \hat{v}^2 = c_0\,s_-\,s_+ \, \prod_{j=1}^m
1809: (\lambda - \lambda_j)^{2K_j}.
1810: \end{equation}
1811: %
1812: Since $s$ is analytic in $\C^*$, both $s_{\pm}$
1813: are analytic in $\C^*$, in fact $s_+(\lambda)$
1814: is entire, and thus also $\hat{v}^2$ is analytic
1815: in $\C^*$. There are two cases depending on the
1816: parity of $N \in \Z$. If $N$ is even, then
1817: making branch-cuts from the odd ordered
1818: roots of $s_+$ to the point at infinity and
1819: branch-cuts from the odd ordered roots of $s_-$ to
1820: the origin gives an open and dense set
1821: $\mathbb S \subset \C^*$ on which $\hat{v}^2$ has
1822: an analytic square root. Since the roots of
1823: $\hat{v}^2$ on $S^1$ are all even, we have that
1824: also $v^2$ has an analytic square root on
1825: $\mathbb S$ and $S^1 \subset \mathbb S$.
1826: Further, $s_+^* = s_-$ implies
1827: $\mathbb S^* = \mathbb S$.
1828:
1829: Assume that $N \in \Z$ is odd. Then $\mathbb S$
1830: is the set obtained as above but with an
1831: additional branch cut from the origin to
1832: the point at infinity. Now $u=\sqrt{\tr^2H-4}$
1833: has an absolutely convergent power series expansion
1834: on $S^1$ and is thus well defined, so there can not
1835: be a branch cut through $S^1$. Hence $N$ is even.
1836: \QED
1837:
1838: \mycorollary{} \label{co:h-split}
1839: Let $h \in \Lambda_r^+ \SL$ for some $0<r\leq 1$
1840: and $H \in \Lambda_{\C^*}\SU$ with non-constant
1841: trace function and let $\mathbb S$ be as in Lemma
1842: \ref{th:S}. If $h\, H \,h^{-1}$
1843: is meromorphic on $\mathbb S \cap A_s$ for some
1844: $0 < s < r$ and in addition we require that
1845: $(h\, H \,h^{-1})^* = (h\, H \,h^{-1})^{-1}$
1846: for all $\lambda \in \mathbb S \cap A_s$, then
1847: %
1848: \begin{equation}
1849: h = U\,C
1850: \end{equation}
1851: %
1852: with meromorphic loops $U,\,C$ on
1853: a cut but connected annulus respectively
1854: a possibly segmented circle $C_l$ for some
1855: $s<l<r$ with $[\,C,\,H\,]=0$ there, and
1856: $U \in \Lambda_{r'} \,\SU$ for some
1857: $r' \in (0,\,1)$.
1858:
1859: \Proof
1860: Note that we have $ S^1 \subset \mathbb S = {\mathbb S}^*$
1861: under our assumptions.
1862: We apply Theorem \ref{th:h-split} and decompose
1863: $h = \mathcal{M} \,\mathcal{C}$ on $C_l$ for some
1864: $s<l<r$ with $\mathcal{M}$ meromorphic on
1865: $\mathbb S \cap A_s$ and
1866: $[\, \mathcal{C} ,\,H ]=0$.
1867: Consequently, we have $h\,H \,h^{-1} =
1868: \mathcal{M}\,H\,\mathcal{M}^{-1}$ on
1869: $\mathbb S \cap A_s$ and
1870: $(h\, H \,h^{-1})^* =
1871: (h\, H \,h^{-1})^{-1}$
1872: is equivalent to
1873: %
1874: \begin{equation}
1875: [\,\mathcal{M}^* \,\mathcal{M} ,\,
1876: H \,]=0
1877: \end{equation}
1878: %
1879: on $\mathbb S\cap A_s$. Next, we seek a loop $L$,
1880: meromorphic on a subset of $\mathbb S \cap A_s$,
1881: of the form $L = f \, \Id + g \, \mathcal{M}^*\mathcal{M}$
1882: with meromorphic functions $f,\,g$ on
1883: $\mathbb S \cap A_s$, such that
1884: %
1885: \begin{equation} \label{eq:Luni}
1886: (\mathcal{M} L^{-1})^* =
1887: (\mathcal{M} L^{-1})^{-1}.
1888: \end{equation}
1889: %
1890: Then $h = \mathcal{M} L^{-1} L \mathcal{C}$
1891: is the desired factorization.
1892: Equation \eqref{eq:Luni} is equivalent to
1893: $\mathcal{M}^* \mathcal{M} = L^* L$ and by setting
1894: $P := \mathcal{M}^* \mathcal{M}$, yields
1895: %
1896: \begin{equation} \label{eq:P} \begin{split}
1897: P &= ff^* \Id + (f^* g + f\,g^*)P + g\,g^* P^2 \\
1898: &= (ff^* - gg^*)\Id + (f^*g + fg^* + gg^* \tr(P))P
1899: \end{split}
1900: \end{equation}
1901: %
1902: since $P^2 = \tr(P)P - \Id$ by Cayley-Hamilton.
1903: The Ansatz $f=f^*=g$ reduces \eqref{eq:P} to
1904: %
1905: \begin{equation}\label{eq:g^2}
1906: g^2 = \frac{1}{2 + \tr P}.
1907: \end{equation}
1908: %
1909: If we denote the entries of $\mathcal{M}$ by
1910: $m_{ij}$, then the function
1911: %
1912: \begin{equation}
1913: \tr P(\lambda) = \sum_{i,j=1}^2 m_{ij}(\lambda)
1914: {m_{ij}^*(\lambda)}
1915: \end{equation}
1916: %
1917: is meromorphic on $\mathbb S \cap A_s$ and
1918: is real and positive on $S^1\subset \mathbb S \cap A_s$ .
1919: With the help of a scalar Birkhoff decomposition of $g^2$
1920: it is straightforward to see that
1921: there exists a well defined square root of \eqref{eq:g^2}
1922: on $S^1$, which extends meromorphically to the set
1923: $\mathbb S_P \subset \mathbb S$, obtained by making branch-cuts
1924: from the odd ordered roots
1925: of $2 + \tr P$ in the usual fashion.
1926: Note that $S^1 \subset \mathbb S_P$ and
1927: $\mathbb S_P^* = \mathbb S_P$.
1928: Now that $g = (2 + \tr P)^{-1/2}$ is
1929: meromorphic on $\mathbb S_P \cap A_s$, we can set
1930: $U:=\mathcal{M}L^{-1}$, which is meromorphic on
1931: $\mathbb S_P \cap A_s$ with $U^* = U^{-1}$,
1932: and $C : = L\,\mathcal{C}$, which is defined on
1933: $C_l \cap \mathbb S_P$ with $[\,C,\,H\,]=0$.
1934: It remains to show that there exists
1935: an $r' \in (0,\, 1)$ such that
1936: $U \in \Lambda_{r'} \,\SU$.
1937:
1938: We choose $0<r'<1$ such that both of the
1939: following conditions hold:
1940:
1941: (i) The only poles of $\mathcal{M}$ in
1942: $A_{r'} \cap \mathbb S$ are on $S^1$,
1943:
1944: (ii) $A_{r'} \cap \mathbb S_P = A_{r'}$.
1945:
1946: Denote the pole set of $\mathcal{M}$ in
1947: $A_{r'}$ by
1948: $\mathcal{P} = \{ p_1,\dots,p_K \} \subset S^1$.
1949: Let $V_j \subset A_{r'}$ be an open
1950: neighborhood of $p_j$ such that
1951: $V_j \cap \mathcal{P} = \{ p_j \}$.
1952: Let $V^*_j = V_j \setminus \{ p_j \}$.
1953: Since $U$ is unitary on
1954: $\mathbb S_P \cap U^*_j \cap S^1$,
1955: its entries are bounded.
1956: Hence $U$ extends analytically to $p_j$
1957: for all $j=1,\dots,K$. Consequently,
1958: $U$ is analytic in $A_{r'}$ with $U^* = U^{-1}$.
1959: Thus $U \in \Lambda_{r'}\,\SU$, concluding this proof.
1960: \QED
1961:
1962: In the following sections we will apply the
1963: results of the present section to the twisted
1964: loop groups.
1965:
1966: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1967: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1968:
1969: \newsection{}
1970: Let $F \in \FID$ and $h \in \LGP$ and let
1971: $\hat F = h \# F$. Then $\hat F \in \FID$.
1972: Let $\tau \in \Delta$ and $\chi(\tau)$,
1973: $\hat\chi(\tau)$ be monodromies of $F$
1974: respectively $\hat F$. Writing $hF = \hat FB$
1975: for an Iwasawa decomposition and
1976: $\tau^*F = \chi(\tau)\,F\,k(\tau)$ and
1977: $\tau^*\hat F=\hat\chi(\tau)\,\hat F\,\hat k(\tau)$,
1978: we obtain
1979: %
1980: \begin{equation} \label{eq:mono_dress1}
1981: \hat\chi(\tau)^{-1} h\,
1982: \chi(\tau) \, h^{-1} \hat F =
1983: \hat F\,\hat k(\tau)\,(\tau^* B)\,
1984: k(\tau)^{-1}B^{-1}.
1985: \end{equation}
1986: %
1987: Since $\hat k(\tau)\,(\tau^* B)\,k(\tau)^{-1}
1988: B^{-1} \in \gauge$,
1989: this proves that
1990: %
1991: \begin{equation} \label{eq:mono_dress2}
1992: \hat \chi (\tau)^{-1}h\,\chi(\tau)\,h^{-1}
1993: \in \mathrm{Iso}_r(\hat F).
1994: \end{equation}
1995: %
1996: Since $F(\tilde{z}_0)= \hat F(\tilde{z}_0)=\Id$,
1997: evaluation of $hF = \hat FB$
1998: at the base-point $\tilde{z}_0$ yields
1999: $h=B(\tilde{z}_0)$. Evaluating equation
2000: \eqref{eq:mono_dress1} at $\tilde{z}_0$
2001: and solving for $\hat \chi(\tau)$ gives
2002: %
2003: \begin{equation} \label{eq:monodromy_dressed1}
2004: \begin{split}
2005: \hat \chi (\tau) &= h\,\chi(\tau)\,
2006: k(\tau,\tilde{z}_0)\,
2007: B\left(\tau(\tilde{z}_0)\right)^{-1}
2008: \hat k(\tau,\tilde{z}_0)^{-1} \\
2009: &= h\, \chi(\tau) \, h^{-1} \,L,\,
2010: L \in \mathrm{Iso}_r(\hat F) \\
2011: &= h\,\chi(\tau)\,b\, h^{-1},\,
2012: b := h^{-1} L \, h \in \mathrm{Iso}_r(F).
2013: \end{split}
2014: \end{equation}
2015: %
2016: This leads us to an application of Theorem
2017: \ref{th:h-split} and shows that if a
2018: dressing matrix preserves the topology, then it
2019: factorizes as in \eqref{eq:h-split}.
2020:
2021: We now denote by $\mathbb S = {\mathbb S}^*$
2022: the set obtained from the eigenvalues of
2023: a monodromy $\chi$.
2024:
2025: \mytheorem{} \label{th:h-split1}
2026: Let $\tau \in \Delta$ and $F \in \FID$
2027: with monodromy $\chi = \chi(\tau)$. We
2028: assume that the trace of $\chi$ is not
2029: constant. Let $h \in \LGP$ such that
2030: $h\,\chi\,h^{-1}$ is meromorphic on
2031: $\mathbb S \cap A_s$ for some $0<s<r$,
2032: where $\mathbb S$ is as in Lemma \ref{th:S}.
2033: If $\hat F = h\#F$ has a monodromy
2034: $\hat{\chi}$ with respect to
2035: $\tau$, and if $\hat{\mathbb S}$ is defined
2036: for $\hat{\chi}$ according to Lemma \ref{th:S},
2037: then $h$ can be factored, on a possibly
2038: segmented circle
2039: $C_l \cap \mathbb S \cap \hat{\mathbb S}$ for
2040: $l \in [s,\,r]$, into
2041: $h = \mathcal{M} \,\mathcal{C}$
2042: where $\mathcal{C} = \mathcal{C}(\tau,\,\lambda)$
2043: is twisted $\SL$--valued and defined on
2044: $C_l \cap \mathbb S \cap \hat{\mathbb S}$ and
2045: $[\, \mathcal{C},\, \chi \,] = 0$ there, and
2046: $\mathcal{M} = \mathcal{M}(\tau,\,\lambda)$
2047: is twisted $\SU$--valued and meromorphic
2048: on the connected open set
2049: $ S^1 \subset \mathbb S \cap
2050: \hat{\mathbb S} \cap A_s$.
2051: %
2052:
2053: \Proof By the above \eqref{eq:monodromy_dressed1},
2054: for the dressed monodromy we have
2055: $\hat{\chi} = h\,\chi\,h^{-1}\,\hat{L}$ with
2056: $\hat{L} \in \mathrm{Iso}_r(\hat F)$. Clearly,
2057: $\hat{L}$ is defined and meromorphic on
2058: $\mathbb S \cap A_s$. Moreover, we can write
2059: $\hat{\chi} = h\,\chi\, L h^{-1}\,$, where
2060: $L \in \mathrm{Iso}_r(F)$.
2061: From this we conclude that the eigenvalue
2062: functions of $\chi\,L$ are the same as those
2063: of $\hat{\chi}$. At this point we undress
2064: every occurring matrix and continue,
2065: until the very end of the proof, to work
2066: in the undressed setting. We note that
2067: the domains of definition of the dressed and
2068: the undressed matrices are in a natural
2069: correspondence and that unitary matrices
2070: are mapped to unitary matrices,
2071: while the "positive matrices" almost correspond.
2072: For simplicity of notation we will use the
2073: same notation as in the twisted situation.
2074: First we note that we can apply Theorem
2075: \ref{th:h-split} to $h\,\chi\,h^{-1}$.
2076: Hence $h = \mathcal{M} \,\mathcal{C}$,
2077: where $\mathcal{M}$ has values in $\SL$
2078: and is defined on $\mathbb S$.
2079: As a consequence, $\hat{\chi} =
2080: \mathcal{M} \chi {\mathcal{M}}^{-1} \hat{L} =
2081: \mathcal{M} \chi \,L'{\mathcal{M}}^{-1}$.
2082: From this expression we infer that $L'$ is
2083: defined and meromorphic on $\mathbb S$.
2084: Now we can apply almost verbatim the proof of
2085: \ref{co:h-split}, if we replace everywhere
2086: $\mathbb S$ by $\mathbb S \cap \hat{ \mathbb S} =
2087: \tilde{\mathbb S}$ and obtain that we can assume
2088: that $\mathcal{M}$ is defined and holomorphic
2089: on some open neighborhood of $S^1$ and unitary
2090: on $S^1$. Finally, twisting every occurring matrix
2091: again we obtain the desired result. \QED
2092: %
2093:
2094: When the associated family of $F \in \FID$
2095: possesses umbilic points, then, since dressing
2096: preserves umbilic points \cite{Wu:dre},
2097: $\rm{Iso}_r(\hat F) = \left\{ \pm \Id \right\}$
2098: for $\hat F = h \# F$ and \eqref{eq:mono_dress2}
2099: implies
2100: %
2101: \begin{equation} \label{eq:monodromy_dressed2}
2102: \hat \chi (\tau) = \pm h\,\chi(\tau)\,h^{-1}
2103: \in \Lhol.
2104: \end{equation}
2105: %
2106: This allows us to restate Corollary
2107: \ref{co:h-split} for the trivial isotropy case:
2108:
2109: %%%%%%%%%%
2110:
2111: \mycorollary{} \label{co:h-split1}
2112: Let $F \in \FID$ with
2113: $\mathrm{Iso}(F) = \{ \pm \Id \}$. Let
2114: $\chi = \chi(\tau) \in \Lhol$ be a
2115: monodromy of $F$ with non-constant trace.
2116: Let $h \in \LGP$ for some $r \in (0,\,1]$.
2117: If $h\#F$ also has a
2118: monodromy with respect to $\tau$, then $h$ can be
2119: factored into $h = U \,C$
2120: with $[\, C,\,\chi \,] = 0$ and
2121: $U = U(\tau,\,\lambda)$ the meromorphic
2122: extension of an element of
2123: $\Lambda_{r^\prime} \SU_\sigma$
2124: for some $r^\prime \in [r,\,1]$. \QED
2125:
2126: We were not able to prove the corresponding result for a
2127: general $F \in \FID$ without the isotropy
2128: condition. It turns out, that we do
2129: have an analogous result for the dressing orbit of
2130: the vacuum, to which we now turn our attention.
2131:
2132: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2133: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2134:
2135: \section{Dressing the vacuum}
2136: %
2137: \newsection{\bf Definitions.}
2138: The standard round cylinder (=''the vacuum'')
2139: has the conformal structure of the punctured
2140: plane $\C^*$. If we identify
2141: $\C^* \cong \C / q\, \Z$ for some
2142: $q \in \C^*$, then
2143: %
2144: \begin{equation}
2145: \exp:\C \to \C^*,\,w \mapsto z = \exp(q\,w)
2146: \end{equation}
2147: %
2148: is the universal covering map. The group of deck
2149: transformations $\Delta \cong \Z$ is generated by
2150: the translation
2151: %
2152: \begin{equation} \label{eq:translation}
2153: \tau_q : w \mapsto w + q.
2154: \end{equation}
2155: %
2156: An extended unitary frame of the associated
2157: family of the vacuum is given by
2158: %
2159: \begin{equation} \label{eq:F_c}
2160: F_c(w,\,\lambda) =
2161: \exp((w \lambda^{-1}
2162: - \overline{w} \lambda)\, A)
2163: \end{equation}
2164: %
2165: where
2166: \begin{equation}
2167: A = \begin{pmatrix}
2168: 0 & 1 \\ 1 & 0 \end{pmatrix}.
2169: \end{equation}
2170: %
2171: A monodromy with respect to $\tau_q$ of
2172: $F_c$ is given by
2173: %
2174: \begin{equation} \label{eq:cylinder_monodromy}
2175: \chi_c (\tau_q) =
2176: \exp((q \lambda^{-1} - \bar{q}
2177: \lambda )\,A).
2178: \end{equation}
2179: %
2180: The map $F_c:\C \to \Lhol$ is obtained in the
2181: DPW framework from the triple
2182: $((\lambda^{-1} + \lambda )A\,dw,\,\Id,\,0 )$.
2183: Note that we have chosen the base-point
2184: $w_0 = 0$. The corresponding holomorphic frame
2185: is
2186: %
2187: \begin{equation} \label{eq:Phi_c}
2188: \Phi_c(w,\,\lambda) =
2189: \exp((\lambda^{-1} + \lambda)\,w\, A)
2190: \end{equation}
2191: %
2192: with monodromy with respect to $\tau_q$ given by
2193: %
2194: \begin{equation} \label{eq:rho_c}
2195: \varrho_c(\tau_q) =
2196: \exp((\lambda^{-1} + \lambda)\,q\, A).
2197: \end{equation}
2198: %
2199: Note that if $q \in i\R$, then
2200: $\varrho_c(\tau_q) \in \Lhol$ and
2201: $\varrho_c(\tau_q) = \chi_c(\tau_q)$.
2202:
2203: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2204: %
2205: \newsection{\bf Commuting flows.}
2206: Denote the abelian sub-algebra of elements of
2207: $\Lambda_{\mbox{\tiny{$\C^*$}}}\sl_\sigma$
2208: that commute with $A$ and have a pole at
2209: $\lambda = 0$ by
2210: %
2211: \begin{equation*}
2212: \mathcal{Z} = \left\{ \phi \,A:
2213: \phi (\lambda) = \Sigma
2214: \phi_j \lambda^j \mbox{ odd },\, j\geq -K,\,
2215: K = 2g +1 \in \N \right\}.
2216: \end{equation*}
2217: %
2218: Let $h \in \LGP,\,\eta \in \mathcal{Z}$ and
2219: define $h \sharp \eta$ via the $r$--Iwasawa
2220: decomposition
2221: %
2222: \begin{equation}
2223: h \exp(\eta) = U \, h \sharp \eta.
2224: \end{equation}
2225: %
2226: Further, for $F = h \# F_c$ define
2227: $F \sharp \eta := (h \sharp \eta) \# F_c$.
2228: By \cite{BurP:dre}, Proposition 4.1, this is
2229: an action on the dressing orbit of $F_c$:
2230: If $\eta_1,\,\eta_2 \in \mathcal{Z}$, then
2231: $F \sharp (\eta_1 + \eta_2) =
2232: (F \sharp \eta_2)\sharp\eta_1 = (F \sharp \eta_1)\sharp\eta_2$ and
2233: if $\eta_1 - \eta_2 \in \Lambda^+_r \sl$, then
2234: $F \sharp \eta_1 = F \sharp \eta_2$.
2235: By construction $\zeta = \phi A \in \mathcal{Z}$
2236: is analytic for $\lambda \in \C^*$ with a pole at
2237: $\lambda = 0$. Expanding $\phi(\lambda) =
2238: \sum_{j\geq -K} \phi_j \lambda^j$, observe that
2239: $\exp(\zeta) \# F_c = \exp(\eta) \# F_c$
2240: where $\eta = \phi^\prime A$ and
2241: $\phi^\prime = \sum_{j=-K}^0\phi_j \lambda^j$ is
2242: analytic on $\CP \setminus \{ 0 \}$. Since
2243: $\eta^* = \phi^* A$ is analytic
2244: in $\C$, $e^{\eta - \eta^*} \in \Lhol$ and
2245: $e^{\zeta}\# F_c = e^{\eta-\eta^*}F_c$.
2246: Hence, when dressing with $e^\zeta$, we restrict
2247: without loss of generality to
2248: $\zeta \in \mathcal{Z}$ of the form
2249: %
2250: \begin{equation} \label{eq:zeta}
2251: \zeta = \sum_{j=1}^K (\phi_j \lambda^{-j} -
2252: \bar{\phi}_j \lambda^j ) A
2253: \end{equation}
2254: %
2255: where $\phi_j \in \C$ and summation is over odd
2256: indices $j=1,\,3,\ldots,K = 2g+1 \in \N$. Then
2257: %
2258: \begin{equation}
2259: \exp (\zeta) \, \# \,
2260: F_c(w,\,\lambda) =
2261: \exp \bigl(\sum_{j=3}^{K} (\phi_j \lambda^{-j} -
2262: \bar{\phi}_j \lambda^j )A \bigr) \,
2263: F_c(w+\phi_1,\,\lambda).
2264: \end{equation}
2265: %
2266: %and in particular, for $K=1$ we have
2267: %$\exp((\phi_0\lambda^{-1}-\bar{\phi}_0\lambda)A)
2268: %\# F_c(z) = F_c(z + \phi_0)$.
2269:
2270: For $\zeta \in \mathcal{Z}$ of the form
2271: \eqref{eq:zeta} we have $e^{t\zeta}\in \Lhol$
2272: for all $t \in \R$.
2273: Hence, for any extended unitary frame
2274: $F \in \mathcal{F}(\MT)$, we have
2275: $e^{t\zeta}\#F = e^{t\zeta}F$ which on the level
2276: of the immersions has the effect
2277: %
2278: \begin{equation*}
2279: f \longmapsto \exp(t\,\zeta)\,f\,\exp(-t\,\zeta) -
2280: \tfrac{1}{2H} i\lambda t
2281: \tfrac{\partial \zeta}{\partial \lambda}.
2282: \end{equation*}
2283: %
2284: Each $\zeta \in \mathcal{Z}$ defines a flow on the
2285: dressing orbit of the standard round cylinder by
2286: %
2287: \begin{equation}
2288: F_t := F \sharp t \zeta = (h\sharp t\zeta)\#F_c
2289: = U(\lambda,\,t)^{-1} h\#e^{t\zeta}F_c
2290: \end{equation}
2291: %
2292: where we have defined $U(\lambda,\,t)$ via the
2293: $r$--Iwasawa decomposition
2294: $he^{t\zeta}=U(\lambda,\,t)(h\sharp t\zeta)$ and
2295: $F = h \# F_c$ for some $h \in \LGP$.
2296: The flow $F_t$ is called
2297: \textit{trivial} if and only
2298: if $F_t = V(t) F V(t)^{-1}$ for a
2299: $\lambda$--independent map $V:\R \to \SU$
2300: and $F$ is of \textit{finite type} if and
2301: only if the subspace
2302: $\mathcal{Z}^\prime \subset \mathcal{Z}$ of
2303: trivial flows has finite co-dimension in
2304: $\mathcal{Z}$.
2305:
2306: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2307:
2308: \newsection{\bf Monodromy.}
2309: Let $h \in \LGP$ and $hF_c = FB$ the
2310: $r$--Iwasawa decomposition of $hF_c$ such that
2311: at the base-point $w_0=0$ we have
2312: $F_c(0,\,\lambda) = F(0,\,\lambda) = \Id$ and
2313: $h(\lambda) = B(0,\,\lambda)$.
2314: Let us assume that $F$ has symmetry
2315: $(\tau_q,\,\chi(\tau_q))$.
2316: From \eqref{eq:monodromy_dressed1} we obtain
2317: %
2318: \begin{equation}\label{eq:dress_hol}
2319: \chi (\tau_q) = h\, \chi _c(\tau_q)\, B(q)^{-1}.
2320: \end{equation}
2321: %
2322: For completeness of exposition we prove the
2323: following result, first obtained in sections
2324: 3.1--3.4 of \cite{DorH:per}.
2325:
2326: \mylemma{}\label{th:f_q}
2327: Let $h \in \LGP$ and $F_c$
2328: be as in \eqref{eq:F_c}.
2329: Assume that $h \# F_c$ has monodromy
2330: $\chi(\tau_q) \in \Lhol$.
2331: Then there exists an odd function
2332: $f_q :C_r \to \C$ with a holomorphic extension
2333: to $I_r$ such that
2334: $B(q)^{-1} = \exp(f_q A) \, h^{-1}$.
2335: Setting
2336: %
2337: \begin{equation} \label{eq:p_q}
2338: p_q(\lambda) = q\lambda^{-1}-\bar{q}\lambda
2339: + f_q(\lambda),
2340: \end{equation}
2341: %
2342: we may write the monodromy of $h \# F_c$ as
2343: %
2344: \begin{equation} \label{eq:dress_monodromy}
2345: \chi (\tau_q) = h\, \exp(p_q A)\, h^{-1}.
2346: \end{equation}
2347: %
2348:
2349: \Proof
2350: Let $\chi_c(\tau_q)$ be the monodromy of
2351: $F_c$ as in \eqref{eq:cylinder_monodromy}.
2352: Then
2353: %
2354: \begin{equation}
2355: \chi_c(\tau_q) F_c = \tau_q^* F_c =
2356: h^{-1} \chi (\tau_q) h F_c B^{-1} \tau_q^* B
2357: \end{equation}
2358: %
2359: in combination with \eqref{eq:dress_hol} gives
2360: $B(q)^{-1}B(0) F_c = F_c \, \tau_q^* B^{-1} B$.
2361:
2362: Hence $[B(q)^{-1}B(0), A] = 0$ and for the matrix
2363: %
2364: \begin{equation}\label{eq:T}
2365: T = \frac{1}{\sqrt{2}} \begin{pmatrix}
2366: 1 & 1 \\ -1 & 1 \end{pmatrix}
2367: \end{equation}
2368: %
2369: we have $TAT^{-1} = \sigma_3$ and
2370: $[TB(q)^{-1}B(0)T^{-1}, \sigma_3] =0$.
2371: Thus
2372: $TB(q)^{-1}B(0)T^{-1} =
2373: \mathrm{diag}[s_q,\, s_q^{-1}]$
2374: for some holomorphic function $s_q:I_r \to \C^*$.
2375: Defining $s_q = \exp(f_q)$ yields
2376: $B(q)^{-1}B(0) = \exp(f_q A)$.
2377: Since $B(0) = h$, this proves the claim.
2378: \QED
2379:
2380: The point of Lemma \ref{th:f_q} is that
2381: we can now write a monodromy of an extended frame
2382: in the dressing orbit of the vacuum as
2383: %
2384: \begin{equation}\label{eq:sinhcosh}
2385: \cosh(p_q) \,\Id + \sinh(p_q) \,h\,A\,h^{-1}
2386: \end{equation}
2387: %
2388: and much of the analysis reduces to the study of
2389: the two scalar functions
2390: %
2391: \begin{equation} \label{eq:alphabeta}
2392: \alpha=\cosh(p_q)\mbox{ and }\beta = \sinh(p_q).
2393: \end{equation}
2394: %
2395: In analogy to Corollary \ref{co:h-split1}
2396: we have the following result for the monodromy
2397: representation of surfaces in the dressing orbit
2398: of the cylinder.
2399: %
2400:
2401: \mytheorem{}\label{th:h-split_cyl}
2402: Let $h \in \LGP$ and $F_c$
2403: be as in \eqref{eq:F_c}.
2404: Assume that $h \# F_c$ has symmetry
2405: $(\tau_q,\,\chi(\tau_q))$.
2406: Then $h = \mathcal{M}(\tau_q)\,\mathcal{C}(\tau_q)$
2407: with $\mathcal{M}(\tau_q) \in \Lambda_{r^\prime}
2408: \SU_\sigma$ for suitable $r \leq r^\prime < 1$ and
2409: $[\,\mathcal{C}(\tau_q),\,A\,]=0$.
2410:
2411: \Proof For $T$ defined in \eqref{eq:T} and by
2412: Lemma \ref{th:f_q}, we may write
2413: %
2414: \begin{equation} \label{eq:chi}
2415: \chi(\tau_q) = h \,T^{-1} \,
2416: \mathrm{diag}[\,\exp(-p_q),\,
2417: \exp(p_q)\,]\, T\, h^{-1}.
2418: \end{equation}
2419: %
2420: for $0< \mid \lambda \mid < r$.
2421: Let $h T^{-1} = \bigl( \begin{smallmatrix}
2422: a & b \\ c & d \end{smallmatrix} \bigr)$.
2423: The diagonal entries of $\chi(\tau_q)$
2424: are analytic functions in $\lambda \in \C^*$
2425: and given by
2426: %
2427: \begin{equation} \begin{split}
2428: a \, d \, \exp(-p_q) - b\,c\,\exp(p_q),\\
2429: a \, d \, \exp(p_q) - b\,c\,\exp(-p_q). \end{split}
2430: \end{equation}
2431: %
2432: Adding these implies that $\alpha = \cosh(p_q)$
2433: is an analytic function on $\C^*$.
2434: More precisely, $\alpha$ has an analytic
2435: extension from $ 0 <\mid \lambda \mid <r$ to $\C^*$.
2436: Solving for $\exp(p_q)$ we obtain
2437: %Then
2438: %
2439: %\begin{equation}
2440: % p_q = \cosh^{-1}(g) =
2441: % \log(g+\sqrt{g^2 -1})
2442: %\end{equation}
2443: %
2444: $\exp(p_q) = \alpha + \sqrt{\alpha^2-1}$.
2445: This is analytic on
2446: the set $\mathbb S \subset \C^*$ obtained by
2447: making branchcuts from the odd--ordered roots of
2448: $\alpha^2-1$ in the usual fashion.
2449: Note that the equation \eqref{eq:chi} above
2450: implies that $\mu_{\pm} = \exp (\pm p_q)$
2451: are the eigenvalues of $\chi (\tau_q)$.
2452: Hence, by Lemma \ref{th:S} we know that
2453: $\mathbb S$ is open, connected, and dense
2454: in $\C^*$. Moreover we have $S^1 \subset \mathbb S$
2455: and $\mathbb S = {\mathbb S }^*$.
2456:
2457: Consequently, $\exp(-p_q)$ is holomorphic on
2458: $\mathbb S$. Multiplying the $(1,1)$ entry of
2459: $\chi(\tau_q)$ by $\exp(p_q)$ and the $(2,2)$
2460: entry of $\chi(\tau_q)$ by $\exp(-p_q)$
2461: and subtracting the resulting equations implies
2462: that $bc \sinh (p_q)$ is holomorphic on $\mathbb S$. Note that $\beta = \sinh (p_q) \not\equiv 0$. Hence
2463: $bc$ is meromorphic on $\mathbb S$.
2464: Similarly one shows that $ad$ is meromorphic on
2465: $\mathbb S$.
2466:
2467: The off--diagonal terms of $\chi(\tau_q)$
2468: are $2 ab \beta$ and $-2 cd \beta$.
2469: Hence both $ab$ and $cd$ are meromorphic on
2470: $\mathbb S$.
2471: If $d \neq 0$ on the circle $C_r$, then as in
2472: \eqref{eq:h-split_case1}, we may write
2473: $hT^{-1} = \hat{h}D_1$ with
2474: $D_1= \mathrm{diag}[1/d,\,d]$
2475: and $\hat{h}$ meromorphic on $\mathbb S$.
2476:
2477: Defining $\mathcal{C}(\tau_q) = T^{-1}D_1 T$,
2478: $\mathcal{M}(\tau_q) = \hat{h} T$ yields
2479: $h = \mathcal{M}(\tau_q) \mathcal{C}(\tau_q)$
2480: with $[\, \mathcal{C}(\tau_q),\,A \,] =0$ and
2481: $\mathcal{M}(\tau_q)$ meromorphic on $\mathbb S$.
2482: Arguing as in the proof of Theorem \ref{th:h-split},
2483: $d \equiv 0$ on an arc $C_s \cap \mathbb S$
2484: implies that both $b,\,c \neq 0$ on $\mathbb S$
2485: and we may write $hT^{-1} = \widetilde{h}D_2$ with
2486: $D_2= \mathrm{diag}[c,\,1/c]$ and
2487: $\widetilde{h}$ meromorphic on $\mathbb S$.
2488: Defining $\mathcal{C}(\tau_q) = T^{-1}D_2 T$,
2489: $\mathcal{M}(\tau_q) = \widetilde{h} T$ yields
2490: $h = \mathcal{M}(\tau_q) \mathcal{C}(\tau_q)$
2491: with $[\,\mathcal{C}(\tau_q),\,A\,] =0$ and
2492: $\mathcal{M}(\tau_q)$ meromorphic on $\mathbb S$
2493: in this case. In either case, on $C_r$ we may write
2494: %
2495: \begin{equation} \label{eq:U}
2496: \chi(\tau_q) = \mathcal{M}(\tau_q)\,
2497: \exp(p_q\,A)\,\mathcal{M}(\tau_q)^{-1}.
2498: \end{equation}
2499: %
2500: Introducing an additional cut from $0$ to $-\infty$
2501: we obtain a simply connected domain
2502: $\hat{\mathbb S} = \hat{\mathbb S}^*$,
2503: on which we can take the logarithm of
2504: the holomorphic function $\exp(p_q)$,
2505: thus defining $p_q$ on $\hat{\mathbb S}$.
2506: Then on $\hat{\mathbb S}$ we combine
2507: $\chi(\tau_q)^* = \chi(\tau_q)^{-1}$
2508: with \eqref{eq:U} to obtain
2509: %
2510: \begin{equation} \label{eq:mstarm}
2511: \mathcal{M}^*(\tau_q) \,\mathcal{M}(\tau_q)
2512: \,\exp(-p_q\,A) = \exp(p^*_q\,A)\,
2513: \mathcal{M}^*(\tau_q) \,\mathcal{M}(\tau_q).
2514: \end{equation}
2515: %
2516: For $T$ defined in \eqref{eq:T}, recall that
2517: $T\,A\,T^{-1} = \mathrm{diag}[1,\,-1]$.
2518: Since $T \in \SU$, the matrix
2519: $X:= T\,\mathcal{M}^*(\tau_q) \,
2520: \mathcal{M}(\tau_q)\,T^{-1}$ is hermitian and
2521: of the form $X = W\,W^*$. In particular, if we write
2522: $X = \bigl( \begin{smallmatrix} x & y \\
2523: \bar{y} & z \end{smallmatrix} \bigr)$,
2524: then $x \neq 0$.
2525: We can rewrite \eqref{eq:mstarm} as
2526: $X\,T\,\exp(-p_q\,A)\,T^{-1} =
2527: T\,\exp(p^*_q\,A)\,T^{-1}\,X$ and obtain
2528: %
2529: \begin{equation}
2530: \begin{pmatrix} x & y \\ \bar{y} & z \end{pmatrix}
2531: \begin{pmatrix} \exp(-p_q) & 0 \\ 0 & \exp(p_q)
2532: \end{pmatrix} = \begin{pmatrix} \exp(p^*_q) & 0 \\
2533: 0 & \exp(-p^*_q) \end{pmatrix}
2534: \begin{pmatrix} x & y \\ \bar{y} & z \end{pmatrix}.
2535: \end{equation}
2536: %
2537: Since $x \neq 0$, then $\exp(-p_q) = \exp(p^*_q)$
2538: on $\hat{\mathbb S}$ and we conclude that
2539: $\cosh(p^*_q) = \cosh(p_q)$ and
2540: $\sinh(p^*_q) = - \sinh(p_q)$. Consequently
2541: $\exp(p^*_q\,A) = \exp(-p_q\,A)$
2542: on $\hat{\mathbb S}$ and equation
2543: \eqref{eq:mstarm} reads
2544: %
2545: \begin{equation}
2546: [\,\mathcal{M}^*(\tau_q) \mathcal{M}(\tau_q),\,
2547: \exp(-p_q\,A)\,] = 0.
2548: \end{equation}
2549: %
2550: From the proof of Corollary \ref{th:S} we
2551: now obtain some $L(\tau_q)$
2552: and $r' \in [r,1)$ such that
2553: $\mathcal{M}(\tau_q)L(\tau_q) \in
2554: \Lambda_{r'}\SU_\sigma$ and
2555: $[\,L^{-1}(\tau_q),\,\mathcal{C}(\tau_q)\,]=0$, and
2556: concludes the proof. \QED
2557: %
2558:
2559:
2560: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2561: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2562: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2563:
2564: Before going on to evaluate the result above
2565: to examples we would like to relate the work
2566: above to \cite{DorH:per}. Of particular
2567: interest to us is equation (3.6.7) there:
2568: %
2569: \begin{equation} \label{eq:x}
2570: h_+ = \frac{i}{2 \sqrt{\tilde{b}}}
2571: \begin{pmatrix}
2572: (x-x^{-1}) \tilde{b} & (x+x^{-1}) \tilde{b} \\
2573: (x+x^{-1}) - (x-x^{-1}) \tilde{a} &
2574: (x-x^{-1}) - (x+x^{-1}) \tilde{a} \end{pmatrix}
2575: \end{equation}
2576: %
2577: where $\tilde{a}$ and $\tilde{b}$
2578: denote the entries of the matrix $hAh^{-1}$
2579: in the $(11)$--position and the $(12)$--position
2580: respectively.
2581: This is the general form of a dressing matrix
2582: having on $\chi$ the same effect as $h$.
2583: The question addressed in the Theorem above
2584: is thus equivalent to the question whether one
2585: can choose $x$ so that $h_+$ is unitary on $S^1$.
2586:
2587: \mylemma{} The matrix $h_+$ is unitary on
2588: $S^1$ if and only if
2589: %
2590: \begin{equation}
2591: |\,x\,|^2 = \frac{1 + \tilde{a} +
2592: \sqrt{|\tilde{b}|^2}}{1-\tilde{a} +
2593: \sqrt{|\tilde{b}|^2}}.
2594: \end{equation}
2595: %
2596: This equation can be solved with some $x$
2597: holomorphic at $\lambda =0$ and well defined
2598: on some sufficiently well cut complex plane.
2599:
2600: \Proof
2601: The unitarity condition for $h_+$ yields
2602: two equations. However, a straightforward
2603: computation shows that
2604: these two equations are equivalent.
2605: It thus remains to consider
2606: %
2607: \begin{equation}
2608: (x^* - (x^*)^{-1}) \sqrt{\tilde{b}^*} =
2609: \frac{1}{\sqrt{\tilde{b}}} \lbrack (x-x^{-1}) -
2610: (x+x^{-1}) \tilde{a} \rbrack.
2611: \end{equation}
2612: %
2613: Splitting this equation into real and imaginary
2614: part yields two equations.
2615: However, again, a straightforward computation
2616: shows that these two equations are equivalent.
2617: It thus suffices to consider
2618: %
2619: \begin{equation}
2620: (Re(x)^* - (Re(x)^*)^{-1}) \sqrt{|\tilde{b}|^2} =
2621: (Re(x)-Re(x)^{-1}) - (Re(x)+ Re(x)^{-1}) \tilde{a}.
2622: \end{equation}
2623: %
2624: Note that we have used here that $\tilde{a}$
2625: is real on $S^1$ by \cite{DorH:per}, Theorem 3.7.
2626: Using $x^{-1} = x/{\mid x \mid}^2$ this
2627: equation rewrites directly into the claim.
2628: \QED
2629: %
2630:
2631: \mycorollary{}
2632: $h_+$ can be chosen diagonal if and only
2633: if $\tilde{a} =0$. In this case the
2634: diagonal entries of $h_+$ are unitary on $S^1$.
2635:
2636: \Proof
2637: Using \eqref{eq:x} for general $x$ we see that
2638: $h_+$ is diagonal only if
2639: $x + x^{-1}$ vanishes identically,
2640: since $\tilde{b}$ cannot vanish by b) and c) of
2641: \cite{DorH:per}, Theorem 3.7.
2642: Inserting this into the second off-diagonal
2643: expression for $h_+$ we obtain $\tilde{a} =0$.
2644: The converse is obvious choosing $x = i$.
2645: Assume now that $h_+$ has been chosen as a
2646: diagonal matrix. Since $\tilde{a}$ vanishes
2647: in this case as just shown, c) and e) of
2648: \cite{DorH:per} imply that
2649: $\tilde{b}$ is unitary on $S^1$.
2650: Finally, in view of \eqref{eq:x}, we obtain
2651: %
2652: \begin{equation}
2653: h_+ = \frac{i}{2 \sqrt{\tilde{b}}}(x - x^{-1})
2654: \mathrm{diag}[b,\,1].
2655: \end{equation}
2656: %
2657: Moreover, $x = -x^{-1}$, whence $x = \pm i$.
2658: The condition $\det h_+ = 1$ is now satisfied.
2659: Thus necessarily
2660: $h_+ = \pm \mathrm{diag}[\sqrt{\tilde{b}},\,
2661: 1/\sqrt{\tilde{b}}]$
2662: and the diagonal entries of $h_+$ are unitary
2663: on $S^1$. \QED
2664:
2665: An immediate consequence of the above discussion leads to
2666: following observation.
2667:
2668: \mycorollary{}
2669: Let $h \in \LGP$ such that
2670: $h\, \exp(p_q A)\, h^{-1} \in \Lhol$ for some
2671: translation $\tau_q$. If $h$ is diagonal,
2672: then $h^4$ is rational on $\CP$ and
2673: $h^2$ is analytic in
2674: $\CP \setminus \mathcal{H}$, where $\mathcal{H}$
2675: is the set obtained by making appropriate
2676: branchcuts from the points lying in the set
2677: %
2678: \begin{equation} \label{eq:branchset}
2679: \mathcal{H}^\prime =
2680: \{ \lambda \in \CP : |\lambda | > r \mbox{ and }
2681: p_q(\lambda) \in \pi i \Z \} \cup \{
2682: \mbox{ singularities of } p_q(\lambda) \}
2683: \end{equation}
2684: %
2685: and $h$ is unitary on $S^1$. \QED
2686:
2687: \myremark{}
2688: Writing $h = \mathrm{diag}[a,\,1/a]$, we have
2689: that $\chi(\tau_q)= \alpha\,\Id +
2690: \beta\,\mathrm{off}[a^2,\,a^{-2}]$
2691: is analytic on $\C^*$.
2692: Hence $a^2$ and $a^{-2}$ are analytic in
2693: $\mathcal{H}$ and even in $\lambda$.
2694: Following the procedure for splitting
2695: a matrix we first untwist $h$.
2696: The above procedure for
2697: constructing the factorisation
2698: $h = \mathcal{MC}$ gives
2699: %
2700: \begin{equation} \label{eq:diagonal_h-split}
2701: \mathcal{M} = \frac{1}{2} \begin{pmatrix}
2702: 1+a^2 & 1-a^2 \\
2703: -1 + a^{-2} & 1 + a^{-2} \end{pmatrix}, \,
2704: \mathcal{C} = \frac{1}{2} \begin{pmatrix}
2705: a+1/a & a-1/a \\ a-1/a & a+1/a \end{pmatrix},
2706: \end{equation}
2707: %
2708: where a really denotes the untwisted function.
2709: Now we need to twist again. This way we obtain
2710: the original function $a$ and in addition
2711: the off-diagonal terms are multiplied by
2712: $\lambda$ (in the $(12-)$position) and
2713: by $\lambda^{-1}$ in the $(21)-$position
2714: respectively. Clearly this leads to a matrix
2715: $\mathcal{M}$ which is twisted as required.
2716: %If $a^{*\,2} = \,a^{-2}$ on
2717: %$\mathcal{H}$, then
2718: %$\mathcal{M}^* = \mathcal{M}^{-1}$.
2719:
2720:
2721: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2722:
2723: \newsection{\bf Delaunay Surfaces.}
2724: \label{sec:Delaunayexample}
2725: As an example, we discuss how to dress the
2726: vacuum into a Delaunay surface and compute
2727: the splitting of this dressing matrix according
2728: to Theorem \ref{th:h-split_cyl}. It is proven in
2729: \cite{Kil:del} that an associated
2730: family of the Delaunay surface with neck radius
2731: $\omega$ is generated by the triple
2732: $(D\, dz/z,\,\Id,\,0 )$, where
2733: %
2734: \begin{equation} \label{eq:delaunay}
2735: D = \begin{pmatrix} 0 & \alpha \lambda^{-1} +
2736: \beta \lambda \\ \beta \lambda^{-1} +
2737: \alpha \lambda & 0 \end{pmatrix},
2738: \end{equation}
2739: %
2740: and $\alpha,\, \beta \in \R$ with
2741: $\alpha + \beta = 1/2$ and
2742: $\omega = \tfrac{1}{2H}
2743: (1 - \sqrt{1 - 16 \alpha \beta})$.
2744: The monodromy of the unitary frame with respect to
2745: the translation $\tau_q$ for $q = 2 \pi i$ is
2746: given by $\exp ( 2 \pi i D )$.
2747: Therefore, to dress the
2748: standard cylinder into a Delaunay surface,
2749: it suffices to determine a diagonal
2750: $h = \mathrm{diag}[a,\, 1/a] \in \LGP$
2751: for some suitable $r>0$ such that
2752: $p_q h A h^{-1} = D$, which is equivalent to
2753: the two equations
2754: %
2755: \begin{equation*}
2756: a^{\pm 2}(q\lambda^{-1}- \bar{q}\lambda + f_q )=
2757: 2 \pi i (\alpha\lambda^{\mp 1} +
2758: \beta \lambda^{\pm 1} ).
2759: \end{equation*}
2760: %
2761: Solving both equations for $a^2$ and equating gives
2762: %
2763: \begin{equation} \label{eq:Del_ex}
2764: 2 \pi i \sqrt{(\alpha + \beta \lambda^2)
2765: (\beta + \alpha \lambda^2)} =
2766: q - \bar{q}\lambda^2 + \lambda f_q,
2767: \end{equation}
2768: %
2769: which in turn yields
2770: $a^2 = \sqrt{(\alpha + \beta \lambda^2)/
2771: (\beta + \alpha \lambda^2)}$.
2772: Evaluating \eqref{eq:Del_ex} at $\lambda = 0$ gives
2773: $q = 2\pi i \sqrt{\alpha\,\beta}$.
2774: If $\rho:= \min \{ |\sqrt{\alpha/\beta}| ,\,
2775: | \sqrt{\beta/\alpha}|\}$,
2776: then $h \in \LGP$ for $0<r<\rho$.
2777: It is easily verified that
2778: $p(\lambda) = 2 \pi i \sqrt{- \det D}$
2779: vanishes at the singularities of $h$.
2780: In summary, the matrix that dresses the vacuum
2781: into a Delaunay surface with
2782: neck radius $\omega$ is given by
2783: %
2784: \begin{equation}
2785: h(\lambda) = \mathrm{diag}[\,
2786: \sqrt[4]{\tfrac{\alpha + \beta \lambda^2}
2787: {\beta + \alpha \lambda^2}},\,
2788: \sqrt[4]{\tfrac{\beta + \alpha \lambda^2}
2789: {\alpha + \beta \lambda^2}}\, ]
2790: \end{equation}
2791: %
2792: Untwisting $h$, factoring
2793: $h = \mathcal{M}\mathcal{C}$ as
2794: in \eqref{eq:diagonal_h-split} and twisting back
2795: gives
2796: %
2797: \begin{equation}
2798: \mathcal{M} = \frac{1}{2} \begin{pmatrix}
2799: 1+ \sqrt{\tfrac{\alpha + \beta \lambda^2}
2800: {\beta + \alpha \lambda^2}} &
2801: \lambda - \lambda \sqrt{\tfrac{\alpha +
2802: \beta \lambda^2}{\beta + \alpha \lambda^2}} \\
2803: \lambda^{-1}\sqrt{\tfrac{\beta +
2804: \alpha \lambda^2}{\alpha + \beta \lambda^2}}
2805: -\lambda^{-1} &
2806: 1 + \sqrt{\tfrac{\beta + \alpha \lambda^2}
2807: {\alpha + \beta \lambda^2}} \end{pmatrix}.
2808: \end{equation}
2809: %
2810: Evidently $\mathcal{M}^* = \mathcal{M}^{-1}$
2811: so that this is actually the decomposition
2812: according to Theorem \ref{th:h-split_cyl}.
2813: The domain of $\mathcal{M}$ is the genus one
2814: hyperelliptic curve $\mu^2 + \det D = 0$,
2815: the \textit{spectral curve}
2816: of the underlying Delaunay surface.
2817:
2818: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2819:
2820: \newsection{\bf Finite Blaschke Products.}
2821: The function $f_q$ occurring in Theorem
2822: \ref{th:f_q} is not always as easy to deal with as
2823: in the above example nor can it generally be
2824: explicitly computed. Nonetheless, even the
2825: trivial case $f_q \equiv 0$ is quite interesting
2826: and the corresponding dressing matrices are of such
2827: a simple form that the involved Iwasawa
2828: decomposition can be explicitly computed.
2829:
2830: \mylemma{}\label{th:f_q=0}
2831: Let $h \in \LGP$ and $F_c$ be the unitary frame
2832: of the round cylinder with monodromy $\chi_c(\tau_q)
2833: = \exp((q\lambda^{-1}-\bar{q}\lambda)A)$ for some
2834: $q \in \C^*$ with
2835: %
2836: \begin{equation} \label{eq:q_condition}
2837: \sqrt{\bar{q}} - \sqrt{q} \notin \pi \,\Z.
2838: \end{equation}
2839: %
2840: Assume $F = h \# F_c$ has a monodromy
2841: $h\,\chi_c(\tau_q)\,\exp(f_q\,A)\,h^{-1} \in \Lhol$
2842: as in equation \eqref{eq:dress_monodromy}.
2843: Then the following are equivalent
2844: %
2845: \begin{enumerate}
2846: \item $h \chi_c(\tau_q) h^{-1} \in \Lhol$
2847: \item $f_q \equiv 0$
2848: \item $h\,A\,h^{-1}$ is rational on $\CP$ with only
2849: simple poles.
2850: \end{enumerate}
2851:
2852: \Proof
2853: We first show that (i) implies (ii).
2854: The assumption (i) allows us to write the
2855: monodromy of $F$ as
2856: $h\,\chi_c(\tau_q)\,h^{-1} \cdot
2857: h\exp(f_q\,A)\,h^{-1}$, where the first group
2858: of factors is unitary and the second group of
2859: factors is contained in $\LGP$ and has first
2860: term $I$. Therefore the condition that the whole
2861: expression is unitary is equivalent to
2862: $h\exp(f_q\,A)\,h^{-1}$ being unitary.
2863: But it is also in $\LGP$. Therefore it is in $\U1$.
2864: But since it starts with $I$, it is identically
2865: equal to $I$. As a consequence $\exp(f_q\,A) = I$.
2866: Hence $f_q \equiv 2 \pi k$ for some $k \in \Z$
2867: and $k =0$, since $f_q$ is an odd function by
2868: Lemma \ref{th:f_q}. This shows that (i) implies (ii).
2869:
2870: Let us prove that (ii) implies (iii). Recall from
2871: \eqref{eq:sinhcosh} and \eqref{eq:alphabeta} that
2872: $h\,\exp(p_q\,A)\,h^{-1} = \alpha\,\Id +
2873: \beta\,h\,A\,h^{-1}$. Since $f_q =0$ we have
2874: $p_q = \lambda^{-1}q - \lambda \bar{q}$ and
2875: therefore
2876: $\beta = \sinh(\lambda^{-1}q-\lambda\bar{q})$ is
2877: holomorphic on $\C^*$.
2878: By assumption, $\beta\,h\,A\,h^{-1}$
2879: is holomorphic on $\C^*$ so that
2880: $h\,A\,h^{-1}$ is meromorphic on $\C^*$.
2881: Let $\lambda_0 \in \C^*$ be a root of
2882: $\beta = \sinh(\lambda^{-1}q-\lambda\bar{q})$,
2883: that is, it lies in the set
2884: %
2885: \begin{equation} \label{eq:spectral_set}
2886: \mathcal{S}_r^q = \left\{ \lambda \in A_r :
2887: \lambda^{-1}q - \lambda\,\bar{q} \in \pi i \Z
2888: \right\}.
2889: \end{equation}
2890: %
2891: Then $\beta'(\lambda_0) = \pm(\lambda_0^{-2}q -
2892: \bar{q}) \neq 0$ if and only if
2893: $\lambda_0 \neq \pm i \sqrt{q/\bar{q}}$. Our
2894: assumption \eqref{eq:q_condition} ensures that
2895: $\pm i \sqrt{q/\bar{q}} \notin \mathcal{S}_r^q$.
2896: Hence all roots
2897: of $\beta$ in $\C^*$ are simple.
2898: Therefore the entries of $h\,A\,h^{-1}$
2899: can only have simple poles in $\C^*$ which
2900: must lie in $\mathcal{S}_r^q$.
2901: Note also that $h\,A\,h^{-1}$ is holomorphic at
2902: $\lambda = 0$. Hence $h\,A\,h^{-1}$ is a holomorphic
2903: germ at the origin with a meromorphic
2904: extension to $\C^*$. In Section 3.5 of
2905: \cite{DorH:per} it is shown that the squares of the
2906: entries of $h\,A\,h^{-1}$ are finite at $\infty$.
2907: Hence $h\,A\,h^{-1}$ is rational on $\CP$ and
2908: proves that (ii) implies (iii).
2909:
2910: Finally we show that (iii) implies (i).
2911: By equations \eqref{eq:sinhcosh} and
2912: \eqref{eq:alphabeta} we write
2913: $\chi = \alpha\,\Id + \beta\,hAh^{-1}$
2914: and use the fact that
2915: $\beta\,h\,A\,h^{-1}$ is holomorphic on $\C^*$.
2916: With assumption (iii) this implies that $\beta$
2917: is meromorphic on $\C^*$.
2918: Since $\alpha$ is holomorphic
2919: on $\C^*$, so is $\alpha^2$ and consequently
2920: $\beta^2 = \alpha^2 - 1$ is holomorphic on
2921: $\C^*$, therefore $\beta$ itself is holomorphic
2922: on $\C^*$. Further, from
2923: $\chi^* = \chi^{-1} = \alpha\,\Id -
2924: \beta\,hAh^{-1}$ we deduce that $\alpha$ is real
2925: on $S^1$, i.e. $\alpha^* = \alpha$.
2926: And since $\beta^2$ is real and non-positive
2927: on $S^1$, and since $\beta$ is holomorphic on
2928: $\C^*$, we obtain $\beta^* = - \beta$. Consequently,
2929: $(hAh^{-1})^* = hAh^{-1}$ and thus
2930: $h\,\chi_c(\tau_q)\,h^{-1} =
2931: \exp((\lambda^{-1}q - \lambda\,q)h\,A\,h^{-1})
2932: \in \Lhol$. This proves that (iii) implies (i)
2933: and concludes the proof of the lemma. \QED
2934:
2935:
2936: \myremark{}
2937: 1. The result above seems to indicate that for
2938: every "type of $f_q$ " one obtains an associated
2939: type of dressing matrix.
2940:
2941: 2. In the case $f_q = 0$ and the setting of
2942: \cite{DorH:per}, the hyperelliptic surface
2943: associated with the dressing is simply the
2944: Riemann sphere $S^2$.
2945: If this class would contain some torus, then one
2946: would have an example for the
2947: "singular tori question" of \cite{Bob:tor}.
2948:
2949: %3. A complete classification for the case
2950: %$f_q =0$ should follow from \cite{DorH:per}.
2951:
2952: 3. The set $\mathcal{S}_r^q$ in
2953: \eqref{eq:spectral_set} is finite for fixed
2954: $r$. Further $\mathcal{S}_r^q \subset \R$ if and
2955: only if $q \in i\R$, in which case $q$ also
2956: satisfies \eqref{eq:q_condition}. Finally,
2957: it is easy to see that if
2958: $\lambda \in \mathcal{S}_r^q$ then also
2959: $\bar{\lambda},\,1/\lambda,\,1/\bar{\lambda}
2960: \in \mathcal{S}_r^q$, and in particular,
2961: $\left(\mathcal{S}_r^q \right)^* = \mathcal{S}_r^q$.
2962:
2963: We apply Lemma \ref{th:f_q=0} to a class of
2964: diagonal dressing matrices for which $f_q \equiv 0$.
2965:
2966: \mycorollary{}
2967: Let $h = \mathrm{diag}[a,\,1/a] \in \LGP$.
2968: Then $h$ has an extension to $A_{r'}$ for some
2969: $r' \in [r,\,1]$ such that $h^* = h^{-1}$ on
2970: $A_{r'}$ and $h \chi_c(\tau_q) h^{-1} \in \Lhol$
2971: if and only if $a^2$ is a finite Blascke product
2972: %
2973: \begin{equation}\label{eq:a^2}
2974: a^2 (\lambda) = \prod_{j=1}^N
2975: \frac{\alpha_j^2 - \lambda^2}
2976: {1 - \bar{\alpha}_j^2 \lambda^2},\,\alpha_j
2977: \in \mathcal{S}_r^q.
2978: \end{equation}
2979: %
2980: \Proof Assume $h$ has a unitary branch on $S^1$
2981: and $h \chi_c(\tau_q) h^{-1} \in \Lhol$. Then by
2982: Lemma \ref{th:f_q=0} we conclude that
2983: $h\,A\,h^{-1} = \mathrm{off}[a^2,\,1/a^2]$
2984: is rational with simple poles and zeroes located in
2985: $\mathcal{S}_r^q$. Hence $a^2$ is of the form
2986: \eqref{eq:a^2}. The converse is proven by direct
2987: verification.
2988: \QED
2989:
2990: Dressing matrices characterised in
2991: the previous proposition are thus of the form
2992: %
2993: \begin{equation} \label{eq:prod_simple}
2994: h = \prod_{j=1}^N \begin{pmatrix}
2995: \sqrt{\tfrac{\alpha_j^2 - \lambda^2}
2996: {1 - \bar{\alpha}_j^2 \lambda^2}} & 0 \\ 0 &
2997: \sqrt{\tfrac{1 - \bar{\alpha}_j^2 \lambda^2}
2998: {\alpha_j^2 - \lambda^2}} \end{pmatrix},\,
2999: \alpha_j \in \mathcal{S}_r^q
3000: \end{equation}
3001: %
3002: and are a special instance of an interesting
3003: class to which we now turn our attention.
3004:
3005: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3006:
3007: \newsection{\bf Simple factors, untwisted case.}
3008:
3009: In this section we will deal primarily with
3010: untwisted loops. The previous discussion has
3011: naturally lead us to a class of special dressing
3012: matrices, the so called \textit{simple factors} of
3013: Uhlenbeck \cite{Uhl} and discussed in similar
3014: context in \cite{TerU} and \cite{Bur:iso}.
3015: The main feature is that dressing with
3016: simple factors is explicit.
3017: This construction, due to Terng and
3018: Uhlenbeck \cite{TerU} goes as follows:
3019:
3020: \par We decompose
3021: $\C^2 = \mathrm{L} \oplus \mathrm{L}^\perp$
3022: for $\mathrm{L} = \C \left( \begin{smallmatrix}
3023: a \\ b \end{smallmatrix}
3024: \right)$. Then for all $A \in \GL$ we have
3025: %
3026: \begin{equation} \label{eq:perp}
3027: \bar{A}^t \mathrm{L} \perp
3028: A^{-1} \mathrm{L}^\perp
3029: \end{equation}
3030: %
3031: The hermitian projection
3032: $\pi_\l:\C^2 \to \mathrm{L}$ onto $\mathrm{L}$ is
3033: given by
3034: %
3035: \begin{equation*}
3036: \pi_\l = \frac{1}{|a|^2+|b|^2} \begin{pmatrix}
3037: |a|^2 & a\bar{b} \\ \bar{a}b & |b|^2
3038: \end{pmatrix}.
3039: \end{equation*}
3040: %
3041: Let $\alpha \in I_1^* = \left\{ \lambda \in \C :
3042: 0 < |\lambda | < 1 \right\}$. Then the map
3043: %
3044: \begin{equation*}
3045: \tau_\alpha(\lambda) = \frac{\alpha - \lambda}
3046: {1-\bar{\alpha}\lambda}
3047: \end{equation*}
3048: %
3049: is invariant under $\varrho$. For given
3050: $\mathrm{L}$ and $\alpha \in I_1$, a
3051: \textit{simple factor} \cite{TerU} is a
3052: loop of the form
3053: %
3054: \begin{equation} \label{eq:psi}
3055: \psi_{\alpha,\l}(\lambda) = \pi_\l +
3056: \tau_\alpha(\lambda) \,\pi_\l^\perp.
3057: \end{equation}
3058: %
3059: By construction, $\psi_{\alpha,\l}: \CP \setminus
3060: \left\{ \alpha,\,1/\bar{\alpha} \right\} \to \GL$
3061: is analytic and since $|\alpha |>0$,
3062: %
3063: \begin{equation}
3064: \psi_{\alpha,\l} \in \Lambda_r^+ \GL
3065: \mbox{ for } r < | \alpha |.
3066: \end{equation}
3067: %
3068: Clearly, simple factors are not twisted.
3069: Further, since $\psi_{\alpha,\l}^* =
3070: \psi_{\alpha,\l}^{-1} =
3071: \pi_\l + \tau_\alpha^{-1} \pi_\l^\perp$,
3072: we have that
3073: %
3074: \begin{equation}
3075: \psi_{\alpha,\l} \in \Lambda_r \mathbf{U}(2)
3076: \mbox{ for } r > | \alpha |.
3077: \end{equation}
3078: %
3079: For later use we also
3080: note the fact that for any $A \in \Utwo$
3081: we have
3082: %
3083: \begin{equation} \label{eq:change}
3084: \psi_{\alpha,A\l} = A \,\psi_{\alpha,\l}\,A^{-1}.
3085: \end{equation}
3086: %
3087:
3088: %
3089: Let $U \in \Lambda_{\mbox{\tiny{$\C^*$}}}\SU,\,
3090: \alpha \in I_1^*$
3091: and $\mathrm{L} \in \C\mathbb{P}^2$ and
3092: $\psi_{\alpha,\l}$ be the corresponding simple
3093: factor. Then it can be shown, see e.g \cite{KilSS},
3094: that
3095: %
3096: \begin{equation}
3097: \psi_{\alpha,\l}\,U\,
3098: \psi_{\alpha,\overline{U(\alpha)}^t \l}^{-1} \in
3099: \Lambda_{\mbox{\tiny{$\C^*$}}}\SU
3100: \end{equation}
3101: %
3102: Consequently, we have an explicit
3103: $r$--Iwasawa decomposition of $\psi_{\alpha,\l}\,U$
3104: for $r < |\alpha |$, given by
3105: %
3106: \begin{equation} \label{sdf}
3107: \psi_{\alpha,\l}\,U = \left( \psi_{\alpha,\l}\,U\,
3108: \psi_{\alpha,\overline{U(\alpha)}^t \l}^{-1}
3109: \right) \psi_{\alpha,\overline{U(\alpha)}^t \l}.
3110: \end{equation}
3111: %
3112: For the geometric applications considered in
3113: this paper it is of great value to have
3114: a simple and explicit formula for the dressed
3115: frame. Simple factors, for which $\mathrm{L}$
3116: depends on $\lambda$ and are of the form
3117: %
3118: \begin{equation}
3119: \psi_{\alpha,\l(\lambda)} = \pi_{\l(\lambda)} +
3120: \tau_\alpha(\lambda) \,(\Id - \pi_{\l(\lambda)})
3121: \end{equation}
3122: %
3123: will be called \emph{generalised simple factors}.
3124: We would like to present an analogous result
3125: to \ref{sdf} for such generalised simple factors:
3126:
3127: \mytheorem{}
3128: Let $\mathrm{L}: \C^* \rightarrow \CP$ be holomorphic and
3129: $\langle \mathrm{L},\,\mathrm{L}^* \rangle \neq 0$ for all
3130: $\lambda \in \C^*$.
3131: Let $\alpha \in I_1^*$ and $\psi_{\alpha,\l(\lambda)}$
3132: be the corresponding generalised simple factor.
3133: Then for any $U \in \Lambda_{\mbox{\tiny{$\C^*$}}}\SU$
3134: %
3135: \begin{equation} \label{gsdf}
3136: \psi_{\alpha,\l(\lambda)}\,\# \,U =
3137: \psi_{\alpha,\l(\lambda)}\,U\,
3138: \psi_{\alpha,\overline{U(\alpha)}^t
3139: \l(\alpha)}^{-1}.
3140: \end{equation}
3141: %
3142: \Proof
3143: Fix $\lambda_0 \in \C^*$ and write
3144: $\mathrm{L}(\lambda) = [a(\lambda) : b(\lambda)]$. Then for
3145: $\mathrm{L}_0 = [1 : 0]$ and the holomorphic map
3146: %
3147: \begin{equation}
3148: W(\lambda) = \begin{pmatrix}
3149: a(\lambda) & -\overline{b(1/\bar{\lambda})} \\
3150: b(\lambda) & \overline{a(1/\bar{\lambda})} \end{pmatrix}
3151: \end{equation}
3152: %
3153: we have $\mathrm{L}(\lambda) = W(\lambda)\,\mathrm{L}_0$ and
3154: $W(\lambda) \in \Lambda_{\mbox{\tiny{$\C^*$}}} \Utwo$.
3155: Applying \ref{eq:change} to
3156: $A = W(\lambda), \,\lambda \in \C^*$, we obtain
3157: $\psi_{\alpha,\l(\lambda)} =
3158: W(\lambda)\,\psi_{\alpha,\l_0}\,W(\lambda)^{-1}$
3159: and
3160: %
3161: \begin{align*}
3162: \psi_{\alpha,\l(\lambda)}\,U(\lambda) &=
3163: W(\lambda)\psi_{\alpha,\l_0}W(\lambda)^{-1}
3164: U(\lambda) \\ &= \left( W(\lambda)\,
3165: \psi_{\alpha,\l_0}\, W(\lambda)^{-1} U(\lambda)\,
3166: \psi_{\alpha,\l_0'}^{-1} \right)
3167: \psi_{\alpha,\l_0'},
3168: \end{align*}
3169: %
3170: where $\mathrm{L}_0' = \overline{(W(\alpha)^{-1}
3171: U(\alpha) )}^t\mathrm{L}_0$.
3172: The right side is an Iwasawa decomposition.
3173: The unitary part can be rewritten in the form
3174: %
3175: \begin{equation*}
3176: W(\lambda)\,\psi_{\alpha,L_0}\,W(\lambda)^{-1}
3177: \,U(\lambda)\,\psi_{\alpha,L_0'}^{-1} =
3178: \psi_{\alpha,L(\lambda)}\,U(\lambda)\,
3179: \psi_{\alpha,L_0'}^{-1}.
3180: \end{equation*}
3181: %
3182: It is straightforward to verify
3183: $\mathrm{L}_0' = \overline{U(\alpha)}^t
3184: \mathrm{L}(\alpha)$, thus proving \eqref{gsdf}. \QED
3185: %
3186:
3187: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3188:
3189: \newsection{\bf Twisting of simple factors}
3190:
3191: Since generalised simple factors are in general
3192: untwisted, we need to
3193: modify the concept so that it can also be used for
3194: the twisted case. We discuss two approaches to
3195: deal with simple factors in the twisted case by
3196: firstly, taking a simple factor and applying the
3197: 'twisting map' and secondly, looking for products
3198: of simple factors which happen to be twisted.
3199: Applying the 'twisting map' \eqref{eq:twist}
3200: to a simple factor we obtain
3201: %
3202: \begin{equation*}
3203: \psi_{\alpha,\l(\lambda)} =D(\lambda)\,
3204: \psi_{\alpha,\l (\lambda^2)}\,D(\lambda)^{-1} =
3205: \psi_{\alpha, D(\lambda) \l (\lambda^2)}.
3206: \end{equation*}
3207: %
3208: \mylemma{}
3209: a) Twisting a generalised simple factor
3210: produces again a generalised simple factor.
3211:
3212: b) Let $\psi$ be a simple factor (L is
3213: independent of $\lambda$). Then the
3214: corresponding twisted simple
3215: factor is again independent of
3216: $\lambda$ if and only if
3217: %
3218: \begin{equation}\label{eq:picase}
3219: D\,\pi_\l \,D^{-1} = \pi_\l \mbox{ or }
3220: \pi_\l^\perp.
3221: \end{equation}
3222: %
3223: \QED
3224:
3225: \myremark{}
3226: Since $D=D(\lambda)$ is diagonal, the only
3227: projections satisfying the condition
3228: \eqref{eq:picase} above are
3229: those onto the canonical basis vectors.
3230:
3231: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3232:
3233: \newsection{\bf Twisted products of simple factors}
3234: As mentioned earlier we also consider products of
3235: two simple factors and determine, when such
3236: matrices are twisted. Recall the two
3237: involutions $\varrho,\,\s$ defined in
3238: \eqref{eq:sigma} respectively \eqref{eq:varrho}.
3239:
3240: \myprop{} \label{th:twist}
3241:
3242: a) Let $g \in \Lambda_r \SL$.
3243: Then $(\s g) g$ is twisted if and
3244: only if $[\,\s g,\,g\,]=0$.
3245:
3246: b) If $g \in \LGU$, then $(\s g)g \in \LGU$.
3247:
3248: c) Let $\psi_{\alpha,\l}$ and
3249: $\psi_{\beta,\widehat{\l}}$ be simple factors.
3250: If $\sigma (\psi_{\alpha,\l})\,
3251: \psi_{\beta,\widehat{\l}}$ is twisted, then
3252: $\psi_{\alpha,\l} = \psi_{\beta,\widehat{\l}}$.
3253:
3254: d) If $\sigma (\psi_{\alpha,\l})\psi_{\alpha,\l}$
3255: is twisted, then either $\sigma(\pi_\l) = \pi_\l$
3256: or $\sigma(\pi_\l) = \pi_\l^\perp$.
3257:
3258: \Proof a) The loop $(\s g) g$ is twisted if and
3259: only if $\s \left( (\s g)\,g \right) = g \,\s g =
3260: (\s g)\,g$, which is equivalent to
3261: $[\s g,\,g] = 0$.
3262:
3263: b) The loop $g \in \LGU$ if and only if
3264: $\varrho \,g = g$. Then
3265: $\varrho \left( (\s g)g \right) =
3266: (\varrho \s g)\,(\varrho g) =
3267: (\s \,\varrho \,g ) \, (\varrho\, g) = (\s g)\,g$,
3268: since $[\varrho,\,\sigma ] = 0$. Hence
3269: $(\s g)\,g \in \LGU$.
3270:
3271: c) The product is twisted if and only if
3272: $\sigma (\psi_{\alpha,L}) \cdot
3273: \psi_{\beta,\widehat{L}} = \psi_{\alpha,L}
3274: \cdot \sigma ( \psi_{\beta,\widehat{L}})$.
3275: Let us abbreviate $A = \psi_{\alpha,\l}$ and
3276: $B = \psi_{\beta,\widehat{\l}}$.
3277: Then the pole of $A$ is at
3278: $1/\overline{\alpha}$, while the pole of
3279: $B$ is at $1/\overline{\beta}$.
3280: On the other hand, the pole of $\sigma (A)$
3281: is at $- 1/\overline{\alpha}$ and the
3282: analogous result holds for $B$.
3283: Thus comparing the two sides we obtain
3284: $\alpha = \beta$. Expanding near a pole
3285: and comparing the factors we derive $\mathrm{L} =
3286: \widehat{\mathrm{L}}$.
3287:
3288: (d) If $(\s \psi_{\alpha,\l} )\psi_{\alpha,\l}$
3289: is twisted, then by part (a),
3290: $[\s \psi_{\alpha,\l},\,\psi_{\alpha,\l}]=0$
3291: which is equivalent to the eigenspaces of
3292: $\s \psi_{\alpha,\l}$ and
3293: $\psi_{\alpha,\l}$ coinciding. Thus there are two
3294: possibilities:
3295: %
3296: \begin{equation}\label{eq:picase1}
3297: \s_3\,\pi_\l \,\s_3^{-1} = \pi_\l \mbox{ or }
3298: \pi_\l^\perp.
3299: \end{equation}
3300: %
3301: This proves (d) and concludes the proof of the
3302: proposition. \QED
3303:
3304: We are going to evaluate the two possibilities
3305: in \eqref{eq:picase1}.
3306:
3307: (i) Let us turn to the first case in
3308: \eqref{eq:picase1}:
3309: For the line
3310: $\mathrm{L} = \C \bigl( \begin{smallmatrix}
3311: 1\\0 \end{smallmatrix} \bigr)$ we have
3312: $\psi_{\alpha,\l}= \mathrm{diag}[1,\,\tau_\alpha]$.
3313: Using $\tau_\alpha(\lambda)\tau_\alpha(-\lambda) =
3314: \tau_{\alpha^2}(\lambda^2)$
3315: we obtain $(\s \psi_{\alpha,\l})\psi_{\alpha,\l} =
3316: \pi_\l + \tau_{\alpha^2}(\lambda^2)\pi^\perp_\l$.
3317: Dividing by the square root of the determinant
3318: we arrive at all diagonal twisted simple factors
3319: %
3320: \begin{equation} \label{eq:simple1}
3321: g_{\alpha,\l}(\lambda) =
3322: \sqrt{\tau_{\alpha^2}^{-1}
3323: (\lambda^2)} \, \pi_\l + \sqrt{\tau_{\alpha^2}
3324: (\lambda^2)}\, \pi^\perp_\l.
3325: \end{equation}
3326: %
3327: Then $g_{\alpha,\l}(\lambda) \in \LGP$ for
3328: $0< r< |\alpha|$ and in matrix form is given by
3329: %
3330: \begin{equation}\label{eq:simple1_matrix}
3331: g_{\alpha,\l}(\lambda) = \begin{pmatrix}
3332: \sqrt{\tfrac{1-\bar{\alpha}^2\lambda^2}
3333: {\alpha^2 - \lambda^2}} & 0 \\
3334: 0 & \sqrt{\tfrac{\alpha^2 - \lambda^2}
3335: {1-\bar{\alpha}^2\lambda^2}}
3336: \end{pmatrix}
3337: \end{equation}
3338: %
3339: Notice that $g_{\alpha,\l}(\lambda)^{-1}$
3340: corresponds to a factor of the matrix
3341: given in \eqref{eq:prod_simple}.
3342:
3343: (ii) Let us turn to the second case in
3344: \eqref{eq:picase1}: $\s_3 \pi_\l \s_3^{-1} =
3345: \pi_\l^\perp$.
3346: For $\mathrm{L} =\C ( a,\,b )^t$ and hermitian
3347: projection $\pi_\l:\C^2 \to \mathrm{L}$ we have
3348: $\sigma \pi_\l = \pi_\l^\perp$ if and
3349: only if $|a|^2=|b|^2 = 1/2$.
3350: Setting $a=\tfrac{1}{\sqrt{2}}\exp(is)$ and
3351: $b=\tfrac{1}{\sqrt{2}}\exp(it)$ and
3352: rescaling, we may assume without loss of
3353: generality that $\mathrm{L} =
3354: \C \bigl( e^{i\theta}, \, 1 \bigr)^t$. Then
3355: $\pi_\l = \tfrac{1}{2}(\Id + \mathrm{off}
3356: [ e^{i\theta},\,
3357: e^{-i\theta}])$ and for
3358: $\psi_{\alpha,\l} = \pi_\l + \tau_\alpha(\lambda)\,
3359: \pi_\l^\perp$
3360: we have $\sigma \psi_{\alpha,\l} = \pi_\l^\perp
3361: + \tau_\alpha(-\lambda)\,\pi_\l$ and consequently
3362: %
3363: \begin{align*}
3364: (\sigma \psi_{\alpha,\l})\psi_{\alpha,\l} &=
3365: \tau_\alpha(\lambda)\,\pi_\l^\perp +
3366: \tau_\alpha(-\lambda)\,\pi_\l \\
3367: &= \tfrac{1}{1-\bar{\alpha}^2\lambda^2} (
3368: (\alpha - \bar{\alpha}\lambda^2)\,\Id
3369: + \lambda (1 - |\alpha|^2)\,
3370: \mathrm{off}[e^{i\theta},\,e^{-i\theta}]).
3371: \end{align*}
3372: %
3373: Normalising, we arrive at the second class of
3374: twisted simple factors in $\LGP$ for
3375: $0 < r < |\alpha|$, given by
3376: %
3377: \begin{equation} \label{eq:simple2}
3378: g_{\alpha,\l}(\lambda) =
3379: \sqrt{\tau_\alpha^{-1}(\lambda)
3380: \,\tau_\alpha (-\lambda)} \, \pi_\l +
3381: \sqrt{\tau_\alpha (\lambda)
3382: \,\tau_\alpha^{-1} (-\lambda)}\, \pi^\perp_\l.
3383: \end{equation}
3384: %
3385: In matrix form, these simple factors look like
3386: %
3387: \begin{equation} \label{eq:simple2_matrix}
3388: g_{\alpha,\l}(\lambda) =
3389: \sqrt{\frac{1-\bar{\alpha}^2\lambda^2}
3390: {\alpha^2 - \lambda^2}}
3391: \begin{pmatrix} \alpha - \bar{\alpha}\lambda^2 &
3392: \lambda (1- |\alpha|^2)e^{i\theta} \\
3393: \lambda (1 - |\alpha|^2)e^{-i\theta} &
3394: \alpha - \bar{\alpha}\lambda^2
3395: \end{pmatrix}.
3396: \end{equation}
3397: %
3398: The simple factors $g_{\alpha,\l}$ derived
3399: in \eqref{eq:simple1} and \eqref{eq:simple2}
3400: are uniquely determined by their 'singularity'
3401: $\alpha \in I_1$ and choice of line
3402: $\mathrm{L} \in \CP$. Note also that in both cases
3403: \eqref{eq:simple1} and \eqref{eq:simple2} we have
3404: that
3405: %
3406: \begin{equation} \label{eq:simpleunitary}
3407: g_{\alpha,\l} \in \LGU
3408: \mbox{ for } |\alpha| < r \leq 1.
3409: \end{equation}
3410: %
3411: The key aspect is that dressing with simple factors
3412: is explicit. This idea is due to
3413: Terng and Uhlenbeck \cite{TerU} and has lead to
3414: variants as in \cite{Bur:iso} and \cite{KilSS}.
3415: The version we need is proven by
3416: twisting Theorem 1.2 in \cite{KilSS} and is as
3417: follows
3418:
3419: \mytheorem{} Let $M$ be a connected Riemann
3420: surface with universal cover $\MT$ and let
3421: $F(z,\lambda) \in \mathcal{F}(\MT)$.
3422: Let $g_{\alpha,\l} \in \LGP$ be a
3423: simple factor. Then
3424: %
3425: \begin{equation} \label{eq:simple_dress}
3426: g_{\alpha,\l} \# F =
3427: g_{\alpha,\l} \,F \,g^{-1}_{\alpha,\l^\prime}
3428: \end{equation}
3429: %
3430: where $\mathrm{L}^\prime =
3431: \overline{F(z,\alpha)}^t \mathrm{L}$ and
3432: $g^{-1}_{\alpha,\l^\prime}$ is, pointwise in
3433: $z \in \MT$, a simple factor of the same
3434: form as $g_{\alpha,\l}$. \QED
3435:
3436: We use this factorisation theorem to characterise
3437: when it is possible to dress an extended unitary
3438: frame with trivial isotropy by a simple factor
3439: while retaining a symmetry.
3440:
3441: \mycorollary{} \label{th:invariance}
3442: Let $\chi(\tau) \in \Lhol$ be a monodromy
3443: of an extended unitary frame
3444: $F \in \mathcal{F}(\MT)$.
3445: Let $g_{\alpha,\l} \in \LGP$ be a simple factor.
3446: Then
3447: $g_{\alpha,\l}\,\chi(\tau) \,g_{\alpha,\l}^{-1}$
3448: is a monodromy of $g_{\alpha,\l} \# F$
3449: with respect to $\tau$ if and only if
3450: %
3451: \begin{equation} \label{eq:invariance}
3452: \overline{\chi(\alpha,\tau)}^t
3453: \mathrm{L} = \mathrm{L}.
3454: \end{equation}
3455: %
3456: \Proof Let $\hat F = g_{\alpha,\l} \# F$.
3457: Then $\hat F = g_{\alpha,\l}\,F\,
3458: g^{-1}_{\alpha,\l^\prime}$ with
3459: $\mathrm{L}^\prime = \overline{F(z,\alpha)}^t
3460: \mathrm{L}$ by \eqref{eq:simple_dress}. Let
3461: $\tau^* F = \chi(\tau)\,F\,k$. Then
3462: %
3463: \begin{align*}
3464: \tau^* \hat F &= g_{\alpha,\l}\,(\tau^* F)\,
3465: g^{-1}_{\alpha,\tau^*\l^\prime} \\
3466: &= g_{\alpha,\l}\,\chi(\tau) \,F\,
3467: k\,g^{-1}_{\alpha,\l''} \mbox{ with }
3468: \mathrm{L}'' = \bar{k}^t\,
3469: \overline{F(\alpha)}^t \,
3470: \overline{\chi(\alpha,\tau)}^t \mathrm{L} =
3471: \tau^* \mathrm{L}' \\
3472: &= g_{\alpha,\l}\,\chi(\tau)\, g_{\alpha,\l}^{-1}
3473: \,g_{\alpha,\l} \,F \, g^{-1}_{\alpha,\l_1 }
3474: k \mbox{ by } \eqref{eq:change}
3475: \end{align*}
3476: %
3477: where $\mathrm{L}_1 = \overline{F(z,\alpha)}^t\,
3478: \overline{\chi(\alpha,\tau)}^t \mathrm{L}$.
3479:
3480: If $g_{\alpha,\l}\,\chi(\tau)\, g_{\alpha,\l}^{-1}$
3481: is a monodromy of $\hat F$ and we write
3482: $\tau^* \hat{F} = g_{\alpha,\l}\,\chi(\tau)\,
3483: g_{\alpha,\l}^{-1} \, \hat{F} \,\hat{k}$
3484: and combine with the above, we obtain
3485: $g_{\alpha,\l_1}^{-1}\,k =
3486: g_{\alpha,\l'}^{-1} \,\hat{k}$.
3487: Hence $g_{\alpha,\l_1}g_{\alpha,\l'}^{-1}$
3488: is independent of $\lambda$ and since there exists a
3489: $\lambda_0 \in \CP$ at which $\tau_\alpha(\lambda_0) = 1$ we obtain
3490: $g_{\alpha,\l_1}g_{\alpha,\l'}^{-1} = \Id$.
3491: This implies that $L' = L_1$ which implies \eqref{eq:invariance}.
3492: The converse is proven by direct verification.
3493: \QED
3494:
3495: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3496: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3497:
3498: \section{Dressing non--finite type surfaces}
3499:
3500: In this section we discuss a family of CMC cylinders
3501: found in \cite{KilMS} that arise as perturbations
3502: of Delaunay surfaces.
3503: The resulting surfaces may possess an arbitrary
3504: number of umbilics and are thus not of finite type and
3505: are CMC cylinders with one Delaunay end \cite{KilKRS}, see Figure 1.
3506: We can dress this class of cylinders with simple factors, having the
3507: effect of adding bubbles to the surface, see Figure 2.
3508:
3509: %
3510: \begin{figure}\label{fig:pert}
3511: \centering
3512: \includegraphics[scale=1.5]{pert.eps}
3513: \includegraphics[scale=1.3]{smythdelbub.eps}\\
3514: \caption{\footnotesize A perturbed round cylinder on the left. A perturbed
3515: and dressed round cylinder on the right.
3516: Images generated with \emph{CMCLab} \cite{Sch:cmclab}. For more images see
3517: \cite{Sch:gallery}.}
3518: \end{figure}
3519: %
3520: %
3521:
3522:
3523: \newsection{\bf Dressing cylinders with umbilics.}
3524: We modify a standard result from the theory of
3525: differential equations with regular singular points.
3526: Since the eigenvalues are $\lambda$--dependent,
3527: we can avoid the assumption that two elements in the
3528: spectrum not differ by an integer by working on an
3529: appropriate $\lambda$--circle $C_r$.
3530: %
3531:
3532: \mytheorem{} \label{th:zap}
3533: Let $U_0^* \subset \C$ be an open neighbourhood of
3534: $z=0$ and $\xi \in \Lambda \Omega(U_0^*)$ with a
3535: simple pole at $z=0$ and residue
3536: $\mathrm{res}_{\,0} \xi = D$
3537: where $D$ has the form \eqref{eq:delaunay}.
3538: Moreover, we assume that $D$
3539: satisfies the first closing condition
3540: \eqref{eq:closing1}. Then there exists
3541: an $r \in (0,1)$ and a solution of
3542: $d\Psi = \Psi \xi$ of the form $\Psi = z^D P$
3543: with $P:U_0 \to \LGC$ holomorphic.
3544: Further, $\Psi$ has Delaunay monodromy and is the
3545: holomorphic frame associated with $(\xi,\,P(1),\,1)$.
3546: %
3547:
3548: \Proof
3549: Let $U_0 = U_0^* \cup \{ 0 \}$ and write
3550: $\xi = D \tfrac{dz}{z} + \eta$ with
3551: $\eta \in \Lambda \Omega_{U_0}$. The differential
3552: equation for $P$ is
3553: $dP = P \eta + [ P,\, D \tfrac{dz}{z}]$.
3554: We will show that this differential equation has a
3555: holomorphic solution. Expanding
3556: $P = \sum_{k = 0}^{\infty} P_k z^k$ and
3557: $\xi = (\tfrac{1}{z}D +\Sigma \eta_k z^k )dz$
3558: at $z=0$, the coefficients are recursively given
3559: by
3560: %
3561: $$
3562: kP_k + [D , P_k ] = \sum_{r+s=k-1} P_r \eta_s.
3563: $$
3564: %
3565: By our Ansatz $P_k = 0$ for $k < 0$ and we are free
3566: to choose $P_0 \in \LGC$ with $[D , P_0 ]=0$.
3567: (Note this freedom only means
3568: $P_0 = \alpha \Id + \beta D$,
3569: since $D$ is a semisimple $2 \times 2-$matrix
3570: with different eigenvalues.)
3571: The eigenvalues of $D$, are of the form $\pm \mu$,
3572: where $\mu$ is a solution to the equation
3573: $\mu^2 = - \det D$. Then
3574: the operators $k \rm{Id} + \rm{ad}_{D}$,
3575: $k \in \Z$, have spectrum
3576: $\sigma = \{ k,\, k \pm 2\mu \}$.
3577: We are only interested in $k \geq 1$, in which case
3578: $0 \notin \sigma$ if and only if
3579: $\mu \notin \tfrac{1}{2} \N$.
3580:
3581: We need to show that there exists an
3582: $r \in (0,\,1)$ for which
3583: $\mu (\lambda) \notin \frac{1}{2}\N$ for every
3584: $\lambda \in C_r$.
3585: A priori, $\mu(\lambda) \in \R$ if and only if
3586: either $\lambda \in S^1$ or $\lambda \in \R$.
3587: Since we seek $r \in (0,\,1)$ we have
3588: $\mu(\lambda) \in \R$ if and only if
3589: $\lambda \in \R$ we need to ensure
3590: $\mu(\pm r)\notin \tfrac{1}{2}\N$.
3591: Since $D$ satisfies the first closing
3592: condition by assumption, we know that
3593: $\mu(1) = n/2$ for some $n \in \Z$
3594: the winding number of the Delaunay surface.
3595: A direct computation shows
3596: $\mu(\pm r) \notin \tfrac{1}{2}\Z$
3597: if and only if
3598: $r^{-2} + r^2 \neq \pm (k^2 - n^2)/(4ab) + 2$
3599: for all $k \in \N$.
3600: The sequences $a^{\pm}_k :=
3601: \pm (k^2 - n^2)/(4ab) + 2$
3602: are monotonic, hence we can always find an
3603: $0 < r <1$ for which $r^{-2} + r^2
3604: \notin \{ a^{\pm}_k \}$.
3605:
3606: This proves the existence of a formal power
3607: series solving the differential equation for $P$.
3608: The coefficients of this formal power series are
3609: defined for all $z \in U_0$. The domain of
3610: analyticity of $P$ is $U_0$ by standard
3611: ODE arguments, see e.g \cite{Har}. It is shown in
3612: \cite{DorW:noids} that $P$ has the twisted $\lambda$
3613: behaviour and that the coefficients of $P$ as
3614: functions of $\lambda$ are in $\mathcal{A}_r$.
3615: To show that $\Psi$ has Delaunay
3616: monodromy, recall from \ref{sec:Delaunayexample}
3617: that the Delaunay monodromy is $\exp(2 \pi i D)$.
3618: For the translation $\tau(z) = z + 2\pi i$
3619: the monodromy of $\Psi$ is $\tau^*\Psi \Psi^{-1} =
3620: \tau^*z^D \tau^*P P^{-1} z^{-D} = \exp(2 \pi i D)$.
3621: \QED
3622: %
3623:
3624: For our purposes, the specific form of $\Psi$
3625: proven above is of great importance to us,
3626: since it gives easy control over the monodromy
3627: singled out by the singularity at $D$.
3628: In particular, if we dress the corresponding
3629: frame by some simple factor it is easy to
3630: determine the dressed monodromy, at least if
3631: the associated surface has umbilical points.
3632: From a geometric point of view this is
3633: particularly interesting, since by a result
3634: of \cite{Mah}, the classical Bianchi--B\"acklund
3635: transformation does correspond to dressing
3636: by some simple factor. Actually, in the twisted
3637: setup, our 'simple factors' are in fact a product
3638: of two simple factors. These correspond to the
3639: two step procedure of the Bianchi--B\"acklund
3640: transformation. As a matter of fact,
3641: by making the right choice for the singularity
3642: of the simple factor, is is possible to control the
3643: topology of the resulting surface.
3644:
3645: \myremark{}
3646: 1) The proof of the Theorem above actually shows
3647: that there is some $0 < r < 1$ such that for all
3648: $0 < r \leq r' <1$ the claim holds.
3649:
3650: 2) Instead of choosing some $0 < r < 1$,
3651: one could ask for what potentials $\xi$
3652: one can obtain a solution of the
3653: form $ \Psi = z^D \, P$ without
3654: singularities on $S^1$. Once this is
3655: achieved, one can conclude as above and
3656: infer statements about the monodromy and
3657: thus one can clear the way for some
3658: understanding for the effect of dressing
3659: by simple factors. Such a condition has
3660: been given in \cite{DorW:noids}.
3661:
3662: In order to dress a CMC surface with non-trivial
3663: topology into a new CMC surface with the same
3664: topology, we recall from \cite{DorH:per} when for a
3665: given triple $(\xi,\Phi_0,\tilde{z}_0)$ one
3666: specific member of the associated family is
3667: invariant under $\Delta$, thus characterising
3668: the period problem in the DPW framework:
3669:
3670: Let $f_\lambda$ be an associated family.
3671: Let $F$ be the extended unitary
3672: frame of the surface with monodromy
3673: $\chi : \Delta \to \Lhol$.
3674: Then there exists a $\lambda_0 \in S^1$ such
3675: that $\tau^* f_{\lambda_0}= f_{\lambda_0}$ for all
3676: $\tau \in \Delta$ if and only if for all
3677: $\tau \in \Delta$, $\chi$ satisfies both
3678: %
3679: \begin{align}
3680: \chi (\tau,\,\lambda_0) =
3681: \pm \rm{Id},&\label{eq:closing1}\\
3682: \left. \tfrac{\partial}{\partial \lambda}
3683: \right|_{\lambda_0}
3684: \chi(\tau,\,\lambda)
3685: = 0.& \label{eq:closing2}
3686: \end{align}
3687: %
3688:
3689: \mycorollary{}
3690: Let $\xi \in \Lambda \Omega(\C^*)$ as in Theorem
3691: \ref{th:zap} and $g_{\alpha,\l} \in \LGP$ be a
3692: simple factor. Choosing $r \in (0,\,1)$ as above
3693: and $\alpha \in \C$ with $|\alpha| \in (r,\,1)$
3694: such that
3695: %
3696: \begin{equation} \label{eq:det}
3697: \det D(\alpha) = \tfrac{n}{4} \mbox{ for some } n \in \Z,
3698: \end{equation}
3699: %
3700: then $(\xi,\, g_{\alpha,\l}\, P(1), \,1)$
3701: generates a CMC cylinder (with winding number $n$).
3702:
3703: \Proof
3704: Let $\Phi = z^D\,P$ and $F$ be the holomorphic
3705: respectively unitary frame generated by
3706: $(\xi ,\, P(1),\, 1)$.
3707: Both have Delaunay monodromy
3708: $\chi (\tau,\,\lambda) = \exp(2\pi iD(\lambda))$
3709: with respect to the translation
3710: $\tau(z) = z + 2\pi i$.
3711: The condition \eqref{eq:det} ensures that
3712: $\chi(\tau,\,\alpha) = \pm \Id$ and
3713: consequently \eqref{eq:invariance} holds
3714: for any choice of line $\mathrm{L} \in \CP$.
3715: Hence, by Corollary \ref{th:invariance},
3716: $\hat\chi(\tau) =
3717: g_{\alpha,\l}\,\chi(\tau)\,g_{\alpha,\l}^{-1}$
3718: is a monodromy of $g_{\alpha,\l} \# F$.
3719: The conditions \eqref{eq:closing1} and
3720: \eqref{eq:closing2} are verified using the facts
3721: $\chi(\tau,1) = \pm \mathrm{Id}$ and $\left.
3722: \partial_{\lambda}\chi(\tau)\right|_{\lambda=1}= 0$.
3723: \QED
3724:
3725:
3726: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3727:
3728: \newsection{\bf Dressing $3$-Noids.}
3729: %
3730: Let us briefly outline how the above theory can be
3731: applied to construct the dressed $3$-Noids of
3732: \cite{KilSS}. Let
3733: $\mathcal{T} = \CP \setminus \{0,\,1,\,\infty\}$
3734: and $f: \mathcal{T} \to \R^3$ a CMC $3$-Noid with
3735: Delaunay ends as constructed in \cite{DorW:noids}
3736: or \cite{Sch:tri}. Let
3737: $F \in \mathcal{F}(\widetilde{\mathcal{T}})$
3738: be an extended frame such that at $\lambda = 1$ the
3739: immersion $f$ is obtained via the Sym--Bobenko
3740: formula \eqref{eq:Sym}.
3741: Then there are three monodromies,
3742: $\chi_1,\, \chi_2,\,\chi_3$, one for each end,
3743: which can be computed in terms of
3744: $\Gamma$--functions \cite{DorW:noids},
3745: \cite{KilSS} and satisfy
3746: $\prod_{j=1}^3 \chi_j = \Id$.
3747:
3748: In \cite{KilSS} it was shown that there exist
3749: values $\alpha \in \C^*,\,|\alpha |<1$ and
3750: invariant subspaces $\mathrm{L} \in \CP$ such that
3751: \eqref{eq:invariance} holds for all three
3752: $\chi_j$'s. Consequently, by Corollary
3753: \ref{th:invariance} the unitary frame
3754: obtained by dressing
3755: with the simple factor $g_{\alpha,\l}$
3756: has monodromies $\hat \chi_j =
3757: g_{\alpha,\l} \, \chi_j \, g_{\alpha,\l}^{-1}$ and
3758: since the closing conditions \eqref{eq:closing1} and \eqref{eq:closing2} are invariant under conjugation,
3759: the resulting surface is again a
3760: CMC $3$--Noid.
3761: %
3762: %.Adding a bubble to the first end can be
3763: %accomplished by dressing with $h^1C^1$, with $h^1 \in \LGP$ as in the
3764: %example of section ?? and $C \in \LGC$ with $[\, C^1, H^1 \, ]=0$.
3765: %If we want to dress from trinoids with embedded ends to trinoids
3766: %with embedded ends, by Theorem \ref{th:mainresult} we need to have
3767: %that $h^1 C^1 H_j (h^1C^1)^{-1} \in \LGU$ for all $r>0$. Moreover,
3768: %the closing conditions \eqref{eq:closing1},\eqref{eq:closing2}
3769: %have to be satisfied. For $j=1$ this all holds by \cite{Kil},
3770: %Theorem 6.2.2.
3771:
3772:
3773: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3774: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3775:
3776: \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
3777: \providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR }
3778: % \MRhref is called by the amsart/book/proc definition of \MR.
3779: \providecommand{\MRhref}[2]{%
3780: \href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2}
3781: }
3782: \providecommand{\href}[2]{#2}
3783: \begin{thebibliography}{10}
3784:
3785: \bibitem{Bob:tor}
3786: A.~I. Bobenko, \emph{All constant mean curvature tori in {$\mathbb{R}^3$},
3787: {$\mathbb{S}^3$}, {$\mathbb{H}^3$} in terms of theta-functions}, Math. Ann.
3788: \textbf{290} (1991), 209--245.
3789: \bibitem{BobPS}
3790: A.~I. Bobenko, T.~V. Pavlyukevich and
3791: B. A. Springborn,
3792: \emph{Hyperbolic constant mean curvature one
3793: surfaces: spinor representation and trinoids in
3794: hypergeometric functions}, Math.Z \textbf{245}, (2003), 63--91.
3795:
3796:
3797: \bibitem{Bun}
3798: L.~Bungart, \emph{On analytic fiber bundles}, Topology \textbf{7} (1968),
3799: 55--68.
3800:
3801: \bibitem{Bur:iso}
3802: F.~Burstall, \emph{Isothermic surfaces: conformal geometry, {C}lifford algebras
3803: and integrable systems.}, Integrable systems, Geometry and Topology, vol.~80,
3804: taiwan, 1995, math.DG/0003096, pp.~353--382.
3805:
3806: \bibitem{BurP:dre}
3807: F.~Burstall and F.~Pedit, \emph{Dressing orbits of harmonic maps}, Duke Math.
3808: J. \textbf{80} (1995), 353--382.
3809:
3810: \bibitem{DorH:per}
3811: J.~Dorfmeister and G.~Haak, \emph{On constant mean curvature surfaces with
3812: periodic metric}, Pacific J. Math. \textbf{182} (1998), 229--287.
3813:
3814: \bibitem{DorH:dre}
3815: \bysame, \emph{Investigation and application of the dressing action on surfaces
3816: of constant mean curvature}, Q. J. Math. \textbf{51} (2000), 57--73.
3817:
3818: \bibitem{DorH:sym2}
3819: \bysame, \emph{On symmetries of constant mean curvature surfaces. {II}.
3820: symmetries in a {W}eierstrass-type representation}, Int. J. Math. Game Theory
3821: and Algebra \textbf{10} (2000), 121--146.
3822:
3823: \bibitem{DorH:cyl}
3824: \bysame, \emph{Construction of non-simply connected {CMC} surfaces via
3825: dressing}, J. Math. Soc. Japan \textbf{55} (2003), no.~2, 335--364.
3826:
3827: \bibitem{DorPW}
3828: J.~Dorfmeister, F.~Pedit, and H.~Wu, \emph{Weierstrass type representation of
3829: harmonic maps into symmetric spaces}, Comm. Anal. Geom. \textbf{6} (1998),
3830: no.~4, 633--668.
3831:
3832: \bibitem{DorW:noids}
3833: J.~Dorfmeister and H.~Wu, \emph{Construction of constant mean curvature
3834: $n$-noids from holomorphic potentials}, in preparation, 2002.
3835:
3836: \bibitem{Har}
3837: P.~Hartman, \emph{Ordinary differential equations}, John Wiley and Sons, Inc.,
3838: New York, 1964.
3839:
3840: \bibitem{Kil:thesis}
3841: M.~Kilian, \emph{Constant mean curvature cylinders}, Ph.D. thesis, Univ. of
3842: Massachusetts, Amherst, 2000.
3843:
3844: \bibitem{Kil:del}
3845: \bysame, \emph{On the associated family of {D}elaunay surfaces},
3846: Proc. AMS, in press.
3847:
3848: \bibitem{KilKRS}
3849: M.~Kilian, S.~Kobayashi, W.~Rossman, and N.~Schmitt, \emph{{CMC} surfaces in
3850: three dimensional space forms}, math.DG/0403366.
3851:
3852: \bibitem{KilMS}
3853: M.~Kilian, I.~McIntosh, and N.~Schmitt, \emph{New constant mean curvature
3854: surfaces}, Experiment. Math. \textbf{9} (2000), no.~4, 595--611.
3855:
3856: \bibitem{KilSS}
3857: M.~Kilian, N.~Schmitt, and I.~Sterling, \emph{Dressing {CMC} n-{N}oids},
3858: Math. Z., \textbf{246}, (2004), 501--519.
3859:
3860: \bibitem{Kob}
3861: S.~Kobayashi, \emph{Bubbletons in 3-dimensional space forms via the {DPW}
3862: method}, Master's thesis, Kobe University, 2001.
3863:
3864: \bibitem{Mah}
3865: A.~Mahler, \emph{{B}ianchi-{B}\"acklund and dressing transformations on
3866: constant mean curvature surfaces}, Ph.D. thesis, Univ. of Toledo, 2002.
3867:
3868: \bibitem{McI}
3869: I.~McIntosh, \emph{Global solutions of the elliptic 2d periodic {T}oda
3870: lattice}, Nonlinearity \textbf{7} (1994), no.~1, 85--108.
3871:
3872: \bibitem{PinS}
3873: U.~Pinkall and I.~Sterling,
3874: \emph{On the classification of constant mean
3875: curvature tori}, Ann. of Math \textbf{130} (1989),
3876: 407--451.
3877:
3878: \bibitem{RuhV}
3879: E.~A.~Ruh and J.~Vilms,
3880: \emph{The tension field of the Gauss map},
3881: Trans. Amer. Math. Soc. \textbf{149} (1970),
3882: 569--573.
3883:
3884: \bibitem{Sch:cmclab}
3885: N.~Schmitt, \emph{{CMCL}ab}, http://www.gang.umass.edu/software.
3886:
3887: \bibitem{Sch:gallery}
3888: \bysame, \emph{{CMC} {G}allery}, http://www.gang.umass.edu/cmcgallery.
3889:
3890: \bibitem{Sch:tri}
3891: \bysame, \emph{Constant mean curvature trinoids}, math.DG/0403036.
3892:
3893: \bibitem{SteW}
3894: I.~Sterling and H.~Wente, \emph{Existence and classification of constant mean
3895: curvature multibubbletons of finite and infinite type}, Indiana Univ. Math.
3896: J. \textbf{42} (1993), no.~4, 1239--1266.
3897:
3898: \bibitem{TerU}
3899: C.~Terng and K.~Uhlenbeck, \emph{B\"{a}cklund transformations and loop group
3900: actions}, Comm. Pure and Appl. Math \textbf{LIII} (2000), 1--75.
3901:
3902: \bibitem{Uhl}
3903: K.~Uhlenbeck, \emph{Harmonic maps into Lie groups
3904: (Classical solutions of the chiral model)},
3905: J. Diff. Geom. \textbf{30} (1989), 1--50.
3906:
3907:
3908: \bibitem{Wu:dre}
3909: H.~Wu, \emph{On the dressing action of loop groups on constant mean curvature
3910: surfaces}, T\^{o}hoku Math. J. \textbf{49} (1997), 599--621.
3911:
3912: \bibitem{ZakS2}
3913: V.~E.~Zakharov and A.~S.~Shabat,
3914: \emph{Integration of the nonlinear equations of
3915: mathematical physics by the method of the inverse
3916: scattering problem II},
3917: Funktsional. Anal. i Prilozhen. \textbf{13} (1978),
3918: 13--22.
3919:
3920: \end{thebibliography}
3921:
3922:
3923: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3924:
3925:
3926: \end{document}
3927:
3928: