1: \def\Datum{October 10, 2006}
2: \magnification=\magstephalf
3: \hsize=16.2truecm
4: \vsize24truecm
5: \parskip4pt plus 1pt
6: \frenchspacing
7: \lineskiplimit=-2pt
8: \vglue6truemm
9:
10: \def \Eins {1}
11: \def \Zwei {2}
12: \def \GP {2.1}
13: \def \LE {2.2}
14: \def \RM {2.3}
15: \def \Drei {3}
16: \def \EL {3.1}
17: \def \CR {3.2}
18: \def \HM {3.3}
19: \def \HN {3.4}
20: \def \UR {3.5}
21: \def \AF {3.6}
22: \def \US {3.7}
23: \def \PY {3.8}
24: \def \ZI {3.9}
25: \def \UQ {3.10}
26: \def \Vier {4}
27: \def \HJ {4.1}
28: \def \DC {4.2}
29: \def \ZU {4.3}
30: \def \ZB {4.4}
31: \def \ZA {4.5}
32: \def \ZV {4.6}
33: \def \PL {4.7}
34: \def \LO {4.8}
35: \def \PS {4.10}
36: \def \PQ {4.11}
37: \def \Fuenf {5}
38: \def \JP {5.1}
39: \def \AQ {5.2}
40: \def \PI {5.3}
41: \def \EB {5.4}
42: \def \NP {5.5}
43: \def \Sechs {6}
44: \def \CO {6.1}
45: \def \PR {6.3}
46: \def \CP {6.4}
47: \def \UH {6.5}
48: \def \UK {6.6}
49: \def \JG {6.7}
50: \def \UL {6.8}
51: \def \PT {6.9}
52: \def \JA {6.10}
53: \def \JC {6.11}
54: \def \KZ {6.12}
55: \def \KX {6.13}
56: \def \KI {6.14}
57: \def \KY {6.15}
58: \def \HA {6.16}
59: \def \HU {6.17}
60: \def \UF {6.18}
61: \def \VZ {6.19}
62: \def \KR {6.20}
63: \def \Sieben {7}
64: \def \IO {7.1}
65: \def \IP {7.2}
66: \def \ZR {7.3}
67: \def \AP {7.4}
68: \def \GO {7.5}
69: \def \KO {7.6}
70: \def \BP {7.7}
71: \def \MO {7.8}
72: \def \KW {7.9}
73: \def \KU {7.10}
74: \def \LP {7.11}
75: \def \HQ {7.12}
76: \def \RU {7.13}
77: \def \Acht {8}
78: \def \EI {8.1}
79: \def \EY {8.2}
80: \def \EZ {8.3}
81: \def \EX {8.4}
82: \def \EV {8.5}
83: \def \NM {8.6}
84: \def \LS {8.7}
85: \def \UA {8.8}
86: \def \UB {8.9}
87: \def \OQ {8.10}
88: \def \IZ {8.11}
89: \def \Neun {9}
90: \def \FJ {9.1}
91: \def \XA {9.2}
92: \def \FA {9.4}
93: \def \CA {9.5}
94: \def \MC {9.6}
95: \def \FC {9.7}
96: \def \FD {9.8}
97: \def \ST {9.9}
98: \def \PA {9.10}
99: \def \FG {9.11}
100: \def \Zehn {10}
101: \def \SS {10.1}
102: \def \LA {10.2}
103: \def \Elf {11}
104: \def \Zwoelf {12}
105: \def \LK {12.1}
106: \def \FM {12.2}
107: \def \EM {12.3}
108: \def \NO {12.4}
109: \def \FN {12.5}
110: \def \FO {12.6}
111: \def \FP {12.7}
112: \def \FQ {12.8}
113: \def \FR {12.9}
114: \def \FS {12.10}
115: \def \ZC {12.11}
116: \def \VB {12.12}
117: \def \Dreizehn {13}
118: \def \VD {13.1}
119: \def \Vierzehn {14}
120: \def \FT {14.1}
121: \def \FU {14.2}
122: \def \XY {14.3}
123: \def \FV {14.4}
124: \def \FW {14.5}
125: \def \Fuenfzehn {15}
126: \def \FI {15.1}
127: \def \FY {15.2}
128: \def \VJ {15.3}
129: \def \FZ {15.4}
130: \def \GQ {15.5}
131: \def \GR {15.6}
132: \def \VF {15.7}
133: \def \VG {15.8}
134: \def \VA {15.9}
135: \def \VK {15.12}
136: \def \VL {15.13}
137: \def \GS {15.14}
138: \def \VM {15.17}
139: \def \VN {15.18}
140: \def \VO {15.20}
141: \def \VP {15.21}
142: \def \Sechzehn {16}
143: \def \VQ {16.1}
144: \def \GT {16.2}
145: \def \GU {16.3}
146: \def \GV {16.5}
147: \def \GX {16.6}
148: \def \ANFR {1}
149: \def \ANHI {2}
150: \def \BELO {3}
151: \def \BHUR {4}
152: \def \BERO {5}
153: \def \BAHR {6}
154: \def \BARZ {7}
155: \def \BOGG {8}
156: \def \BOUR {9}
157: \def \BUSH {10}
158: \def \CART {11}
159: \def \CHMO {12}
160: \def \DAYA {13}
161: \def \DOKR {14}
162: \def \EAEZ {15}
163: \def \EBFT {16}
164: \def \EBEN {17}
165: \def \FELS {18}
166: \def \FEKA {19}
167: \def \GAME {20}
168: \def \HOCH {21}
169: \def \KNAP {22}
170: \def \HART {23}
171: \def \HORM {24}
172: \def \ISMI {25}
173: \def \KAZT {26}
174: \def \LOBO {27}
175: \def \LOBP {28}
176: \def \LOBQ {29}
177: \def \MENA {30}
178: \def \PALA {31}
179: \def \SAKA {32}
180: \def \SEVL {33}
181: \def \TANA {34}
182:
183:
184: \input cls.def\immediate\openout\aux=\jobname.aux
185:
186:
187: \arxtrue
188:
189: \ifdraft
190: \footline={\hss\sevenrm \Datum\hss}
191: \else\nopagenumbers\fi
192:
193: \headline={\ifnum\pageno>1\sevenrm\ifodd\pageno Homogeneous Levi
194: degenerate CR-manifolds
195: \hss{\tenbf\folio}\else{\tenbf\folio}\hss{Fels-Kaup}
196: \fi\else\hss\fi}
197:
198:
199: \centerline{\Gross Classification of Levi degenerate homogeneous}
200: \bigskip
201: \centerline{\Gross CR-manifolds in dimension 5}
202: \bigskip\bigskip
203: \centerline{Gregor Fels\quad and\quad Wilhelm Kaup}
204: \bigskip
205: {\parindent0pt\footnote{}{\ninerm 2000 Mathematics
206: Subject Classification: 32M17, 32M25, 32V25.
207: }}
208:
209: \def\Partskip{\vskip9mm}
210:
211: \bigskip\bigskip
212:
213: \ifarx
214: {\narrower\narrower\Klein{\bf\noindent Abstract:} In this paper we present new
215: examples of homogeneous $2$-nondegenerate CR-manifolds of dimension 5
216: and give, up to local CR-equivalence, a full classification of all
217: CR-manifolds of this type.\par}
218:
219: \bigskip\bigskip
220:
221: {\def \Einsa {1}
222: \def \Zweia {6}
223: \def \Dreia {7}
224: \def \Viera {11}
225: \def \Fuenfa {14}
226: \def \Sechsa {16}
227: \def \Siebena {22}
228: \def \Achta {26}
229: \def \Neuna {29}
230: \def \Zehna {33}
231: \def \Elfa {36}
232: \def \Zwoelfa {38}
233: \def \Dreizehna {41}
234: \def \Vierzehna {43}
235: \def \Fuenfzehna {44}
236: \def \Sechzehna {49}
237: \def \Refa {52}
238: \SkIp=70pt
239: \advance\baselineskip8pt\Klein\advance\hsize-60pt
240: \line{\hbox to \SkIp{\hfill\Eins. }Introduction\dotfill\Einsa}
241: \vskip8pt
242: \line{\hbox to 90pt{\hfill}\elfrm PART 1: Levi degenerate CR-manifolds\hfill}
243: \vskip1pt
244: \line{\hbox to \SkIp{\hfill\Zwei. }Preliminaries\dotfill\Zweia}
245: \line{\hbox to \SkIp{\hfill\Drei. }Tube manifolds\dotfill\Dreia}
246: \line{\hbox to \SkIp{\hfill\Vier. }Tube manifolds over cones\dotfill\Viera}
247: \line{\hbox to \SkIp{\hfill\Fuenf. }Some examples\dotfill\Fuenfa}
248: \line{\hbox to \SkIp{\hfill\Sechs. }Levi degenerate CR-manifolds associated with an
249: endomorphism\dotfill\Sechsa}
250: \line{\hbox to \SkIp{\hfill\Sieben. }Homogeneous 2-nondegenerate manifolds of CR-dimension 2\dotfill\Siebena}
251: \line{\hbox to \SkIp{\hfill\Acht. }Homogeneous 2-nondegenerate CR-manifolds in dimension 5\dotfill\Achta}
252: \vskip8pt
253: \line{\hbox to 90pt{\hfill}\elfrm PART 2: The classification\hfill}
254: \vskip1pt
255: \line{\hbox to \SkIp{\hfill\Neun. }Lie-theoretic characterization of locally homogeneous
256: CR-manifolds\dotfill\Neuna}
257: \line{\hbox to \SkIp{\hfill\Zehn. }The Lie algebra $\7g$ has small semisimple part\dotfill\Zehna}
258: \line{\hbox to \SkIp{\hfill\Elf. }The cases $\7s\cong\so(2,3)$ and $\7s\cong\so(1,3)$\dotfill\Elfa}
259: \line{\hbox to \SkIp{\hfill\Zwoelf. }The case $\7s\cong\7{sl}(2,\RR)$\dotfill\Zwoelfa}
260: \line{\hbox to \SkIp{\hfill\Dreizehn. }The case $\7s\cong\su(2)$\dotfill\Dreizehna}
261: \line{\hbox to \SkIp{\hfill\Vierzehn. }Reduction to the case where $\7g$ is solvable and of dimension
262: 5\dotfill\Vierzehna}
263: \line{\hbox to \SkIp{\hfill\Fuenfzehn. }The existence of a 3-dimensional abelian ideal in
264: $\7g$ suffices\dotfill\Fuenfzehna}
265: \line{\hbox to \SkIp{\hfill\Sechzehn. }The final steps\dotfill\Sechzehna}
266: \vskip4pt
267: \line{\hbox to \SkIp{\hfill}References\dotfill\Refa}
268: }\fi
269:
270: \newwrite\lst\immediate\openout\lst=\jobname.lst
271:
272:
273: \KAP{Eins}{Introduction}
274:
275: The main topic of this paper is the study of real-analytic
276: CR-manifolds $M$ with everywhere degenerate Levi form. In particular,
277: for homogeneous manifolds of this type, we develop methods for the
278: computation of the Lie algebra s$\hol(M,a)$ of infinitesimal
279: CR-transformations at every $a\in M$. We also classify up to local
280: CR-equivalence all locally homogeneous degenerate
281: CR-manifolds in dimension 5.
282:
283: In this context, a well studied example of a homogeneous Levi degenerate
284: CR-manifold is the quadratic hypersurface
285: $$
286: \5M:=\{z\in\CC^{3}: (\Re z_{1})^{2}+(\Re z_{2})^{2}=(\Re
287: z_{3})^{2},\;\Re z_{3} >0\} \;,
288: $$ compare
289: e.g. \Lit{EBFT}, \Lit{FEKA}, \Lit{KAZT} and \Lit{SEVL}. This
290: 5-dimensional CR-manifold has several remarkable properties and
291: serves as motivation for various considerations in this paper.
292: Notice that $\5M$ can also be written as tube manifold
293: $$\5M=\5F+i\RR^{3}\subset\CC^{3},\steil{where}\5F:=\{x\in\RR^{3}:
294: x^{2}_{1}+x^{2}_{2}=x^{2}_{3},\,x_{3}>0\}$$ is the future light cone
295: in 3-dimensional space-time. A glance at this description shows that
296: $\5M$ is homogeneous under a group of complex-affine
297: transformations. It is less obvious that the Lie algebras of {\sl
298: global} and {\sl local} infinitesimal CR-transformations at $a\in\5M$,
299: $\hol(\5M)$ and $\hol(\5M,a)$, are both isomorphic to $\so(2,3)$, and
300: hence have dimension 10, compare \Lit{KAZT}. Also the following
301: `globalization' is known: The group $\SO(2,3)$ acts on the complex
302: quadric $\fett\3Q_3\subset\PP_{4}(\CC)$ by biholomorphic
303: transformations and has a hypersurface orbit that contains $\5M$ as a
304: dense domain.
305:
306: \ifarx\medskip\fi
307:
308: The cone $\5F$ clearly is a disjoint union of affine
309: half-lines. Therefore, $\5M$ is a disjoint union of complex
310: half-planes, actually $\5M$ is a fiber bundle with typical fiber
311: $\HH^{+}:=\{z\in\CC:\Re(z)>0\}$. However, the total space $\5M$ is not
312: even locally CR-equivalent to a product $\HH^+\Times M'$ with $M'$ a
313: CR-manifold. Notice that the Levi form in both cases, i.e., for $\5M$
314: and for a product of $\HH^{+}$ with a Levi nondegenerate 3-dimensional
315: CR-manifold $M'$, has exactly one non-zero eigenvalue at every
316: point. Hence, one needs further invariants to distinguish those
317: CR-manifolds. While every product $M'\times\CC$ is holomorphically
318: degenerate, the crucial fact here is that the light cone tube $\5M$ is
319: nondegenerate in a higher order sense: To be precise, $\5M$ is
320: 2-nondegenerate at every point, and we refer to \Lit{BAHR} and also to
321: \Lit{BERO}, 11.1, for the notion of $k$-nondegeneracy.
322:
323: \ifarx\medskip\fi
324:
325: In the non-homogeneous setting, for every $k\in \NN$ and fixed
326: manifold dimension it is not difficult to construct large classes of
327: CR-manifolds, even hypersurfaces, which are $k$-nondegenerate at some
328: points, but are Levi nondegenerate in a dense open subset. It seems
329: to be much harder to construct CR-manifolds which are
330: $k$-nondegenerate everywhere for $k\ge 2$. Note that the CR-dimension
331: of a {\sl homogeneous} $M$ is an upper bound for the degree $k$ of
332: nondegeneracy. Hence, the lowest manifold dimension for which
333: everywhere $2$-nondegenerate CR-manifolds can exist is 5. which
334: raises the intriguing question whether besides the light cone tube
335: there exist further 2-nondegenerate homogeneous CR-manifolds in
336: dimension 5. So far, all known examples in dimension finally turned
337: out to be locally CR-isomorphic to $\5M$, compare e.g. \Lit{EBEN},
338: \Lit{KAZT}, \Lit{GAME}, \Lit{FEKA}, \Lit{BELO}, and the belief arose
339: that there are no others.
340:
341: \ifarx\medskip\fi
342:
343: The main objective of this paper is to show that there are actually
344: infinitely many locally mutually non-equivalent examples and to
345: provide a full classification. Starting point is the following simple
346: observation: Suppose that $F\subset\RR^{n}$ is an affinely homogeneous
347: (connected) submanifold of dimension say $d<n$. Then the tube
348: $M:=F+i\RR^{n}$ is a generic CR-submanifold of $\CC^{n}$ of
349: CR-dimension $d$ and is homogeneous under a group of complex-affine
350: transformations. Indeed, every real-affine transformation leaving $F$
351: invariant extends to a complex-affine transformation leaving $M$
352: invariant and, furthermore, $M$ is invariant under all translations
353: $z\mapsto z+iv$ with $v\in\RR^{n}$. Clearly, the crucial question is,
354: when is $M$ $k$-nondegenerate and when are two tubes $M,M'$ of this
355: type locally CR-equivalent?
356:
357: \ifarx\medskip\fi
358:
359: The classification of all affinely homogeneous surfaces
360: $F\subset\RR^{3}$ can be found in \Lit{DOKR} and \Lit{EAEZ}. In
361: particular, a complete list (up to local affine equivalence and in a
362: slightly different formulation) of all degenerate types that are not a
363: cylinder, is given by the following examples (1) -- (3).
364:
365: \item{(1)} $F=\5F$ the future light cone as above.
366:
367: \item{(2a)} $F=\{\,r\,(\cos t,\sin t,e^{\omega
368: t})\in\RR^{3}:r\in\RRp,t\in\RR\}$ with $\omega>0$ arbitrary.
369:
370: \item{(2b)} $F=\{\,r\,(1,t,e^{t})\in\RR^{3}:r\in\RRp,t\in\RR\}$.
371:
372: \item{(2c)} $F=\{\,r\,(1,e^{t},e^{\theta
373: t})\in\RR^{3}:r\in\RRp,t\in\RR\}$ with $\theta>2$ arbitrary.
374:
375: \item{(3)} $F=\{\,c(t)+r\re1c'(t)\in\RR^{3}: r\in\RRp,t\in\RR\}$,
376: where $c(t):=(t,t^{2},t^{3})$ parameterizes the {\sl twisted cubic}
377: $\{(t,t^{2},t^{3}):t\in\RR\}$ in $\RR^{3}$ and $c'(t)=(1,2t,3t^{2})$.
378:
379: \noindent Notice that the limit case $\omega=0$ in (2a) gives the
380: future light cone $\5F$, while the limit case $\theta=2$ in (2c) gives
381: the linearly homogeneous surface
382: $\{x\in\RR^{3}:x_{1}x_{3}=x_{2}^{2},\,x_{1}>0,\,x_{2}>0\}$,
383: which is locally, but not globally linearly equivalent to $\5F$. In
384: fact, $\5F$ is linearly equivalent to the cone\nline
385: $\{x\in\RR^{3}:x_{1}x_{3}=x_{2}^{2},\,x_{1}+x_{3}>0\}$.
386:
387: \ifarx\vfill\eject\else\medskip\fi
388:
389: As our first main result we show, compare \ruf{UA} and \ruf{UB} for
390: details:
391:
392: \noindent{\parskip0pt\bf Theorem I: \sl For every surface $F$ in {\rm(1) -- (3)}
393: the corresponding tube manifold $M:=F+i\RR^{3}$ is a homogeneous
394: $2$-nondegenerate CR-submanifold of $\CC^{3}$ and any two of them are
395: pairwise locally CR-inequivalent. Furthermore, for every $F$ in
396: {\rm(2a) -- (3)} and every $a\in M=F+i\RR^{3}$ the following holds: \0
397: The Lie algebra $\hol(M,a)$ is solvable and has dimension $5$. \1 The
398: stability group $\Aut(M,a)$ is trivial. \1 Every homogeneous
399: real-analytic CR-manifold $M'$, that is locally CR-equivalent to $M$,
400: is already globally CR-equivalent to $M$.\par}
401:
402: \smallskip\noindent Notice that, a priori, there is no reason why the
403: $F$ in (1) -- (3), although known to be locally affinely inequivalent,
404: should have locally CR-inequivalent tubes (for nondegenerate affinely
405: homogeneous surfaces in $\RR^{3}$, for instance, there are
406: counterexamples).
407:
408: \medskip We actually prove an analog of Theorem I in every dimension
409: $n\ge3$, where the same trichotomy occurs as above: Consider the
410: following surfaces $F\subset\RR^{n}$
411:
412: \item{(1')} $F=\5F^{n}:=\{x\in\RR^{n}:x_{1}>0,\,x_{2}>0\steil{and}
413: x_{j}=x^{j-1}_{2}/x_{1}^{j-2} \steil{for}3\le j\le n\}$.
414:
415: \item{(2')}
416: $F=\{r\re1e^{t\phi}(a)\in\RR^{n}:r\in\RRp,t\in\RR\}$, where
417: $\phi$ is an endomorphism of $\RR^{n}$ having $a\in\RR^{n}$ as cyclic
418: vector (i.e., the iterates $\phi^{k}(a)$, $k\ge0$ span $\RR^{n}$) and
419: the $n$ eigenvalues of $\phi$ do not form an arithmetic progression in
420: $\CC$.
421:
422:
423: \item{(3')} $F=\{c(t)+r\re1c'(t):r\in\RRp,t\in\RR\}$, where
424: $c(t):=(t,t^{2},\dots,t^{n})\in\RR^{n}$ parameterizes the twisted
425: $n$-ic in $\RR^{n}$.\par
426:
427: \noindent In Sections \ruf{Sechs} and \ruf{Sieben} we show among other
428: statements: For every $F,F'$ from (1') -- (3') the tube manifolds
429: $M:=F+i\RR^{n}$, $M':=F'+i\RR^{n}$ are affinely homogeneous generic
430: $2$-nondegenerate submanifolds of $\CC^{n}$ with CR-dimension $2$.
431: Furthermore, $M$, $M'$ are locally CR-equivalent if and only if $F$,
432: $F'$ are globally affinely equivalent, and this holds if and only if
433: for given $a\in M$, $a'\in M'$ the Lie algebras $\hol(M,a)$,
434: $\hol(M',a')$ are isomorphic. In case $F=\5F^{n}$ the Lie algebra
435: $\hol(M,a)$ contains a copy of $\gl(2,\RR)$ and hence is not
436: solvable. In all other cases, i.e. $F$ comes from (2') or (3'), the
437: Lie algebra $\hol(M,a)$ is solvable of dimension $n+2$ and the
438: stability group $\Aut(M,a)$ has order at most $2$.
439:
440:
441: \medskip Let us briefly comment on the proof of Theorem I: Once
442: defining equations for an $F$ under consideration are explicitly known
443: (this is quite obvious for the types (1) -- (3), compare Section
444: \ruf{Acht}, but seems to be hard for the types (2'), (3')\hskip1pt),
445: one can compute by standard methods the order $k$ of
446: nondegeneracy. However, the amount of calculation necessary to
447: determine $k$ in such a way grows very fast with $k$ and with the
448: dimension or codimension of $M\subset \CC^n$. This is one of the
449: reasons, especially with an eye on possible generalizations, to choose
450: a different approach, which does not use explicit equations. For
451: instance, given an arbitrary submanifold $F\subset \RR^n$ which is
452: (locally) {\sl affinely homogeneous}, we present a simple criterion
453: (Proposition \ruf{US}) which allows to determine quickly the order $k$
454: of nondegeneracy for the corresponding tube manifold.
455:
456: \noindent The hard part of the proof is to show that the various tubes
457: $F+i\RR^3$ are actually locally inequivalent as CR-manifolds. Recall
458: that for real-analytic, not necessary homogeneous, {\sl hypersurfaces
459: with nondegenerate Levi form} there exist local invariants which
460: determine each $M$ up to local CR-equivalence due to the
461: fundamental work of Cartan, Tanaka and Chern-Moser. However, an
462: analogues approach is not available for $M$ of higher codimensions or
463: when the Levi form is degenerate. To distinguish the various tubes
464: $F+i\RR^3,$ we develop a method (valid also in greater generality)
465: which enables us to determine explicitly the Lie algebras $\hol(M,a)$
466: of infinitesimal CR-transformations of the various CR-germs $(M,a)$
467: (see Section \ruf{Zwei} for basic definitions and Sections \ruf{Sechs}
468: for further details).
469:
470: \medskip The CR-manifolds occurring in the previous theorem are quite
471: special as they all are tube manifolds. Moreover, all but one (namely
472: the twisted cubic case in (3)\hskip1pt) are actually conical. In the
473: Levi nondegenerate case, many (homogeneous) examples are known which
474: are not locally CR-equivalent to any tube manifold. For instance, the
475: unit sphere subbundle of $TS^3$ with its canonical CR-structure is
476: such an example. Therefore, our second main result came quite
477: unexpected to us:
478:
479: \medskip\noindent {\bf Theorem II.} {\sl Every 5-dimensional locally
480: homogeneous 2-nondegenerate CR-manifold $M$ is locally
481: CR-equivalent to $F+i\RR^3$ with $F$ being one of the surfaces in {\rm
482: (1) -- (3)}}.
483:
484: \medskip\noindent For the precise definition of local homogeneity we
485: refer to Section \ruf{Zwei}. A priori, locally homogeneous
486: CR-manifolds might exist which are not locally CR-equivalent to any
487: globally homogeneous one. As a by-product of the above 2 results we
488: can see that such a pathology does not happen in the case under
489: consideration. Theorem II gives a classification of all
490: (abstract) 2-nondegenerate locally homogeneous CR-manifolds in
491: dimension 5 up to local CR-equivalence. In fact, using Cartan's
492: classification \Lit{CART} of the 3-dimensional Levi nondegenerate
493: homogeneous CR-manifolds, the following result holds.
494:
495: \medskip\noindent {\bf Classification.} {\sl Every 5-dimensional
496: locally homogeneous CR-manifold $M$ with degenerate Levi form is
497: locally CR-equivalent to one of the following: \0 $M=F+i\RR^3\subset
498: \CC^3$ and $F$ is one of the surfaces {\rm (1) -- (3)} from above. \1
499: $M=\CC\times M'$, where $M'$ is one of the 3-dimensional Levi
500: nondegenerate CR-manifolds from Cartan's list in \Lit{CART}. \1
501: $M=\CC^2\times \RR$ or $M=\CC\times \RR^3$ with $\RR^3$ totally real.
502:
503: \noindent The manifolds in {\rm (i)} are all $2$-nondegenerate and in
504: {\rm (ii) -- (iii)} are holomorphically degenerate. Also, the
505: manifolds in {\rm (iii)} are just the Levi flat ones.}
506:
507: \smallskip With Theorem II the following question arises naturally for
508: higher codimension: {\sl Are there, up to local CR-equivalence,
509: further locally homogeneous $2$-nondegenerate CR-manifolds of
510: CR-dimension 2 besides those that are tubes over the surfaces $F$ in
511: {\rm (1') -- (3')} above?} Notice in this context that every locally
512: homogeneous $2$-nondegenerate CR-manifold of dimension $6$ necessarily
513: has CR-dimension $2$.
514:
515: \smallskip For 5-dimensional CR-hypersurfaces with {\sl nondegenerate}
516: Levi form, that is, when Chern-Moser invariants are available, there
517: already exists a partial classification: Locally homogeneous
518: CR-hypersurfaces in $\CC^3$ with stability groups of positive
519: dimension have been classified by Loboda in terms of local equations
520: in normal form, see \Lit{LOBO}, \Lit{LOBP}, \Lit{LOBQ}.
521:
522: \medskip The paper is organized as follows. After recalling some
523: necessary preliminaries in Section \ruf{Zwei} we discuss in Section
524: \ruf{Drei} tube manifolds $M=F\oplus i\RR^{n}\subset\CC^{n}$ over
525: real-analytic submanifolds $F\subset\RR^{n}$. It turns out that the
526: CR-structure of $M$ is closely related to the real-affine structure of
527: the base $F$. For instance, the Levi form of $M$ is essentially the
528: sesquilinear extension of the second fundamental form of the
529: submanifold $F\subset\RR^{n}$. Generalizing the notion of the second
530: fundamental form we define higher order invariants for $F$ (see
531: \ruf{HN}). In the uniform case these invariants precisely detect the
532: $k$-nondegeneracy of the corresponding CR-manifold $M=F\oplus i\RR^n$.
533: It is known that the (uniform) $k$-nondegeneracy of a real-analytic
534: CR-manifold $M$ together with minimality ensures that the Lie algebras
535: $\hol(M,a)$ are finite-dimensional, and is equivalent to this in the
536: special class of locally homogeneous CR-manifolds. For submanifolds
537: $F\subset\RR^{n}$ which in addition are homogeneous under a group of
538: affine transformations, a simple criterion for $k$-nondegeneracy of
539: the associated tube manifold $M$ is given in Proposition \ruf{US}.
540:
541:
542: In Section \ruf{Vier} these results are applied to the case where $F$
543: is conical in $\RR^n$, that is, locally invariant under dilations
544: $z\mapsto tz$ for $t$ near $1\in\RR$. In this case $M$ is always {\sl
545: Levi degenerate.} Assuming that $\hol(M,a)$ is finite dimensional
546: (which automatically is the case for minimal and finitely
547: nondegenerate CR-manifolds) we develop some basic techniques for the
548: explicit computation of $\hol(M,a)$. The main results in Section
549: \ruf{Vier} are the following: We prove that under the finiteness
550: assumption $\hol(M,a)$ consists only of polynomial vector fields and
551: carries a natural graded structure, see Proposition \ruf{DC}. We prove
552: (under the same assumptions, see Proposition \ruf{ZB}.ii) that local
553: CR-equivalences between two such tube manifolds are always rational
554: maps (even if these manifolds are not real-algebraic). Furthermore,
555: jet determination estimates are provided (\ruf{ZB}.iv). In the special
556: situation where $\hol(M,a)$ consists of affine germs only, these
557: results are further strengthened.
558:
559: In Section \ruf{Fuenf} we illustrate by examples how our methods can
560: be applied. In Example \ruf{JP} we present to every $c\ge 1$ and
561: $k\in \{2,3,4\}$ a homogeneous submanifold of $\CC^{n}$ which is
562: $k$-nondegenerate, of codimension $c$ and of CR-dimension $k$. We
563: close this section investigating the tubes $M^\alpha_{p,q}$ over the
564: cones $F^\alpha_{p,q}:=\{x\in \RR^{p{+}q}_+: \sum \epsilon_j
565: x_j^\alpha=0\}$ with $p$ positive and $q$ negative $\epsilon_j$'s and
566: $\alpha\in \RR^{*}$. Using our methods from the previous section we
567: explicitly determine the Lie algebras $\hol(M^\alpha_{p,q},a)$ for
568: arbitrary integers $\alpha\ge2$ and $p\ge q\ge 1$.
569:
570: In Section \ruf{Sechs}, we construct homogeneous CR-submanifolds
571: $M=M^{\phi,d}\subset\CC^{n}$ of tube type, depending on the choice of
572: an endomorphism $\phi\in \End(\RR^n)$, an integer $1<d<n$ and a cyclic
573: vector $a\in\RR^{n}$ for $\phi$ in the following way: The powers
574: $\phi^0,\phi^1,\ldots,\phi^{d-1}$ span an abelian Lie algebra $\7h$
575: and, in turn, give a cone $F:=\exp(\7h)a\subset \RR^n$. For $\phi$
576: which admits a cyclic vector $a\in \RR^n$ the corresponding tube
577: manifold $M^{\phi,d}=F+i\RR^{n}$ is 2-nondegenerate, minimal and of
578: CR-dimension $d$. The key result here is the explicit determination
579: of the CR-invariant $\hol(M^{\phi,d},a)$ for $\phi$ in 'general
580: position' (Propositions \ruf{UH} and \ruf{HU}). The precise meaning
581: for $\phi$ of being in 'general position' is stated in Lemma \ruf{UL}.
582: Further results are, again for $\phi$ in general position, that the
583: tube manifold $M$ is simply connected and has trivial stability group
584: at every point. As a consequence, the manifolds $M$ of this type have
585: the following remarkable property: {\sl Every homogeneous
586: (real-analytic) CR-manifold locally CR-equivalent to $M$ is globally
587: CR-equivalent to $M$}.
588:
589: In Section \ruf{Sieben} the results from Section \ruf{Sechs} are
590: further refined in the case of homogeneous CR-manifolds
591: $M^\phi:=M^{\phi,2}:=F^\phi+i\RR^n\subset \CC^n$ of CR-dimension 2 but
592: without restrictions on the codimension. In fact, every minimal and
593: locally homogeneous tube CR-manifold $M:=S+i\RR^n$ with a conical
594: 2-dimensional $S\subset \RR^n$ is locally CR-equivalent to
595: $M^{\phi,2}$ for some cyclic $\phi$. In this section also the case
596: when $\phi$ is not in 'general position', i.e., the characteristic
597: roots do form an arithmetic progression (see Lemma \ruf{UL})
598: is treated. The main results here are: Whether a cyclic $\phi$ is in
599: general position or not, the Lie algebras $\hol(M^\phi,a)$
600: (Proposition \ruf{ZR}) and the global automorphism groups
601: $\Aut(M^\phi)$ (Proposition \ruf{GO}) are determined. Furthermore,
602: the problem of local and global CR-equivalence among the $M^\phi$'s is
603: solved (Propositions \ruf{KO} and \ruf{BP}) and a moduli space is
604: constructed (Subsection \ruf{MO}).
605:
606: Part I of the paper is concluded with Section \ruf{Acht} where the
607: examples (1) -- (3) from Theorem I are presented in more detail. The
608: results of the preceeding section are applied to this case of
609: 5-dimensional tube manifolds $M$. In particular Propositions \ruf{UA}
610: and \ruf{IZ} contain some additional information to that stated in
611: Theorem I and also complete the proof of Theorem I.
612:
613: Part II of the paper is mainly devoted to prove Theorem II. In the
614: preliminary section \ruf{Neun} we explain how the geometric properties
615: such as $k$-nondegeneracy, minimality or the CR-dimension of an
616: arbitrary locally homogeneous CR-germ $(M,o)$ can be encoded in a pure
617: Lie algebraic object, the associated CR-algebra $(\7g,\7q)$. This is
618: the key for our classification and is based on results taken from
619: \Lit{FELS}. Specified to 5-dimensional CR-germs, we formulate the
620: precise algebraic conditions on a CR-algebra $(\7g,\7q)$ ensuring that
621: the associated CR-germ $(M,o)$ is 2-nondegenerate.
622:
623: Once the classification of 2-nondegenerate 5-dimensional locally
624: homogeneous CR-germs $(M,o)$ is reduced to a classification problem of
625: certain CR-algebras, we begin in Section \ruf{Zehn} to carry out the
626: details of the proof. It is subdivided into several sections, lemmata
627: and claims and will only be completed in Section \ruf{Sechzehn}. Our
628: proof relies on a quite subtle analysis of Lie algebraic properties of
629: CR-algebras and uses basic structure theory of Lie algebras and Lie
630: groups. The methods are quite general and can be adapted to handle
631: also higher dimensional cases.
632:
633: A more detailed outline of the proof can be found in the first part of
634: Section \ruf{Zehn}.
635:
636:
637:
638: \ifarx\vfill\eject\else\Partskip\fi
639:
640: \centerline{\twelverm PART 1: Levi degenerate CR-manifolds}
641: \bigskip
642:
643: \KAP{Zwei}{Preliminaries}
644:
645: In the following let $E$ always be a complex vector space of finite
646: dimension and $M\subset E$ an immersed connected real-analytic
647: submanifold. In most cases $M$ will be locally closed in $E$. Due to
648: the canonical identifications $T_aE=E,$ for every $a\in M$ we consider
649: the tangent space $T_{a}M$ as an $\RR$-linear subspace of $E$. Then
650: $H_{a}M:=T_{a}M\cap iT_{a}M$ is the largest complex linear subspace of
651: $E$ contained in $T_{a}M$. The manifold $M$ is called a {\sl
652: CR-submanifold} if $\dim_{\CC}\!H_{a}M$ does not depend on $a\in
653: M$. This dimension is called the {\sl CR-dimension} of $M$ and
654: $H_{a}M$ is called the {\sl holomorphic tangent space} at $a$, compare
655: \Lit{BERO} as general reference for CR-manifolds. Given a further
656: real-analytic CR-submanifold $M'$ of a complex vector space $E'$, a
657: smooth mapping $g:M\to M'$ is called {\sl CR} if for every $a\in M$
658: the differential $dg_{a}:T_{a}M\to T_{ga}M'$ maps the corresponding
659: holomorphic tangent spaces in a complex-linear way to each
660: other. Keeping in mind the identification $T_aE=E$, a vector field on
661: $M$ is a mapping $f:M\to E$ with $f(a)\in T_{a}M$ for all $a\in
662: M$. For better distinction we also write $\xi=f(z)\dd z$ instead of
663: $f$ and $\xi_{a}$ instead of $f(a)$, compare the convention (2.1) in
664: \Lit{FEKA}.
665:
666: An {\sl infinitesimal CR-transformation} on $M$ is by definition a
667: real-analytic vector field $f(z)\dd z$ on $M$ such that the
668: corresponding local flow consists of CR-transformations. Let us denote
669: by $\hol(M)$ the space of all such vector fields, which is a real Lie
670: algebra with respect to the usual bracket. For every $f(z)\dd
671: z\in\hol(M)$ and every $a\in M$ there exist an open neighbourhood
672: $U\subset M$ of $a$ with respect to the manifold topology on $M$, an
673: open neighbourhood $W$ of $a$ with respect to $E$ and
674: a holomorphic mapping $h:W\to E$ with $f(z)=h(z)$
675: for all $z\in U\cap W$, compare \Lit{ANHI} or 12.4.22 in
676: \Lit{BERO}.
677:
678: Further, for every $a\in M$ we denote by $\hol(M,a)$ the Lie algebra
679: of all germs of infinitesimal CR-transformations defined on arbitrary
680: open neighbourhoods of $a$ with respect to the manifold topology of
681: $M$. For simplicity and without essential loss of generality we always
682: assume that the CR-submanifold $M$ is {\sl generic} in $E$, that is,
683: $E=T_{a}M\oplus iT_{a}M$ for all $a\in M$. This assumption allows us to
684: consider $\hol(M,a)$ in a canonical way as a real Lie subalgebra of
685: the complex Lie algebra $\hol(E,a)$, what we always do in the
686: following. The CR-manifold $M$ is called {\sl holomorphically
687: nondegenerate} at $a$ if $\hol(M,a)$ is totally real in $\hol(E,a)$,
688: that is, if $\hol(M,a)\cap i\,\hol(M,a)=0$ in $\hol(E,a)$. This
689: condition together with the minimality of $M$ at $a$ implies $\dim
690: \hol(M,a)<\infty$, see 12.5.3 in \Lit{BERO}. Here, the CR-submanifold
691: $M\subset E$ is called {\sl minimal} at $a\in M$ if $T_{a}R=T_{a}M$
692: for every submanifold $R\subset M$ with $a\in R$ and $H_{z}M\subset
693: T_{z}R$ for all $z\in R$. By Proposition 15.5.1 in \Lit{BERO}, $\dim
694: \hol(M,a)<\infty$ implies that $M$ is holomorphically nondegenerate at
695: $a$. The CR-manifold $M$ is called {\sl locally homogeneous} at the
696: point $a\in M$ if there exists a Lie subalgebra $\7g\subset\hol(M,a)$
697: of finite dimension such that the canonical evaluation map $\7g\to
698: T_{a}M$ is surjective. In case $M$ is locally homogeneous at $a$ the
699: condition $\dim \hol(M,a)<\infty$ is equivalent to $M$ being
700: holomorphically nondegenerate and minimal at $a$.
701:
702: By $\aut(M,a):=\{\xi\in\hol(M,a):\xi_{a}=0\}$ we denote the {\sl
703: isotropy subalgebra} at $a\in M$. Clearly, $\aut(M,a)$ has finite
704: codimension in $\hol(M,a)$. Furthermore, we denote by $\Aut(M,a)$ the
705: group of all {\sl germs} of real-analytic CR-isomorphisms $ h:W\to
706: \tilde W$ with $ h(a)=a$, where $W,\tilde W$ are arbitrary open
707: neighbourhoods of $a$ in $M$. It is known that every germ in
708: $\Aut(M,a)$ can be represented by a holomorphic map $U\to E$, where
709: $U$ is an open neighbourhood of $a$ in $E$, compare e.g. 1.7.13 in
710: \Lit{BERO}. Furthermore, $\Aut(M)$ denotes the group of all global
711: real-analytic CR-automorphisms $h:M\to M$ and $\Aut(M)_a$ its isotropy
712: subgroup at $a$. There is a canonical group monomorphism
713: $\Aut(M)_a\hookrightarrow\Aut(M,a)$ as well as an exponential map
714: $\exp:\aut(M,a)\to\Aut(M,a)$ for every $a\in M$.
715:
716: By $\aff(M)\subset\hol(M)$ we denote the Lie subalgebra of all
717: (complex) affine infinitesimal CR-transformations on $M$. For every
718: $a\in M$ furthermore $\aff(M,a)\subset\hol(M,a)$ is the Lie subalgebra
719: of all affine germs. The canonical embedding
720: $\aff(M)\hookrightarrow\aff(M,a)$ is an isomorphism for every $a\in
721: M$.
722:
723: Suppose that $g\colon U\to U'$ is a biholomorphic mapping between open
724: subsets $U,U'\subset E$. Then
725: $$
726: g_{*}\big(f(z)\dd z\big)=g'(g^{-1}z)\big(f(g^{-1}z)\big)\,\dd
727: z\Leqno{GP}
728: $$
729: defines a complex Lie algebra isomorphism
730: $g_{*}\colon\hol(U)\to\hol(U')$, where $g'(u)\in\End(E)$ for every
731: $u\in U$ is the derivative of $g$ at $u$. For real-analytic
732: CR-submanifolds $M,M'\subset E$ every CR-isomorphism
733: $g\colon(M,a)\to(M',a')$ of manifold germs induces a Lie algebra
734: isomorphism $g_{*}\colon\7g\to\7g'$, where $\7g\!:=\hol(M,a)$ and
735: $\7g'\!:=\hol(M',a')$. From \Ruf{GP} it is clear that $g_{*}$ extends
736: to a complex Lie algebra isomorphism $\7l\to\7l'$, where the sums
737: $\7l:=\7g\oplus i\7g\subset\hol(E,a)$ and
738: $\7l':=\7g'\oplus i\7g'\subset\hol(E,a')$ are not necessarily direct. In
739: particular, $g\mapsto g_{*}$ defines a group homomorphism
740: $\Aut(M,a)\to\Aut(\7g)$.
741:
742: \medskip A basic invariant of a CR-manifold is the (vector valued)
743: {\sl Levi form}. Its definitions found in the literature may differ by
744: a constant factor. Here we choose the following one: It is
745: well-known that for every point $a$ in the CR-manifold $M$ there is a
746: well defined alternating $\RR$-bilinear map
747: $$
748: \omega_{a}:H_{a}M\times H_{a}M\;\to\; E/H_{a}M
749: $$
750: satisfying $\omega_{a}(\xi_{a},\zeta_{a})\equiv[\xi,\zeta]_{a}
751: \steil{mod}H_{a}M,$ where $\xi,\zeta$ are arbitrary smooth vector
752: fields on $M$ with $\xi_{z},\zeta_{z}\in H_{z}M$ for all $z\in M$. We
753: define the Levi form
754: $$
755: \5L_{a}:H_{a}M\times H_{a}M\;\to\; E/H_{a}M\Leqno{LE}
756: $$
757: to be twice the sesquilinear part of $\omega_{a}$. By {\sl
758: sesquilinear} we always mean `conjugate linear in the first and
759: complex linear in the second variable', that is,
760: $$
761: \5L_{a}(v,w)=\omega_{a}(v,w)+i\omega_{a}(iv,w)\qquad \hbox{for all }
762: v,w\in H_{a}M\,.
763: $$
764: In particular, the vectors $\5L_{a}(v,v),$ $v\in H_aM,$ are contained
765: in $iT_{a}M/H_{a}M$, which can be identified in a canonical way with
766: the normal space $E/T_{a}M$ to $M\subset E$ at $a$. The following
767: remark follows immediately from the way the Levi form is defined:
768:
769: \Remark{RM} Suppose $Z$ is a complex manifold, $\phi:Z\to M$ is a smooth
770: CR-mapping and $a=\phi(c)$ for some $c\in Z$. Then every vector $v\in
771: d\phi_{c}(T_{c}Z)\subset H_{a}M$ satisfies $\5L_{a}(v,v)=0$. In
772: general, $v$ is not contained in the {\sl Levi kernel}
773: $$
774: K_{a}M:=\{v\in H_{a}M:\5L_{a}(v,w)=0\steil{for all}w\in H_aM\}\,.
775: $$
776: \Formend The CR-manifold $M$ is called {\sl Levi nondegenerate} at $a$
777: if $K_{a}M=0$. Generalizing that, the notion of {\sl
778: $k$-nondegeneracy} for $M$ at $a$ has been introduced for every
779: integer $k\ge1$ (see \Lit{BAHR}, \Lit{BERO}). As shown in 11.5.1 of
780: \Lit{BERO} a real-analytic and connected CR-manifold $M$ is
781: holomorphically nondegenerate at $a$ (equivalently: at every $z\in M$)
782: if and only if there exists a $k\ge1$ such that $M$ is
783: $k$-nondegenerate at some point $b\in M$. For $k=1$ this notion is
784: equivalent to $M$ being Levi nondegenerate at $b\in M$.
785:
786: \bigskip In the second part of our paper we also need a more general
787: notion of a (real-analytic) CR-manifold. This is a connected
788: real-analytic manifold $M$ together with a subbundle $HM\subset TM$
789: (called the `holomorphic subbundle') and a bundle endomorphism $J$ of
790: $HM$ with $J^2= - \id$ such that $(HM,J)$ is involutive, compare 7.4
791: in \Lit{BOGG}. By a theorem in \Lit{ANFR} there exists an embedding
792: $M\into Z$ into a complex manifold $Z$, such that $H_zM$ corresponds
793: to $T_{z}M\cap\,iT_{z}M$, where $T_zZ\to T_zZ$ is simply the
794: multiplication with the imaginary unit $i$ (here the real-analyticity
795: is necessary). The bundle homomorphism $J:HM\to HM$ is then the
796: restriction of that multiplication with $i$ to $H_{z}M$ for every
797: $z\in M$. For local considerations one can always assume that $Z$ is
798: (an open subset of) $\CC^n$.
799:
800: \KAP{Drei}{Tube manifolds}
801:
802: Let $V$ be a real vector space of finite dimension and $E:=V\,\oplus\,
803: iV$ its complexification. Let furthermore $F\subset V$ be a connected
804: real-analytic submanifold and $M:=F+iV\subset E$ the
805: corresponding {\sl tube manifold.} $M$ is a generic CR-submanifold of
806: $E$, invariant under all translations $z\mapsto z+iv$, $v\in V$. In
807: case $V'$ is another real vector space of finite dimension, $E'$ its
808: complexification, $F'\subset V'$ a real-analytic submanifold and
809: $\phi:V\to V'$ an affine mapping with $\phi(F)\subset F'$, then
810: clearly $\phi$ extends in a unique way to a complex-affine mapping
811: $E\to E'$ with $\phi(M)\subset M'$. However, it should be noted that
812: higher order real-analytic maps $\psi:F\to F'$ also extend locally to
813: holomorphic maps $\psi:U\to E'$, $U$ open in $E$. But in contrast to
814: the affine case we have in general $\psi(M\cap U)\not\subset M'$. We
815: may therefore ask how the CR-structure of $M$ is related to the real
816: affine structure of the submanifold $F\subset V$.
817:
818: For every $a\in F$ let $T_{a}F\subset V$ be the {\sl tangent space}
819: and $N_{a}F:=V/T_{a}F$ the {\sl normal space} to $F$ at $a$. Then
820: $T_{a}M=T_{a}F\oplus iV$ for the corresponding tube manifold $M$, and
821: $N_{a}F$ can be canonically identified with the normal space
822: $N_{a}M=E/T_{a}M$ of $M$ in $E$. Define the map $\ell_{a}:T_{a}F\times
823: T_{a}F\to N_{a}F$ in the following way: For $v,w\in T_{a}F$ choose a
824: smooth map $f:V\to V$ with $f(a)=v$ and $f(x)\in T_{x}F$ for all $x\in
825: F$ (actually it suffices to choose such an $f$ only in a small
826: neighborhood of $a\,$). Then put
827: $$
828: \ell_{a}(v,w):=f'(a)(w)\;\mod\;T_{a}F\,,\Leqno{EL}
829: $$
830: where the linear operator $f'(a)\in\End(V)$ is the derivative of $f$
831: at $a$. One shows that $\ell_a$ does not depend on the choice of $f$
832: and is a symmetric bilinear map. We mention that if $V$ is provided
833: with a flat Riemannian metric and $N_aF$ is identified with
834: $T_aF^\perp$ then $\ell$ is nothing but the second fundamental form of
835: $F$ (see the subsection II.3.3 in \Lit{SAKA}). The form $\ell_{a}$ can
836: also be read off from local equations for $F$, more precisely, suppose
837: that $U\subset V$ is an open subset, $W$ is a real vector space and
838: $h:U\to W$ is a real-analytic submersion with $F=h^{-1}(0)$. Then the
839: derivative $h'(a):V\to W$ induces a linear isomorphism $N_{a}F\cong W$
840: and modulo this identification $\ell_{a}$ is nothing but the second
841: derivative $h''(a):V\times V\to W$ at $a$, restricted to $T_{a}F\times
842: T_{a}F$. By
843: $$
844: K_{a}F:=\{w\in T_{a}F:\ell_{a}(v,w)=0\steil{for all}v\in T_{a}F\}
845: $$
846: we denote the {\sl kernel} of $\ell_{a}$. The manifold $F$ is called
847: (affinely) {\sl nondegenerate} at $a$ if $K_{a}F=0$ holds. The
848: following statement follows directly from the definition of $\ell_a$:
849:
850: \Lemma{CR} Suppose that $\phi\in\End(V)$ satisfies $\phi(x)\in T_{x}F$
851: for all $x\in F$. Then $\phi(a)\in K_{a}F$ if and only if
852: $\phi\big(T_{a}F\big)\subset T_{a}F$.\Formend
853:
854: \noindent Lemma \ruf{CR} applies in particular for $\phi=\id$ in case
855: $F$ is a {\sl cone,} that is, $rF=F$ for all real $r>0$. More
856: generally, we call the submanifold $F\subset V$ {\sl conical} if $x\in
857: T_{x}F$ for all $x\in F$. Then $\RR a\subset K_{a}F$ holds for all
858: $a\in F$.
859:
860: \medskip
861:
862: In the remaining part of this section we explain how the CR-structure
863: of the tube manifold $M$ is related to the real objects $\ell_a,\,
864: TF,\, KF$ and $ K^r\!F$, to be defined below, which depend only on the
865: affine geometry of $F$. In general it needs some effort to check
866: whether a given CR-manifold $M$ is $k$-nondegenerate at a point $a\in
867: M$ (in sense of \Lit{BHUR}). For affinely homogeneous tube manifolds,
868: however, there are simple criteria, see Propositions \ruf{UR} and
869: \ruf{US}. We start with some preparations. For every $a\in F\subset M$
870: $$
871: H_{a}M=T_{a}F\oplus iT_{a}F\,\subset\, E
872: \Leqno{HM}
873: $$
874: is the holomorphic
875: tangent space at $a$, and $E/H_{a}M$ can be canonically identified
876: with $N_{a}F\oplus iN_{a}F$. It is easily seen that the Levi form
877: $\5L_{a}$ of $M$ at $a$, compare \Ruf{LE}, is nothing but the
878: sesquilinear extension of the form $\ell_{a}$ from $T_{a}F\times
879: T_{a}F$ to $H_{a}M\times H_{a}M$. In particular,
880: $$
881: K_{a}M=K_{a}F\oplus iK_{a}F
882: $$
883: is the {\sl Levi kernel} of $M$ at $a$. In case the dimension of
884: $K_{a}F$ does not depend on $a\in F$ these spaces form a subbundle
885: $KF\subset TF$. In that case to every $v\in K_{a}F$ there exists a
886: smooth function $f:V\to V$ with $f(a)=v$ and $f(x)\in K_{x}F$ for all
887: $x\in F$, i.e., the tangent vector $v$ extends to a smooth section in
888: $KF$. In any case, let us define inductively linear subspaces
889: $K_{a}^{r}F$ of $T_{a}F$ as follows:
890:
891: \Definition{HN} For every real-analytic submanifold $F\subset V$,
892: every $a\in F$ and every $r\in\NN$ put \0 $K_{a}^{0}F:=T_{a}F$ and
893: define \1 $K_{a}^{r+1}\!F$ to be the space of all vectors $v\in
894: K_{a}^{r}F$ such that there is a smooth mapping $f:V\to V$ with
895: $f'(a)(T_{a}F)\subset K_{a}^{r}F$, $f(a)=v$ and $f(x)\in K_{x}^{r}F$
896: for all $x\in F$.
897:
898: \medskip\noindent It is clear that $K_{a}^{1}F=K_{a}F$ holds. Let us
899: call $F$ of {\sl uniform degeneracy} or {\sl uniformly degenerate} if
900: for every $r\in\NN$ the dimension of $K_{a}^{r}F$ does not depend on
901: $a\in F$. In that case it can be shown that for every $v\in
902: K_{a}^{r}F$ the outcome of the condition $f'(a)(T_{a}F)\subset
903: K_{a}^{r}F$ in (ii) does not depend on the choice of the smooth
904: mapping $f:V\to V$ satisfying $f(a)=v$ and $f(x)\in K_{x}^{r}F$ for
905: all $x\in F$. For instance, $F$ is of uniform degeneracy if $F$ is
906: locally affinely homogeneous, that is, if there exists a Lie algebra
907: $\7a$ of affine vector fields on $V$ such that every $\xi\in\7a$ is
908: tangent to $F$ and such that the canonical evaluation map $\7a\to
909: T_{a}F$ is surjective for every $a\in F$. Clearly, if $F$ is locally
910: affinely homogeneous in the above sense then the corresponding tube
911: manifold $M=F+iV$ is locally homogeneous as CR-manifold.
912:
913: \medskip Recall our convention that every smooth map $f:V\to V$ is
914: considered as the smooth vector field $\xi=f(x)\dd x$ on $V$. Our
915: computations below are considerably simplified by the obvious fact
916: that every smooth vector field $\xi$ on V has a unique smooth
917: extension to $E$ that is invariant under all translations $z\mapsto
918: z+iv$, $v\in V$. In case $\xi$ is tangent to $F\subset V$ the
919: extension satisfies $\xi_{z}\in H_{z}M$ for all $z\in M$.
920:
921: In case the submanifold $F\subset V$ is uniformly degenerate in a
922: neighbourhood of $a\in F$ we call $F\,$ {\sl affinely
923: $k$-nondegenerate} at $a$ if $K_{a}^{k}F=0$ and $k$ is minimal with
924: respect to this property. It can be seen that `affine
925: $k$-nondegeneracy' is invariant under affine coordinate changes. As a
926: consequence of \Lit{KAZT}, compare the last 5 lines in the Appendix
927: therein, we state:
928:
929: \Proposition{UR} Suppose that $F$ is uniformly degenerate in a
930: neighbourhood of $a\in F$. Then the corresponding tube manifold
931: $M=F+iV$ is $k$-nondegenerate as CR-manifold at $a\in M$ if and
932: only if $F$ is affinely $k$-nondegenerate at $a$.\Formend
933:
934: \Corollary{AF} Suppose $\dim F \ge2$ and $K_{x}F=\RR x$ for all $x\in
935: F$. Then $F$ is affinely $2$-nondegenerate at every point.
936:
937: \Proof The map $f=\id$ has the property $f(x)\in K_xF$ for every $x\in
938: F$. Hence, the relation $f'(x)(T_{x}F)=T_{x}F\not\subset K_{x}F$
939: implies $x\notin K_{x}^{2}F$ and thus $K_{x}^{2}F=0$ as well as
940: $x\ne0$. In particular, $F$ is uniformly degenerate.\qed
941:
942: \medskip For locally affinely homogeneous submanifolds $F\subset V$
943: the spaces $K_{a}^{r}F$ can easily be characterized. For each affine
944: vector field $\xi=h(x)\dd x$ on $V$ denote by
945: $\xi^{\lin}:=h-h(0)\in\End(V)$ the {\sl linear part} of $\xi$.
946:
947: \Proposition{US} Suppose that $\7A$ is a linear space of affine vector
948: fields on $V$ such that every $\xi\in\7A$ is tangent to $F$ and the
949: canonical evaluation mapping $\7A\to T_{a}F$ is a linear
950: isomorphism. Then, given any $r\in\NN$, the vector $v\in K_{a}^{r}F$
951: is in $K_{a}^{r+1}\!F$ if and only if $\xi^{\lin}(v)\in K_{a}^{r}F$
952: for every $\xi\in\7A$.
953:
954: \Proof By the implicit function theorem, there exist open
955: neighbourhoods $Y$ of $0\in\7A$ and $X$ of $a\in M$ such that
956: $g(y):=\exp(y)a$ defines a diffeomorphism $g:Y\to X$. Define the
957: smooth map $f:X\to V$ by $f(x)=\mu_{y}(v)$, where $\mu_{y}$ for
958: $y:=g^{-1}(x)$ is the linear part of the affine transformation
959: $\exp(y)$. Then $f(a)=v$ and $f(x)\in K_{x}^{r}F$ for every $x\in
960: X$. A simple computation shows that\medskip
961: \centerline{$f'(a)\big(g'(0)\xi\big)=\xi^{\lin}(v)\steil{for every}
962: \xi\in\7A$.} \nline In view of Definition \ruf{HN}.ii this identity
963: implies the claim. \qed
964:
965: \medskip It is easily seen that a necessary condition for $M$ being
966: minimal as CR-manifold is that $F$ is not contained in an affine
967: hyperplane of $V$. For later use in Proposition \ruf{PS} we state the
968: following sufficient condition.
969:
970: \Proposition{PY} Suppose that $\7A$ has the same properties as in
971: Proposition \ruf{US} and denote by $\Lambda\subset\End(V)$ the {\rm
972: associative} real subalgebra generated by
973: $\{\xi^{\lin}:\xi\in\7A\}$. Then the tube manifold $M=F+iV$ is
974: minimal at $a$ if $V$ is the linear span of all vectors $\lambda(v)$
975: with $v\in T_{a}F$ and $\lambda\in\Lambda$.
976:
977: \Proof Without loss of generality we assume that the canonical
978: evaluation mapping $\7A\to T_{x}F$ is a linear isomorphism for every
979: $x\in F$. We also assume that $V$ is the linear span of all
980: $\lambda(v)$ as above. Define inductively for every integer $r\ge1$
981: the subbundle $H^{r}M\subset TM$ in the following way: $H^{1}M:=HM$
982: and every $H^{r+1}_{z}M$, $z\in M$, is the linear span of $H^{r}_{z}M$
983: together with all vectors $[\xi,\eta]_{z}$, where $\xi,\eta$ are
984: arbitrary smooth sections in $H^{r}M$ over $M$. For the proof it is
985: enough to show
986: $T_{a}M=H^{\infty}_{a}M:=\bigcup_{r\ge1}\!T_{a}M$.\nline From
987: $T_{a}F\oplus iT_{a}F\,\subset\, H_{a}^{r}M\,\subset\, T_{a}F\oplus
988: iV$ we see that for every $1\le r\le\infty$ there is a unique linear
989: subspace $H^{r}_{a}F\subset V$ with $H^{r}_{a}M=T_{a}F\oplus
990: iH^{r}_{a}F$ and $H^{1}_{a}F=T_{a}F$. Therefore it is enough to show
991: $H^{\infty}_{a}F=V$. We claim that $\xi^{\lin}(H^{r}_{a}F)\subset
992: H^{r+1}_{a}F$ holds for all $\xi\in\7A$. To see this fix an arbitrary
993: $w\in H^{r}_{a}F$ and an arbitrary vector field $\xi\in\7A$. Choose a
994: smooth section $\eta$ over $F$ in the bundle $iH^{r}_{a}F$ with
995: $\eta_{a}=iw$ and extend $\eta$ as well as $\xi$ in the unique way to
996: smooth vector fields on $M$ that are invariant under all translations
997: $z\mapsto z+iv$ with $v\in V$. Then $\xi,\eta$ are sections in
998: $H^{r}M$, and $[\xi,\eta]_{a}=i\xi^{\lin}(w)\in H^{r+1}_{a}M$ implies
999: $\xi^{\lin}(w)\in H^{r+1}_{a}F$ as required. Now define inductively
1000: the linear subspaces $W^{r}\subset V$ by $W^{1}:=H^{1}_{a}F=T_{a}F$
1001: and $W^{r+1}$ as the linear span of $W^{r}$ together with all
1002: $\xi^{\lin}(W^{r})$, $\xi\in\7A$. Then $V=\bigcup_{r\ge1}W^{r}$ by
1003: assumption and $W^{r}\subset H^{r}_{a}F$ by induction gives $V\subset
1004: H^{\infty}_{a}M\subset V$ as desired.\qed
1005:
1006: \medskip
1007:
1008: \Lemma{ZI} Suppose that $F\subset V$ is a submanifold such that for
1009: every $c\in V$ with $c\ne0$ there exists a linear transformation
1010: $g\in\GL(V)$ with $g(F)=F$ and $g(c)\ne c$ (this
1011: condition is automatically satisfied if $F$ is a cone). Then for
1012: $M=F+iV$ the CR-automorphism group $\Aut(M)$ has trivial center.
1013:
1014:
1015: \Proof Let an element in the center of $\Aut(M)$ be given and let
1016: $h:U\to E$ be its holomorphic extension to an appropriate connected
1017: open neighbourhood $U$ of $M$. Since $h$ commutes with every
1018: translation $z\mapsto z+iv$, $v\in V$, it is a translation itself:
1019: Indeed, for $a\in F$ fixed and $c:=h(a)-a$ the translation
1020: $\tau(z):=z+c$ coincides with $h$ on $a+iV$ and hence on $U$ by the
1021: identity principle. For every $g\in\GL(V)\cap\Aut(M)$ the identity
1022: $gh=hg$ implies $g(c)=c$. This forces $c=0$ by our
1023: assumption, that is, $h(z)\equiv z$.\qed
1024:
1025: \Proposition{UQ} Suppose that the homogeneous CR-manifold $M$ is
1026: simply connected and that $\Aut(M)$ has trivial center. In case the
1027: stability group $\Aut(M,a)$ is trivial for some (and hence every)
1028: $a\in M$, the following properties hold: \0 Let $M'$ be an arbitrary
1029: homogeneous CR-manifold and $D\subset M$, $D'\subset M'$ non-empty
1030: domains. Then every real-analytic CR-isomorphism $h:D\to D'$ extends
1031: to a real-analytic CR-\-isomorphism $M\to M'$. \1 Let $M'$ be an
1032: arbitrary locally homogeneous CR-manifold and $D'\subset M'$ a domain
1033: that is CR-equivalent to $M$. Then $D'=M'$.
1034:
1035: \Proof ad (i): Fix a point $a\in D$. To every $g\in G:=\Aut(M)$
1036: with $g(a)\in D$ there exists a transformation $g'\in G':=\Aut(M')$
1037: with $hg(a)=g'h(a)$. Because of $\Aut(M',a')=\{\id\}$ the
1038: transformation $g'$ is uniquely determined by $g$ and satisfies
1039: $hg=g'h$ in a neighbourhood of $a$. Since the Lie group $G$ is simply
1040: connected $g\mapsto g'$ extends to a group homomorphism $G\to G'$ and
1041: $h$ extends to a CR-covering map $h:M\to M'$. The deck transformation
1042: group $\Gamma:=\{g\in G:gh=h\}$ is in the center of $G$ and hence is
1043: trivial by assumption. Therefore, $h:M\to M'$ is a
1044: CR-isomorphism. \nline ad (ii): The proof is essentially the same
1045: as for Proposition 6.3 in \Lit{FEKA}.\qed
1046:
1047: The condition `locally homogeneous' in Proposition \ruf{UQ}.ii cannot
1048: be omitted. A counterexample is given for every integer $k\ge3$
1049: by the tube $M'\subset\CC^{3}$ over the cone
1050: $$
1051: F':=\{x\in\RR^{3}:x_{2}^{k}=x_{1}^{k-1}x_{3},\;x_{1}^{2}+
1052: x_{2}^{2}>0\}\,.
1053: $$ Then with $\;\RRp:=e^{\RR}$ the tube $M$ over
1054: $F:=F'\cap(\RRp)^{3}$ is the Example \ruf{EX} below for $\theta=k$,
1055: and $M,M'$ satisfy for $D'=M$ the assumption of Proposition
1056: \ruf{UQ}.ii.
1057:
1058: \ifarx\vfill\eject\fi
1059:
1060: \KAP{Vier}{Tube manifolds over cones}
1061:
1062: In this section we always assume that the submanifold $F\subset V$ is
1063: conical (that is, $x\in T_xF$ for every $x\in F$) and that
1064: $a\in F$ is a given point. Then, for $M:=F+iV$, the Lie algebra
1065: $\7g:=\hol(M,a)$ contains the Euler vector field $\delta:=z\dd z$.
1066: Denote by $\7P$ the complex Lie algebra of all polynomial holomorphic
1067: vector fields $f(z)\dd z$ on $E$, that is, $f:E\to E$ is a polynomial
1068: map. Then $\7P$ has the $\ZZ$-grading
1069: $$
1070: \7P=\bigoplus_{k\in\ZZ}\7P_{k}\,,\qquad[\7P_{k},\7P_{l}]\subset
1071: \7P_{k+l}\,,\Leqno{HJ}
1072: $$
1073: where $\7P_{k}$ is the $k$-eigenspace of $\ad(\delta)$ in $\7P$. Then
1074: $\7P_{k}$ is the subspace of all $({k+1})$-homogeneous vector fields
1075: in $\7P$ if $k\ge-1$ and is $0$ otherwise. Define
1076: $\7g_k:=\7g\cap\7P_k.$ Clearly, $\bigoplus_{k\in \ZZ}\7g_k$ is a
1077: graded, in general proper subalgebra of $\7g$, and
1078: $\aff(M,a)=\7g_{-1}\oplus\7g_{0}$.
1079:
1080: \Proposition{DC} Retaining the above notation, suppose that
1081: $\7g=\hol(M,a)$ has finite dimension. Then \0 $\7g\subset\7P$, that
1082: is, every $f(z)\dd z\in\7g$ is a polynomial vector field on $E=V\oplus
1083: iV$. Furthermore, $f(iV)\subset iV$.\vskip3pt \1 $\displaystyle
1084: \7g=\bigoplus\limits_{k\ge-1}\7g_{k}\,,\quad[\7g_{k},\7g_{l}]\subset
1085: \7g_{k+l}\Steil{and}\7g_{-1}=\{iv\dd z: v\in V\}\;$.\vskip2pt \1 For
1086: every $z\in M$ the canonical map $\hol(M)\to\hol(M,z)$ is a Lie
1087: algebra isomorphism. \1 $\7g_{k}=0$ for some $k\in\NN$ implies
1088: $\7g_{j}=0$ for every $j\ge k$.
1089:
1090: \Proof Consider $\7l:=\7g\oplus i\7g\subset\hol(E,a)$, which
1091: contains the vector field $\eta:=(z-a)\dd z$. We first show
1092: $\7l\subset \7P$: Fix an arbitrary $\xi:=f(z) \dd z\in\7l$. Then in a
1093: certain neighbourhood of $a\in E$ there exists a unique expansion
1094: $\xi=\sum_{k\in\NN}\xi_{k}$, where $\xi_{k}=p_{k}(z-a)\dd z$ for a
1095: $k$-homogeneous polynomial map $p_{k}:E\to E$. It is easily verified
1096: that the vector field $\ad(\eta)\xi\in\7l$ has the expansion
1097: $\ad(\eta)\xi=\sum_{k\in\NN}(k{-}1)\xi_{k}$. Now assume that for
1098: $d:=\dim \7l $ there exist indices $k_{0}<k_{1}<\dots<k_{d}$ such that
1099: $\xi_{k_{l}}\ne0$ for $0\le l\le d$. Since the Vandermonde matrix
1100: $\big((k_{l}-1)^{j}\big)$ in non-singular, we get that the vector
1101: fields $(\ad\eta)^{j}\xi=\sum_{k\in\NN}(k{-}1)^{j}\xi_{k}$, $0\le j\le
1102: d$, are linearly independent in $\7l$, a contradiction. This implies
1103: $\xi\in\7P$ as claimed.\nline Since $\7g\subset\7P$ has finite
1104: dimension, every $\eta\in\7g$ is a finite sum
1105: $\eta=\sum_{k=-1}^{m}\eta_{k}$ with $\eta_{k}\in\7P_k$ and $m\in\NN$
1106: not depending on $\eta$. For every polynomial $p\in\RR[X]$ then
1107: $p(\ad\delta)\eta=\sum_{k=-1}^{m}p(k)\eta_{k}$ shows
1108: $\eta_{k}\in\7g_{k}$ for all $k$, that is, $\7g=\bigoplus\7g_{k}$. The
1109: identity $\7g_{-1}=\{iv\dd z: v\in V\}$ follows from the fact that
1110: $\7g_{-1}$ is totally real in $\7P_{-1}$and this implies $f(iV)\subset
1111: iV$ for all $f(z)\dd z\in\7g_{k}$ by
1112: $[\7g_{-1},\7g_{k}]\subset\7g_{k-1}$ and induction on $k$. For the
1113: proof of the last claim assume $\7g_{k}=0$ for some $k\ge0$. Then
1114: $[\7P_{-1},\7g_{k+1}]\subset (\7g_{k}\oplus i\7g_{k})=0$ implies
1115: $\7g_{k+1}\subset\7g_{-1}$ and hence $\7g_{k+1}=0$. \qed
1116:
1117: \Corollary{ZU} In case $\7g=\hol(M,a)$ has finite dimension the
1118: CR-manifold $M=F+iV$ is locally homogeneous at $a$ if and only
1119: if $F$ is locally linearly homogeneous at $a\in F$.
1120:
1121: \Proof From Proposition \ruf{DC} follows that
1122: $W:=\{\xi_{a}:\xi\in\bigoplus_{k\in\NN}\7g_{2k}\}$ is a subspace of
1123: $V$ while $\{\xi_{a}:\xi\in\bigoplus_{k\in\NN}\7g_{2k-1}\}=iV.$ Hence,
1124: $M$ is locally homogeneous at $a$ if and only if $W=T_{a}F$. But for
1125: every $k\in\NN$ and every $\xi\in\7g_{2k}$ the vector field
1126: $\eta:=(\ad ia\dd z)^{2k}\xi$ is in $\7g_{0}$ and satisfies
1127: $\eta_{a}=(-1)^{k}(2k+1)!\,\xi_{a}$, that is,
1128: $W=\{\xi_{a}:\xi\in\7g_{0}\}$.\qed
1129:
1130: Notice that the conclusion $\7g\subset\7P$ together with an eigenspace
1131: decomposition as in Proposition \ruf{DC} can for $V=\RR^{n}$ be
1132: obtained in the same way if instead of `$F$ conical' it is only
1133: assumed for $M=F+i\RR^{n}$ that $\7g=\hol(M,a)$ contains a
1134: vector field $\alpha_{1}z_{1}\dd{z_{1}}+
1135: \alpha_{2}z_{2}\dd{z_{2}}+\dots+\alpha_{n}z_{n}\dd{z_{n}}$ with
1136: $\alpha_{k}>0$ for all $k$, compare e.g. \Ruf{RU}.
1137:
1138: \smallskip
1139:
1140:
1141: \Proposition{ZB} Assume that $\7g:=\hol(M,a)$ has finite dimension and
1142: that $F'\subset V$ is a further conical submanifold with tube manifold
1143: $M'=F'+iV$ and $\7g':=\hol(M',a')$ for some $a'\in F'$. Assume
1144: that the CR-germs $(M,a)$ and $(M',a')$ are isomorphic and let
1145: $g,\tilde g:(M,a)\to(M',a')$ be arbitrary CR-isomorphisms. Then \0
1146: $\dim\7g_{k}=\dim\7g_{k}'$ for all $k\in\ZZ$, where
1147: $\7g^{}_{k},\7g'_{k}$ are given by the decomposition \ruf{DC}.ii. \1
1148: $g$ is represented by a rational transformation on $E$. \1 In case
1149: $\7g_{1}=0$, $\,g$ is represented by a linear transformation in
1150: $\GL(V)\subset\GL(E)$ mapping every $K_{a}^{r}F$ onto
1151: $K_{a'}^{r}F'$. \1 $g=\tilde g$ if and only if $g,\tilde g$ have the
1152: same $d$-jet at $a$, where $d:=\min\{k\in\NN:\7g_k=0\}$.
1153:
1154: \Proof Let $\7l:=\7g\oplus i\7g$ and $\7l_{k}:=\7l\cap\7P_{k}$ for all
1155: $k$. The Lie algebra automorphism $\Psi:=\exp(\ad a\dd z)$ of $\7l$
1156: maps every $f(z)\dd z$ to $f(z+a)\dd z$. For every $k$ denote by
1157: $\7l^{k}\subset\7l$ the subspace of all vector fields that vanish of
1158: order at least $k{+}1$ at $a$. Then $\Psi(\7l^{k})=\bigoplus_{j\ge
1159: k}\7l_{j}$ implies $\dim\7l_{k}=\dim\7l^{k}/\7l^{k+1}$. As a
1160: consequence, $\dim_{\RR}\7g_{k}=\dim_{\CC}\!\7l_{k}$ is a CR-invariant
1161: of the germ $(M,a)$ for every $k$.\nline For the proof of (ii), (iii)
1162: put $\7l':=\7g'+i\7g'$ and extend $g$ to a biholomorphic mapping
1163: $g:U\to U'$ with $g(a)=a'$ and $g(U\cap M)=U'\cap M'$ for suitable
1164: connected open neighbourhoods $U,U'$ of $a,a'\in E$. Consider
1165: the induced Lie algebra isomorphism $ g_{*}:\7l\to\7l'$, compare
1166: \Ruf{GP}. Its inverse $\Theta:=g_{*}^{-1}$ is given by
1167: $$
1168: \Theta\big(f(z)\dd z\re1\big)=g'(z)^{-1}\!f\big(g(z)\big)\re2\dd
1169: z\,.\Leqno{ZA}
1170: $$
1171: Since $\7l$ consists of polynomial vector fields (Proposition
1172: \ruf{DC}) there exist {\sl polynomial} maps $p:E\to E$ and $
1173: q:E\to\End(E)$ such that
1174: $$
1175: \Theta\big(z\dd z\big)=p(z)\dd z\Steil{and}\Theta\big(e\dd
1176: z\big)=\big(q(z)e\big)\dd z
1177: $$
1178: for all $e\in E$. Then \Ruf{ZA} implies $g'(z)^{-1}=q(z)$ and
1179: $g'(z)^{-1}g(z)=p(z)$, that is,
1180: $$
1181: g(z)=q(z)^{-1}p(z)\Leqno{ZV}
1182: $$
1183: in a neighbourhood of $a\in E$ and, in particular, $g$ is
1184: rational.\nline Now suppose $\7g_{1}=0$. Then also
1185: $\7g_{k}=\7g'_{k}=0$ for all $k\ge1$ by \ruf{DC}.iv and (i). Clearly
1186: $\Theta(\7l'_{a'})=\7l_{a}$, where $\7l_{a}:=\{\xi\in\7l:\xi_{a}=0\}$
1187: and similarly $\7l'_{a'}\subset\7l'$ are the isotropy subalgebras at
1188: $a,a'$. Also, $\delta_{a}:=(z-a)\dd z$ is the unique element in
1189: $\7l_{a}$ such that $\ad(\delta_{a})$ induces the negative identity on
1190: the factor space $\7l/\7l_{a}$. Since $\delta_{a'}$ has the same
1191: uniqueness property for $\7l'_{a'}$ we must have
1192: $\Theta(\delta_{a'})=\delta_{a}$. Since $\delta=z\dd z$ is the
1193: $\7g$-component of $\delta_{a}$ as well as of $\delta_{a'}$ in
1194: $\7g\oplus i\7g$ we actually get $\Theta(\delta)=\delta$, that is,
1195: $p(z)\equiv z$. Also $\7l_{-1}$ is $\Theta$-invariant, implying that
1196: $q$ is constant. Therefore $g=q(a)^{-1}$. \nline For the proof of
1197: (iv) we assume without loss of generality that $M=M'$, $a=a'$ and that
1198: $g,\id\in\Aut(M,a)$ have the same $d$-jet at $a$. This implies that
1199: $(g(z)-z)$ vanishes of order $>d$ and $(g'(z)-\id)$ vanishes of order
1200: $\ge d$ at $a$. Therefore also $(g'(z)^{-1}-\id)=(q(z)-\id)$ vanishes
1201: of order $\ge d$ at $a$. Since $q$ is a polynomial of degree $\le d$,
1202: there is a $d$-homogeneous polynomial $s:E\to\End(V)$ with
1203: $q(z)=\id+s(z-a)$. Consider the vector field $\eta:=(z-a)\dd z\in\7l$
1204: and define the holomorphic mappings $h,r:U\to E\,$ by
1205: $$
1206: h(z):=q(z)\big(g(z)-a\big)=(z-a)+r(z)\,.
1207: $$
1208: Then $\Theta\big(\eta)=h(z)\dd z\;\in\;\7g$ shows that $h$ and $r$ are
1209: polynomials of degree $\le d$. But
1210: $$
1211: r(z)=\big(g(z)-z\big)+s(z-a)\big(g(z)-a\big)
1212: $$
1213: vanishes of order $>d$ at $a$, that is, $r=0$ and
1214: $\Theta(\eta)=\eta$. This implies $\Theta(\7l_{-1})=\7l_{-1}$ since
1215: $\7l_{-1}$ is the $(-1)$-eigenspace of $\ad(\eta)$. Therefore $g$ is
1216: an affine transformation on $E$. From $g(a)=a$ and $g'(a)=\id$ we
1217: finally get $g=\id$.\qed
1218:
1219: \Corollary{PL} Let $M:=F+iV$, $M':=F'+iV$ with conical
1220: submanifolds $F,F'\subset V$ and let $a\in F$, $a'\in F'$ be arbitrary
1221: points. Assume furthermore that $\hol(M,a)=\aff(M,a)$ holds. Then the
1222: following conditions are equivalent: \0 The manifold germs
1223: $(M,a),\,(M',a')$ are CR-equivalent. \1 The manifold germs
1224: $(M,a),\,(M',a')$ are affinely equivalent. \1 The manifold germs
1225: $(F,a),\,(F',a')$ are linearly equivalent.\Formend
1226:
1227:
1228:
1229: \bigskip\noindent Recall that $\aut(M,a)\subset\7g=\hol(M,a)$ is
1230: defined as the isotropy subalgebra at $a$ and $\Aut(M,a)$ is the
1231: CR-automorphism group of the manifold germ $(M,a)$, also called the
1232: {\sl stability group} at $a\in M$.
1233:
1234: \Proposition{LO} Let $F\subset V$ be conical and $M=F+iV$. In
1235: case $\7g=\hol(M,a)$ has finite dimension, the following conditions
1236: are equivalent: \0 $\7g_{1}=0$. \1 $\7g=\aff(M,a)$. \1 The tangential
1237: representation $\;g\mapsto g'(a)$ induces a group monomorphism
1238: $\Aut(M,a)\;\hookrightarrow\;\GL(V)$. \smallskip\noindent Each of
1239: these conditions is satisfied if $\aut(M,a)=0$.
1240:
1241: \Proof Let $\7l:=\7g\oplus i\7g\subset\hol(E,a)$ and
1242: $\7l_{k}:=\7g_{k}\oplus i\7g_{k}$ for all $k$. \nline\To12 Follows
1243: from the last claim in Proposition \ruf{DC}. \nline\To23 By
1244: Proposition \ruf{ZB}.iii every $g\in\Aut(M,a)$ is represented by a
1245: linear transformation on $V$. \nline\To31 Let $\xi\in\7g_{1}$ be an
1246: arbitrary vector field. Then there exists a unique symmetric bilinear
1247: map $b:E\times E\to E$ with $\xi=b(z,z)\dd z$. Now $(\ad ia\dd
1248: z)^{2}\xi=-2b(a,a)\dd z\in\7g$, that is, $\eta:=h(z)\dd z$ is in
1249: $\aut(M,a)$, where $h(z):=b(z,z)-b(a,a)$. For every $t\in\RR$
1250: therefore the transformation $\psi_{t}:=\exp(t\eta)\in\Aut(M,a)$ has
1251: derivative $\psi_{t}'(a)=\exp(th'(a))\in\GL(E)$ in $a$. But
1252: $\psi_{t}'(a)\in\GL(V)$ by \ruf{ZB}.iii and thus $2b(a,v)=h'(a)v\in V$
1253: for all $v\in V$. On the other hand $b(a,v)\in iV$ by Lemma \ruf{ZI},
1254: implying $\psi_{t}'(a)=\id$ for all $t\in\RR$. By the injectivity of
1255: the tangential representation therefore $\eta_{t}$ does not depend on
1256: $t$ and we get $\xi=0$. This proves (i) and thus the equivalence of
1257: (i) -- (iii).\nline Suppose $\aut(M,a)=0$ and that there exists a
1258: non-zero vector field $\xi\in\7g_{1}$. Then $\xi_{a}\in iV$ and there
1259: exists an $\eta\in\7g_{-1}$ with $\xi-\eta\in\aut(M,a)$, a
1260: contradiction.\qed
1261:
1262: \Remark{} Notice that the condition (iii) in Proposition \ruf{LO}
1263: states that the tangential representation takes its values in the
1264: subgroup $\GL(V)\subset\GL(E)$. In general, the tangential
1265: representation is not injective and also takes values outside
1266: $\GL(V)$. The tube $\5M$ over the future light cone can serve as a
1267: counterexample for both of these phenomena.\Formend
1268:
1269:
1270: \Proposition{PS} Suppose that $M=F+iV$, $F\subset V$ a conical
1271: submanifold, is locally homogeneous and that
1272: $\7g=\7g_{-1}\oplus\7g_{0}$ for $\7g=\hol(M,a)$. Then the tangential
1273: representation at $a$ induces a group isomorphism
1274: $$
1275: \Aut(M,a)\;\cong\;\{g\in\GL(V):g\7g_{0}g^{-1}=\7g_{0} \steil{and}g(a)=
1276: a\}\,,
1277: $$
1278: where $\7g_{0}$ is considered in the canonical way as linear subspace
1279: of $\,\End(V)$.
1280:
1281: \Proof The assumptions imply that $H(a)\cap F$ is a neighbourhood of
1282: $a$ in $F$ for $H:=\exp(\7g_{0})\subset\GL(V)$. Let $g\in\GL(V)$ be an
1283: arbitrary linear transformation with $g(a)=a$ and
1284: $g\7g_{0}g^{-1}=\7g_{0}$. Then $gHg^{-1}=H$ and hence
1285: $gH(a)=Hg(a)=H(a)$, that is $g\in\Aut(M,a)$. Conversely, by \ruf{ZB}.iii
1286: every element of $\Aut(M,a)$ can be represented by some
1287: $g\in\GL(V)$ with $g(a)=a$. The Lie algebra automorphism
1288: $\Theta=g_{*}$ of $\7g$ leaves $\delta$ and thus also $\7g_{0}$
1289: invariant. As a consequence, $\Theta(\phi)=g\phi g^{-1}$ for every
1290: $\phi\in\7g_{0}$, that is, $g\7g_{0}g^{-1}=\7g_{0}$.\qed
1291:
1292:
1293: \medskip The real Lie algebra structure of $\hol(M,a)$ is a
1294: CR-invariant for the manifold germ $(M,a)$. For certain classes of
1295: conical tube manifolds this gives a complete invariant:
1296:
1297: \Proposition{PQ} Let $F,F'\subset V$ be conical submanifolds for which
1298: the corresponding tubes $M:=F+iV$, $M':=F'+iV$ are locally
1299: homogeneous CR-manifolds. Let $a\in F$, $a'\in F'$ be arbitrary points
1300: and assume that for $\7g:=\hol(M,a)$ the spaces $\7g_{k}$ occurring in
1301: the gradation \ruf{DC}.ii satisfy $\7g_{k}=[\7g_{0},\7g_{0}]=0$. Then
1302: the following conditions are equivalent: \0 The germs
1303: $(M,a),\,(M',a')$ are CR-equivalent. \1 $\hol(M,a)$, $\,\hol(M',a')$
1304: are isomorphic as real Lie algebras.
1305:
1306: \Proof Only the implication \To21 is not obvious. Suppose that
1307: $\Theta:\7g\to\7g':=\hol(M',a')$ is a Lie algebra isomorphism. We use
1308: the same symbol for the complex linear extension $\Theta:\7l\to
1309: \7l'$. \nline Our first step is to show that $\Theta$ can be modified
1310: in such a way that it respects the gradations. To begin with,
1311: $[\7g',\7g']$ is abelian since our assumption implies that
1312: $[\7g,\7g]=\7g_{-1}$ has this property. With $\7g'=\bigoplus \7g_k'$
1313: being the gradation for $\7g'$ as in \ruf{DC}.ii assume that there
1314: exists a minimal integer $k\ge1$ with $\7g'_{k}\ne0$. Then
1315: $\7g'_{-1}=[\delta,\7g'_{-1}]$ and $\7c:=[\7g'_{-1},\7g'_{k}]\ne0$
1316: imply $\7g'_{-1},\7c\subset[\7g',\7g']$ together with
1317: $[\7g'_{-1},\7c]\ne0$, a contradiction. Therefore
1318: $\7g'=\7g'_{-1}\oplus\7g'_{0}$ with $[\7g',\7g']=\7g'_{-1}$, and as a
1319: consequence,
1320: $\Theta(\7g_{-1})=\Theta([\7g,\7g])=[\7g',\7g']=\7g_{-1}'$. Since
1321: $\ad:\7g'_{0}\to \gl(\7g'_{-1})$ is injective, $\delta+\7g'_{-1}$ is
1322: precisely the set of all $\xi\in\7g'$ such that $\ad\xi$ induces on
1323: $\7g'_{-1}$ the negative identity. Therefore, there exists
1324: $\eta\in\7g'_{-1}$ with $\Theta(\delta)=\delta-\eta$. Replacing
1325: $\Theta$ by $\exp(\ad \eta)\Theta=(\id+\ad \eta)\Theta,$ we get
1326: $\Theta(\delta)=\delta$ and finally $\Theta(\7g_{k})=\7g'_{k}$ for all
1327: $k$.\nline There exists a linear operator
1328: $\theta\in\GL(V)\subset\GL(E)$ with $\Theta(e\dd z)=\theta(e)\dd z$
1329: for all $e\in E=V\oplus iV$. We claim that $H'=\theta H\theta^{-1}$
1330: holds for the abelian subgroups $H:=\exp \7g_0$ and $H':=\exp\7g_0'$
1331: of $\GL(V)$. Indeed, application of $\Theta$ to $\big[e\dd
1332: z,\lambda(z)\dd z\big]=\lambda(e)\dd z\;$ yields
1333: $$
1334: \tilde\lambda \theta=\theta\lambda\Steil{ for all}\lambda(z)\dd
1335: z\in\7g_{0}\steil{and}\tilde\lambda(z)\dd z:=\Theta\big(\lambda(z)\dd
1336: z\big)\in\7g'_{0}\,.
1337: $$
1338: We may therefore assume (possibly after replacing $F$ by $\theta F$
1339: and $a$ by $\theta a$) that $H=H'$, $F=H(a)$ and $F'=H(a')$. Denote by
1340: $\Lambda\subset\End(V)$ the associative subalgebra generated by all
1341: $\lambda\in\End(V)$ with $\lambda(z)\dd z\in\7g_{0}$. Then $\Lambda$
1342: is abelian, contains the identity of $\End(V)$ and
1343: $H\subset\Lambda$. Since $\dim\7g<\infty$, the CR-manifold $F+
1344: iV$ is minimal and consequently $F$ cannot be contained in a
1345: hyperplane of $V$. This implies $\Lambda(a)=V$ and thus the existence
1346: of a $g\in\Lambda$ with $a'=g(a)$. From $gH=Hg$ we get
1347: $F'=g(F)$. Since $g(F)$ also cannot be contained in a hyperplane of
1348: $V$ finally $g\in\GL(V)$ follows.\qed
1349:
1350: In Propositions \ruf{UH} an d \ruf{HU} a large class of linearly
1351: homogeneous conical submanifolds $F\subset V$ is given for which the
1352: corresponding tubes $M$ satisfy the condition
1353: $\7g_{1}=[\7g_{0},\7g_{0}]=0$ in Proposition \ruf{PQ}.
1354:
1355: We note that we do not know a single example with $\dim\7g<\infty$ and
1356: $\dim\7g_{k}>\dim\7g_{-k}$ for some $k\in\NN$. We also do not know
1357: any pair $M,M'$ of holomorphically nondegenerate conical tube
1358: manifolds, that are locally CR-equivalent but are not locally affinely
1359: equivalent. For Levi nondegenerate tube manifolds (which necessarily
1360: cannot be conical) such examples can be found in \Lit{DAYA}. In
1361: \Lit{ISMI} even two affinely homogeneous examples are contained which
1362: are locally affinely non-equivalent but whose associated tube
1363: manifolds are locally CR-equivalent.
1364:
1365:
1366:
1367: \KAP{Fuenf}{Some examples}
1368:
1369: In this section we present two classes of examples. To our knowledge,
1370: the only known example of a homogeneous $k$-nondegenerate CR-manifold
1371: with $k\ge3$ occurs in \Lit{FELS} for the case $k=3$: That is a
1372: hypersurface $M$ in a $7$-dimensional compact complex manifold, on
1373: which the simple Lie group $\SO(3,4)^{0}$ acts by biholomorphic
1374: transformations with orbit $M$. In our first example we give for
1375: arbitrary CR-codimenion $c\ge1$ a minimal homogeneous
1376: $3$-nondegenerate as well as a minimal homogeneous $4$-nondegenerate
1377: CR-manifold. The second class of examples deals with tubes $M$ over
1378: cones of the form $\{x\in(\RRp\!)^{n}:\sum_{j\le p}
1379: x_{j}^{\alpha}=\sum_{j>p} x_{j}^{\alpha}\}$ with $1\le p<n$ and
1380: $\alpha\ne0,1$. Using results from the preceding section, we
1381: explicitly determine all Lie algebras $\7g=\hol(M,a)$ for certain
1382: $\alpha$. Among these are all hyperquadrics (that is $\alpha=2$) of
1383: signature $(p,q)$ with $q:=n-p$, where $\7g$ turns out to be
1384: isomorphic to $\so(p{+}1,q{+}1)$.
1385:
1386: The first example of CR-manifolds introduced below consists of tubes
1387: $M=\Gamma(a)+iV$ over certain group orbits $\Gamma(a)$, where the
1388: connected group $\Gamma:=\{g\in\GL(2,\RR):\det(g)>0\}$ acts linearly
1389: on a real vector space $V$. In that way we obtain homogeneous
1390: $k$-nondegenerate CR-manifolds for $k\in\{2,3,4\}$. For dimension
1391: reasons it is impossible to construct CR-manifolds of higher
1392: nondegeneracy, employing the group $\Gamma=\GL(2,\RR)^{0}$. We do not
1393: know how to construct $k$-nondegenerate homogeneous tube manifolds with
1394: $k\ge5$ (should these exist) with suitable other groups, either.
1395:
1396: \Example{JP} For fixed integers $k\in\{2,3,4\}$ and $c\ge1$ let
1397: $V\subset\RR[u_{1},u_{2}]$ be the subspace of all homogeneous
1398: polynomials of degree $m:=k+c-1$. Then the group $\Gamma$ (see above)
1399: acts irreducibly on $V$ by $p\mapsto p\circ g^{-1}$ for all
1400: $g\in\Gamma$ and has the subgroup $\{g\in\RR\id:g^{m}=\id\}$ as kernel
1401: of ineffectivity. For $a:=\sum^{k-2}_{j=0}u_{1}^{j}u_{2}^{m-j}\in V$
1402: the orbit $F=\5F^{k,c}:=\Gamma(a)$ is a connected conical submanifold
1403: of dimension $k$ in $V$. With Proposition \ruf{US} it is easily seen
1404: that
1405: $$
1406: K^{r}_{a}F=\sum^{k-1-r}_{j=0}\RR\,u_{1}^{j}u_{2}^{m-j}\steil{for
1407: all}r\ge0\,.
1408: $$
1409: In particular, the tube manifold $\5M^{k,c}:=F+iV$ is a
1410: $k$-nondegenerate homogeneous CR-manifold of CR-dimension $k$ and
1411: CR-codimension $c$. The manifolds $\5M^{2,c}$ will be discussed in
1412: more detail in Section \ruf{Sieben}. In particular, $\5M^{2,1}$
1413: is linearly equivalent to the future light cone tube $\5M$. For
1414: easier handling let us identify $\RR^{m+1}$ with $V$ via
1415: $(x_{0},x_{1},\dots,x_{m})
1416: \;\leftrightsquigarrow\;\sum^{m}_{j=0}x_{j}u_{1}^{j}u_{2}^{m-j}\;.$
1417: Since $\Gamma$ acts on $\5M^{k,c}$ by (linear) CR-automorphisms, the
1418: linear part $\7g_0$ of $\7g=\hol(\5M^{k,c},a)$ (compare \ruf{DC}.ii)
1419: contains a copy of $\gl(2,\RR)$. More explicitly, this subalgebra is
1420: spanned by the vector fields
1421: $$
1422: \eqalign{ \zeta_{1}:&=\sum^{m}_{j=0}jz_{j}\dd{z_{j}}\,,\mskip53.6mu
1423: \xi^{-1,1}:=\sum^{m-1}_{j=0}(m-j)z_{j+1}\dd{z_{j}}
1424: \cr
1425: \xi^{1,-1}:&=\sum^{m}_{j=1}jz_{j-1}\dd{z_{j}} \,,\mskip59mu
1426: \zeta_{2}:=\sum^{m}_{j=0}(m-j)z_{j}\dd{z_{j}}
1427: \;.\cr}\Leqno{AQ}
1428: $$
1429: In particular, the vector fields $\xi^{1,-1},\,\xi^{-1,1}$ and
1430: $\xi^{0,0}:=\zeta_{1}-\zeta_{2}=\sum^{m}_{j=0}(2j-m)z_{j}\dd{z_{j}}$
1431: span a copy of $\7{sl}(2,\RR)$ in $\7g_{0}$. In case $k=2$ a
1432: straightforward computation shows that
1433: $\Gamma_{a}=\Big\{\Matrix\alpha\beta0\epsilon\in\Gamma:
1434: \epsilon^{m}=1\Big\}$ is the isotropy subgroup at $a\in\5M^{2,c}$.
1435: Hence,
1436: $\Big\{\Matrix\alpha\beta{-\beta}\alpha\in\Gamma\Big\}\cong\CC^{*}$
1437: acts transitively\vadjust{\vskip4pt} on $F$, that is, $\5M^{2,c}$ is
1438: diffeomorphic to $\RR^{n}\times\CC^{*}$, where $n:=m+1=\dim V$.
1439:
1440: \medskip\noindent It seems quite hard to find explicit global
1441: equations for $\5M^{k,c}$ in case $k>2$. For $k=2$ see \Ruf{IP}. Only
1442: for $k=3$ and $c=1$ we have the following: Consider in $\RR^{4}$ the
1443: algebraic hypersurface given by the homogeneous equation
1444: $$
1445: S:=\{x\in\RR^{3}:x^{2}_{0}x_{3}^2+4x_{0}x_{2}^{3}-6x_{0}x_{1}x_{2}x_{3}-
1446: 3x_{1}^{2}x_{2}^{2}+4x_{1}^{3}x_{3}=0\}\,.\Leqno{PI}
1447: $$ It is obvious that $S$ contains the point $a=(1,1,0,0)$. Since the
1448: application of every vector field in $\Ruf{AQ}$ to the defining
1449: function of $S$ gives a multiple of this function, $S$ is invariant
1450: under the group $\Gamma$. Therefore the $3$-nondegenerate homogeneous
1451: CR-submanifold $\5M^{3,1}=\Gamma(a)+i\RR^{4}$ of $\CC^{4}$ is an
1452: open piece of the tube $S+i\RR^{4}$.
1453:
1454:
1455: \bigskip\Example{EB} Fix integers $p\ge q\ge1$ with $n:=p+q\ge3$ and a
1456: real number $\alpha$ with $\alpha^2\ne\alpha$. Then
1457: $$
1458: F=F_{p,q}^{\alpha}:=\Big\{x\in(\RRp)^{n}:\sum_{j=1}^{p}
1459: x_{j}^{\alpha}=\sum_{j=p+1}^{n}
1460: x_{j}^{\alpha} \Big\}\,,
1461: $$
1462: is a hypersurface in $V:=\RR^{n}$. Furthermore, $F$ is a cone and
1463: therefore $\dim K_{a}F\ge1$ for every $a\in F$. On the other hand, the
1464: second derivative at $a$ of the defining equation for $F$ gives a
1465: non-degenerate symmetric bilinear form on $V\times V$, whose
1466: restriction to $T_{a}F\times T_{a}F$ then has a kernel of dimension
1467: $\le1$. Therefore $\dim K_{a}F=1$ for every $a\in F$ and by Corollary
1468: \ruf{AF} the CR-manifold $M=M_{p,q}^{\alpha}:=F_{p,q}^{\alpha}+
1469: iV$ is everywhere $2$-nondegenerate (compare Example 4.2.5 in
1470: \Lit{EBFT} for the special case $n=\alpha=3$). Since $M$ as
1471: hypersurface is also minimal, $\7g=\hol(M,a)$ has finite dimension.
1472:
1473: For the special case $\alpha=2$ and $q=1$ the above cone
1474: $F=F_{n-1,1}^{2}$ is an open piece of the future light cone
1475: $$
1476: \{x\in\RR^{n}:x_{1}^{2}+\dots+x_{n-1}^{2}=x_{n}^{2},\;x_{n}>0\}
1477: $$
1478: in $n$-dimensional space-time, which is affinely homogeneous. In
1479: \Lit{KAZT} it has been shown that for the corresponding tube manifold
1480: $M$ the Lie algebra $\7g=\hol(M,a)$ is isomorphic to $\so(n,2)$ for
1481: every $a\in M$. In case $q>1$ the following result seems to be new:
1482:
1483: \noindent{\fett Case $\alpha=2$:} Consider on $\CC^{n}$ the
1484: symmetric bilinear form $\langle
1485: z|w\rangle:=\sum\epsilon_{j}z_{j}w_{j}$. Then $F$ is an open piece of
1486: the hypersurface $\{x\in\RR^{n}:x\ne0\,,\;\langle
1487: x|x\rangle=0\}$, on which the reductive group $\RR^{*}\cd\O(p,q)$ acts
1488: transitively. Therefore, $\7s_{0}^{}:=\RR\delta\oplus\so(p,q)$ is
1489: contained in $\7g_{0}$. One checks that
1490: $$
1491: \7s:=\7g_{-1}\oplus\7s_{0}^{}\oplus\7s_{1}^{}\,,\qquad\7s_{1}^{}:=
1492: \big\{\big(2i\langle c|z\rangle z-i\langle z|z\rangle c\big)\dd
1493: z:c\in\RR^{n}\big\}
1494: $$
1495: is a Lie subalgebra of $\7g$. The radical $\7r$ of $\7s$ is
1496: $\ad(\delta)$-invariant and hence of the form
1497: $\7r=\7r_{-1}\oplus\7r_{0}\oplus\7r_{1}$ for
1498: $\7r_{k}^{}:=\7r\cap\7g_{k}$. From $\so(p,q)$ semisimple we
1499: conclude $\7r_{0}\subset\RR\delta$. But $\delta$ cannot be in $\7r$
1500: since otherwise $\7g_{-1}\subset\7r$ would give the false statement
1501: $[\7g_{-1},\7s_{1}^{}]\subset\RR\delta$. Therefore $\7r_{0}^{}=0$, and
1502: $[\7g_{-1},\7r_{1}^{}]=[\7r_{-1}^{},\7s_{1}^{}]=0$ implies
1503: $\7r=0$. Now Proposition 3.8 in \Lit{KAZT} implies $\7g=\7s$, and, in
1504: particular, that $\7g$ has dimension ${n+2\choose2}$. In fact, it can
1505: be seen that $\7g$ is isomorphic to $\so(p+1,q+1)$.
1506:
1507: \medskip\noindent{\fett Case $\alpha$ an integer $\ge3$:} Then $F$
1508: is an open piece of the real-analytic submanifold
1509: $$
1510: S:=\Big\{x\in\RR^{n}:x\ne0\Steil{and}\sum_{j=1}^{n}\epsilon_{j}
1511: x_{j}^{\alpha}= 0\Big\}\Leqno{NP}
1512: $$
1513: which is connected in case $q>1$ and has two connected components
1514: otherwise. For every $x\in\RR^{n}$ let $d(x)\in\NN$ be the cardinality
1515: of the set $\{j:x_{j}=0\}$. It is easily seen that $\dim
1516: K_{x}S=1+d(x)$ holds for every $x\in S$. Now consider the group
1517: $$
1518: \GL(F):=\{g\in\GL(V):g(F)=F\}\,.
1519: $$
1520: Every $g\in\GL(F)$ leaves $S$ and hence also
1521: $H:=\{x\in\RR^{n}:d(x)>0\}$ invariant, that is, $g$ is the product of
1522: a diagonal with a permutation matrix. Inspecting the action of
1523: $\GL(F)$ on $\{c\in\overline F:d(c)=n-2\}$ we see that $\GL(F)$ as
1524: group is generated by $\RRp\!\cd\id$ and certain coordinate
1525: permutations. As a consequence, $\7g_{0}=\RR\delta$. Now suppose that
1526: there exists a non-zero vector field $\xi\in\7g_{1}$. Then
1527: $\xi=q(z,z)\dd z$ for some symmetric bilinear map
1528: $q:\CC^{n}\times\CC^{n}\to\CC^{n}$ with $q(c,z)\dd z\in\CC\delta$ for
1529: every $c\in\CC^{n}$. Because of $q\ne0$ symmetric therefore any two
1530: vectors in $\CC^{n}$ must be linearly dependent, which contradicts
1531: $n\ge3$. Therefore $\7g_{1}=0$ and hence $\7g_{k}=0$ for all $k\ge1$
1532: by Proposition \ruf{ZB}.iv. In particular, $\dim\7g=n+1<\dim M$ for
1533: the tube manifold $M=F+iV$, that is, $M$ is not locally
1534: homogeneous. For $n=3$ this gives an alternative proof for Proposition
1535: 6.36 in \Lit{EBEN}.
1536:
1537: \Proposition{} Let $a,a'\in F=F^{\alpha}_{p,q}$ be arbitrary
1538: points. Then in case $3\le\alpha\in\NN$ the CR-manifold germs $(M,a)$
1539: and $(M,a')$ are CR-equivalent if and only if $a'\in\GL(F)(a)$.
1540:
1541: \Proof Suppose that $g:(M,a)\to(M,a')$ is an isomorphism of
1542: CR-manifold germs. From $\7g'=\7g_{-1}'\oplus\,\RR\delta$ for
1543: $\7g':=\hol(M',a')$ Proposition \ruf{ZB}.iii implies that $g$ is
1544: represented by a linear transformation in $\GL(V)$ that we also denote
1545: by $g$. But then $g(F)\subset S$ with $S$ defined in \Ruf{NP}. Because
1546: $g(F)$ has empty intersection with $H$ we actually have $g(F)\subset
1547: F$. Replacing $g$ by its inverse we get the opposite inclusion, that
1548: is $g\in\GL(F)$.\qed
1549:
1550:
1551: \KAP{Sechs}{Levi degenerate CR-manifolds associated with an
1552: endomorphism}
1553:
1554: The lowest CR-dimension for which there exist homogeneous CR-manifolds
1555: that are Levi degenerate but not holomorphically degenerate is
1556: $2$. The construction recipe below will give, up to local affine
1557: equivalence, all affinely homogeneous conical tube submanifolds of
1558: $\CC^{n}$ with CR-dimension $2$. Indeed, it is based on the following
1559: simple observation: Suppose that $F\subset V:=\RR^{n}$ is a conical
1560: locally linearly homogeneous submanifold of dimension $2$. Denote by
1561: $\7a$ the Lie algebra of all linear vector fields on $V$ that are
1562: tangent to $F$. Then, fixing a point $a\in F$, there exists a
1563: $\phi\in\End(V)$ with $\phi(x)\dd x\in\7a$ and $T_{a}F=\RR
1564: a\oplus\RR\phi(a)$. Therefore, the orbit $H(a)$ under the subgroup
1565: $H:=\{\exp(r\id+t\,\phi):r,t\in\RR\}$ is an (immersed) surface in $V$
1566: having the same germ at $a$ as $F$.
1567:
1568:
1569: \Joker{CO}{Construction recipe} Throughout this section let $1<d<n$ be
1570: arbitrary integers and $V$ a real vector space of dimension $n$. Let
1571: furthermore $\phi\in\End(V)$ be a fixed endomorphism and
1572: $\7h\subset\End(V)$ the linear span of all powers $\phi^{k}$ for
1573: $k=0,1,\dots,d-1$. Then $H:=\exp(\7h)\subset\GL(V)$ is an abelian
1574: subgroup, and for given $a\in V$ the orbit $F:=H(a)$ is a cone and an
1575: immersed submanifold of $V$ (not necessarily locally closed in case
1576: $n\ge4$). Furthermore, the tube $M=F+iV\subset E$ is an immersed
1577: CR-submanifold of $E$.
1578:
1579: \Joker{}{Cyclic endomorphisms and vectors} A vector $a\in V$ is called
1580: {\sl cyclic} with respect to $\phi\in \End(V)$ if the $\phi^k(a)$,
1581: $k\ge 0$, span $V$. This is equivalent to
1582: $a,\phi(a),\ldots,\phi^{n-1}(a)$ being a basis of $V$. We call
1583: $\phi\in\End(V)$ {\sl cyclic} if it has a cyclic vector and denote
1584: by $\Cyc(V)\subset\End(V)$ the subset of all cyclic endomorphisms. If
1585: $a,b\in V$ both are cyclic vectors of $\phi$ then there exists a
1586: transformation $g\in\RR[\phi]\subset\End(V)$ with $b=g(a)$. But $g$
1587: commutes with every element of the group $H=\exp(\7h)$ and hence maps
1588: the orbit $H(a)$ onto $H(b)$. In particular, the CR-isomorphism type
1589: of $M=H(a)+iV$ only depends on $\phi$ and $d$, but not on the
1590: choice of the cyclic vector $a$. To emphasize this dependence we also
1591: write $M^{\phi,d}$ for $M$ and $F^{\phi,d}$ for $H(a)$, but only if
1592: $\phi$ is cyclic. We tacitly assume that a choice for a cyclic
1593: vector $a$ has been made. In case $d=2$ we even write $M^{\phi}$ and
1594: $F^{\phi}$ instead of $M^{\phi,2}$ and $F^{\phi,2},$ respectively.
1595: The following proposition shows the relevance of these manifolds in
1596: our discussion.
1597:
1598: \Proposition{PR} For $F=H(a)$ and $M=F+iV$ as in \ruf{CO} the
1599: following conditions are equivalent: \0 $\hol(M,a)$ has finite
1600: dimension. \1 $a$ is a cyclic vector of $\phi$.\par\noindent If these
1601: conditions are satisfied, $M$ is a minimal $2$-nondegenerate
1602: homogeneous CR-manifold with CR-dimension $d$ and Levi kernel
1603: $K_{a}M=\CC a$.
1604:
1605: \Proof \To12 Condition (i) together with the homogeneity of $M$
1606: implies that $M$ is minimal. Let $W\subset V$ be the linear span of
1607: all vectors $\phi^{k}(a)$, $k\ge0$. Then $H\subset\RR[\phi]$ implies
1608: $H(a)\subset W$ and hence $W=V$ by the minimality of $M$. Therefore,
1609: $a$ is a cyclic vector, and the $\phi^{k}(a)$, $0\le k<d$, form a
1610: basis of the tangent space $T_{a}F$. In particular, $F$ has dimension
1611: $d$, which is also the CR-dimension of $M$. \nline \To21 Suppose that
1612: $a$ is a cyclic vector of $\phi$. Lemma \ruf{CR} gives $\RR a\subset
1613: K_{a}F$ since $F$ is a cone in $V$. For the proof of the opposite
1614: inclusion fix an arbitrary $w\in T_{a}F$ with $w\notin\RR a$. Then
1615: $w=\sum_0^m c_j \phi^j(a)$ with $c_m\ne0$ for some $1\le m<d$ and
1616: $\phi^{d-m}(w)\notin T_{a}F$ shows $w\notin K_{a}F$ by Proposition
1617: \ruf{US}. Therefore, $M$ is $2$-nondegenerate by Corollary \ruf{AF}
1618: and $K_{a}M=\CC a$. It remains to show that $M$ is minimal at $a$. But
1619: this immediately follows from Proposition \ruf{PY}.\qed
1620:
1621:
1622: For the manifolds $M=M^{\phi,d}$ it is possible to compute the Lie
1623: algebras $\7g=\hol(M,a)$ in `most cases'. Clearly, the Lie algebra
1624: $\7g_{0}$ contains the $d$-dimensional Lie algebra $\7h$ (as before,
1625: $\7g_{0}$ is canonically identified with a linear subspace of
1626: $\End(V)\,$). In the Propositions \ruf{UH} and \ruf{HU} we will show
1627: that the equality $\7g=\7g_{-1}\oplus\7h$ holds for all $\phi$ in
1628: `general position'.
1629:
1630: \noindent Suppose that $\phi\in\End(V)$ has the cyclic vector $a\in V$
1631: and that the integer $d$ satisfies $1<d<n$. Let
1632: $$
1633: \chi^{}_{\phi}:=\prod_{j=1}^{s}(X-\lambda_{j})^{n_{j}}\;\in\;\RR[X]\,,\quad
1634: n_{j}\ge1\,,\Leqno{CP}
1635: $$ with mutually distinct eigenvalues
1636: $\lambda_{1},\dots,\lambda_{s}\in\CC$ be the characteristic polynomial
1637: of $\phi$. Let furthermore $\alpha_{1},\dots,\alpha_{n}$ be the
1638: family of all roots of $\chi_{\phi}$, that is, each $\lambda_{j}$
1639: occurs $n_{j}$-times in this string. As usual,
1640: $\alpha_{1},\dots,\alpha_{n}$ is called an {\sl arithmetic
1641: progression} in $\CC$ if there exists a $\beta\in\CC$ such that, after
1642: a suitable permutation, $\alpha_{j}=\alpha_{1}+(j-1)\beta$ for all
1643: $1\le j\le n$.
1644:
1645: \Proposition{UH} Let $M=M^{\phi,d}$ and $\7g:=\hol(M,a)$ for a cyclic
1646: vector $a\in V\cap M$ of $\phi$. Then $\7g_{0}=\7h$ holds if one of the
1647: following conditions is satisfied. \0 $d=2$ and the characteristic
1648: roots $\alpha_{1},\dots,\alpha_{n}$ of $\phi$ do not form an {\sl
1649: arithmetic progression}. \1 $d\ge3$ and $s>d$.\Formend
1650:
1651: \noindent For the proof we need several preparations. To simplify the
1652: notation at various places let us introduce
1653: $$
1654: S:=\{1,\dots,s\},\quad m:=d-1\Steil{and}m_{k}:=n_{k}-1\steil{for
1655: all}k\in S\,.\Leqno{UK}
1656: $$
1657: For every $K\subset S$ define furthermore
1658: $$
1659: \Delta_{K}:= \big\{\beta_{jk}:j\in
1660: S,k\in K\big\}\steil{with}\beta_{jk}:=(\lambda_{j}^{}-\lambda_{k}^{},
1661: \lambda_{j}^{2}- \lambda_{k}^{2},
1662: \dots,\lambda_{j}^{m}-\lambda_{k}^{m})\in\CC^{m}\Leqno{JG}
1663: $$
1664: and denote by $\6K$ the set of all non-empty subsets $K\subset S$ such
1665: that there exist subsets $P\subset Q\subset (S\setminus K)$ with the
1666: following two properties, where the maximum over the empty set here is
1667: defined to be $0\,$: \0 $\sum_{k\in K}n_{k}\;+\;\sum_{q\in Q}n_q
1668: \;\;\ge\;\; d\;+\;\max\re2\{n_{p}-1:p\in P\}\,.$ \1 To every
1669: $\beta\in\Delta_{Q}\!\setminus\Delta_{K}$ there exist uniquely
1670: determined $j\in S$ and $q\in Q$ with $\beta=\beta_{jq}$ such that
1671: $n_{j}<n_{q}$ and $q\in P$.
1672:
1673: \noindent Notice that $\6K$ contains every subset $K\subset S$ with
1674: $\;\sum_{k\in K}n_{k}\ge d\;$ (just take $Q=\emptyset$). In the
1675: following Lemma we show that (i) or (ii) in Proposition \ruf{UH} will
1676: follow from a more general technical condition that will allow us to
1677: give a uniform proof of \ruf{UH} for all $d$.
1678:
1679: \Lemma{UL} Suppose that $ s\ge d$ and $\bigcap_{K\in\6K}\!\Delta_{K}
1680: \,=\,\{0\}$ holds. Then one of the conditions (i) and (ii) in
1681: Proposition \ruf{UH} is satisfied.
1682:
1683: \Proof {\fett $d=2$:} Assume that $\alpha_{1},\dots,\alpha_{n}$ is not
1684: an arithmetic progression and that there exists a non-zero
1685: $\beta\in\Cap_{K\in\6K}\Delta_{K}$. We claim that
1686: $\lambda_{1},\dots,\lambda_{s}$ is an arithmetic progression:
1687: Otherwise $s\ge3$ and there exists a $k$ with $1<k<s$ such that,
1688: without loss of generality, $\lambda_{j}=\lambda_1+(j-1)\beta$ for all
1689: $1\le j\le k$ and $\lambda_1-\beta\ne\lambda_{r}\ne \lambda_k+\beta$
1690: for all $r>k$. For every $j>k$ the set $\Delta_{\{k,j\}}$ contains the
1691: number $\beta$, that is, there is an $r\in S$ with
1692: $\lambda_{r}=\lambda_{k}+\beta$ or
1693: $\lambda_{r}=\lambda_{j}+\beta$. The first possibility violates our
1694: assumptions. In the second case necessarily $r>k$ must hold since
1695: $r\le k$ would imply $\lambda_{j}=\lambda_1-\beta$ and thus the second
1696: possibility cannot be true for {\it all}\/ $j$ with $k<j\le s$. This
1697: proves that $\lambda_1,...\,,\lambda_s$ is an arithmetic progression
1698: and also $s<n$ by assumption. In particular, $n_j>1$ for some $j\in
1699: S$. Next we claim that $\{1\}\in\6K$ and $\{s\}\in\6K$: To see that
1700: $\{1\}\in\6K,$ let $k\in \{1,...\,, s\}$ be minimal with
1701: $n_{k}>1$. With $P:=Q:=\{k\}\setminus\{1\}$ condition \Ruf{JG}.ii is
1702: fulfilled and the first part of the claim follows. A similar argument
1703: proves also $\{s\}\in\6K$. But then
1704: $\Delta_{\{1\}}\cap\Delta_{\{s\}}=\{0\}$ gives a contradiction.\nline
1705: {\fett $d\ge3$:} Suppose that $0\ne \beta_{jk}\in
1706: \Cap_{K\in\6K}\Delta_{K}$ and that $s>d$. Then
1707: $L:=S\setminus\{k\}\in\6K$ (with $Q=\emptyset$) and there are $\ell\in
1708: L$, $r\in S$ with $\beta_{jk}=\beta_{r\ell}$. Since $d\ge 3,$ the
1709: equation $\beta_{jk}=\beta_{r\ell}$ implies $k=\ell$, a
1710: contradiction.\phantom{XXXX}\qed
1711:
1712: \medskip\noindent{\fett Proof of
1713: Proposition \UH:} It is enough to assume that the assumption of Lemma
1714: \ruf{UL} is satisfied. Consider the decomposition
1715: $E=E_{1}\oplus\cdots\oplus E_{s}$ with $E_{k}$ the kernel of
1716: $(\phi-\lambda_{k})^{n}$ for every $k\in S$. Denote by
1717: $\pi_{k}:E\to E_{k}$ the canonical projection and by
1718: $\epsilon_{k}:E_{k}\to E$ the canonical injection. Then
1719: $a_{k}:=\pi_{k}(a)$ is a cyclic vector for
1720: $\phi_{k}:=\pi_{k}\phi\re1\epsilon_{k}\in\End(E_{k})$. Furthermore, if
1721: we put $a^{j}_{k}:=(\phi_{k}-\lambda_{k})^{j}(a_{k})$ for all $j\ge0$,
1722: then $a^{0}_{k},\dots,a^{m_{k}}_{k}$ is a basis of $E_{k}$. With
1723: $m=d-1$ as defined above let
1724: $\Phi:=(\phi^{1},\dots,\phi^{m})\in\End(V)^{m}$.
1725:
1726: \noindent For every real (or complex) vector space $W$ and all tuples
1727: $t=(t_{1},\dots,t_{m})\in\RR^{m}$, $w=(w_{1},\dots,w_{m})\in W^{m}$
1728: let us write as shorthand $t\cd w:=\sum_{j}t_{j}w_{j}$. For every
1729: $t\in\RR^{m}$ the point $e^{t\cd\Phi}(a)$ is contained in $F=M\cap V$.
1730:
1731: \noindent Now fix an arbitrary $\mu\in\7g_{0}\subset\End(V)$. Since
1732: the vector field $\mu$ is tangent to $F$, to every
1733: $t\in\RR^{m}$ there exist real coefficients
1734: $r_{0},r_{1},\dots,r_{m}$ with
1735: $$
1736: \mu e^{t\cd\Phi}(a)\;=\;Re^{t\cd\Phi}(a)\Steil{for}R:=
1737: \sum^{m}_{\ell=0}r^{}_{\li2\ell}\phi^{\ell}\,.\Leqno{PT}
1738: $$
1739: Actually, every $r_{\ell}$ has to be considered as a real valued function
1740: on $\RR^{m}$. Put
1741: $$
1742: \mu_{k}:=\pi_{k}\mu\Steil{and}N_{k}:=\big((\phi_{k}-\lambda_{k}),
1743: \dots,(\phi^{m}_{k}-\lambda_{k}^{m})\big)\;\in\;\End(E)^{m}
1744: $$
1745: for all $k\in S$. Applying $\pi_{k}$ to \Ruf{PT}
1746: gives
1747: $$
1748: \eqalign{ R\,e^{t\cd N_{k}}(a_{k})\;&=\;e^{t\cd N_{k}}R(a_{k})=\;
1749: \sum_{j=1}^{s}e^{t\cd\beta_{jk}} \mu_{k}e^{t\cd N_{j}}(a_{j})
1750: \,\quad\hbox{with}\cr
1751: R(a_{k})&=\;\sum^{m_{k}}_{j=0}\rho_{k,j}a^{j}_{k}\Steil{for}
1752: \rho_{k,j}(t):=\sum^{m}_{\ell=j}\textstyle{\ell\choose
1753: j}\lambda^{\ell-j}_{k}r^{}_{\ell}(t)\;\;.\cr }\Leqno{JA}
1754: $$
1755: For every subset $B\subset\CC^{m}$ denote by
1756: $\6F(B)\subset\5C(\RR^{m},\CC)$ the smallest linear subspace
1757: containing all functions $h(t)e^{t\cd\beta}$ with $\beta\in B$ and
1758: $h\in\CC[t]=\CC[t_{1},\dots,t_{m}]$ a polynomial function on
1759: $\RR^{m}$. Then it is well known that every $f\in\6F(B)$ has a unique
1760: representation $f=\sum_{\beta\in B}f^{[\beta]}e^{t\cd\beta}$ with
1761: $f^{[\beta]}\in\CC[t]$ and $\{\beta\in B:f^{[\beta]}\ne0\}$ finite.
1762: Since $t\cd N_{j}$ is nilpotent, i.e., $e^{\pm t\cd N_j}$ are
1763: polynomial, as a consequence of the identities \Ruf{JA} we get for
1764: every $K\subset S$
1765: $$
1766: \rho_{k,j}\;\in\;\6F(\Delta_{K})\Steil{for every}k\in K\steil{and}0\le
1767: j\le m_{k}\,.\Leqno{JC}
1768: $$
1769: Denote by $\6B$ the set of all subsets $B\subset\CC^{m}$ with
1770: $r_{\ell}\in\6F(B)$ for all $\ell$.
1771:
1772: \noindent{\bf Claim 1:} \quad {\sl $\bigcap_{K\in
1773: \6K}\Delta_K\in\6B$, that is, $r_{\ell}\in\CC[t]$ for all
1774: $\ell$.}
1775:
1776: \ProofC Fix an arbitrary $K\in\6K$ and let $P\subset Q$ be as in the
1777: definition of $\6K$. Since $\6B$ is closed under intersections it is
1778: enough to show $\Delta_{K}\in\6B$. Assume to the contrary that this is
1779: not true. Consider the linear system of equations for the $r_{\ell}$,
1780: compare also \ruf{HA},
1781: $$
1782: \rho_{k,j}=\sum^{m}_{\ell=j}\textstyle{\ell\choose
1783: j}\lambda^{\ell-j}_{k}r^{}_{\ell}\;\in\;\6F(\Delta_{K\cup
1784: Q})\Steil{for all}k\in K\cup Q\steil{and}0\le j\le
1785: m_{k}\,.\Leqno{KZ}
1786: $$
1787: The coefficient matrix is of generalized Vandermonde type and hence
1788: has rank $d=m+1$ since by the definition of $\6K$ the number of
1789: equations is at least $d$. This implies
1790: $r_{\ell}\in^{}\6F(\Delta_{K}\cup\Delta_{Q})$ for all $\ell$. Since
1791: by assumption not all $r_{\ell}$ are in $\6F(\Delta_{K})$ there is a
1792: $\beta\in\Delta_{Q}\!\setminus\Delta_{K}$ such that the
1793: $\beta$-components $r^{[\beta]}_{\ell}\in\CC[t]$ do not vanish for all
1794: $\ell$ simultaneously. By the definition of $\6K$ there are uniquely
1795: determined $p\in S$ and $q\in P$ with $\beta=\beta_{pq}$. Define
1796: $L:=K\cup Q\setminus \{q\}$. Since $\beta\not\in\Delta_L$ we get from
1797: \Ruf{JA} the following linear system
1798: $$
1799: \rho^{[\beta]}_{k,j}=\sum^{m}_{\ell=j}\textstyle{\ell\choose
1800: j}\lambda^{\ell-j}_{k}r^{[\beta]}_{\ell}\;=0\,,\steil{where}k
1801: \steil{runs through}L\steil{and}0\le j\le m_{k}\,\,.\Leqno{KX}
1802: $$
1803: By the very definition of $K$, $P$ and $Q$ it follows that the above
1804: linear system consists of at least $d-1$ equations. Consequently we
1805: can write it in the form:
1806: $$
1807: \sum^{m-1}_{\ell=j}\textstyle{\ell\choose
1808: j}\lambda^{\ell-j}_{k}r^{[\beta]}_{\ell}\;\in\CC
1809: r^{[\beta]}_{m}\Steil{for all}k\in L \steil{and}\;0\le j\le m_{k}\;.
1810: $$
1811: Since its coefficient matrix is of generalized Vandermonde type, every
1812: $r^{[\beta]}_{\ell}$ is a complex multiple of $r^{[\beta]}_{m}$ and,
1813: in particular, $r^{[\beta]}_{m}\ne0$. We claim
1814: $\rho^{[\beta]}_{q,0}\ne0$. Indeed, otherwise we could add the
1815: equation $\rho^{[\beta]}_{q,0}=0$ to the linear system \Ruf{KX}, which
1816: then has a coefficient matrix of generalized Vandermonde type with
1817: rank $d$ contradicting $r^{[\beta]}_{m}\ne0$. Denote by $D$ the
1818: degree of
1819: $\rho^{[\beta]}_{q,0}=\sum_{\ell}\lambda^{\ell}_{q}r^{[\beta]}_{\ell}$. Then
1820: $D=\deg \rho^{[\beta]}_{q,0}=\deg r_{m}^{[\beta]}\ge 0$ and all
1821: $r_{\ell}$ and $\rho^{[\beta]}_{q,j}$ have degree $\le D$. The
1822: equations \Ruf{JA} imply (replace $k$ by $q$, form the
1823: $\beta$-components for $\beta:=\beta_{pq}$ and carry out the
1824: multiplication with $e^{t\cd N_{q}}$)
1825: $$
1826: \mu_{q}e^{t\cd N_{p}}\re1(a_{p})\;=\;e^{t\cd
1827: N_{q}}\sum^{m_{q}}_{j=0}\rho^{[\beta]}_{q,j}a^{j}_{q}=\sum^{m_{q}}_{j=0}
1828: \big(f_{j}\rho^{[\beta]}_{q,0}+g_{j}\big)a^{j}_{q}\Leqno{KI}
1829: $$
1830: for certain polynomials $f_{j},g_{j}\in\CC[t]$ with
1831: $\deg(g_{j})<\deg(f_{j})=j$. Comparing degrees on both sides in \Ruf{KI}
1832: and keeping in mind $n_{p}<n_{q}$ (by the definition of $\6K$) we get
1833: $$
1834: D+m_{q}=\deg(f_{m_{q}}\rho^{[\beta]}_{q,m_{q}}+g_{m_{q}})\le
1835: m_{p}<m_{q}\,.
1836: $$
1837: This contradicts $r_m^{[\beta]}\ne 0$ and Claim 1 is proved.\qd
1838:
1839:
1840: \noindent{\bf Claim 2: \sl Every $r_{\ell}$ is a constant polynomial.}
1841:
1842: \ProofC Fix a $k\in S$ with $D:=\deg(\rho_{k,0})=\max_{j\in
1843: S}\deg(\rho_{j,0})$. With $s\ge d$ a Vandermonde argument applied to
1844: the linear system $\rho_{j,0}=\sum_{\ell}\lambda^{\ell}_{j}r_{\ell}$,
1845: $\,j\in S$, gives that every $r_{\ell}$ has degree $\le D$. As in
1846: \Ruf{KI} we have
1847: $$
1848: \mu_{k}e^{t\cd N_{k}}\re1(a_{k})\;=\;e^{t\cd
1849: N_{k}}\sum^{m_{k}}_{j=0}\rho_{k,j}a^{j}_{k}=\sum^{m_{k}}_{j=0}
1850: \big(f_{j}\rho_{k,0}+g_{j}\big)a^{j}_{k}\Leqno{KY}
1851: $$ for polynomials $f_{j},g_{j}\in\CC[t]$ with
1852: $\deg(g_{j})<\deg(f_{j})=j$. All coefficient polynomials in \Ruf{KY}
1853: in front of the $a^{j}_{k}$ have degree $\le m_{k}$, that is
1854: $D+m_{k}\le m_{k}$ and hence $D\le0$. This proves Claim 2.\qd
1855: \nline The proof of \ruf{UH} now is complete: Indeed, since
1856: $F$ contains a basis of $V$, the endomorphism $\mu$ is uniquely
1857: determined by the function tuple $(r_{\ell})$, that is, $\dim\7g_{0}\le
1858: d=\dim \7h $.\qed
1859:
1860: \smallskip \Remark{HA} For given tuples
1861: $\lambda_{1},\dots,\lambda_{s}\in\CC$ and $\,n_{1},\dots,n_{s}\in\NN$
1862: with $n_{k}\ge1$ and $n:=\sum\!n_{k}$ let $L:=\{(k,j):1\le k\le
1863: s,\,0\le j<n_{k}\}$ endowed with the lexicographic order. Then it can
1864: be seen that the following $n\Times n$-matrix (every entry with $\ell<
1865: j$ is zero)
1866: $$
1867: \Big(\textstyle{\ell\choose
1868: j}\lambda^{\ell-j}_{k}\Big)_{0\le\ell<n,\re1(k,j)\in L}\Steil{has
1869: determinant}\displaystyle\prod\limits_{p<q}
1870: (\lambda_{q}-\lambda_{p})^{n_{p}n_{q}}\;.
1871: $$
1872:
1873:
1874: Notice from the proof of \ruf{UH} that the condition
1875: $\bigcap_{K\in\6K}\!\Delta_{K} \,=\,\{0\}$ guarantees that every
1876: $\mu\in\7g_{0}$ leaves every generalized eigenspace $E_{k}$ of $\phi$
1877: invariant, while $s\ge d$ guarantees that every such $\mu$ actually is
1878: in $\7h$. In the next section we will see that the condition (i) in
1879: Proposition \ruf{UH} for $d=2$ is optimal, compare Proposition \ruf{ZR}.
1880:
1881: \Proposition{HU} Let $M=M^{\phi,d}$ and assume that $\7g_{0}=\7h$ for
1882: $\7g=\hol(M,a)$. Then $\7g=\aff(M,a)$ and $\aut(M,a)=0$.
1883:
1884: \Proof For the proof of $\7g=\aff(M,a)$ it is enough to show
1885: $\7g_{1}=0$ by Proposition \ruf{DC}.iv. This is more easily done in
1886: the more general complex setting, compare the following Lemma
1887: \ruf{UF}. Finally, counting dimensions yields $\aut(M,a)=0$.\qed
1888:
1889: \medskip\noindent It remains to show the next Lemma. As before,we
1890: identify the spaces $\End(E)$ and $\7P_{0}$, see \Ruf{HJ}.\vskip-30pt
1891:
1892: \Lemma{UF} Let $\phi\in\Cyc(E)$ be an arbitrary cyclic
1893: endomorphisms. For given integer $d<n=\dim E$ let furthermore $\7h$ be
1894: the complex linear span of all powers $\phi^{j}$, $0\le j<d$. Then
1895: $$
1896: \big\{\xi\in\7P_{1}:[\7P_{-1},\xi]\subset\7h\big\}\;=0\;\,.
1897: $$
1898: \Proof We identify $E$ with $\CC^{n}$ in such a way that the
1899: matrix $\Phi$ of $\phi$ is in Jordan normal form, more precisely:
1900: $\Phi$ is a block diagonal matrix with Jordan blocks
1901: $J_{1},\dots,J_{s}$, where each block $J_{l}$ is lower triangular, has
1902: the eigenvalue $\lambda_{l}$ on its main diagonal and is of size
1903: $n_{l}\Times n_{l}$ for some $n_{l}\ge1$. We also introduce an
1904: equivalence relation on $\{1,\dots,n\}$ in the following way: Put
1905: $j\sim k$ if the $j^{\th}$ row and the $k^{\th}$ column in $\Phi$
1906: intersect in one of the Jordan blocks. \nline Now suppose that there
1907: exists a non-zero vector field $\xi\in \7P_{1}$ with
1908: $[\7P_{-1},\xi]\subset\7h$. This $\xi$ has a unique representation
1909: $$
1910: \xi=\sum_{j,k,p=1}^{n}c_{p}^{jk}z_{j}z_{k}\dd{z_{p}}\Steil{with}
1911: c_{p}^{jk}=c_{p}^{kj}\in\CC\,.
1912: $$
1913: For every $j\le n$ the vector field $\dd z_{j}$ is contained in
1914: $\7P_{-1}$. Therefore
1915: $$
1916: \xi_{j}:={1\over2}\big[\dd{z_{j}},\xi\big]=\sum_{k,p=1}^{n}
1917: c^{jk}_{p}z_{k} \dd{z_{p}}
1918: $$
1919: is contained in $\7h$ implying $c_{p}^{jk}=0$ if $k\not\sim p$ and, by
1920: symmetry, $c_{p}^{jk}=0$ if $j\not\sim p$. This implies
1921: $$
1922: \xi_{j}=\sum_{p\sim j}\sum_{k\sim j}c^{jk}_{p}z_{k}
1923: \dd{z_{p}}\,.\Leqno{VZ}
1924: $$ By assumption $\xi\ne0$ and therefore $\xi_{r}\ne0$ for some $r\le
1925: n$. Without loss of generality we may assume $r\sim1$. Replacing
1926: $\phi$ by $(\phi-\lambda_{1})$ we also may assume $\lambda_{1}=0$.
1927: Put $b:=n_{1}$, that is, the Jordan block $J_{1}$ has size $b\times b$
1928: and is nilpotent. Define for all $1\le p,k\le b$
1929: $$
1930: \eta_{p}:=z_{1}\dd{z_{p}}+z_{2}\dd{z_{p+1}}+\dots+z_{b+1-p}\dd{z_{b}}
1931: \,,
1932: $$
1933: $$
1934: \7D_{k}:=\sum_{j=k}^{b}\CC\re1\eta_{j}\Steil{and}\psi_{k}:=
1935: \phi^{k-1}\prod_{l=2}^{s}(\phi-\lambda_{l})^{n_{l}}\;
1936: \in\;\End(E)\,.
1937: $$
1938: From \Ruf{VZ} we know
1939: $\xi_{j}\in\7D_{1}\cap\7h$ for all $j\le b$. In particular,
1940: $$ \xi_{j}=\sum_{p=1}^{b}c^{j1}_{p}\eta_{p}\,,\qquad c_{p}^{jk}=
1941: \cases{c^{j1}_{p+1-k}&if $k\le p$\cr0&otherwise\cr}\qquad\hbox{and hence}
1942: $$
1943: $$
1944: \xi_{j}=c^{11}_{1}\eta^{}_{j}+c^{11}_{2} \eta^{}_{j+1}+
1945: \dots+c^{11}_{b+1-j}\eta^{}_{b}
1946: $$ for all $j$ due to the symmetry of $c^{jk}_{p}$ in the upper
1947: indices. As a consequence there exists a minimal $q\le b$ with
1948: $q\ge1$ and $c^{11}_{q}\ne0$. But then
1949: $\xi_{b+1-q}=c^{11}_{q}\eta^{}_{b}$ implies
1950: $\eta^{}_{b}\in\7D_{b}\cap\7h$. We show that this cannot be
1951: true:\nline Since $\psi_{1}$ as polynomial in $\phi$ has constant term
1952: $\lambda_{2}\lambda_{3}\dots\lambda_{s}\ne0$, we get that $\psi_{k}$
1953: spans $\7D_{k}$ over $\7D_{k+1}$ for every $k\ge1$. This implies
1954: $\eta_{b}=\sum_{j=1}^{n}e_{j}\phi^{j-1}$ for suitable real
1955: coefficients $e_{j}$ with $e_{n}\ne0$ and thus
1956: $\eta_{b}\notin\7h$.\phantom{XXX}\qed
1957:
1958:
1959: \bigskip For the application of Proposition \ruf{UQ} to manifolds of
1960: the type $M=M^{\phi,d}$ it is necessary to know when $M$ is simply
1961: connected and when $\Aut(M,a)$ is the trivial group. \nline A
1962: sufficient condition for $M^{\phi,d}$ (and $F^{\phi,d}=H\Kl a$) to be
1963: simply connected is the following: {\sl There exist eigenvalues
1964: $\lambda_{1},\dots,\lambda_{d}$ of $\phi$ such that ${\det(A+\overline
1965: A)}\ne0$, where $A=(\lambda_{j}^{k-1})_{1\le j,k\le d}$ is the
1966: corresponding Vandermonde matrix}. \nline To get a partial answer to
1967: the second question consider to given $\epsilon\in\RR^{*}$ a $g\in\GL(V)$
1968: with $g(a)=a$ and $g\phi g^{-1}=\epsilon\phi$. From
1969: $g\phi^{k}g^{-1}=(\epsilon\phi)^{k}$ we get
1970: $g(\phi^{k}(a))=\epsilon^{k}\phi^{k}(a)$ for all $k\ge0$, that is, $g$
1971: is uniquely determined in $\GL(V)$ by the above assumptions since
1972: $\phi$ is cyclic. A further consequence is
1973: $g\re1\exp(t\phi^{k})\re2g^{-1}=\exp(t\epsilon^{k}\phi^{k})$ for all
1974: $t\in\RR$ and $k\ge0$. This means $gHg^{-1}=H$ for the group
1975: $H=\exp(\7h)$. But then $g(F)=F$ for $F=H(a)$ and consequently
1976: $g\in\Aut(M,a)$.
1977:
1978: Notice that we always may assume without loss of generality that
1979: $\phi\in\Cyc(V)$ has trace $0$ (otherwise replace $\phi$ by
1980: $(\phi-c\,\id)\in\End(V)$ for $c:=n^{-1}\tr(\phi)$, since this
1981: procedure does not change the algebra $\7h)$.
1982:
1983: \Lemma{KR} Suppose that $2d\le n+1$ for $n=\dim V $ and that $\phi$
1984: is trace-free. Suppose in addition that $\aut(M,a)=0$ holds for the
1985: cyclic vector $a\in V$ and $M:=M^{\phi,d}$. Then
1986: $$\Aut(M,a)=\big\{g\in\GL(V):g(a)=a\steil{and}g\phi
1987: g^{-1}=\pm\phi\big\}\,.$$ In particular, $\Aut(M,a)$ always has order
1988: $\le 2$ and is trivial, for instance, if the spectrum of $\phi$ in
1989: $\CC$ is not symmetric with respect to the origin of $\CC$.
1990:
1991:
1992: \Proof By Proposition \ruf{LO} the assumption $\aut(M,a)=0$ implies
1993: $\7g_{1}=0$ and $\7g_{0}=\7h$. As a consequence of Proposition
1994: \ruf{PS} therefore
1995: $\Aut(M,a)=\big\{g\in\GL(V):g(a)=a\steil{and}g\7hg^{-1}= \7h\big\}$
1996: holds. Let $g\in\Aut(M,a)$ be an arbitrary automorphism. Then $g\phi
1997: g^{-1}=\sum^{m}_{j=0}c_{j}\phi^{j}$ for real coefficients $c_{j}$ and
1998: $m:=d-1$. We show by induction on $k$ that $c_{j}=0$ holds for all
1999: $j>m/k$ and all $1\le k\le m$. For $k=1$ this is obvious. So fix a
2000: $k>1$ with $k\le m$. By induction hypothesis $g\phi^{k}
2001: g^{-1}=\big(\sum c_{j}\phi^{j}\big)^{k}
2002: =\sum^{2m}_{\ell=0}e_{\ell}\phi^{\ell}\in\7h$ with real coefficients
2003: $e_{\ell}$. Since $2m<n$ by assumption, we must have $e_{\ell}=0$ for
2004: all $\ell>m$, that is $c_{j}=0$ for all $j>m/k$. For $k=m$ this
2005: implies $c_{j}=0$ for all $j>1$. Taking traces finally gives $g\phi
2006: g^{-1}=\epsilon\phi$ for $\epsilon:=c_{1}$. By the above example
2007: $\phi$ cannot be nilpotent, that is, $\phi$ has non-zero spectrum in
2008: $\CC$. Since this spectrum is invariant under multiplication with
2009: $\epsilon$ necessarily $\epsilon=\pm1$ holds.\qed
2010:
2011:
2012:
2013: \KAP{Sieben}{Homogeneous 2-nondegenerate manifolds of CR-dimension 2}
2014:
2015: In this section we specialize to homogeneous tube manifolds $M=F+iV$
2016: in $E=V\oplus iV$ of CR-dimension $2$, that is, where $F\subset V$ is
2017: a surface of dimension $2$. We begin with manifolds of type
2018: $M^{\phi}=M^{\phi,2}$ that are obtained by the construction recipe
2019: \ruf{CO}. Since Propositions \ruf{UH} and \ruf{HU} of the preceding
2020: section do not cover the case where the characteristic roots
2021: $\alpha_{1},\dots,\alpha_{n}$ of $\phi\in\Cyc(V)$, $n=\dim V$, form an
2022: arithmetic progression, let us discuss this case first:\nline Possibly
2023: after replacing $\phi$ by $\phi-r\id$ with an appropriately chosen
2024: constant $r$ we may assume without loss of generality that $\phi$ is
2025: trace-free. Since multiplication of $\phi$ by any non-zero real number
2026: does not change the algebra $\7h=\RR\id\oplus\RR\phi$, there are
2027: essentially 3 different cases for $\phi$ with characteristic roots
2028: forming an arithmetic progression -- either $\phi$ is nilpotent or
2029: $\phi$ has pairwise different characteristic roots in $\RR$ or in
2030: $i\RR$. Consider the manifold $M:=\5M^{2,n-2}=\5F^{2,n-2}+iV$ from
2031: Example \ruf{JP}. As already remarked in \ruf{JP} the conformal
2032: subgroup $H^{\i}:=\RRp\cd\SO(2)\subset\GL(2,\RR)$ acts transitively on
2033: $\5F^{2,n-2}$. The corresponding Lie algebra is
2034: $\7h_{\i}:=\RR\id\oplus\RR\phi^{}_{\i}$, where
2035: $\phi_{\i}:=\xi^{1,-1}-\xi^{-1,1}\in\7g_{0}$ is a trace-free
2036: semisimple endomorphism with eigenvalues in $i\RR$, see \Ruf{AQ} for
2037: the notation. Next consider the subgroup
2038: $H^{\r}:=\RRp\cd\SO(1,1)\subset\GL(2,\RR)$ with Lie algebra
2039: $\7h^{\r}:=\RR\id\oplus\RR\phi^{}_{\r}$, where
2040: $\phi_{\r}:=\xi^{1,-1}+\xi^{-1,1}\in\7g_{0}$ is a trace-free
2041: semisimple endomorphism with real eigenvalues. In particular,
2042: $H^{\r}(a)$ is open in $H^{\i}(a)$. Finally, consider the solvable
2043: subgroup
2044: $$
2045: H^{\z}:=\Big\{\Matrix\alpha0\beta\alpha\in\GL(2,\RR):\alpha>0\Big\}
2046: \steil{with Lie algebra}\7h^{\z}:=\RR\id\oplus\RR\phi^{}_{\z}\,,
2047: $$
2048: where $\phi_{\z}:=\xi^{1,-1}$ only has zero eigenvalues. Again, the
2049: orbit $H^{\z}(a)$ is open in $H^{\i}(a)$. This implies that the
2050: manifolds $M^{\phi_{\r}}$ and $M^{\phi_{\z}}$ are open subsets of
2051: $M^{\phi_{i}}=\5M^{2,n-2}$ and hence that all three of them are
2052: locally CR-equivalent. Since $\5M^{2,n-2}$ is minimal as CR-manifold
2053: the endomorphisms $\phi_{\i},\phi_{\r}$ and $\phi_{\z}$ have $a$ as
2054: cyclic vector by Proposition \ruf{PR} (what also easily can be
2055: verified). By construction, the characteristic roots of the
2056: endomorphisms $\phi_{\i}$, $\phi_{\r}$ and $\phi_{\z}$ form an
2057: arithmetic progression and represent the three types with imaginary,
2058: real and zero characteristic roots. The estimate $\dim \7g_{0}\ge
2059: 4>\dim \7h^\psi$ for $\psi=\phi_{\i},\phi_r,\phi_n$ together with
2060: Propositions \ruf{UH} and \ruf{HU} implies that the characteristic
2061: roots of all 3 endomorphisms form an arithmetic progression. Summing
2062: up we have proved the following result.
2063:
2064: \Proposition{IO} Let $\phi,\phi'\in\End(V)$ be endomorphisms with cyclic
2065: vectors $a,a'\in V$. Assume that for both endomorphisms the families
2066: of characteristic roots $\alpha_{1},\dots,\alpha_{n}$ and
2067: $\alpha'_{1},\dots,\alpha'_{n}$ form an arithmetic progression. Then
2068: the germs $(M^{\phi},a)$ and $(M^{\phi'},a')$ are
2069: CR-equivalent.\Formend
2070:
2071: \medskip We can use Proposition \ruf{IO} to get explicit global
2072: equations for every $M^{\phi}$ where the characteristic roots of
2073: $\phi\in\Cyc(V)$ form an arithmetic progression: Indeed, we may take
2074: $\phi:=\xi^{1,-1}$ from \Ruf{AQ} on $\CC^{m+1}$ with coordinates
2075: $(z_{0},z_{1},\dots,z_{m})$ and $a:=(1,0,\dots,0)$. Then $\phi$ is
2076: nilpotent and
2077: $S:=^{}\exp(\RR\phi)(a)=\{(1,t,t^{2},\dots,t^{m}):t\in\RR\}$. Consequently,
2078: $M^{\phi}=F+iV$ is the tube over the cone $F$ generated by $S$ (that
2079: is $F=\RRp\!\cd S$). As a consequence, $F$ is an open piece of the
2080: algebraic surface given by the following explicit system of quadratic
2081: equations on $\RR^{m+1}$ with coordinates $(x_{0},x_{1},\dots,x_{m})$
2082: $$
2083: x_{0}x_{j+1}=x_{1}x_{j}\Steil{for}0<j<m\,.\Leqno{IP}
2084: $$
2085: This can be reformulated also in the following slightly different
2086: form: Let $C:=\{(t,t^{2},\dots,t^{n}):t\in\RR\}$ in $\RR^{n}$ be the
2087: {\sl twisted $n$-ic} (also called twisted cubic, quartic etc, see
2088: \Lit{HART} for interesting properties of these curves). Then the cone
2089: $\RRp\!\cd C$ generated by $C$ is a nonsingular surface outside the
2090: origin and the corresponding tube manifold is locally CR-equivalent to
2091: $M^{\phi}$ with $\phi$ as in \ruf{IO}. The twisted $n$-ic will also
2092: show up in another type of examples, compare \ruf{KU}.
2093:
2094: \medskip Next we extend Propositions \ruf{UH} and \ruf{HU} to the case
2095: $d=2$ where the characteristic roots of $\phi$ do form an arithmetic
2096: progression. Recall that for the light cone tube $\5M\cong\5M^{2,1}$
2097: the Lie algebra $\7g=\hol(\5M,a)$ is isomorphic to $\so(2,3)$, compare
2098: \Lit{KAZT}, \Lit{FEKA}. In particular, $\7g_{0}\cong\gl(2,\RR)$ and
2099: $\dim\7g_{1}=3$ in this case.
2100:
2101: \Proposition{ZR} Assume that the characteristic roots of
2102: $\phi\in\Cyc(V)$ form an arithmetic progression. Then for
2103: $M=M^{\phi}$, $\,n=\dim V$ and $\7g=\hol(M,a)$ the following
2104: properties hold. \0 $\7g_{0}$ is isomorphic to $\gl(2,\RR)$ and hence
2105: has dimension $4$. \1 $\7g=\aff(M,a)$ in case $n\ge4$. In particular,
2106: $\dim\7g=n+4$ in this case.
2107:
2108: \Proof ad (i): We may assume that $\lambda_{j}=(1-n)/2+(j-1)$ for
2109: all $j\in S$ using the notation in \Ruf{UK} and \Ruf{JG}. Then
2110: $\bigcap_{K\in\6K}\Delta_{K}=\{-1,0,1\}$. Solving \Ruf{PT} for
2111: $r_{0},r_{1}$ gives that all pairs
2112: $$
2113: \eqalign{r_{0}&=(n-1)(ue^{-t}+v_{0}-we^{t})\cr
2114: r_{1}&=2(ue^{-t}+v_{1}+we^{t})\,,\cr}\Leqno{AP}
2115: $$ $u,v_{0},v_{1},w\in\RR$ arbitrary, form the solution space. This
2116: implies $\dim\7g_{0}=4$. Since $M$ is locally CR-equivalent to
2117: $\5M^{2,n-2}$ the Lie algebra $\7g_{0}$ contains a copy of
2118: $\gl(2,\RR)$, that is $\7g_{0}\cong\gl(2,\RR)$.\nline ad (ii): Let
2119: $n\ge4$. We may assume that for $m=n-1$ the Lie algebra $\7g_{0}$ is
2120: the linear span of the vector fields \Ruf{AQ}. For every
2121: $\nu\in\ZZ^{2}$ let
2122: $\7g^{\nu}:=\{\xi\in\7g:[\zeta_{j},\xi]=\nu_{j}\xi\steil{for}j=1,2\}\,.$
2123: Then $\7g^{\nu}\subset\7g_{k}$ for $k=(\nu^{}_{1}+\nu^{}_{2})/m$ and
2124: $$
2125: \7g=\bigoplus_{\nu\in\ZZ^{2}}\7g^{\nu}\Steil{with}[\7g^{\nu},
2126: \7g^{\mu}]\subset\7g^{\nu+\mu}\,,
2127: $$ compare also (3.5) in \Lit{FEKA}.
2128: Clearly $i\dd{z_{k}}\in\7g^{-k,k-m}\steil{for all}0\le k\le m$.
2129: Because of Proposition \ruf{DC}.iv it is enough to show $\7g_{1}=0$.
2130: Assume to the contrary that there exists a non-zero $\xi\in\7g_{1}$.
2131: Then we may assume without loss of generality that $\xi\in\7g^{k,m-k}$
2132: for some $k\in\ZZ$. Let $c$ be the cardinality of $\{0\le j\le
2133: m:[\dd{z_{j}},\xi]\ne0\}$. From
2134: $\7g_{0}=\7g^{-1,1}\oplus\7g^{0,0}\oplus\7g^{1,-1}$ and
2135: $[i\dd{z_{j}},\xi]\in\7g^{k-j,j-k}$ we see $c\le3$. Assume $c=3$,
2136: which implies $0<k<m$. From $[\dd{z_{j}},\xi]\ne0$ for $j=k\pm1$ and
2137: the special form of $\xi^{-1,1}$, $\xi^{1,-1}$ in \Ruf{AQ} we see that
2138: $\xi$ must depend on all $n$ variables, a contradiction to $n>3$. But
2139: $c\le2$ also gives a contradiction since in all spaces $\7g^{-1,1}$,
2140: $\7g^{0,0}$, $\7g^{1,-1}$ every non-zero vector field must depend on
2141: at least $n-2$ variables.\qed
2142:
2143: \medskip Recall that $\Aut(M)$ is the group of all global
2144: CR-automorphisms and $\Aff(M)$ is the subgroup of all of affine
2145: transformations of $M$.
2146:
2147: \Proposition{GO} Let $\phi\in\End(V)$ be a trace-free cyclic
2148: endomorphism. Then the groups $\Aut(M)$ and $\Aff(M)$
2149: coincide. Furthermore, with $n=\dim V$ the following dimension
2150: estimates hold.\0 $\dim\Aut(M)=n+4$ if the characteristic roots of
2151: $\phi$ are pairwise distinct and form an arithmetic progression in
2152: $i\RR$.\1 $\dim\Aut(M)=n+3$ if $\phi$ is nilpotent.\1
2153: $\dim\Aut(M)=n+2$ in all other cases.
2154:
2155: \Proof Let $F:=M\cap V$ and denote by $\7a\subset\7g=\hol(M,a)$ the
2156: Lie algebra of $\Aut(M)$. Since $\7a$ contains the Euler vector field
2157: we have $\7a=\7a_{-1}\oplus\7a_{0}\oplus \7a_{1}$ for
2158: $\7a_{j}:=\7a\cap\7g_{j}$. We first determine $\dim\7a_{0}$. Because
2159: of $\7h:=\RR\id\oplus\RR\phi\subset\7a_{0}$ we have
2160: $\dim\7a_{0}\ge2$.\nline In case (i) $M$ is CR-equivalent to
2161: $\5M^{2,n-2}$, compare Example \ruf{JP}. Therefore $\GL(2,\RR)$ acts
2162: transitively on $F$ and $\dim\7a_{0}=4$ by Proposition \ruf{ZR}.\nline
2163: Next consider the case (ii), that is, $\phi$ is nilpotent. Then
2164: $\7a_{0}$ consists of all $\xi\in\7g_{0}\subset\End(V)$ with
2165: $\xi(c)\in\RR c$ for all $c\in\partial F:=\overline F\setminus F$,
2166: where $\overline F$ is the closure of $F$ in $V$. We may assume that
2167: $a=(1,0,\dots,0)\in\RR^{n}$ and $\phi=\xi^{1,-1}$ in the notation of
2168: \Ruf{AQ}. This implies
2169: $F=\{e^{s}(1,t,t^{2},\dots,t^{n-1})\in\RR^{n}:s,t\in\RR\}$ and hence
2170: $\partial F=\RR c$ for $c:=(0,\dots,0,1)$. Therefore, $\7a_{0}$ is the
2171: linear span of $\zeta_{1}$, $\zeta_{2}$ and $\xi^{1,-1}$ and
2172: $\dim\7a_{0}=3$ in this situation.\nline Next consider the case
2173: $V=V_{1}\oplus V_{2}\oplus\dots\oplus V_{n}$ where every $V_{j}$ is
2174: the $\big((1-n)/2+(j-1)\big)$ eigenspace of $\phi$, that is, the
2175: characteristic roots of $\phi$ form an arithmetic progression in
2176: $\RR$. Here $\partial F=V_{1}\cup V_{n}$ is easily verified. The
2177: vector fields in $\7g_{0}$ are characterized by the function tuples
2178: $(r_{0},r_{1})$ in \Ruf{AP}. The condition $\xi(V_{j})\subset V_{j}$
2179: for $j=1,n$ implies that $r_{0},r_{1}$ are constant for every
2180: $\xi\in\7a_{0}$, that is $\7a_{0}=\7h$. On the other hand, if the
2181: characteristic roots of $\phi$ do not form an arithmetic progression,
2182: then also $\7g_{0}=\7h$ by Proposition \ruf{UH}.i, that is,
2183: $\dim\7a_{0}=2$ always holds in case (iii).\nline Next we show
2184: $\7a_{1}=0$ in all cases: For $n\ge4$ this follows from $\7g_{1}=0$,
2185: see Proposition \ruf{ZR}.ii. In case (iii) we have $\7a_{0}=\7h$ and
2186: the claim follows with Lemma \ruf{UF}. Therefore we only have to
2187: consider the cases (i) and (ii) for $n=3$. In case (i) $M$ is the
2188: future light cone tube $\5M$ and $\Aut(\5M)=\Aff(\5M)$ follows as a
2189: special case of Proposition 6.9 in \Lit{KAZT}. In case (ii) $M$ is a
2190: proper domain in $\5M$: We realize $\5M=\5F+i\RR^{3}$ in
2191: $\CC^{3}$ with coordinates $(z_{0},z_{1},z_{2})$ as
2192: $\5F=\{x\in\RR^{3}:x_{0}x_{2}=x^{2}_{1}\steil{and}x_{0}+x_{2}>0\}$.
2193: Then $\hol(\5M)$ is the linear span of the vector fields (3.5) and
2194: (3.7) in \Lit{FEKA}. We may assume without loss of generality that
2195: $\phi=\xi^{-1,1}=2z_{1}\dd{z_{0}}+z_{2}\dd{z_{1}}$. This implies
2196: $\5M\setminus M=\RRp\!\cd c+i\RR^{3}$ for $c:=(1,0,0)\in\5F$.
2197: We know already that $\7a$ is the linear span of the vector fields
2198: $\zeta_{1}$, $\zeta_{1}$ and $\xi^{-1,1}$. From Figure 1 and (3.7) in
2199: \Lit{FEKA} we therefore derive that either $\7a_{1}=0$ or
2200: $\7a_{1}=\RR\xi^{0,2}$ for
2201: $\xi^{0,2}:=iz^{2}_{1}\dd{z_{0}}+iz_{1}z_{2}
2202: \dd{z_{1}}+iz^{2}_{2}\dd{z_{2}}$. The latter possibility cannot occur
2203: since $\xi^{0,2}$ is not tangent to $\5M\setminus M$: Check, for
2204: instance, the point $(1,i,0)$. This proves $\7a=\7g_{-1}\oplus\7a_{0}$
2205: in all cases. As in the proof of Proposition \ruf{ZB}.iii it is shown
2206: that this implies $\Aut(M)=\Aff(M)$. The above dimension estimates for
2207: $\7a_{0}$ imply the dimension estimates in (i) -- (iii).\qed
2208:
2209: \medskip Next we solve the local as well as the global CR-equivalence
2210: problem for all manifolds $M^{\phi}$. Because of Proposition \ruf{IO}
2211: in the local situation only the case has to be considered where the
2212: characteristic roots of $\phi$ do not form an arithmetic
2213: progression. Recall that without loss of generality we always may
2214: assume that $\phi$ is trace-free.
2215:
2216:
2217: \Proposition{KO} Let $\phi,\phi'\in\End(V)$ be trace-free cyclic
2218: endomorphisms with characteristic roots $\alpha_{1},\dots,\alpha_{n}$
2219: and $\alpha'_{1},\dots,\alpha'_{n}$ respectively. Suppose that the
2220: $\alpha_{1},\dots,\alpha_{n}$ do not form an arithmetic
2221: progression. Then for given cyclic vectors $a,a'\in V$ and
2222: corresponding $M=M^{\phi}$, $M'=M^{\phi'}$ the Lie algebras
2223: $\hol(M,a)$ and $\aff(M,a)$ coincide. Furthermore, the following
2224: conditions are equivalent. \0 The Lie algebras $\hol(M,a)$ and
2225: $\hol(M',a)$ are isomorphic . \1 The germs $(M,a)$ and $(M',a')$ are
2226: CR-equivalent. \1 $g\phi'g^{-1}=r\phi$ for suitable $g\in\GL(V)$ and
2227: $r\in\RR^{*}$. \1 There exists a permutation $\pi\in\7S_{n}$ and an
2228: $r\in\RR^{*}$ with $\alpha'_{j}=r\alpha_{\pi(j)}$ for all $j$.
2229:
2230: \Proof $\7g:=\hol(M,a)=\aff(M,a)$ and $\dim\7g=n+2$ follows from
2231: Propositions \ruf{UH} and \ruf{HU}.\nline \To12 With $\7g$ also
2232: $\7g':=\hol(M',a')$ has dimension $n+2$. Therefore
2233: also $\alpha'_{1},\dots,\alpha'_{n}$ do not form an
2234: arithmetic progression, see Proposition \ruf{ZR}. This implies
2235: $\7g_{0}=\7h$, $\7g'_{0}=\7h'$ and (ii) follows with Proposition
2236: \ruf{PQ}. \nline\To23 Let $g$ be a CR-isomorphism
2237: $(M,a)\to(M',a')$. Then $g\in\GL(V)$ as a consequence of Proposition
2238: \ruf{ZB}.iii. and clearly $g\7h g^{-1}=\7h'$. Since
2239: $\RR\phi'\subset\7h'$ is precisely the subset of all trace-free
2240: endomorphisms (iv) follows. \nline The remaining implications are easy
2241: to check and left to the reader.\qed
2242:
2243: \ifarx\eject\fi
2244:
2245: \Proposition{BP} Let $\phi,\phi'\in\End(V)$ be trace-free cyclic
2246: endomorphisms and $M:=M^{\phi}$, $M':=M^{\phi'}$. Then the following
2247: conditions are equivalent. \0 The groups $\Aff(M)$ and $\Aff(M')$ are
2248: isomorphic. \1 $M$ and $M'$ are globally CR-equivalent. \1
2249: $g\phi'g^{-1}=r\phi$ for suitable $g\in\GL(V)$ and $r\in\RR^{*}$.
2250: \Proof \To13 Suppose that (i) holds. Because of Proposition \ruf{GO}
2251: we may assume $\dim\Aut(M)=n+2$ without loss of generality. Then
2252: (iii) follows from Proposition \ruf{KO} if at least for one of the
2253: $\phi,\phi'$ the characteristic roots do not form an arithmetic
2254: progression. In the remaining cases the claim follows from Proposition
2255: \ruf{GO} since for both endomorphisms the characteristic roots form an
2256: arithmetic progression in $\RR$ and are pairwise distinct.\nline\To32
2257: $\Rightarrow$ (i) is obvious since $\Aut(M)=\Aff(M)$ and
2258: $\Aut(M')=\Aff(M')$ by Proposition \ruf{GO}.\qed
2259:
2260: \ifarx\bigskip\else\medskip\fi
2261: \Joker{MO}{Some moduli spaces} For fixed $n=\dim V$ let $\6M$ be
2262: the space of all global CR-equivalence classes $[M^{\phi}]$ of
2263: manifolds $M^{\phi}$ with $\phi\in\Cyc(V)$, that is, every
2264: $[M^{\phi}]$ is the set of all $M^{\psi}$ that are globally
2265: CR-equivalent to $M^{\phi}$. Furthermore put
2266: $$\Phi:=\{\phi\in\Cyc(V):\tr(\phi)=0\}\Steil{and}\6M^{*}:=
2267: \big\{[M^{\phi}]\in\6M: \phi\in\Phi,\,\phi^{n}\ne0\big\}\,.$$ The
2268: reductive group $\RR^{*}\Times\GL(V)$ acts on $\End(V)$ by
2269: $\phi\mapsto rg\phi g^{-1}$ for every $(r,g)\in\RR^{*}\Times\GL(V)$
2270: and leaves the cone $\Phi$ invariant. By Proposition \ruf{BP} $\6M$
2271: can be identified as set with the quotient
2272: $\Phi/(\RR^{*}\Times\GL(V))$. This quotient can be built in several
2273: steps: For every $j$ let $\sigma_{j}(\phi)\in\RR$ be the $j^{\th}$
2274: elementary symmetric function in $n$ variables evaluated on the
2275: characteristic roots of $\phi$, that is,
2276: $$X^{n}+\sum^{n}_{j=2}(-1)^{j}\sigma_{j}(\phi)X^{n-j}\;\in\;\RR[X]$$
2277: is the characteristic polynomial of $\phi\in\Phi$. Let $W:=\RR^{n-1}$
2278: with coordinates $(x_{2},\dots,x_{n})$ and denote by $\sigma:\Phi\to
2279: W$ the mapping given by
2280: $\phi\mapsto(\sigma_{2}(\phi),\dots,\sigma_{n}(\phi))$. Since every
2281: real polynomial factors in a product of linear and quadratic real
2282: polynomials, the map $\sigma$ is surjective and $\6M$ can be
2283: canonically identified as set with the quotient $W/\RR^{*}$, where
2284: $\RR^{*}$ acts on $W$ by
2285: $(x_{2},x_{3},\dots,x_{n})\mapsto(t^{2}x_{2},t^{3}x_{3},\dots,t^{n}x_{n})$
2286: for every $t\in\RR^{*}$. The subgroup $\{\pm1\}\subset\RR^{*}$ leaves
2287: the sphere $S^{n-2}=\{x\in W:\sum x_{j}^{2}=1\}$ invariant and
2288: $\6M^{*}$ can be identified with the quotient
2289: $Q^{n-2}:=S^{n-2}/\{\pm1\}$. In general, $Q^{n-2}$ can be stratified
2290: into a finite number of manifolds. For instance, $Q^{1}$ is a compact
2291: line segment and $Q^{2}$ is homeomorphic to the sphere $S^{2}$. At
2292: this point a word of caution is necessary: We do not give a topology
2293: to $\6M$, the topology on $Q^{n-2}$ only serves for the readers
2294: imagination.
2295:
2296: \ifarx\bigskip\fi
2297: Instead of $\6M$ we can also consider the space of {\sl local }
2298: CR-equivalence classes for manifolds of type $M^{\phi}$. By our
2299: results this space is of the form $\6M/\Sim\,$, where the equivalence
2300: relation $\Sim$ on $\6M$ just identifies the $3$ equivalence classes
2301: $[M^{\phi}]\in\6M$ such that the characteristic roots of $\phi$ form
2302: an arithmetic progression. Clearly, $\6M/\Sim=\6M^{*}/\Sim$ can be
2303: obtained by identifying two points in $Q^{n-2}$. In the spacial case
2304: $n=3$ the endpoints of the line segment $Q^{1}$ have to be identified,
2305: that is, $\6M$ can be thought of in this case as the circle
2306: $\overline{\RR}:=\RR\cup\infty$ (without topology) where the point
2307: $\infty$ corresponds to the class represented by the future light cone
2308: tube $\5M$. To be more specific, call in case $n=3$ for every
2309: $\phi\in\Phi$
2310: $$
2311: \mu\big(M^{\phi}\big):=-{\sigma_{2}(\phi)^{3}\over
2312: \sigma_{3}(\phi)^{2}} \;\;\in\;\;\overline{\RR}\Leqno{KW}
2313: $$
2314: the {\sl modulus} of the CR-manifold $M^{\phi}$ (with $t/0:=\infty$
2315: for all $t\in\RR$). It is clear that $M^{\phi}$ and $M^{\psi}$ in case
2316: $n=3$ are locally CR-equivalent if and only if they have the same
2317: modulus. A special meaning has the modulus $\mu_{0}:=27/4$:
2318: For real moduli $>\mu_{0}$ the endomorphism $\phi$ has $3$ distinct
2319: real eigenvalues while in case of real moduli $<\mu_{0}$ the
2320: endomorphism $\phi$ has one real and two purely imaginary eigenvalues.
2321:
2322: \ifarx\vfill\eject\fi
2323: \bigskip\medskip\Joker{KU}{Another type of examples} For $k=3$ and
2324: $c\ge1$ let $V$, $\;\Gamma=\GL(2,\RR)^{0}$ and
2325: $a=u^{m}_{2}+u_{1}u^{m-1}_{2}\in V$ be as in Example
2326: \ruf{JP}. Consider the subgroup
2327: $$
2328: \Sigma:=\Big\{\Matrix\alpha0\beta1:\alpha\in\RRp,\beta\in\RR\Big\}
2329: \;\subset\;\Gamma\,.\Leqno{LP}
2330: $$
2331: Then $F:=\Sigma(a)$ is a homogeneous surface in
2332: $V$. For $o:=u_{2}^{m}$ the orbit $C:=\Sigma(o)$ is a homogeneous
2333: curve. Identify $V$ and $\RR^{m+1}$ with coordinates
2334: $(x_{0},x_{1},\dots,x_{m})$ as in Example \ruf{JP}. Then
2335: $a=(1,1,0,\dots,0),\,o=(1,0,\dots,0)\in\RR^{m+1}$ and the Lie algebra
2336: of $\Sigma$ corresponds to the linear span of the two vector fields
2337: $\zeta_{1}$ and $\xi^{1,-1}$ in \Ruf{AQ}. In particular
2338: $$
2339: C=\{(1,t,t^{2},\dots,t^{m}):t\in\RR\}
2340: $$
2341: and $T_{o}C=\RR\cd b$ with $b:=(0,1,0,\dots,0)$ for the tangent space
2342: at $o\in C$. On the other hand, the affine {half}line $o+\RRp\!\cd b$
2343: is contained in $F$. The geometric meaning of this is the following:
2344: The development $S:=\bigcup_{c\in C}(c+T_{c}C)$ of the curve $C$ is
2345: divided by $C$ in two $\Sigma$-orbits, one of which is $F$ (compare
2346: \Lit{EAEZ}, \p45 for the special case $m=3$). Now identify the
2347: $\Sigma$-invariant hyperplane $W:=\{x\in V:x_{0}=1\}$ with $\RR^{m}$
2348: by 'dropping the coordinate' $x_{0}$. Then $C$ becomes the twisted
2349: $m$-ic $\{(t,t^{2},\dots,t^{m}):t\in\RR\}$, $o$ becomes the origin and
2350: $a$ the first basis vector $(1,0,\dots,0)$ in $\RR^{m}$. In the
2351: coordinates of $\RR^{m}$ the vector fields $\zeta_{1}$ and
2352: $\xi^{1,-1}$ are affine and have the forms
2353: $$
2354: \zeta_{1}=\sum^{m}_{j=1}jz_{j}\dd{z_{j}}\Steil{and}
2355: \xi^{1,-1}=\dd{z_{1}}+\sum^{m}_{j=2}jz_{j-1}\dd{z_{j}}\,.\Leqno{HQ}
2356: $$
2357: With Proposition \ruf{US} it is easily verified that
2358: $K^{r}_{a}F=\{x\in\RR^{m}:x_{j}=0\steil{if}j+r>2\}$ for all
2359: $r\ge0$. Since the twisted $m$-ic is not contained in any hyperplane
2360: of $\RR^{m}$ we therefore get that the tube $M:=C+i\RR^{m}$ is a
2361: homogeneous minimal $2$-nondegenerate submanifold of $\CC^{m}$ with
2362: CR-dimension $2$ and CR-codimension $m-2$. Notice that for the cone
2363: $\RRp\!\cd F$ generated by $F$ in $V$ the tube $\RRp\!\cd
2364: F+iV$ is an open piece of $\5M^{3,n-3}$, $n:=m+1$, and hence is
2365: $3$-nondegenerate.\nline Denote for every integer $j$ by $\7g^ {(j)}$
2366: the $k$-eigenspace of $\ad(\zeta_{1})$ in $\7g:=\hol(M,a)$. Then every
2367: $\xi\in\7g^{(j)}$ is a complex linear combination of monomial vector
2368: fields $z_{1}^{\nu_{1}}z_{2}^{\nu_{2}}\cdots
2369: z_{m}^{\nu_{m}}\dd{z_{p}}$ with
2370: $\nu_{1}+2\nu_{2}+\dots+m\nu_{m}=j+p$. As in the proof of Proposition
2371: \ruf{DC}, it is shown that $\7g$ has the $\ZZ$-gradation
2372: $$
2373: \7g=\bigoplus_{j\ge-m}\7g^{(j)}\,.\Leqno{RU}
2374: $$ It can be seen that $\7g$ is the linear span of all $i\dd{z_{j}}$,
2375: $1\le j\le m$, as well as $\zeta_{1}$ and $\xi^{1,-1}$. In particular,
2376: $\7g$ is a solvable Lie algebra of dimension $m+2$, coincides with
2377: $\aff(M,a)$ and has commutator subgroup of dimension $m+1$. A proof
2378: will be sketched for the special case $m=3$ in Example \ruf{EV}.
2379:
2380:
2381:
2382: \KAP{Acht}{Homogeneous 2-nondegenerate CR-manifolds in dimension 5}
2383: \medskip
2384:
2385: In this section we specialize the examples of the previous section to
2386: the case $V=\RR^{3}$. We start with manifolds of type
2387: $M=M^{\phi}=F^{\phi}+i\RR^{3}$ in $\CC^{3}$, Then the local
2388: CR-equivalence classes of these manifolds are parameterized by the
2389: modules $\mu(M)\in\overline{\RR}$, compare \Ruf{KW}. The $\phi$
2390: occurring in the following examples not necessarily are trace-free but
2391: easily could be transformed to. As defined in the previous section
2392: let $\mu_{0}:=27/4$.
2393:
2394:
2395: \Example{EI}{\fett ($\mu=\infty$) } Let
2396: $F:=\{x\in\RR^{3}:x_{1}^{2}+x_{2}^{2}=x^{2}_{3},\,x_{3}>0\}$ be the
2397: future light cone. This surface occurs as $F^{\phi}$ for
2398: $\phi:=x_{2}\dd{x_{1}}-x_{1}\dd{x_{2}}$ having spectrum $\{\pm i,0\}$.
2399:
2400:
2401: \Example{EY}{\fett ($\mu<\mu_{0}$) } For $\omega>0$ let
2402: $F\subset\RR^{3}$ be the orbit of $(1,0,1)$ under the group of all
2403: linear transformations $x\mapsto r(\cos t\,x_{1}-\sin t\,x_{2},\sin
2404: t\,x_{1}+\cos t\,x_{2},e^{\omega t}x_{3})$, $r\in\RRp,t\in\RR$.\nline
2405: With $r:=(x_{1}^{2}+x_{2}^{2})^{1/2}$, the manifold $F$ is given in
2406: $\{x\in\RR^{3}:r>0\}$ by the explicit equations
2407: $$
2408: x_{3}=r\exp\big(\omega\cos^{-1}(x_{1}/r)\big)=r\exp\big
2409: (\omega\sin^{-1}(x_{2}/r)\big)\,,
2410: $$
2411: where locally always one of these suffices. A suitable choice is
2412: $\phi=x_{1}\dd{x_{2}}-x_{2}\dd{x_{1}}+\omega x_{3}\dd{x_{3}}$ with
2413: spectrum $\{\pm i,\omega\}$.
2414:
2415: \medskip
2416:
2417: \Example{EZ}{\fett ($\mu=\mu_{0}$) } Let $F\subset\RR^{3}$ be the
2418: orbit of $(1,0,1)$ under the group of all linear transformations
2419: $x\mapsto r(x_{1},x_{2}+tx_{1},e^{t}x_{3})$ with
2420: $r\in\RRp,t\in\RR$, that is,
2421: $$
2422: F=\big\{x\in\RR^{3}:x_{1}>0,\;x_{3}^{}= x_{1}^{}\re1e^{ x_{2}/x_{1}}
2423: \big\}\,.
2424: $$
2425: Here $\phi=x_{1}\dd{x_{2}}+x_{3}\dd{x_{3}}$ has characteristic roots
2426: $0,0,1$.
2427:
2428: \medskip
2429:
2430: \Example{EX}{\fett ($\mu>\mu_{0})$ } For $\theta>2$ let
2431: $$
2432: F:=\{x\in(\RRp)^{3}:x_{3}=x_{1}(x_{2}/x_{1})^{\theta}\}\,.
2433: $$
2434: Here $\phi=x_{2}\dd{x_{2}}+\omega x_{3}\dd{x_{3}}$ has eigenvalues
2435: $\{0,1,\omega\}$.
2436:
2437: \bigskip
2438:
2439: The following is Example \ruf{KU} specialized to $m=3$.
2440:
2441: \Example{EV} Let $\Sigma$ be the group generated by the following two
2442: one-parameter groups
2443: $$
2444: x\mapsto(e^{t}x_{1},e^{2t}x_{2},e^{3t}x_{3})\,,\qquad
2445: x\mapsto(x_{1}+t,x_{2}+2tx_{1}+t^{2},x_{3}+3tx_{2}+3t^{2}x_{1}+
2446: t^{3})\Leqno{NM}
2447: $$
2448: of affine transformations on $\RR^{3}$. Then $\Sigma$ is isomorphic to
2449: the group defined in \Ruf{LP}. For $a:=(1,0,0)$ the orbit
2450: $F:=\Gamma(a)$ is
2451: $$
2452: F\;=\;\{(t,t^{2},t^{3})+r(1,2t,3t^{2})\in\RR^{3}:r\in\RRp,t\in\RR\}\,.
2453: $$
2454: The tube $M:=F+i\RR^{3}$ is an affinely homogeneous
2455: $2$-nondegenerate CR-manifold. The Lie algebra of $\Sigma$ is spanned
2456: by the affine vector fields
2457: $$
2458: \zeta_{1}:=z_{1}\dd{z_{1}}+2z_{2}\dd{z_{2}}+3z_{3}\dd{z_{3}}
2459: \Steil{and}\xi^{1,-1}:=\dd{z_{1}}+z_{1}\dd{z_{2}}+z_{2}\dd{z_{3}}
2460: \,,
2461: $$
2462: compare also \Ruf{HQ}. The Lie algebra $\7g:=\hol(M,a)$ is of finite
2463: dimension and has the gradation \Ruf{RU} for $m=3$, where $\7g^ {(k)}$
2464: is the $k$-eigenspace of $\ad(\zeta_{1})$. We claim that $\7g$ has
2465: dimension $5$ and coincides with $\aff(M,a)$. The proof consists of
2466: several elementary steps which we only sketch here. To begin with,
2467: define $\7a^{(k)}\subset\7g^{(k)}$ by
2468: $$
2469: \7a^{(-3)}:=\RR\, i\dd{z_{3}}\,,\; \7a^{(-2)}:=\RR
2470: i\dd{z_{2}}\,,\;\7a^{(-1)}:=\RR
2471: i\dd{z_{1}}\oplus\RR\phi\,,\;\7a^{(0)}:=\RR\zeta
2472: $$
2473: and $\7a^{(k)}:=0$ for all other $k$. By induction on $k$ it is seen
2474: that $\7g^{(k)}=\7a^{(k)}$ holds for all $k$: For $k\le-3$ this is
2475: obvious. For $k=-2$ suppose there exists a
2476: $\xi\in\7g^{(-2)}\setminus\7a^{(-2)}$. Then
2477: $\xi=\alpha\dd{z_{2}}+\beta z_{1}\dd{z_{3}}$ for some
2478: $\alpha,\beta\in\CC$. From $[\xi,\7a^{(-1)}]\subset\7a^{(-3)}$ we get
2479: $\beta\in\RR$ and then $\xi_{a}\in T_{a}M$ implies $\beta=0$. But then
2480: $\xi\in\7a^{(-2)}$ since $\7a^{(-2)}\!\oplus
2481: i\7a^{(-2)}\not\subset\7g^{(-2)}$, a contradiction. For $k\ge-1$ the
2482: procedure is as follows. Suppose there exists a
2483: $\xi\in\7g^{(k)}\setminus\7a^{(k)}$. Then write $\xi$ as complex
2484: linear combination of monomial vector fields as mentioned above and
2485: subtract from $\xi$ a suitable element of $\7a^{(k)}$ thus killing as
2486: many coefficients in front of monomial terms of the form
2487: $f(z)\dd{z_{1}}$ as possible. By induction hypothesis
2488: $[\7g^{(k)},\7a^{(j)}]\subset\7a^{(k+j)}$ holds for all $j<0$ and
2489: gives $\xi=0$, a contradiction. This proves the claim and also the
2490: first part of the following statement.
2491:
2492:
2493: \Lemma{LS} Let $M:=F+i\RR^{3}$ for $F=H(a)\subset\RR^{3}$ as in
2494: Example \ruf{EV}. Then $\7g=\hol(M,a)$ is a solvable Lie algebra of
2495: dimension $5$ with commutator algebra $[\7g,\7g]$ of dimension
2496: $4$. Furthermore, $\Aut(M,a)=\{\id\}$.
2497:
2498: \Proof Fix $a:=(1,0,0)\in F$ and write $\xi_j:=i\ddz{j}$ for
2499: $j=1,2,3$. Let $h\in \Aut(M,a)$ and $\Theta$ the induced Lie algebra
2500: automorphism of $\;\7l:=\7g\oplus i\7g$. With $\7g$ the subspaces
2501: $$
2502: \7n:=[\7g,\7g]=\bigoplus_{k<0}\7g^{(k)}\;,\qquad
2503: [\7n,\7n]=\RR\xi_{2}\oplus
2504: \RR\xi_{3}\Steil{and}[\7n,[\7n,\7n]]=\RR\xi_{3}
2505: $$
2506: are stable under $\Theta$ and hence also
2507: $$
2508: \7n\cap\;K_aM = \RR\xi_1\Steil{and}[\7n,\7n]\cap H_aM= \RR \xi_2\,,
2509: $$
2510: where $\7g$ and the tangent space $T_aM$ are identified via the
2511: evaluation map. In particular, $\Theta(\7P_{-1})=\7P_{-1}$ and
2512: $h(z)=g(z)+c$ for a diagonal matrix $g\in\GL(3,\RR)$ with $c=g(a)-a$,
2513: compare \Ruf{GP}. Taking commutators of $h$ with all elements in the
2514: second $1$-parameter group of \Ruf{NM} and then taking the derivative
2515: by $t$ at $t=0$ gives $c\dd z\in\7g$. From $\7g\cap\;i\7g=0$ we
2516: conclude $c=0$ and thus $g_{11}=1$ for the diagonal matrix $g$. Now
2517: $\psi$ is the unique vector field in $\7g$ that has the value $\dd z$
2518: at both points $0,a$, implying $\Theta(\psi)=\psi$. Therefore $g$ is
2519: the unit matrix and $h$ is the identity in $\Aut(M,a)$.\qed
2520:
2521:
2522: \smallskip For every $M=F+i\RR^{3}$ with $F\subset\RR^{3}$ a
2523: cone from Example \ruf{EI} -- \ruf{EX} the commutator of $\hol(M,a)$
2524: has either dimension $10$ (Example \ruf{EI}) or dimension $3$ (all the
2525: others). As a consequence of Proposition \ruf{KO} and Lemma \ruf{LS}
2526: we therefore get:
2527:
2528: \Proposition{UA} The CR-manifolds $M=F+i\RR^{3}$ with
2529: $F\subset\RR^{3}$ occurring in the examples \ruf{EI} -- \ruf{EV}, are
2530: all homogeneous and $2$-nondegenerate. Furthermore, they are mutually
2531: locally CR-inequivalent.\Formend
2532:
2533:
2534: \medskip With an argument from \Lit{FEKA} together with 2.5.10 in
2535: \Lit{HORM} a holomorphic extension property for {\sl global}
2536: continuous CR-functions $f$ on $M=F\oplus i\RR^{3}\subset\CC^{3}$, $F$
2537: one of the cones from examples \ruf{EI} -- \ruf{EV}, can be obtained:
2538: Every such $f$ has a unique continuous extension to the convex hull
2539: $\hat{M}$ of $M$ in $\CC^{3}$ that is holomorphic on the interior of
2540: $\hat{M}$ with respect $\CC^{3}$. Since $M$ is completely contained in
2541: the interior of $\hat{M}$ in case $F$ belongs to Example \ruf{EY},
2542: every global continuous CR-function on such an $M$ is real-analytic.
2543:
2544: \bigskip For the tube $\5M$ over the future light cone (that is
2545: Example \ruf{EI}) there exist many (even simply-connected) homogeneous
2546: CR-manifolds that are all locally CR-equivalent to $\5M$ but are
2547: mutually non-diffeomorphic, compare \Lit{KAZT}. In contrast to this,
2548: using already Theorem II from Section \ruf{Neun}, we can state the
2549: following global result:
2550:
2551:
2552: \Proposition{UB} Let $M$ be a homogeneous $2$-nondegenerate
2553: CR-manifold that is not locally CR-equiva\-lent to the tube $\5M$ over
2554: the future light cone. Then $M$ is simply connected and $\Aut(M)$ is a
2555: solvable Lie group of dimension 5 acting transitively and freely on
2556: $M$. For every $a\in M$ the stability group $\Aut(M,a)$ is trivial and
2557: every homogeneous real-analytic CR-manifold $M'$, that is locally
2558: CR-equivalent to $M$, is already globally CR-equivalent to $M$.
2559:
2560: \Proof By Theorem II $M=F+iV$ for $F$ as in one of the
2561: examples \ruf{EY} -- \ruf{EV}. It is easily checked that $F$ and hence
2562: $M$ is simply connected. In case $F$ is a cone, the claim follows with
2563: Lemma \ruf{KR}. Therefore we may assume that $F$ is the
2564: submanifold of Example \ruf{EV}. But then $\Aut(M,a)$ is the trivial
2565: group by Lemma \ruf{LS} and $\Aut(M)$ has trivial center by
2566: Proposition \ruf{ZI}. But then the claim follows from Proposition
2567: \ruf{UQ}.~ \qed
2568:
2569: \bigskip The affinely homogeneous surfaces $F\subset\RR^{3}$ of
2570: examples \ruf{EI} -- \ruf{EV} occur already in \Lit{EAEZ} \p43. There
2571: the surfaces are presented in their affine normal forms: Our Example
2572: \ruf{EI} ($\mu=\infty$) corresponds to P4, the Examples \ruf{EY} --
2573: \ruf{EX} ($\mu\in\RR$) to P3 and Example \ruf{EV} to P1. The remaining
2574: degenerate types in \Lit{EAEZ}, types P2$^{\pm}$ and P5, do not show
2575: up among our examples since the associated tube manifolds are
2576: holomorphically degenerate.
2577:
2578: Let us consider the type P3 in \Lit{EAEZ} \p43 a little bit closer.
2579: This is the family of local surfaces in $\RR^{3}$ given by the local
2580: equations in affine normal form
2581: $$x_{3}=x_{1}^{2}+x_{1}^{2}x_{2}+x_{1}^{2}x_{2}^{2}+x_{1}^{5}+
2582: x_{1}^{2}x_{2}^{3}+4x_{1}^{5}x_{2}+x_{1}^{2}x_{2}^{4}+ax_{1}^{7}
2583: +10x_{1}^{5}x_{2}^{2}+x_{1}^{2}x_{2}^{5}+{\rm O}(8),\Leqno{OQ}$$ where
2584: $a\in\RR$ is an arbitrary parameter and ${\rm O}(8)$ for every fixed
2585: $a$ is a convergent power series in $x=(x_{1},x_{2},x_{3})$ vanishing
2586: of order $\ge8$ at the origin and uniquely determined by the
2587: requirement, that \Ruf{OQ} defines near the origin of $\RR^{3}$ a
2588: locally affinely homogeneous surface. Different values of $a\in\RR$
2589: give locally affinely inequivalent surfaces and there tubes in
2590: $\CC^{3}$ correspond in a 1-1-way to our Examples \ruf{EY} --
2591: \ruf{EX}. It is not difficult to see that the modulus $\mu$ of every
2592: such surface is related to the parameter $a$ in \Ruf{OQ} by the
2593: formula $100\mu=(28a)^{3}$.
2594:
2595: \medskip In \Lit{DOKR}, \Lit{EAEZ} all locally affinely homogeneous
2596: surfaces in $\RR^{3}$ have been classified up to local affine
2597: equivalence. Inspecting the degenerate surfaces in these classifications
2598: gives together with our results the following
2599:
2600: \Proposition{IZ} Let $F$ be a locally affinely homogeneous surface in
2601: $\RR^{3}$ and assume that the corresponding tube $M:=F+i\RR^{3}$
2602: in $\CC^{3}$ is $2$-nondegenerate. Then \0 $M$ is locally
2603: CR-equivalent to a manifold occurring in the Examples \ruf{EI} --
2604: \ruf{EV}, and \1 for every further locally affinely homogeneous
2605: surface $F'$ in $\RR^{3}$ the corresponding tubes $M$ and
2606: $M':=F'+i\RR^{3}$ are locally CR-equivalent if and only if $F$
2607: and $F'$ are locally affinely equivalent.\Formend
2608:
2609:
2610: \Partskip
2611:
2612: \ifarx\bigskip\fi
2613:
2614: \centerline{\twelverm PART 2: The classification}
2615:
2616:
2617: \KAP{Neun}{Lie-theoretic characterization of locally homogeneous
2618: CR-manifolds}
2619:
2620: In this part of the paper we classify
2621: all homogeneous 5-dimensional 2-nondegenerate CR-manifolds up to
2622: local CR-equivalence, that is, we
2623: carry out the proof of
2624:
2625:
2626: \medskip\noindent
2627: {\bf Theorem II.} {\sl
2628: Let $M$ be a locally homogeneous $2$-nondegenerate
2629: real-analytic CR-manifold of dimension $5$. Then $M$ is locally
2630: CR-equivalent to a tube $F+ i\RR^{3}\;\subset\;\CC^{3}$, where
2631: $F\subset\RR^{3}$ is one of the affinely homogeneous surfaces
2632: occurring in the Examples \ruf{EI} -- \ruf{EV}.}
2633:
2634: \smallskip\noindent We call (in accordance with Section \ruf{Zwei}) a
2635: real-analytic CR-manifold $M$ {\sl locally homogeneous} at a point
2636: $o\in M$ if there exists a Lie subalgebra $\7g\subset\hol(M,o)$ of
2637: finite dimension such that the canonical evaluation map $\7g\to
2638: T_{o}M$ is surjective, that is, such that the tangent vectors
2639: $\xi_{o}$, $\xi\in\7g$, span the tangent space $T_{o}M$. If this is
2640: the case we also call the corresponding CR-germ $(M,o)$ {\sl locally
2641: homogeneous}. If a particular locally transitive $\7g\subset\hol(M,o)$
2642: has been fixed we also say that the germ $(M,o)$ is $\7g$-homogeneous.
2643:
2644: The proof of Theorem II relies on a natural equivalence (see
2645: \Lit{FELS}, Prop. 4.1) between the category of CR-manifold germs with
2646: a locally transitive Lie algebra action and a certain purely
2647: algebraically defined category. Before we briefly outline the main
2648: steps of our proof, we recall the notion of a CR-algebra, taken from
2649: \Lit{MENA}, and introduce some notation.
2650:
2651: \Definition{FJ} A CR-algebra is a pair $(\7g,\7q)$, where $\7g$ is a
2652: real Lie algebra of {\sl finite} dimension and $\7q$ is a complex Lie
2653: subalgebra of the complexification $\7l:=\7g\oplus i\7g$. The
2654: CR-algebra $(\7g,\7q)$ is called {\sl effective} if $0$ is the only
2655: ideal of $\7g$ contained in $\7g\cap\7q$. \Formend
2656:
2657: \Remarks{XA} \0 In \Lit{MENA} also the case is allowed where $\7g$ has
2658: infinite dimension, but $\7q$ has to have finite codimension in $\7l$.
2659: In this part of the paper however, only finite-dimensional Lie
2660: algebras occur. \1 The CR-algebras form in an obvious way a category:
2661: A morphism $(\7g,\7q)\to(\7g',\7q')$ of CR-algebras is a Lie algebra
2662: homomorphism $\7g\to\7g'$ in the usual sense whose complex linear
2663: extension $\7l\to\7l'$ maps $\7q$ to $\7q'$. Unfortunately, the
2664: resulting notion of isomorphism between CR-algebras is too strict for
2665: our purposes. We therefore mainly work with the coarser notion of
2666: geometric equivalence between CR-algebras to be introduced later. \1
2667: The geometric situation behind the notion of a CR-algebra is the
2668: following: Let $Z$ be a complex manifold homogeneous under a complex
2669: Lie group $L$, let $o\in Z$ be a point with isotropy subgroup
2670: $Q\subset L$ at $o$ and $G\subset L$ a connected real form of $L$,
2671: that is, the connected identity component of the fixed point set
2672: $L^\sigma$ for an involutive antiholomorphic automorphism $\sigma$ of
2673: $L$. Then each $G$--orbit $M$ in $Z$ is a generic (immersed)
2674: CR-submanifold. Let $\7g,\7q$ be the Lie algebras of $G$ and $Q$,
2675: respectively. Then $(\7g,\7q)$ is a CR-algebra that completely
2676: describes the CR-germ $(M,o)$ together with the local action of $G$
2677: near $o$. \1 In general, however, not every CR-algebra $(\7g,\7q)$ can
2678: be obtained from a global situation as described in (iii), only from a
2679: more general local setting. Nevertheless, the set \CRG of all
2680: CR-equivalence classes of locally homogeneous CR-germs and the set
2681: \CRA of all geometric equivalence classes of CR-algebras stand in a
2682: canonical 1-1-correspondence, compare also Section 4 in
2683: \Lit{FELS}. For convenience of the reader, we briefly describe below
2684: this correspondence in both directions. As a reference for a general
2685: discussion of `local' and `infinitesimal' actions we refer to the
2686: original paper of Palais, \Lit{PALA}.\Formend
2687:
2688: \medskip\noindent In the following let $M$ always be a locally
2689: homogeneous real-analytic CR-manifold with base point $o\in M$ that
2690: {\sl locally} can be embedded in some $\CC^{n}$ (equivalently, $M$
2691: is an (abstract) real-analytic {\sl involutive} CR-manifold as defined
2692: at the end of Section \ruf{Zwei}). Each such CR-manifold
2693: can globally and generically be embedded into a complex manifold $Z$
2694: \Lit{ANFR}. Since we are only interested in the local structure of $M$
2695: at $o$ and therefore mostly deal with CR-germs $(M,o)$, we may assume
2696: without loss of generality that $M$ is embedded in a complex vector
2697: space $E\cong \CC^n$ as a locally closed generic CR-submanifold.
2698:
2699: Next we describe the interplay between locally homogeneous CR-germs
2700: and CR-algebras more closely. In particular, we give two canonical
2701: constructions that induce the 1-1-correspondence between the sets \CRG
2702: and \CRA mentioned in \ruf{XA}.iv and also allow the precise
2703: definition of `geometric equivalence' for CR-algebras:
2704:
2705: Let $(M,o)$ be a CR-germ and $\7g\subset\hol(M,o)$ a locally
2706: transitive Lie subalgebra of finite dimension. Then an effective
2707: CR-algebra $(\7g,\7q)$ can be associated in the following way: To
2708: begin with, realize $\hol(M,o)$ in the canonical way as real Lie
2709: subalgebra of the complex Lie algebra $\hol(E,o)$. This is possible,
2710: since we assumed $M$ to be generic in $E$ and we can use Proposition
2711: 12.4.22 in \Lit{BERO}. As in Definition \ruf{FJ} we always denote by
2712: $\7l=\7g^{\CC}=\7g\oplus i\7g$ the formal complexification of
2713: $\7g$. Now $\Xi(\xi+ i\eta):=\xi+J\eta$ defines a Lie algebra
2714: homomorphism $\Xi:\7l\to\hol(E,o)$, where $J$ denotes the complex
2715: structure tensor $J:TE\to TE$. We will not make a notational
2716: distinction between the complex structures in $\7l$ or $TZ$, and write
2717: `$i$' for it. The homomorphism $\Xi$ is in general not injective (it
2718: is, if $M$ is holomorphically nondegenerate). Let $\7q\subset\7l$ be
2719: the $\Xi$-preimage of the isotropy subalgebra
2720: $\{\xi\in\hol(E,o):\xi_{o}=0\}$. Then the CR-algebra $(\7g,\7q)$ is
2721: called a {\sl CR-algebra associated} to the locally homogeneous
2722: CR-germ $(M,o)$. It is obvious that $\7g\cap\7q$ is nothing but the
2723: isotropy subalgebra $\7g_{o}=\{\xi\in\7g:\xi_{o}=0\}$ and the tangent
2724: space $T_{o}M$ can be canonically identified with $\7g/\7g_{o}$. Also
2725: the holomorphic tangent space $H_{o}M$ and the partial complex
2726: structure $J:H_oM\to H_oM$ (equivalently: the decomposition
2727: $H^{1,0}_oM\oplus H^{0,1}_oM$ of the complexification
2728: $H^{\CC}_oM=H_oM\otimes \CC$) can be read off the CR-algebra
2729: $(\7g,\7q)$: Let $\sigma$ be the conjugate linear involution of $\7l$
2730: with $\7l^{\sigma}:=\Fix(\sigma)=\7g$. Then it is easily verified that
2731: $\7H:=(\7q+\sigma\7q)^{\sigma}$ coincides with $\{\xi\in\7g:\xi_{o}\in
2732: iT_{o}M\}$, that is, $\7H/\7g_{o}$ is canonically isomorphic to
2733: $H_{o}M$ (the capital letter for $\7H$ is chosen to indicate that
2734: $\7H$ in general is not a Lie algebra, only a linear
2735: subspace). Further, $H^{0,1}_oM=\qu{\7q}{\7q\!\cap\!\sigma\7q}$ and
2736: $H^{1,0}_oM=\qu{\sigma\7q}{\7q\!\cap\! \sigma\7q}$. In \Lit{FELS} it
2737: has been shown that the geometric properties of the CR-structure of
2738: the CR-germ $(M,o)$ like minimality, $k$-nondegeneracy, holomorphic
2739: degeneracy can be read off every CR-algebra $(\7g,\7q)$ associated
2740: with $(M,o)$. The facts relevant for our classification will be
2741: discussed below.
2742:
2743: There is also a canonical way to associate a locally homogeneous
2744: CR-manifold germ $(M,o)$ to a given CR-algebra $(\7g,\7q)$ (not
2745: necessarily effective): Choose a complex Lie group $L$ with Lie
2746: algebra $\7l=\7g\oplus i\7g$ and a complex linear subspace
2747: $E\subset\7l$ with $\7l=\7q\oplus E$. Then there exist open
2748: neighbourhoods $U$ of $0\in E$, $V$ of $0\in\7q$ and $R$ of $\id\in L$
2749: such that $(u,v)\mapsto\exp(u)\exp(v)$ defines a biholomorphic mapping
2750: $\phi:U\times V\to R$. Choose an open neighbourhood $P$ of $0\in\7g$
2751: with $\exp(P)\subset R$. Then, in our particular situation, with
2752: $\pi:U\times V\to U$ being the canonical projection, the mapping
2753: $\psi:=\pi\circ\phi^{-1}\circ\exp:P\to U$ has constant rank. Without
2754: loss of generality we therefore may assume that $M:=\psi(P)$ is a
2755: connected real-analytic submanifold of $E$ containing the origin $0\in
2756: E$. The Lie algebra $\7l$ can be identified with the Lie algebra of
2757: all right-invariant vector fields on $L$ and every $\xi\in\7l$ can be
2758: projected along $\pi\circ\phi^{-1}:R\to U$ to a holomorphic vector
2759: field $\tilde\xi\in\hol(U)$. Thus the real subalgebra
2760: $\tilde\7g:=\{\tilde\xi:\xi\in\7g\}\subset\hol(U)$ is a homomorphic
2761: image of $\7g$ and spans at every $x\in M$ the tangent space
2762: $T_{x}M$. In particular, $M$ is a generic CR-submanifold of $E$ and
2763: $\tilde\7g$ is a locally transitive subalgebra. It is not difficult so
2764: see that $\tilde\7g$ is obtained from $\7g$ by factoring out the
2765: kernel of ineffectivity, more precisely, let $\7j$ be the largest
2766: ideal in $\7g$ with $\7j\subset \7g\cap\7q $. Then $\tilde\7g$ is
2767: isomorphic to $\7g/\7j$. If there exists a Lie group $L$ with Lie
2768: algebra $\7l$ such that the subalgebra $\7q$ corresponds to a {\sl
2769: closed} complex subgroup $Q\subset L$ then we may take $L/Q$ for $U$
2770: and the $G$--orbit through $[Q]\in L/Q$ for $M.$ We call $(M,o)$ the
2771: CR-{\sl germ associated} to the CR-algebra $(\7g,\7q)$.
2772:
2773: \Definition{} The CR-algebras $(\7g,\7q)$ and $(\7g',\7q')$ are called
2774: {\sl geometrically equivalent} if the associated CR-germs are
2775: CR-equivalent.
2776:
2777: Notice that CR-algebras are always geometrically equivalent if they
2778: are isomorphic in the categorical sense of \ruf{XA}.ii, but not
2779: conversely in general. Notice also that every CR-algebra is
2780: geometrically equivalent to an effective one. In the following
2781: paragraphs \ruf{FA}--\ruf{FD}, we fix
2782:
2783: \vskip5pt\centerline{\sl for the rest of the paper}
2784:
2785: \noindent the basic setup and notation, which are mainly taken from
2786: \Lit{FELS}:
2787:
2788: \Notation{FA} Given a CR-algebra $(\7g,\7q)$, let $(M,o)$ be the
2789: associated CR-germ. Write $\7l:=\7g^{\CC}=\7g\oplus i\7g$ for the
2790: complexification and $\sigma:\7l\to \7l$ for the complex conjugation
2791: with $\7l^\sigma=\7g$. Then \0 $\7g_o : =\7g\cap \7q$ is called the
2792: {\sl real isotropy subalgebra}. Define
2793: $\7l_o:=\7q^{(\infty)}:=\7q\cap\sigma\7q$ and note that $\7l_o$ is
2794: the complexification $(\7g_o)^{\CC}$ of $\7g_{o}$. \1 $\7g/\7g_{o}$
2795: and $\7H/\7g_{o}\subset\7g/\7g_{o}$ for
2796: $\7H:=(\7q+\sigma\7q)^\sigma\subset\7g$ are called the {\sl real} and
2797: the {\sl holomorphic tangent space} respectively. \1 The descending
2798: chain $\7q^{(0)}\supset\7q^{(1)}\supset\7q^{(2)}\supset\cdots\supset
2799: \7q^{(\infty)}$ of complex subalgebras is inductively defined by
2800: $\7q^{(0)}:=\7q$, $\7q^{(\infty)}:=\7q\cap \sigma\7q$ and
2801: $\7q^{(k+1)}:=\{w\in\7q^{(k)}:[w,\sigma\7q]\subset\7q^{(k)}\li1+
2802: \sigma\7q\}$, for $k\in \NN$.
2803:
2804: \noindent If $(M,o)$ is the manifold germ associated to the CR-algebra
2805: in \ruf{FA} then the holomorphic tangent space
2806: $\7H/\7g_{o}\subset\7g/\7g_{o}$ in the sense of (ii) can be
2807: canonically identified with the holomorphic tangent space $H_{o}M$ in
2808: the geometric sense. As shown in \Lit{FELS}, the mapping $\7q\to\7H$,
2809: $w\mapsto w+\sigma w$, induces a complex linear isomorphism
2810: $\7q/\7q^{(\infty)}\cong\7H/\7g_{o}\,$. The former quotient is
2811: canonically isomorphic to $H^{0,1}_oM$ (similarly, $H^{1,0}_oM\cong
2812: \sigma\7q/\7q^{(\infty)}$). The Levi kernel $K_{o}M$ and its higher
2813: order analogues $K^{r}_{o}M$ can be considered as complex linear
2814: subspaces of $\7q/\7q^{(\infty)}$. We will make extensive use of some
2815: of the main results of \Lit{FELS}, such as Theorem 5.10:
2816:
2817: \Joker{CA}{Algebraic characterization of $k$-nondegeneracy} For every
2818: $r\ge0$ the space $\7q^{(r)}$ is a Lie subalgebra of $\7q$ and the
2819: $r^{{\th}}$ Levi kernel $K^{r}_{o}M$ is isomorphic to
2820: $\7q^{(r)}/\7q^{(\infty)}$. In particular, for every $k\ge1$ the
2821: locally homogeneous CR-manifold $M$ is $k$-nondegenerate if and only
2822: if
2823:
2824: \vskip5pt\centerline{$\7q^{(k-1)}\ne\;\7q^{(k)}=\;\7q^{(\infty)}$.}\Formend
2825:
2826: \medskip\noindent To handle the $5$-dimensional case we also introduce
2827: the following abbreviations \1 $\7f:=\7q^{(1)}=\{v\in
2828: \7q:[v,\sigma\7q]\subset \7q+\sigma\7q\}$ and
2829: $\7F:=(\7f+\sigma\7f)^\sigma$.
2830:
2831: \noindent Then $\7g_o\subset\7F\subset \7H\subset\7g$ are real
2832: subspaces stable under $\ad(\7g_o)$ and $\7l_o\subset\7f\subset\7q$
2833: are complex subalgebras. In Lemma 5.9. of \Lit{FELS} it has been shown
2834: that actually $\7F$ is a Lie algebra and coincides with
2835: $N_{\7g}(\7H)\cap \7H$ (here, given $W\subset \7g$, $N_\7g(W):=\{v\in
2836: \7g:[v,W]\subset W\}$).
2837:
2838: \smallskip Summarizing the above discussion, the following result is
2839: the key for our classification.
2840:
2841: \Proposition{MC} Let $(M,o)$ be an $\7g$-homogeneous CR-germ and let
2842: $(\7g,\7q)$ be the corresponding CR-algebra. Then $M$ is
2843: 5-dimensional, minimal and 2-nondegenerate if and only if
2844: $$
2845: \codim_{\7g}(\7g_{0})=5\Steil{and}
2846: \7q\ne\7f\ne\7q^{(2)}=\7q^{(\infty)}:=\7q\cap\sigma\7q\;.
2847: $$
2848:
2849: \ifarx\eject\fi
2850:
2851: {\openup2pt\Lemma{FC} The Lie algebraic terms in \Ruf{MC} are
2852: equivalent to the following set of conditions: \o
2853: $\dim\qu{\7F}{\7g_o}=\dim \qu{\7H}{\7F}=2,\quad
2854: \dim_{\RR}\qu{\7g}{\7H}=1=
2855: \dim_{\CC}\qu{\7f}{\7l_o}=\dim_{\CC}\qu{\7q}{\7f}$
2856:
2857: \l $[\7g_o,\7F]\subset\7F,\quad[\7F,\7F]\subset\7F,\quad
2858: [\7F,\7H]\subset\7H$
2859:
2860: \l $[\7q,\sigma\7q]\not\subset\7q\,{+}\kern1pt\sigma\7q$\hfill\hbox
2861: to 7truecm{\Klein\sl $M$ is not Levi flat\hss}
2862:
2863: \l $[\7f,\sigma\7q]\,\subset \7q\,{+}\kern1pt\sigma\7q,\quad
2864: \7f\ne\7l_o$\hfill\hbox to 7truecm{\Klein\sl $M$ is Levi
2865: degenerate\hss}
2866:
2867: \l $[\7f,\sigma\7q]\,\not\subset\,\7f\,{+}\,\sigma\7q\,$\hfill\hbox to
2868: 7truecm{\Klein\sl $M$ is 2-nondegenerate.\hss}
2869: }
2870: \Formend
2871:
2872: \smallskip\noindent We will frequently use the fact that the condition
2873: `$\,[\7F,\7H]\subset \7F\,$' (instead of $\subset \7H$) violates
2874: condition \25.
2875:
2876: \medskip By the canonical bijection between the classes \CRG and \CRA
2877: mentioned in \ruf{XA}.iv and made precise above our classification
2878: problem is transferred to the classification of certain effective
2879: CR-algebras up to geometric equivalence. Unfortunately, to a given
2880: locally homogeneous CR-germ $(M,o)$ there may be associated many
2881: CR-algebras $(\7g,\7q)$ for which the $\7g$'s are non-isomorphic. For
2882: instance, if $M=\5M$ is the tube over the future light cone, then to
2883: $(\5M,o)$ there are associated CR-algebras $(\7g,\7q)$ with
2884: $\7g\cong\so(2,3)$, $\so(1,3)$, $\so(2,2)$ together with a bunch of
2885: other Lie algebras that are not semisimple, see \Lit{FEKA} for
2886: explicit realizations. Therefore, the best we can do in the following
2887: is to consider only CR-algebras $(\7g,\7q)$ such that $\dim\7g$ is
2888: minimal in the geometric equivalence class of $(\7g,\7q)$. But, also
2889: then, we are still left in the example $(\5M,o)$ with several
2890: non-isomorphic solvable Lie algebras of dimension 5 as well as one
2891: non-solvable Lie algebra of dimension $5$ with $2$-dimensional
2892: non-abelian radical and Levi part $\cong\7{sl}(2,\RR)$.
2893:
2894:
2895: \smallskip \Joker{FD}{Fundamental Assumption} In the following, every
2896: CR-algebra $(\7g,\7q)$ under consideration is assumed to satisfy the
2897: condition \Ruf{MC} (equivalently \21 -- \25\hskip1pt) as well as the
2898: following additional condition. \l For every CR-algebra $(\7g',\7q')$,
2899: which is geometrically CR-equivalent to $(\7g,\7q)$, the dimension
2900: estimate $\dim\7g'\ge\dim\7g$ holds.
2901:
2902: \smallskip\noindent Condition \26 implies, in particular, the
2903: following two conditions.
2904:
2905: \item{\26$_{1}$} $(\7g,\7q)$ is effective.
2906:
2907: \item{\26$_{2}$} There is no proper subalgebra $\7g'\subset \7g$ with
2908: $\7g'+\7g_o=\7g$ for $\7g_o=\7g\cap\7q$.
2909:
2910: \noindent Indeed, $(\7g',\7q')$ with $\7q':=(\7g'+i\7g')\cap\7q$ is a
2911: CR-algebra that is geometrically equivalent to $(\7g,\7q)$ in case
2912: $\7g'+\7g_o=\7g$.
2913:
2914: \bigskip\Joker{ST}{Basic structure theory} In the following we
2915: frequently use standard facts concerning reductive Lie algebras and
2916: parabolic subalgebras, see \Lit{KNAP}, Chap.VI-VII, as a general
2917: reference. To fix our notation, let $\7s$ be a complex reductive Lie
2918: algebra and $\7t$ a Cartan subalgebra of $\7s$. Denote by
2919: $\Phi=\Phi(\7s,\7t)\subset \7t^*$ the corresponding root system and by
2920: $\Pi\subset\Phi$ the subset of simple roots. A subalgebra
2921: $\7r\subset\7s$ is called {\sl parabolic} if it contains a maximal
2922: solvable subalgebra (also called a {\sl Borel subalgebra}) $\7b$ of
2923: $\7s$. Conjugacy classes of parabolic subalgebras in $\7s$ are
2924: parameterized by the subsets of $\Pi$: For every $\6P\subset \Pi$ and
2925: $\langle\kern-2pt\langle \6P \rangle\kern-2pt\rangle:=\Phi\cap
2926: \bigoplus_{\alpha\in \6P} \ZZ \alpha$ the corresponding parabolic
2927: subalgebra is defined by
2928: $$
2929: \7r=\7r_\6P:=\7r^{\red}\oplus
2930: \7r^{\nil}\Steil{with}\7r^{\red}:=\7t\;\oplus\!
2931: {\textstyle\bigoplus\limits_{\alpha\in\langle\!\langle
2932: \6P\rangle\!\rangle}}\7s_\alpha\Steil{and}\7r^{\nil}:=\!
2933: {\textstyle\bigoplus\limits_{\alpha\notin\langle\!\langle
2934: \6P\rangle\!\rangle}}\7s_\alpha \;\;.\Leqno{PA}
2935: $$
2936: The case of a reductive {\sl real} Lie algebra $\7s$ is a little bit
2937: more sophisticated. In contrast to the complex situation there may
2938: exist several conjugacy classes of Cartan subalgebras. Among these the
2939: class most suitable for our purposes consists of the so-called {\sl
2940: maximally split Cartan subalgebras} $\7t\subset\7s$, defined as
2941: follows: Select a Cartan decomposition $\7s=\7k\oplus\7p$ and let
2942: $\7a$ be a maximal abelian subalgebra of $\7p$. Consider the
2943: centralizer $\7m:=C_\7k(\7a)$ and put $\7t:=\7a\oplus \7t_\7m$, where
2944: $\7t_\7m\subset \7m$ is a maximal abelian subalgebra. The conjugacy
2945: classes of real parabolic subalgebras in $\7s$ are parameterized by
2946: the subsets of the simple roots $\Pi\subset \Phi(\7s,\7a)$ in the
2947: restricted root system $\Phi(\7s,\7a)\subset \7a^*$. Each parabolic
2948: subalgebra $\7r$ in $\7s$ has the decomposition
2949: $\7r=\7r^{{\red}}\ltimes \7r^{\nil}$ into the reductive and nilpotent
2950: parts (once a maximally split Cartan subalgebra $\7t\subset\7r$ is
2951: selected, the reductive factor with $\7r^{\red}\supset \7t$ is
2952: unique). Following a common convention the roots $\lambda$ occurring
2953: in the root space decomposition of the nilpotent ideal
2954: $\7r^{\nil}=\bigoplus_{\lambda\in\Lambda} \7s_\lambda$ are negative
2955: and we write $\7r^{-\n}:=\7r^{\nil}$, $\;\7r^{\r}:=\7r^{\red}$ and
2956: $\Phi^{-\n}:=\Phi(\7r^{\nil},\7a)=\Lambda$. Each parabolic subalgebra
2957: $\7r$ containing $\7t$ determines the decomposition
2958: $$
2959: \7s=\7r^{\n}\oplus \7r^{\r}\oplus \7r^{-\n}\Steil{with}
2960: \7r^{\n}:=\!\!{\textstyle\bigoplus\limits_{\lambda\in\Phi(\7r^
2961: {\nil},\7a)}}\li{12}\7s_{-\lambda}\;,\Steil{and we
2962: put}\7r^{\opp}:=\7r^{\r}\ltimes \7r^{\n}\;.\Leqno{FG}
2963: $$
2964: \vskip-6pt\noindent We call $\rk(\7s):=\dim \7t$ the {\sl rank} of
2965: $\7s$ and $\rk_{\RR}(\7s):=\dim \7a$ with $\7a\subset \7p$ as above
2966: the {\sl real rank} of $\7s.$
2967:
2968: \KAP{Zehn}{The Lie algebra $\7g$ has small semisimple part}
2969:
2970:
2971: In the preceding section we have explained how 2-nondegeneracy can be
2972: expressed in pure Lie algebraic terms. In this section we start with
2973: the actual proof of Theorem II. Since this proof will be quite
2974: involved, we subdivide it into several sections, lemmata and claims
2975: until the final step is completed in Section \ruf{Sechzehn}. For the
2976: convenience of the reader we briefly outline the main steps.
2977:
2978: As explained above the classification can be reduced to the
2979: determination of all CR-algebras satisfying the fundamental assumption
2980: \ruf{FD}. Once all possible CR-algebras are known, we have to
2981: identify the underlying CR-germs. In general, it may happen that
2982: algebraically non-equivalent CR-algebras give rise to equivalent
2983: CR-germs. For this last part of the proof we use results from Section
2984: \ruf{Acht}.
2985:
2986: Our proceeding will be to show that the assumption \ruf{FD} severely
2987: restricts the possibilities for $(\7g,\7q)$. This will be achieved by
2988: a detailed structural study of the Lie algebras $\7g$ occurring in
2989: $(\7g,\7q)$. Recall that every Lie algebra $\7h$ has a Levi-\Malcev
2990: decomposition $\7h=\7h^{\s}\ltimes\rad(\7h)$, where $\7h^{\s}$ is
2991: semisimple and is uniquely determined up to an inner automorphism of
2992: $\7h$. Furthermore, $\rad(\7h)$ is the radical of $\7h$, i.e., the
2993: unique maximal solvable ideal in $\7h$. In the first part of the proof
2994: we investigate the various possibilities for $\7g^{\s}$ where
2995: $(\7g,\7q)$ satisfies certain conditions stated in the previous
2996: section. To be precise: {\sl For the rest of the paper the fundamental
2997: assumption \ruf{FD} remains in force for {\rm all} CR-algebras
2998: $(\7g,\7q)$ under consideration}. In particular, every $(\7g,\7q)$ is
2999: effective (condition \26$_{1}$, and therefore $\7g$ can be considered
3000: as a transitive Lie subalgebra of $\hol(M,o)$). Furthermore, condition
3001: \26$_{2}$ states that there is no proper Lie subalgebra of $\7g$ that
3002: also is transitive on $(M,o)$.
3003:
3004:
3005: Let an arbitrary CR-algebra $(\7g,\7q)$ subject to \ruf{FD} be given
3006: and let $\7g^{\s}\ltimes \rad(\7g)$ be a Levi-\Malcev decomposition
3007: of $\7g$. In this and in the following few sections we assume that
3008: $\7g^{\s}\ne 0$ and investigate which simple factors can occur in
3009: $\7g^{\s}$. Thereby we use the following notation: We fix a simple
3010: ideal $\7s$ in $\7g^{\s}$ and denote by $\7s'$ the corresponding
3011: complementary ideal, i.e.,
3012: $$
3013: \7g=\7g^{\s}\ltimes \rad{\7g}\Steil{and} \7g^{\s}=\7s\times \7s'\;.
3014: \Leqno{SS}
3015: $$
3016:
3017: In this section we show that $\7g^{\s}$ only can contain simple
3018: factors isomorphic to one of the Lie algebras $\so(2,3),\;
3019: \so(1,3),\;\su(2)$ and $\7{sl}(2,\RR)$. This result is obtained by
3020: analyzing which simple real Lie algebras can contain proper
3021: subalgebras of very low codimensions. In the next sections we exclude
3022: further possibilities for $\7s$: In Section \ruf{Elf} we show that the
3023: factors $\so(2,3)$ and $\so(1,3)$ cannot occur in $\7g^{\s}$ as it
3024: will turn out that their existence in the Levi factor would violate
3025: the minimality assumption \26. Nevertheless notice that there exist
3026: CR-algebras $(\so(2,3),\7q)$ and $(\so(1,3),\7q)$ satisfying \21 --
3027: \25 and \26$_1$. All the underlying CR-germs of such CR-algebras are
3028: locally CR-equivalent to the light cone tube $\5M$. In Section
3029: \ruf{Zwoelf} we find first examples of CR-algebras which satisfy
3030: \ruf{FD}. In these cases $\7g^{\s}$ contains a simple factor $\7s\cong
3031: \7{sl}(2,\RR)$, and then necessarily $\7g\cong \7{sl}(2,\RR)\times
3032: \7r$, where $\7r$ is a 2-dimensional non-abelian Lie algebra. Also all
3033: such CR-algebras give rise only to CR-germ locally CR-equivalent to
3034: the tube over the light cone. In Section \ruf{Dreizehn} we finally
3035: eliminate the possibility $\7s\cong \su(2)$ for the simple factor
3036: $\7s$ in $\7g^{\s}$. At that stage of the proof we will have proved
3037: the following dichotomy: Let $(\7g,\7q)$ be an arbitrary CR-algebra
3038: which obeys \ruf{FD}. If $\7g^{\s}\ne 0$ then $\7g^{\s}\cong
3039: \7{sl}(2,\RR)$ and furthermore the underlying CR-germ is locally
3040: CR-equivalent to the light cone tube $\5M$. Or, $\7g^{\s}=0$, i.e.,
3041: $\7g$ is solvable.
3042:
3043: For that reason, from Section \ruf{Vierzehn} on we only consider
3044: CR-algebras $(\7g,\7q)$ with $\7g$ solvable and show that then
3045: necessarily $\dim \7g=5=\dim M$. In Section \ruf{Fuenfzehn} we look
3046: closer at the nilcenter $\7z$ of $\7g$ and find out that $\dim
3047: \7z\in \{1,3\}$. The Main Lemma \ruf{GS}, which might be of interest
3048: for itself, gives a sufficient condition for a CR-algebra to be
3049: associated to a tube $F+i\RR^{3}\subset\CC^{3}$ over an affinely
3050: homogeneous surface $F\subset\RR^{3}$. In the last section we show by
3051: ad-hoc methods that indeed every 5-dimensional solvable Lie algebra
3052: $\7g$ occurring in the CR-algebra $(\7g,\7q)$ under consideration
3053: fulfills the assumption of the Main Lemma. Hence, due to the
3054: aforementioned assertions and since (up to local affine equivalence)
3055: all affinely homogeneous surfaces in $\RR^{3}$ occur among the
3056: examples \ruf{EI} -- \ruf{EV}, this completes the proof of the
3057: classification theorem.
3058:
3059: \medskip
3060:
3061: We now begin with the proof of Theorem II:
3062:
3063:
3064:
3065: \Lemma{LA} Let $(\7g,\7q)$ be a CR-algebra subject to \Ruf{FD}.
3066: Then the simple Lie subalgebra $\7s\subset \7g^{\s}$ can only
3067: be isomorphic to
3068: $\;\so(2,3)\;,\;\so(1,3)\,,\,\7{sl}(2,\RR)\;$ or $\;\su(2)$.
3069:
3070: \Proof The proof is carried out in several reduction steps. To begin
3071: with, we write as shorthand
3072: $$
3073: \7h_o:=\7g_o\cap \7h\,,\quad\7h_\7F:=\7F\cap
3074: \7h\steil{and}\7h_\7H:=\7H\cap\7h
3075: $$
3076: for every subalgebra $\7h\subset\7g$. Notice that $\7h_o\subset
3077: \7h_{\7F}$ are subalgebras and $\7h_{\7H}$ is a linear
3078: $\ad(\7h_{\7F})$--stable subspace of $\7h.$ \nline Consider the
3079: subalgebras $\7s_o\subset \7s_\7F$ of $\7s$. The case
3080: $\7s_o=\7s_\7F=\7s$, that is $\7s\subset\7g_o$, can be ruled out since
3081: then $\7s'\oplus\rad(\7g)$ would be a proper locally transitive
3082: subalgebra of $\7g$, contradicting assumption \26$_{2}$. Therefore, at
3083: least one of the inclusions $\7s_{o}\subset \7s_\7F\subset\7s$ is
3084: proper. Consequently, there is always a proper subalgebra of
3085: codimension $\le3$ in $\7s$. Indeed, in case $\7s_\7F\ne\7s$ the
3086: subalgebra $\7s_\7F$ has this property and in case $\7s_\7F=\7s$ the
3087: proper subalgebra $\7s_{o}$ has codimension $\le2$. Hence, there
3088: exists a maximal proper subalgebra $\7h$ of $\7s$ with either
3089: $\7s_{\7F}\subset\7h$ or $\7s_o\subset \7h.$ Such a maximal subalgebra
3090: $\7h$ has codimension $\le3$ in $\7s$. Due to \Lit{BOUR}, Chap.~VIII,
3091: \S 10, Cor.~1, every maximal proper subalgebra of $\7s$ is either
3092: reductive or parabolic. In the following claims we list all simple Lie
3093: algebras $\7s$ which admit proper maximal subalgebras of such low
3094: codimensions. We discuss the reductive and parabolic case separately.
3095:
3096: \noindent{\fett Claim 1: } \sl Let $\7k$ be a simple real algebra and
3097: $\7h\subset \7k$ reductive with
3098: $0<\codim_\7s\7h\le 3$. Then $\7k$ can only be isomorphic to
3099: $\7{sl}(2,\RR),\;\su(2)$ or $\so(1,3)$.
3100:
3101: \ProofC Let $\7h=\7h_1\times \cdots \times \7h_p\times \7z$ be the
3102: decomposition of the reductive subalgebra $\7h$ into the simple
3103: factors $\7h_j$ and the center $\7z.$ Weyl's Theorem implies the
3104: existence of an $\ad(\7h)$-stable complement $\7v\subset \7s.$ Let
3105: $\rho:\7h\to \gl(\7v)$ be the induced adjoint representation. Every
3106: restriction $\rho_j:\7h_j\to \gl(\7v)$ must be faithful since
3107: otherwise $\7h_j$ would be an ideal in $\7s$. The crucial condition
3108: here is $\dim \7v\le 3$ which, in turn, implies that each $\7h_j$ is
3109: isomorphic either to $\7{sl}(2,\RR)$, $\so(3)$ or $\7{sl}(3,\RR)$. As
3110: a consequence
3111: $$
3112: \dim \7h\le \cases{8k&$r=2k$\cr8k+3&$r=2k+1$\cr}
3113: \Steil{for}r:={\rk}(\7h)\le{\rk}(\7s)\,.
3114: $$
3115: On the other hand, a
3116: glance at the classification of simple Lie algebras shows
3117: $$
3118: \dim \7s \ge \cases{2k^2+4k=\dim_{\RR} \7{sl}(k+1,\CC)& $r=2k$\cr
3119: 4k^2+8k+3=\dim\7{sl}(2k+2,\RR) &$r=2k+1\,.$}
3120: $$
3121: Putting both inequalities together and bearing in mind $\dim\7s-\dim
3122: \7h\le 3$ shows that the rank of $\7s$ can only be one of the numbers
3123: $1,2,4$. Since $\7{sl}(2,\RR)$ and $\su(2)$ are the only simple real
3124: Lie algebras of rank $1$ we may assume that $\7s$ has rank $2$ or $4$.
3125: The case $\rk(\7s)=4$ can be ruled out in the following way: Consider
3126: first the situation where $\7s$ is {\sl of complex type,} that is,
3127: $\7s$ is the underlying real Lie algebra of a complex simple Lie
3128: algebra $\7c$ of (complex) rank $2$. Then $\7c$ is either
3129: $\7{sl}(3,\CC)$ or $\so(5,\CC)$. But in both cases a proper reductive
3130: real subalgebra has at least (real) codimension 4 (in fact, 8). If
3131: $\7s$ is {\sl of real type} then the above estimates give
3132: $\dim\7h\le16$ and $\dim\7s\ge24=\dim \7{sl}(5,\RR)$. For the remaining
3133: case $\rk(\7s)=2$ either $\7s\cong \so(1,3)$, which is in the list of
3134: the claim, or $\7s$ is isomorphic to a real form of $\7{sl}(3,\CC)$ or
3135: $\so(5,\CC)$. In both cases every proper reductive complex subalgebra
3136: has at least codimension 4. This proves Claim 1.\qd
3137:
3138: \noindent{\fett Claim 2:} \sl Let $\7k$ is a simple real Lie algebra
3139: and $\,\7h\subset \7k$ parabolic with $0<\codim_\7s\7h\le 3$. Then
3140: $\7s$ is isomorphic to $\so(1,3)$, $\7{sl}(4,\RR)$, $\su(2)$ or to a
3141: non-compact real form of $\7{sl}(3,\CC) $ or $\so(5,\CC)$.
3142:
3143: \ProofC Since every parabolic subalgebra of a compact Lie algebra is
3144: trivial, apart from $\7s\cong \su(2)$ we only have to consider the
3145: case where $\7s$ is non-compact. The estimate
3146: $\rk(\7s)\le\dim\7s-\dim\7h$, compare \Ruf{ST}, implies
3147: $\rk(\7s)\le3$. We work out the various cases separately.
3148:
3149: \noindent {\fett $\rk\7s=3\,$:} The complexification $\7s^{\CC}$ of
3150: $\7s$ is one of the Lie algebras $\7{sl}(4,\CC)$, $\7{so}(7,\CC)$ or
3151: $\7{sp}(3,\CC)$. The latter two can be immediately ruled out since a
3152: glance at the corresponding Satake diagrams shows that every proper
3153: parabolic subalgebra of them is at least of codimension 4. In the
3154: remaining case $\7s^{\CC}\cong \7{sl}(4,\CC)$ only the normal real
3155: form $\7{sl}(4,\RR)$ has a parabolic subalgebra of codimension 3. (A
3156: glance at the Satake diagrams for the remaining non-compact real forms
3157: of $\7{sl}(4,\CC)$ excludes further possibilities).
3158:
3159: \noindent {\fett $\rk\7s=2\,$ or $1$:} The rank conditions imply that
3160: $\7s$ is the underlying real Lie algebra of $\7{sl}(2,\CC)$ or a real
3161: form of $\7{sl}(3,\CC)$ or $\so(5,\CC).$ The Lie algebra $\so(1,3)$ as
3162: well as every non-compact real form of $\7{sl}(3,\CC)$ or $\so(5,\CC)$
3163: contains a parabolic subalgebra of codimension less or equal 3. Since
3164: $\dim \7{sl}(2,\RR)=\dim\su(2)=3$, all simple Lie algebras of rank one
3165: also contain parabolic subalgebras of the required codimension.
3166:
3167: \noindent In order to further reduce the list of possibilities for
3168: $\7s$ we have to look at the particular real forms obtained in Claim 1
3169: and 2 but not in the list of Lemma \ruf{LA} more closely:
3170:
3171: \noindent {\fett Elimination of $\7s\cong \7{sl}(4,\RR)$:} Up
3172: to an automorphism of $\7{sl}(4,\RR)$ there is only one such parabolic
3173: subalgebra $\7h$ of codimension $3$. Let $\7t=\7a\subset \7h$ be the
3174: split Cartan subalgebra and $\7h=\7h^{\r}\ltimes \7h^{-\n}$ the
3175: decomposition as in \Ruf{FG}. Note that here the reductive factor
3176: $\7h^{\r}\cong \gl(3,\RR)$ acts irreducibly on $\7h^{\pm\n}\cong
3177: \RR^3$. The only possibility for the flag $\7s_o\subset \7s_\7F\subset
3178: \7s_\7H\subset \7s$, which cannot be trivially excluded, is when
3179: $\7s_o\subset \7s_\7F=\7h\subsetneq \7s_\7H\subsetneq \7s$ and
3180: $\codim_\7h(\7s_o)\le 2.$ This cannot be true since
3181: $\dim\qqu{\7s_\7H}{ \7s_\7F}=2$, $[\7s_\7F:\7s_\7H]\subset \7s_\7H$
3182: (condition \22) but $\7h=\7s_\7F$ acts irreducibly on the
3183: 3-dimensional space $\qqu{\7s}{\7s_\7F}\cong \7h^{\n}.$
3184:
3185:
3186: \noindent{\fett Elimination of $\7s\cong\7{su}(1,2)\,$:} In this case
3187: of real rank 1, there exists up to conjugacy only one parabolic
3188: subalgebra $\7h$: This is the minimal one $\7h=\7h_\emptyset
3189: =\7m\oplus \7a\oplus\7n$ which is solvable with $\codim_{\7s}\7h
3190: =3$. The reductive part $\7h^{\r}=\7m\oplus \7a=:\7t$ of $\7h$ is a
3191: maximally split Cartan subalgebra. As before, the only situation which
3192: cannot be trivially disposed of is when $\7s_o\subset \7s_\7F=\7h$ and
3193: $c:=\codim_\7h(\7s_o)\in \{0,1,2\}.$ Recall that $\7h$ yields the
3194: decomposition $\7s=\7h^{\n}\oplus \7t\oplus \7h^{-\n}$. If $c=0,$ that
3195: is $\7s_o=\7h$, then $\7h^{\n}\oplus\rad(\7g)$ would be a proper
3196: locally transitive subalgebra of $\7g$ in contradiction to assumption
3197: \26$_{2}$. If $c=1$ then either $\7t\cap \7s_o\ne \7t$, and then
3198: $\7h^{{\opp}}\oplus\rad(\7g)$ would be a proper locally transitive
3199: subalgebra of $\7g$ (again contradicting \26$_{2}$) or $\7t\subset
3200: \7s_o.$ But taking into account the particular structure of
3201: $\7h^{-\n}=\7s_{-\lambda}\oplus \7s_{-2\lambda}$ this also does not
3202: occur since otherwise $\7s_o\cap \7h^{-\n}$ would be a $\7t$-stable
3203: 2-dimensional subalgebra, which is impossible since $\ad(\7t)$ acts
3204: irreducibly on the 2-dimensional root spaces $\7s_{\pm\lambda}$ and
3205: $[\7s_{\lambda},\7s_{\lambda}]=\7s_{2\lambda}.$ \nline It remains the
3206: case $c=2,$ that is, $\7s=\7g\cong \su(1,2)$ is locally
3207: transitive. Then $\7l=\7s^{\CC}\cong\7{sl}(3,\CC)$ and $\7q$ is a
3208: subalgebra of complex dimension $5$. Consequently $\7q$ is contained
3209: in a maximal (6-dimensional) parabolic subalgebra
3210: $\7h\cong\7{gl}(2,\CC)\ltimes\CC^2$ of $\7l$, that is, either $\7q$
3211: coincides with a Borel subalgebra $\7b$ or is conjugate to a
3212: subalgebra $\7j\cong \7{sl}(2,\CC)\ltimes \CC^2$. In both cases the
3213: subgroups $Q$ of $L=\SL(3,\CC)$ corresponding to $\7b$ or $\7j$ are
3214: closed and the underlying CR-germ $(M,o)$ is locally CR-equivalent to
3215: an $\SU(1,2)$-orbit either in $L/B\cong\FF(\CC^3)$, the complex
3216: manifold of full flags in $\CC^3$, or in the $\CC^*$-principal bundle
3217: $L/J$ over $\PP_2(\CC)$. A direct check shows that in both cases there
3218: do not exist 2-nondegenerate $\SU(1,2)$-orbits.
3219:
3220: \noindent{\fett Elimination of $ \7s\cong\7{sl}(3,\RR)$.} Then
3221: $3\le\dim\7s_o\le6$ holds except for the trivial case $\7s_o=\7s$. If
3222: $\dim\7s_o=3$ and hence $\7s$ is locally transitive, then as in the
3223: previous situation each CR-germ $(M,o)$ associated with a CR-algebra
3224: $(\7{sl}(3,\RR),\7q)$ is locally CR-equivalent to an
3225: $\SL(3,\RR)$-orbit either in $\SL(3,\CC)/B$ or $\SL(3,\CC)/J$ (as
3226: discussed above, with $\SL(3,\RR)$ in place of $\SU(1,2)$). Again,
3227: none of these orbits is 2-nondegenerate. It remains the case $\dim
3228: \7s_o\ge 4$. But then there always exists a parabolic (proper)
3229: subalgebra $\7h\subset\7s$ with $\7s_o+\7h=\7s$, that is,
3230: $\7h\oplus\rad(\7g)$ is a proper locally transitive subalgebra of
3231: $\7g$ excluding the case $\7s^{\CC}\cong\7{sl}(3,\CC)$.
3232:
3233: \noindent{\fett Elimination of $ \7s\cong\so(1,4)$.} There exists up
3234: to conjugacy a unique parabolic subalgebra
3235: $\7h=\7h^{\r}\ltimes\7h^{-\n}\subset\7s$ of codimension $3$, and we
3236: have to investigate the cases $\7s\supsetneq \7s_\7H\supsetneq
3237: \7s_\7F=\7h\supset \7s_o$ only. A close look at the minimal (and
3238: maximal proper) parabolic subalgebra $\7h=\7m\oplus \7a\oplus
3239: \7s_{\lambda}$ shows that $\7m\cong \so(3)$ and $\7m$ acts irreducibly
3240: on the 3-dimensional nilpotent ideal $\7h^{-\n}=\7s_{-\lambda}$.
3241: Consequently $\7h$ acts irreducibly on $\7s/\7h =\7s/\7s_\7F\cong
3242: \RR^3.$ This leads to a contradiction as
3243: $[\7s_{\scriptscriptstyle\7F},\7s_{\scriptscriptstyle\7H}]\subset
3244: \7s_{\scriptscriptstyle\7H}$, i.e.,
3245: $\qqu{\7s_{\scriptscriptstyle\7H}}{\7s_{\scriptscriptstyle\7F}}$ would
3246: be a 2-dimensional stable subspace. The proof of Lemma \ruf{LA} is now
3247: complete.\qed
3248:
3249: \medskip
3250: In Theorem 6 of \Lit{BELO} it is claimed that for every
3251: $2$-nondegenerate real-analytic hypersurface $M\subset\CC^{3}$ the Lie
3252: algebra $\hol(M,a)$ has dimension $\le11$ at every point. Using this
3253: result would save few arguments in the proof of Lemma \ruf{LA}.
3254: Instead, we preferred to present a self-contained proof of the
3255: proposition.
3256:
3257:
3258: \KAP{Elf}{The cases $\7s\cong\so(2,3)$ and $\7s\cong\so(1,3)$}
3259:
3260:
3261: We continue the proof of Theorem II. So far we have proved that for
3262: the CR-algebra $(\7g,\7q)$ under consideration the simple factor $\s$
3263: of $\7g^{\s}$, see \Ruf{SS}, can only be isomorphic to one of the
3264: simple Lie algebras listed in Lemma \ruf{LA}. In this section we show
3265: that from these the possibilities $\7s\cong \so(2,3)$ and $\7s\cong
3266: \so(1,3)$ cannot occur. Here and in the following upper case roman
3267: numerals refer to the conditions around \ruf{FD}.
3268:
3269:
3270: We like to mention that for the tube $\5M$ over the future light cone
3271: $\hol(\5M,o)\cong\so(2,3)$ holds and that there exists a copy of
3272: $\so(1,3)$ in $\hol(\5M,a)$ that is also locally transitive. Since
3273: these Lie algebras have dimensions 10 and 6 and since, on the other
3274: hand, there are transitive subalgebras of $\hol(\5M,o)$ of dimension 5
3275: these two Lie algebras do not satisfy the minimality condition \26.
3276: In the following we consider both cases separately:
3277:
3278: \noindent{\fett $\7s\cong\so(2,3)$.} The only Lie subalgebras $\7h$ in
3279: the normal real form $\so(2,3)$ with $\codim_\7s\7h\le 3$ are the
3280: 3-codimensional maximal parabolic subalgebras. We need to take a
3281: closer look at the structure of these subalgebras. There are up to
3282: isomorphisms of $\7s$ only two such parabolic subalgebras: If
3283: $\Pi(\7a)=\{\alpha,\beta\}$ is a basis of the root system $\Phi(\7a)$
3284: ($\7a$ split Cartan subalgebra, $\alpha$ long, $\beta$ short) then
3285:
3286:
3287: \smallskip $\displaystyle \eqalign{\7h_1:&=(\7a\oplus \7s_\alpha\oplus
3288: \7s_{-\alpha}) \;\;\oplus \;\; \7s_{-\beta}\oplus
3289: \7s_{-\alpha-\beta}\oplus \7s_{-\alpha-2\beta}=\7h_1^{\r}\ltimes
3290: \7h_1^{-\n} \cr \7h_2:&=(\7a\oplus \7s_\beta\oplus
3291: \7s_{-\beta})\;\;\oplus\;\; \7s_{-\alpha}\oplus
3292: \7s_{-\alpha-\beta}\oplus \7s_{-\alpha-2\beta}=\7h_2^{\r}\ltimes
3293: \7h_2^{-\n}}$
3294:
3295: \smallskip\noindent are representatives of the corresponding isomorphy
3296: classes. The only instance where the filtration $\7s_o\subset
3297: \7s_{\scriptscriptstyle\7F}\subset \7s_{\scriptscriptstyle\7H}\subset
3298: \7s$ could be nontrivial (recall that $\7s_o,\;
3299: \7s_{\scriptscriptstyle\7F}$ are subalgebras,
3300: $\7s_{\scriptscriptstyle\7H}$ is a
3301: $\ad(\7s_{\scriptscriptstyle\7F})$-stable subspace) arises when
3302: $\7s_o\subset \7s_{\scriptscriptstyle\7F}=\7h_j\subset
3303: \7s_{\scriptscriptstyle\7H}\subset \7s$ for $j=1,2.$ The possibility
3304: `$\,\7h_2=\7s_{\scriptscriptstyle \7F}$' cannot occur since the
3305: adjoint representation of $\7h_2^{\r}$ on $\7h^{\n}_2\cong
3306: \qu{\7s}{\7s_{\scriptscriptstyle\7F}}$ is irreducible, contradicting
3307: the existence of a 2-dimensional
3308: $\ad(\7s_{\scriptscriptstyle\7F})$--stable subspace
3309: $\qqu{\7s_{\scriptscriptstyle\7H}}{\7s_{\scriptscriptstyle\7F}}.$ It
3310: remains the possibility ${\7s_{\scriptscriptstyle\7F}}=\7h_1\supset
3311: \7s_o$. From now on $\7h:=\7h_1$ and we analyze the various
3312: possibilities for $\dim {\7s_{\scriptscriptstyle\7F}}-\dim
3313: \7s_o\in\{0,1,2\}.$ The equation ${\7s_{\scriptscriptstyle\7F}}=\7s_o$
3314: contradicts condition \26$_{2}$ since in that case the proper
3315: subalgebra $(\7h^{\n}\oplus \7s')\ltimes \rad(\7g)$ would be
3316: transitive on $(M,o).$ The case when $\7s_o$ is of codimension 1 in
3317: ${\7s_{\scriptscriptstyle\7F}}=\7h$ can also be ruled out: Either
3318: $\7h^{-\n}\subset \7s_o$ and then $(\7h^{\opp}\oplus \7s')\ltimes
3319: \rad(\7g)$ would be a transitive proper subalgebra of $\7g$, or the
3320: intersection of $\7s_o$ with $\7h^{-\n}=(\7s_{-\beta}\oplus
3321: \7s_{-\beta-\alpha}) \oplus \7s_{-\alpha-2\beta}$ is a 2-dimensional
3322: subalgebra. In such a case the image $\pi(\7s_o)$ of the projection
3323: $\pi:\7h=\7h^{-n}\rtimes \7h^{\r}\to \7h^{\r}$ coincides with
3324: $\7h^{\r}.$ But this also leads to a contradiction: Neither the
3325: intersection $\7s_o\cap \7h^{-n}$ can coincide with
3326: $\7s_{-\beta}\oplus \7s_{-\beta-\alpha}$ (since it is not a
3327: subalgebra), nor $(\7s_{-\beta}\oplus \7s_{-\beta-\alpha})\cap \7s_o$
3328: can be 1-dimensional (since $\7h^{\r}$ acts irreducibly on
3329: $(\7s_{-\beta}\oplus \7s_{-\beta-\alpha})$). Finally we are left with
3330: the case $\dim {\7s_{\scriptscriptstyle\7F}}-\dim \7s_o=2, $ i.e.,
3331: $\7s$ is transitive on $(M,o).$ Thus $\7g=\7s$ by assumption \26. But
3332: then $\dim\7g_{o}=5$ and there always exist proper subalgebras
3333: $\7g'\subset \7g$ with $\7g'+\7g_{0}=\7g$, a contradiction to
3334: condition \26$_{2}$.
3335:
3336: \medskip \noindent{\fett $\7s\cong\so(1,3)\cong\7{sl}(2,\CC) $.} We
3337: work out this case by investigating various possibilities for $\dim
3338: \7s_o.$ Assume first that \nline $\Bullet\;\dim \7s_o\ge 3.$ We claim
3339: that then there exists a solvable subalgebra $\7r\subset \7s$ such
3340: that $\7s_o+\7r=\7s:$ The case $\dim \7s_o\ge 4$ is easily settled as
3341: all 4-dimensional subalgebras in $\7s$ are maximal, i.e., they are
3342: Borel subalgebras of $\7{sl}(2,\CC)$ and consequently have nilpotent
3343: complementary subalgebras. There do not exist (real) subalgebras of
3344: $\7{sl}(2,\CC)$ of dimension 5. If $\dim \7s_o=3$ then $\7s$ is either
3345: semisimple or solvable. In the semisimple case we work with an
3346: explicit matrix realization $\7s\subset \CC^{2\times 2}$: Either
3347: $\7s_o \cong \su(2)$ or $\cong\su(1,1),$ i.e.,
3348: $$
3349: \7s_o\cong \Big\{\!\!\pmatrixe{it & \epsilon\overline z \cr z &
3350: -it}\!\!:t\in \RR,z\in \CC
3351: \Big\}\Steil{for}\epsilon=1\steil{or}\epsilon=-1\;.
3352: $$ In both cases the upper triangular Borel subalgebra $\7b^+\subset
3353: \7{sl}(2,\CC)\subset \CC^{2\times 2}$ forms a linear complement of
3354: $\7s_o.$ This cannot happen, since then condition \26$_{2}$ would be
3355: violated. \nline If $\7s_o$ is solvable then $\7s_o$ is contained as
3356: certain 1-codimensional (real) subalgebra in a (complex) Borel
3357: subalgebra $\7b=\7t\ltimes \7b^{\nil}$ of $\7s\cong\7{sl}(2,\CC)$. We
3358: claim that $\7s_o\supset \7b^{\nil}:$ Otherwise (i.e., if
3359: $\dim^{}_{\RR }\7s_o\cap \7b^{\nil}=1$) we would have
3360: $\pi(\7s_o)=\7t$, where $\pi:\7b\to \7t$ is the projection
3361: homomorphism, and consequently $\7s_o$ would contain a certain complex
3362: Cartan subalgebra $\7t'$ of $\7b$ which acts $\RR$-irreducibly on
3363: $\7b^{\nil}$. \nline However, from our claim it follows that the
3364: opposite Borel subalgebra is a complementary subspace of $\7s_o$ in
3365: $\7s$. The case `$\dim \7s_o=3$' now is completely ruled out. Next, we
3366: investigate the case \nline {$\Bullet\;\dim \7s_o=2.$} It follows that
3367: $\7s_o$ is solvable and either complex or totally real. Since it is
3368: immediate that every complex subalgebra in $\7{sl}(2,\CC)$ has a
3369: complementary subalgebra (simple check), it remains only to deal with
3370: the totally real case, i.e., with $\7s_o$ being a real form of a Borel
3371: subalgebra $\7b\subset \7{sl}(2,\CC).$ Let $\tau:\7b\to \7b$ be the
3372: conjugation with $\7b^\tau=\7s_o.$ It is well-known that there exists
3373: a $\tau$-stable Cartan subalgebra $\7t\subset \7b$ and consequently we
3374: have the $\tau$-stable decomposition $\7b=\7t\ltimes
3375: [\7b,\7b]=:\7t\ltimes \7n.$ Select $\4h\in\7t,\;\4e\in \7n$ such that
3376: $[\4h,\4e]=2\4e.$ By construction $\tau(\4h)=a\cd \4h$ and
3377: $\tau(\4e)=b\cd \4e$ for suitable $a,b\in\CC$. Since $\tau$ is an
3378: automorphism, $a=1.$ Since $\tau$ is an involution, $|b|=1.$ This
3379: shows that up to a conjugation we may assume that $\7b$ is the
3380: upper-triangular Borel subalgebra of $\7{sl}(2,\CC)$ and the real
3381: forms $\7s_o\subset\7b^+$ have the following realizations:
3382: $$
3383: \7s_o\cong \Big\{\!\!\pmatrixn{t & sc\cr 0& -t}:t,s\in
3384: \RR\Big\}\steil{for some}c\in\CC^{*}\;.
3385: $$ For every $c\in\CC^{*}$ at least one of the Borel subalgebras
3386: $\CC\big({0\kern5pt 1 \atop 1\kern5pt 0}\big)\oplus \CC\big({1\kern2pt
3387: -1 \atop 1\kern2pt -1}\big)$ or $\CC\big({0\kern2pt -i \atop i\kern7pt
3388: 0}\big)\oplus \CC\big({1\kern7pt i \atop i\kern2pt -1} \big)$ then is
3389: complementary to $\7s_o$ in $\7{sl}(2,\CC).$ Finally, we need to deal
3390: with the case \nline $\Bullet\;\dim \7s_o= 1,$ i.e., $\7s$ itself is
3391: locally transitive and thus $\7g=\7s$ by the minimality condition
3392: \26$_{2}$. Consider the following matrix realization:\vskip-20pt
3393: $$
3394: \displaylines{\7l=\7{sl}(2,\CC)\times\7{sl}(2,\CC) \;\subset
3395: \;\CC^{2\times 2}\!\times \CC^{2\times 2} \cr
3396: \sigma(\4x,\4y)=(\overline{\4y},\overline{\4x})\,, \qquad
3397: \7g=\{(\4x,\overline{\4x}):\4x\in \7{sl}(2,\CC)\}\subset \7l\;.}
3398: $$
3399: The 3-dimensional complex subalgebra $\7q\subset\7l$ is either simple
3400: or solvable: \nline {\sl $\7q$ simple:} Then $\7q\cong \7{sl}(2,\CC)$
3401: and $\7q$ is either one of the two factors of $\7{sl}(2,\CC)\times
3402: \7{sl}(2,\CC)$ or $\7q$ is conjugate in $\7l$ to $\7g$. The first case
3403: can be ruled out immediately since then $\7q\cap \7g=0=\7g_o$ which is
3404: absurd. In the second case, all simple subalgebras $\7q\subset \7l$
3405: which are not ideals are conjugate to each other. We may select the
3406: particular subalgebra $\7q=\{(\4x,-\4x^t):\4x\in
3407: \7{sl}(2,\CC)\}\subset \7l$, where $\4x^t$ is the transpose of
3408: $\4x$. The corresponding Lie subgroup $Q=\{(\4q,(\4q^{-1})^t):\4q\in
3409: \SL(2,\CC)\}\subset L:=\SL(2,\CC)\times\SL(2,\CC)$ is closed and the
3410: map $(\4x,\4y)\mapsto \4x\cd \4y^t$ identifies the quotient with the
3411: affine quadric $\SL(2,\CC)\subset\CC^{2\Times2}$ on which
3412: $S:=\SL(2,\CC)$ (considered as real Lie group) acts by the holomorphic
3413: transformations $(\4s,\4z)\mapsto \4s\4z\overline{\4s}{}^t$. Hence, in
3414: this situation every CR-germ associated with a CR-algebra $(\7s,\7q)$
3415: is globalizable. It is well-known (see, for instance \Lit{FEKA}) that
3416: the hypersurface $S$-orbits in $L/Q$ are either Levi-nondegenerate or
3417: locally CR-equivalent to the tube $\5M$ over the light cone. However,
3418: $\dim \7g=6$ and in this case $(\7g,\7q)$ satisfies \21 -- \25,
3419: \26$_{1}$ and \26$_{2}$, but not \26.\nline {\sl $\7q$ solvable:} We
3420: will show that also this case contradicts the fundamental assumption
3421: \ruf{FD}: \nline If $\7l_o=\CC(\4t,\overline{\4t})$ is ad-semisimple
3422: (and without loss of generality $\4t=\big({t \kern1em \atop \kern.7em
3423: -t}\big )$ for some $t\in \CC^{*}$) then $\7l_o$ is a regular torus in
3424: $\7l.$ Consequently, the centralizer $C_{\7l}(\7l_o)=:\7t_1\times
3425: \7t_2$ is a $\sigma$-stable Cartan subalgebra. Either
3426: $C_{\7l}(\7l_o)\subset \7q$ (but then $\7q+\sigma\7q$ is of
3427: codimension 2 in $\7l$) or $C_{\7l}(\7l_o)\cap \7q=\7l_o.$ In the
3428: latter case denote by $\7l^{\pm\alpha_1},$ $\7l^{\pm\alpha_2}$ the
3429: root spaces with respect to $\7t_1\times \7t_2.$ A direct check shows
3430: that $\7q$ is the direct sum of $\7l_o$ with two further root spaces
3431: (also if $\overline{\4t}=\4t$). Since $\7q$ is solvable, there are 4
3432: possibilities of choosing such pairs of root spaces. In all 4 cases
3433: either $\7q+\sigma\7q$ is too small or $\7f=\7l_o,$ i.e., the
3434: corresponding CR-germ is Levi nondegenerate. \nline It remains to
3435: discuss the case when $\7l_o=\CC(\4n,\overline{\4n})$ is
3436: ad-nilpotent. Since $\7q$ is solvable, it is contained in a Borel
3437: subalgebra $\7b$ of $ \7l.$ Consequently, $\7l_o\subset
3438: \7b^{\nil}=C_\7l(\7l_o)\cong \CC\4n \times \CC\overline{\4n}.$ Let
3439: $\pi:\7b\to \7b/\7b^{\nil}$ be the canonical projection. The image
3440: $\pi(\7q)$ cannot be surjective since there is no 3-dimensional
3441: subalgebra $\CC(\4n,\overline{\4n})\subset \7q \subset \7b$ with this
3442: property. It remains only the possibility $\7b^{\nil}\subset \7q$ but
3443: then $\7q\cap\sigma\7q \supset \7b^{\nil},$ which is too big.
3444:
3445: Summarizing, we have for the CR-algebra $(\7g,\7q)$ satisfying
3446: \ruf{FD} that $\7g^{\s}$ must be a finite direct sum of copies of
3447: $\7{sl}(2,\RR)$ and $\su(2)$. In the next section we direct our
3448: attention to factors of type $\7{sl}(2,\RR)$.
3449:
3450:
3451: \KAP{Zwoelf}{The case $\7s\cong\7{sl}(2,\RR)$}
3452:
3453: In this section we continue the proof of Theorem II and consider only
3454: CR-algebras $(\7g,\7q)$ subject to \ruf{FD}, for which the simple
3455: factor $\7s$ of $\7g^{\s}$ is isomorphic to $\7{sl}(2,\RR)$, compare
3456: \Ruf{SS}. As already proved, the remaining simple factors in $\7s'$
3457: (if there are any) are isomorphic to $\su(2)$ or $\7{sl}(2,\RR)$.
3458:
3459: From $\hol(\5M,o)\cong\so(2,3)$ for the light cone tube $\5M$ it is
3460: clear that $\hol(\5M,o)$ contains copies of
3461: $\so(2,2)\cong\7{sl}(2,\RR)\Times\7{sl}(2,\RR)$. Fixing a subalgebra
3462: $\7r\subset\7{sl}(2,\RR)$ (which is necessarily non-abelian) the
3463: 5-dimensional Lie algebra $\7{sl}(2,\RR)\times\7r$ can be embedded
3464: into $\hol(\5M,o)$, and this can be done in such a way that the image
3465: is a transitive subalgebra. Therefore, the simple factor
3466: $\7{sl}(2,\RR)$ of $\7g^{\s}$ cannot avoided in the classification
3467: proof. However, we will show that this factor only occurs in the
3468: instance described above, that is, in connection with $\5M$.
3469:
3470: Our first result is that in $\7g^{\s}$ the simple factor
3471: $\7{sl}(2,\RR)$ can occur at most once. Here and in the following we
3472: repeatedly use the basic fact that each proper subalgebra of
3473: $\7{sl}(2,\RR)$ has a solvable complementary subalgebra. Furthermore,
3474: through this subsection we denote by $\pi:\7g=(\7s\times \7s')\ltimes
3475: \rad(\7g) \to \7s$ the canonical projection. We analyze various
3476: possibilities for the image of the isotropy subalgebra $\7g_{o}$ under
3477: $\pi$. \nline If $\pi(\7g_o)\ne 0$ then there exists a solvable
3478: complement $\7r\subset \7s$ to $\pi(\7g_o)$ and $(\7r\times
3479: \7s')\ltimes\rad(\7g)$ is a proper transitive subalgebra of $\7g$
3480: violating \26$_{2}$.\nline If $\pi(\7g_o)= 0$ then $\7g_o\subset
3481: \7s'\ltimes \rad(\7g)$ and is there of codimension 2. It follows that
3482: there is no factor $\7h\cong\7{sl}(2,\RR)$ in $\7s'$: Otherwise,
3483: counting dimensions we would have $\dim \7h \cap\7s_o\ge 1$ which
3484: implies that there is a proper transitive subalgebra of $\7g$,
3485: violating \26$_{2}$. This proves that $\7s'$ is a product of factors
3486: which all are isomorphic to $\su(2)$. Finally, we show that $\7s'$
3487: consists of at most {\sl one} such simple factor: Indeed, suppose that
3488: $\7s'=\7h\times\7s''$ for some ideal $\7h\cong\su(2)$ of $\7s'$.
3489: Denote by $\pi_{\7h}:=(\7s\times\7h\times\7s'')\ltimes
3490: \rad(\7g)\to\7h$ the canonical projection. Then $\pi_{\7h}(\7g_{o})$
3491: has codimension $\le2$ in $\7h$. Since $\su(2)$ does not have a
3492: subalgebra of dimension $2$, the dimension of $\pi_{\7h}(\7g_{o})$ can
3493: only be $1$ or $3$. But dimension $3$ violates \26$_{2}$ since then
3494: $\7g=\7g_{o}+\ker(\pi_{\7h})$. It remains the case
3495: $\dim\pi_{\7h}(\7g_{o})=1$. Then $\pi_{\7h}(\7g_o)$ is a torus in
3496: $\7h$ and $\pi(\7g_o)=\7h\cap \7g_o$. Since $\7g':=\7s\Times \7h$ is a
3497: transitive subalgebra of $\7g$ we must have $\7g'=\7g$ by
3498: \26$_{2}$. But this is not possible:
3499:
3500: \Lemma{LK} $\7s\cong\7{sl}(2,\RR)$ implies
3501: $\7g\;=\;\7s\!\ltimes\!\rad(\7g)$ and $\7g_o\subset\rad(\7g)\,.$
3502:
3503: \Proof By the above discussion we only have to rule out the case
3504: $\7g=\7s\times\7s'$ with $\7g_{o}\subset\7s'=\su(2)$. The key point
3505: here is that the isotropy subalgebra $\7g_o$ is toral in $\su(2)$ and
3506: that there exists a unique $\ad(\7g_o)$--stable subspace
3507: $\7p^\star\subset\su(2)$ on which the adjoint representation of
3508: $\7g_o$ is irreducible. Let $\pi_{\su}:\7s\oplus \su(2)\to \su(2)$ be
3509: the projection onto the second factor. Either $\pi_{\su}(\7F)=\su(2)$
3510: or $\pi_{\su}(\7F)=\7g_o.$ In the first case, we actually have
3511: $\7F=\su(2)$ as $[\7g_o,\7F]\subset \7F$ and $[\7s,\7g_o]=0$. But this
3512: cannot be true: Independently of what exactly $\7H=W\oplus \su(2)$
3513: would be, we always would have $[\7H,\7F]\subset \7F$. But this
3514: violates the nondegeneracy condition \25. \nline It remains the case
3515: $\pi_{\su}(\7F)=\7g_o$, i.e., $\7F=\7b\oplus \7g_o$, where
3516: $\7b=\7F\cap \7s\subset\7s$ is a 2-dimensional real subalgebra. By a
3517: dimension argument $\dim \7H\cap \su(2) \ge 2.$ But this intersection
3518: must be $\ad(\7g_o)$-stable, which implies $\7H=\7b\oplus \su(2)$,
3519: i.e., $\7H$ is a subalgebra of $\7g=\7s\oplus \su(2)$. Clearly, this
3520: violates condition \23.\qed
3521:
3522: \smallskip In the next Lemma we show that the semidirect product
3523: $\7s\ltimes\rad(\7g)$ in \ruf{LK} actually is a direct product
3524: $\7s\Times\rad(\7g)$ and that the radical has dimension 2, i.e. $\7g$
3525: has dimension 5. Recall the obvious fact that, up to isomorphy, there
3526: exist precisely two Lie algebras of dimension 2, the abelian Lie algebra
3527: $\RR^{2}=\RR\Times\RR$ and a non-abelian one.
3528:
3529: \Lemma{FM} $\7s\cong\7{sl}(2,\RR)$ implies $\7g\;=\;\7s\Times\7r$ with
3530: $\7r:=\rad(\7g)$ being non-abelian and of dimension 2.
3531:
3532: \Proof Let $\pi:\7g\to \7s$ be the canonical projection. We use the
3533: same symbol also for the complex extension $\7l\to
3534: \7s^{\CC}$. Because of \21 and $\7g_o\subset\7r$ the image
3535: $\pi(\7F)$ has dimension $\le2$.
3536:
3537: \noindent $\Bullet\dim \pi(\7F)=0$ implies $\7F=\7r,$ i.e., $\7F$
3538: is an ideal in $\7g.$ But this violates \25.
3539:
3540: \noindent $\Bullet\dim \pi(\7F)=1$ implies
3541: $\pi(\7f)=\pi(\sigma\7f)$ and $\dim\pi(\7f)=1$. Consequently
3542: $\pi(\7q),\pi(\sigma\7q)$ are 2-dimensional (Borel) subalgebras
3543: which generate $\7{sl}(2,\CC)$ as a linear space (otherwise $(M,o)$
3544: would be Levi flat). It follows $\pi^{-1}(\pi(\sigma\7q))\cap
3545: \7q=\7f.$ But this implies $[\7f,\sigma\7q]\subset \7f+\sigma\7q,$
3546: a contradiction to \25.
3547:
3548: \noindent $\Bullet\dim \pi(\7F)=2.$ Note that $\pi(\7f),\;
3549: \pi(\sigma\7f)$ are 1-dimensional since $\7l_o\subset \7r^{\CC}$.
3550: Since $\pi(\7f)+\pi(\sigma\7f)$ is a 2-dimensional subalgebra,
3551: $\7t:=\pi(\7f),$ $\7t':=\pi(\sigma\7f)=\sigma\7t$ are tori as follows
3552: with the elementary structure theory of $\7{sl}(2,\CC)$. The case
3553: `$\pi(\7f)=\pi(\7q)$' can be excluded since otherwise we would have
3554: $\pi(\7q+\sigma\7q)=\pi(\7f+\sigma\7f)$ which is a subalgebra, a
3555: contradiction to \23. Hence, the only possibility remaining is
3556: $\pi(\7f)\ne \pi(\7q)$. This implies $\7q\cap \rad= \sigma\7q\cap
3557: \rad=\7l_o$. Furthermore, $\7l_o$ is an ideal in $\7q,\sigma\7q,$ and,
3558: since $[\7q,\sigma\7q]=\7l,$ even an ideal of $\7l.$ By the
3559: effectivity assumption \26$_{1}$, this implies $\7g_o=0,$ i.e., $\7g$
3560: has dimension $5$ and thus
3561: $$\7g\cong\7{sl}(2,\RR)\ltimes_\rho\RR^2\Steil{ or }\7g\cong
3562: \7{sl}(2,\RR)\times \7r\steil{with}\dim\7r=2\,,\Leqno{EM}$$ where
3563: $\rho:\7{sl}(2,\RR)\to \End(\RR^2)$ is the canonical inclusion.
3564:
3565: \Claim{NO} The first case in \Ruf{EM}, that is
3566: $\7g\cong\7{sl}(2,\RR)\ltimes_\rho\RR^2$, cannot occur.
3567:
3568: \ProofC Let $\pi:\7l\to \7{sl}(2,\CC)$ be the canonical projection and
3569: $\sigma:\7l\to \7l$ the complex conjugation defining the real form
3570: $\7g$ of $\7l$. The possibility \nline $\Bullet\;\pi(\7q)=0\,$
3571: ($=\pi(\sigma\7q)\,$) can be excluded since then
3572: $\7q=\rad(\7l)=\sigma\7q$, violating \23. Also
3573:
3574: \noindent $\Bullet\dim \pi(\7q)=1$ can be ruled out: Then
3575: $\pi(\7q)\ne\pi(\7f)$ since otherwise
3576: $\pi(\7q+\sigma\7q)=\pi(\7f+\sigma\7f)$ would be a subalgebra,
3577: violating \23. Hence, $\7f=\7q\cap \rad(\7l)$. But this contradicts
3578: \25 since then $[\7f,\sigma\7q]\subset\rad(\7l)=\7f\oplus
3579: \sigma\7f$.
3580:
3581: \noindent $\Bullet\dim \pi(\7q)=2$ (i.e., $\7q\cap \rad(\7l)=0$) is
3582: the most involved case. Then $\pi(\7q)=\7b$ is a Borel subalgebra of
3583: $\7{sl}(2,\CC)$. To rule out also this case we need some
3584: preparations. First realize $\7l$ as \vskip-20pt
3585: $$
3586: \7{sl}(2,\CC)\ltimes_{\rho} \CC^2=\{(X,w): X\in \7{sl}(\CC^2), \;w\in
3587: \CC^2 \}\steil{with}[(X,w),(Y,u)]=([X,Y]\;, \;Xu-Yw)
3588: $$ and complex conjugation $\sigma$ given by $(X,w)\mapsto(\overline
3589: X,\overline w)$. Since $(M,o)$ is not Levi flat (compare \23), we
3590: necessarily have $\7b+\sigma\7b=\7{sl}(2,\CC),$ and the 1-dimensional
3591: intersection $\7b\cap \sigma\7b$ is a toral subalgebra
3592: $\7t\subset\7{sl}(2,\CC)$ (we use here the well-known fact that the
3593: [$\sigma$--stable] intersection of any two Borel subalgebras contains
3594: a [$\sigma$--stable] Cartan subalgebra). In the next two paragraphs we
3595: recall some elementary facts from the representation theory of
3596: $\7{sl}(2,\CC)$ which we need to complete our proof.
3597:
3598: \Joker{FN}{$\sigma$-adapted $\7{sl}_2$-triples} Let $\4h\in \7t\subset
3599: \7{sl}(2,\CC)$ be the element for which $[\4h,E]=2E$ and $[\4h,F]=-2F$
3600: for every $E\in \7b^{\nil},$ $F\in
3601: \sigma(\7b^{\nil})=(\sigma\7b)^{\nil}.$ Since $\sigma$ interchanges
3602: the eigenspaces of $\ad(\4h),$ we have $\sigma(\4h)=-\4h$. There are
3603: crucial technical points here: We claim that there exists $\4e\in
3604: \7b^{\nil}$ such that $[\4e,\sigma(\4e)]=\4h,$ i.e., $(\4e,\4h,\4f)$
3605: with $\4f:=\sigma(\4e)$ being an $\7{sl}_{2}$-triple in
3606: $\7{sl}(2,\CC)$ (i.e.,
3607: $[\4h,\4e]=2\4e,\;[\4h,\4f]=-2\4f,\;[\4e,\4f]=\4h$). \nline{\bf
3608: Construction:} We have to be careful here about signs: Since $\sigma$
3609: defines a {\sl non-compact} real form of $\7{sl}(2,\CC),$ the
3610: nondegenerate Hermitian 2-form $\kappa(\cdot\,,\sigma(\,\cdot\, ))$
3611: has precisely one negative eigenvalue, where $\kappa$ denotes the
3612: Killing form. Since $\kappa(\4h,\sigma(\4h))=\kappa(\4h,-\4h)<0$, it
3613: follows $\kappa(E,\sigma(E))>0$ and consequently $[\lambda
3614: E,\sigma(\lambda E)]=\4h$ for an appropriately chosen $\lambda\in
3615: \CC^*$ (keeping in mind the general formula $[E,F]=\kappa(E,F)H$ where
3616: $H$, the coroot, is a positive multiple of $\4h$). Define then
3617: $\4e:=\lambda E$ and $\4f:=\sigma(\4e)$. Each $\7{sl}_{2}$-triple
3618: $\4e',\4h',\4f'\in \7s^{\CC}\cong \7{sl}(2,\CC)$ with
3619: $\sigma(\4e')=\4f'$ is called $\sigma$--{\sl adapted} in the
3620: following.
3621:
3622: \Joker{FO}{\hskip-3pt} Complexifying $\7{sl}(2,\RR)\ltimes_\rho \RR^2$, we
3623: briefly discuss how the 2-dimensional abelian radical $\CC^2,$
3624: considered as an $\7{sl}(2,\CC)$-module, is related to the real
3625: structure. Let a $\sigma$-adapted $\7{sl}_2$-triple $\4e,\4h,\4f$ be
3626: given. Let $\CC^2=V_+\oplus V_-$ be the decomposition into $(\pm 1)$
3627: $\4h$-eigenspaces. They are interchanged by $\sigma$. We claim that
3628: there exists a $v_+\in V_+$ with
3629: $v_-:=\sigma(\4e)v_+=\sigma(v_+)\ne0$: Indeed, choose $w_+\in V_+\sm
3630: \{0\}$ arbitrarily. Then there is a $c\in \CC^*$ with $\4f
3631: w_+=c\sigma(w_+)$, and a direct check shows $|c|=1.$ Choose a
3632: $b\in\CC$ with $b^{2}=\overline c$ and put $v_+:=b\cd w_+$. \nline We
3633: use the linear basis $v_+,v_-$ of the radical $\rad(\7l)\cong\CC^2$ in
3634: the following computations.
3635:
3636: \noindent We now resume the proof of Claim \ruf{NO}. For short,
3637: write $\7r^{\CC} :=\rad(\7l)$
3638: for the abelian radical.
3639: Since $\7q\cap \7r^{\CC}=0$, we can write
3640: $\7q=\CC\cd(\4e,w_1)\oplus \CC\cd(\4h,w_2)$ for suitable $w_1,w_2\in
3641: \CC^2.$ Clearly, the $w_j$'s are not arbitrary: Write
3642: $w_1=\lambda_1v_+ +\mu_1v_-$ and $w_2=\lambda_2v_+ +\mu_2v_-.$ Note
3643: that $w_2\ne -\overline w_2$ (otherwise $\7q\cap\sigma\7q\ne 0$),
3644: i.e., $\lambda_2\ne -\overline \mu_2.$ The fact that $\7q$ is a
3645: subalgebra imposes a further condition on the coefficients
3646: $\lambda_j,\mu_j:$ A simple computation shows that $\mu_1=0$ and
3647: $\mu_2=-\lambda_1,$ i.e., for each $\lambda,\mu \in \CC$ with
3648: $\lambda\ne \overline \mu$ we have the two 2-dimensional complex
3649: subalgebras $\7q=\7q_{\lambda,\mu}$ and $\sigma\7q:$
3650:
3651: \noindent \hfil $\7q=\CC\cd(\4e,\lambda v_+)\oplus
3652: \CC\cd(\4h,\mu v_+-\lambda v_-),\quad ~ \sigma\7q=\CC\cd(\4f,\overline
3653: \lambda v_-)\oplus \CC\cd(-\4h,-\overline \lambda v_++\overline\mu
3654: v_-)\,. $
3655:
3656: \noindent Recall the definition $\7f=\{u\in
3657: \7q:[u,\sigma\7q]\subset\7q+\sigma\7q\}$ from \ruf{CA}.iv. A
3658: straightforward calculation shows that in all cases $\7f=0$
3659: holds. This contradicts \24 and proves \ruf{NO}. \qd
3660:
3661: \noindent We proceed the proof of Lemma \ruf{FM} by restricting the
3662: radical of $\7g\cong\7{sl}(2,\RR)\times\7r$ further:
3663:
3664: \Claim{FP} The Lie algebra $\7g$ cannot be isomorphic to
3665: $\7{sl}(2,\RR)\times \RR^2$.
3666:
3667: \ProofC Assume to the contrary that $\7g=\7{sl}(2,\RR)\times\RR^{2}$
3668: and denote by $\pi_{1}:\7l\to\7{sl}(2,\CC)$, $\pi_{2}:\7l\to\CC^{2}$
3669: the canonical projections. As in the proof of the previous claim, the
3670: cases $\dim \7q\cap \rad(\7l)\ge 1$ can be ruled out immediately. We
3671: therefore only have to exclude the case $\7q\cap\CC^{2}=0$: In that
3672: case, let $\7b$ be the 2-dimensional $\pi_{1}$-image in
3673: $\7{sl}(2,\CC)$. It follows $\sigma(\7b)\ne \7b$ since otherwise
3674: $\pi_\7s(\7q+\sigma\7q)\subset \7b$ would contradict \23. Consequently
3675: there exist $\4e\in \7b,$ $\4h\in \7b\cap \sigma\7b$ and
3676: $\4f=\sigma(\4e)\in \sigma(\7b)=\pi(\sigma\7q)$ with the properties
3677: described in \ruf{FN}. The fact that $\7q,\sigma\7q$ are subalgebras
3678: and that $\pi(\7q)=\CC\4h \oplus \CC \4e$ determine $\7q,\sigma\7q$ as
3679: follows: $\7q=\CC\cd(\4h,w)\oplus\CC\cd (\4e,0)\steil{and}
3680: \sigma\7q=\CC\cd (\4h,-\overline w) \oplus \CC\cd (\4f,0)\steil{for a}
3681: w\in \CC^2\,.$\nline We may further assume that $w,\overline w$ are
3682: linearly independent, otherwise $\pi^{}_{2}(\7q+\sigma\7q)\ne\CC^{2}$
3683: would contradict \23. As a consequence, $(\4h,0)\not\in
3684: \7q+\sigma\7q$. The definition of $\7f$ then shows $
3685: \7f=\CC\cd(\4h,w)\;$. On the other hand, for this $\7f$ we have
3686: $[\7f,\sigma\7q]\subset \7f\oplus \sigma\7q$ in contradiction to
3687: \25. This proves the claim and Lemma \ruf{FM}.\qed
3688:
3689: \medskip So far we know that under the assumption $\7s\cong
3690: \7{sl}(2,\RR)$
3691: necessarily $\7g\cong\7{sl}(2,\RR)\times\7r$, where $\7r$ is the
3692: 2-dimensional non-abelian Lie algebra. To determine the full CR-algebra
3693: $(\7g,\7q)$ we still have to find out how the complex subalgebra
3694: $\7q\subset\7l$ sits inside $\7{sl}(2,\CC)\times\7r^{\CC}$.
3695:
3696: \Claim{FQ} Up to a CR-algebra automorphism of $(\7g,\7q)$ the
3697: subalgebra $\7q\subset\7l$ is obtained in the following way: Fix a
3698: linear basis $\4x,\4z$ of $\,\7r^{\CC}$ with $[\4x,\4z]=\4z$ and a
3699: $\sigma$--adapted $\7{sl}_{2}$-triple $\4e,\4h,\4f\in\7{sl}(2,\CC)$
3700: (see \ruf{FN} for the definition). Then
3701: $\7q=\CC(\4h\,,2\4x+\mu\4z)\;\oplus\;\CC(\4e\;, \nu\4z)\;$ for
3702: suitable $\mu,\nu\in\CC$ with $\Im(\mu)=\pm2$ and $|\nu|=1$.
3703:
3704: \ProofC The case $\7q=\7r^{\CC}$ can clearly be
3705: excluded. If the intersection $\7q\cap \7r^{\CC}$ would be
3706: 1-dimensional then $\7q\cap \7r^{\CC}\;\oplus\;\sigma\7q\cap
3707: \7r^{\CC}=\7r^{\CC}$ (as the sum $\7q+\sigma\7q$ must be
3708: direct). Since $\7f$ must also be 1-dimensional, we have
3709: $\7f=\7r^{\CC}\cap \7q$. But then $[\7f,\sigma\7q]\subset
3710: \7r^{\CC}=\7f\oplus \sigma\7f$, violating condition \25. \nline It
3711: remains the case $\7q\cap \7r^{\CC}=0,$ i.e., $\pi_\7s(\7q)$ is a
3712: Borel subalgebra $\CC \4e \oplus \CC \4h,$ and
3713: $\pi(\sigma\7q)=\CC\4h\oplus \CC \4f$ where $\4h,\4e,\4f\in
3714: \7{sl}(2,\CC)$ are chosen as explained in \ruf{FN}. Since
3715: $\7q\subset \7s^{\CC}\times \7r^{\CC}$ is a Lie sub{\sl algebra}, it
3716: necessarily has the following form:
3717: $$ \diagram{\7q=\CC\cd (\4h,\lambda \4x+\mu \4z)\oplus \CC\cd (\4e,
3718: \nu \4z)\Steil{with}\sigma\7q= \CC\cd (-\4h,\overline\lambda
3719: \4x+\overline \mu \4z)\oplus \CC\cd (\4f, \overline \nu \4z) \cr
3720: \hbox{and} \quad \lambda,\mu,\nu\in \CC\Steil{satisfying}
3721: \lambda=2\steil{if}\nu\ne 0\,.}\Leqno{FR}
3722: $$ {\fett$\nu =0$:} Then $\lambda \4x+\mu \4z$ and $\overline \lambda
3723: \4x+\overline \mu \4z$ must be linearly independent in $\7r^{\CC}$,
3724: i.e., $\lambda\overline\mu\ne \overline\lambda\mu$ (otherwise
3725: $\pi_\7r(\7q{+}\sigma\7q)\ne \7r^{\CC}$). Observe that
3726: $(\4h,0)\not\in\7q+\sigma\7q$ and $\lambda\ne0$. A direct verification
3727: shows $\7f=\CC\cd(\4h,\lambda \4x+\mu \4z)$ if $\lambda\in i\RR$ and
3728: $\7f=0$ otherwise. But then $\7f\oplus \sigma\7q$ is a subalgebra in
3729: contradiction to \25.\nline {\fett $\nu\ne0$:} Possibly after
3730: replacing $\4z$ by $|\nu|\4z$ we may assume $|\nu|=1$ in
3731: \Ruf{FR}. Employing the definition of $\7f$ in \ruf{CA}.iv, a
3732: simple calculation shows
3733: $$
3734: \7f=\cases{\CC(\4h\mp2i\overline\nu\,\4e\,,\,2\4x +\Re(\mu)\4z)\quad
3735: & $\Im(\mu) =\pm 2$\cr 0 & otherwise,}\Leqno{FS}
3736: $$
3737: that is, $\7f$ is 1-dimensional only if $\Im\mu =\pm 2$. This proves
3738: the claim.\qd
3739:
3740: \Lemma{ZC} For the CR-algebra $(\7g,\7q)$ in \ruf{FQ} the associated
3741: CR-germ $(M,o)$ is CR-equivalent to $(\5M,a)$, where $\5M$ is the tube
3742: over the future light cone.
3743:
3744: \Proof We start by giving a particular representing manifold for the
3745: associated germ, compare Example 6.6 in \Lit{FEKA}. Let $\mu,\nu$ be
3746: the constants occurring in \ruf{FQ} and consider the affine quadric
3747: $Z:=\SL(2,\CC)\subset \CC^{2\times 2}$, on which the group $\widehat
3748: L:=\SL(2,\CC)\times \SL(2,\CC)$ acts holomorphically by $z\mapsto
3749: gzh^{-1}$ for all $(g,h)\in\widehat L$. Then $\widehat
3750: G:=\SL(2,\RR)\times \SL(2,\RR)$ is a real form of $\widehat L$. Via
3751: $$
3752: \eqalign{ \4x:=&{1\over 2}\pmatrix{ 0 &1 \cr 1 & 0},\quad \4z:={1\over
3753: 2}\pmatrix{1 &\mkern-10mu -1 \cr 1 &\mkern-10mu -1}\quad
3754: \in\;\7r\qquad\cr \4e:=&{\nu\over 2}\pmatrix{1 & i\cr i &
3755: \mkern-12mu -1},\quad \4h:=\pmatrix{0 & \mkern-10mu -i \cr i &
3756: 0},\quad \4f:=\sigma(\4e)\;\;\;\in\;\;\7s^{\CC} }\Leqno{VB}
3757: $$ we consider $\7l=\7s^{\CC}\times\7r^{\CC}$ as a complex subalgebra
3758: of $\widehat\7l=\7{sl}(2,\CC)\times\7{sl}(2,\CC)$ and $\7g$ as a
3759: subalgebra of $\widehat{\7g}$. \nline We recall some basic facts,
3760: compare \Lit{FEKA}: The polynomial function $\psi(z):=\det(z+\overline
3761: z)-2$ on $Z$ is invariant under the action of the group $\widehat
3762: G$. Furthermore, the nonsingular part $M$ of the algebraic subset
3763: $S:=\psi^{-1}\big(\{\pm2\}\big)$ is a (non-connected) hypersurface in
3764: $Z$, locally CR-equivalent to $\5M$. On the other hand, the singular
3765: part of $S$ is a totally real submanifold of $Z$. Consider the point
3766: $$
3767: o:={e^{-\pi i/4}\over4}\pmatrix{4+\mu & -\mu \cr i\mu
3768: &i(4-\mu)}\;\in\;Z
3769: $$
3770: which actually is in $S$ because of $\psi(o)=\Im(\mu)$. A direct
3771: computation shows that the isotropy subalgebra $\7l_{o}$ of
3772: $\7l\subset\widehat\7l$ at $o\in Z$ is $\CC(\4h,2\4x+\mu\4z)\oplus \CC
3773: (\4e,\nu\4z)$ and that the isotropy subalgebra of
3774: $\7g\subset\widehat\7g$ at $o$ is trivial. This implies $a\in M$ and
3775: also that the CR-algebra $(\7g,\7q)$ is associated to the germ
3776: $(M,o)$. Since $M$ is locally CR-equivalent to the tube $\5M$ over the
3777: future light cone the proof for \ruf{ZC} is complete.\qed
3778:
3779:
3780:
3781:
3782: \KAP{Dreizehn}{The case $\7s\cong\su(2)$}
3783:
3784:
3785: The status of our proof so far can be summarized as follows: For the
3786: CR-algebra $(\7g,\7q)$ satisfying \ruf{FD} and $\7g^{\s}\ne 0$ the
3787: simple ideal $\7s$ of $\7g^{\s}$ can only by isomorphic to
3788: $\7{sl}(2,\RR)$ or $\su(2)$, Furthermore, the case
3789: $\7s\cong\7{sl}(2,\RR)$ only occurs if the associated CR-germ is
3790: equivalent to $(\5M,o)$ with $\5M$ being the light cone tube, and then
3791: $\7g\cong \7{sl}(2,\RR)\times \7r$ with nonabelian 2-dimensional $\7r$.
3792: Consequently, only the situation needs to be investigated when
3793: $\7g^{\s}$ only contains simple ideals isomorphic to $\su(2)$. We
3794: assume this throughout this section and state as main lemma:
3795:
3796: \Lemma{VD} There is no simple factor of $\7g^{\s}$ isomorphic to
3797: $\su(2)$.\Formend
3798:
3799: \noindent The {\sl proof} will be subdivided into several claims.
3800: Note that contrary to $\7{sl}(2,\RR),$ the only non-trivial proper
3801: subalgebras of $\su(2)$ are 1-dimensional tori. As before let $\7s$
3802: be a fixed simple factor of $\7g^{\s}$ and denote by $\pi:\7g\to\7s$
3803: the canonical projection. Then the image $\pi(\7g_o)$ must be a proper
3804: subalgebra since otherwise $\7g'=\pi^{-1}(0)$ would violate
3805: \26$_{2}$. Our first observation is
3806: \Claim{} For $\7g$ and $\7r:=\rad(\7g)$ only the following cases may
3807: occur\vskip-23pt
3808: $$
3809: \displaylines{ \7g\cong\su(2)\ltimes\7r,\qquad \7g\cong\su(2)\times
3810: \su(2) \qquad\hbox{\rm or} \cr \7g\cong\big(\su(2)\times
3811: \su(2)\big)\ltimes\7r \Steil{ with }\7r\ne
3812: 0\steil{and}\7g_o=(\RR\kern1pt \4t_1 \times \RR\kern1pt \4t_2)
3813: \ltimes(\7g_o\!\cap\7r)\steil{for}\4t_{j}\in\su(2)\;.}
3814: $$
3815:
3816: \ProofC Write $\7g=(\7s_1\oplus \7s_2\oplus \7s'')\ltimes \7r$ where
3817: $\7s_1,\7s_2\cong\su(2)$, $\7s''$ is the complementary ideal in
3818: $\7g^{\s}$ and $\pi_j$ denotes the projection onto $\7s_1$ or $\7s_2$,
3819: respectively. Only one of the following possibilities could occur:
3820: \nline $\Bullet\pi_1(\7g_o)=0$. Then $\7s_1\oplus \7s_2$ is already a
3821: transitive subalgebra of $\7g$ and $\7s''=\7r=0$. \nline
3822: $\Bullet\pi_1(\7g_o)\ne 0\ne \pi_2(\7g_o)$. Then either
3823: $\7g=\7s_1\oplus\7s_2$ or $\7g_o\cap (\7s_1\oplus \7s_2)=
3824: (\7g_o\cap\7s_1)\oplus (\7g_o\cap \7s_2)= \RR \4t_1 \oplus \RR \4t_2$
3825: for suitable nonzero $\4t_1,\7t_2\in \su(2)$. In the latter case
3826: $\7g_o\cap (\7s''\ltimes \7r)$ is of codimension 1 in $\7s''\ltimes
3827: \7r$. But then we conclude that $\7s''=0$ as there is no
3828: 1-codimensional subalgebras in $\su(2)$. \qd
3829:
3830: \noindent For the proof of \ruf{VD} we only need to investigate the
3831: above three types of $\7g$. We repeatedly use the fact that each 1- or
3832: 2-dimensional representation of $\su(2)\cong\so(3)$ is trivial and
3833: that each toral subalgebra $\7t\subset \su(2)$ acts irreducibly on
3834: $\qu{\su(2)}{\7t}.$
3835:
3836:
3837: \Claim{} $\7g\cong(\su(2)\times \su(2))\ltimes\7r$ implies
3838: $\7r=0$.
3839:
3840: \ProofC Let $\pi_1,\pi_2$ be the projection onto the first and the
3841: second simple factor, respectively. As observed in the preceeding
3842: Claim, $\7g_o=(\7t_1\times \7t_2)\ltimes \7r_o$ with $\7r_o:=\7g_o\cap
3843: \7r$ and 1-dimensional toral subalgebras $\7t_j.$ Recall that we have
3844: the $\7g_o$-stable filtration $\7g\supset \7H\supset \7F\supset
3845: \7g_o.$ At least one of the images $\pi_j(\7F)$ coincides with
3846: $\7s_j,$ say, for $j=2.$ Since $\7t_j$ acts irreducibly on
3847: $\qu{\7s_j}{\7t_j},$ it follows $\7F=\7s_2\ltimes \7r_o$. Since then
3848: $\7H\cap \7s$ is at least 2-dimensional, the irreducibility of the
3849: action of $\7t_1$ implies $\7H=\7s \times \7s_2\ltimes \7r_o.$ But
3850: this would imply $[\7H,\7H]\subset \7H,$ contradicting \23.\qd
3851:
3852: \Claim{} $\7g\not\cong\su(2)\times\su(2)$.
3853:
3854: \noindent\rm {\sl Proof of the Claim.} Suppose to the contrary that
3855: $\7g=\su(2)\times\su(2)$. Since $\7F\subset\7g$ is a subalgebra of
3856: dimension 3 and there is no solvable subalgebra of this dimension,
3857: necessarily $\7F\cong \su(2)$. Consequently, either $\7F$ is one of
3858: the simple factors of $\7g$ or $\7F$ is the graph of an automorphism
3859: of $\su(2)$. Both possibilities lead to a contradiction: In the first
3860: case $\7F$ is an ideal in $\7g,$ contradicting \25, and in the second
3861: case $\7F$ acts irreducibly on $\qu{\7g}{\7F}$, violating the
3862: existence of the $\ad(\7F)$-stable proper subspace $
3863: \qu{\7H}{\7F}\subset\qu{\7g}{\7F}$.\qd
3864:
3865: \noindent The remaining case $\7g=\su(2)\ltimes\7r$ with
3866: $\7r:=\rad(\7g)$ is the most involved. In this situation the
3867: projection of $\7g_o$ under the canonical projection
3868: $\pi:\7g\to\su(2)$ is of dimension $\le1$. As usual we denote the
3869: canonical projection $\7l\to\7{sl}(2,\CC)$ by the same symbol
3870: $\pi$. We investigate the various possibilities for the $\pi$-images
3871: of the Lie subalgebras $\7l_o\subset\7f\subset\7q$ defined in
3872: \Ruf{FA}.\nline $\Bullet\pi(\7l_o)=\pi(\7f)=\pi(\7q)$. Since
3873: $\pi(\7l_o)$ is a $\sigma$-stable torus, it would follow
3874: $\pi(\7q+\sigma\7q)=\pi(\7l_o).$ Counting dimensions, this cannot
3875: happen. \nline $\Bullet\pi(\7l_o)\subsetneq\pi(\7f)=\pi(\7q)$. Here,
3876: we have to rule out the 2 possibilities $\dim \pi(\7l_o)\le1$. \nline
3877: If $\dim\pi(\7l_o)=0$ then $\pi(\7F)$ is a 1-dimensional toral
3878: subalgebra in $\su(2),$ and consequently
3879: $\pi(\7f)=\pi(\sigma\7f)=\pi(\7q)=\pi(\sigma\7q)$ is
3880: 1-dimensional. This contradicts the fact that $\7q+\sigma\7q$ is a
3881: hyperplane in $\7l.$ It remains the case when $\pi(\7l_o)$ is
3882: 1-dimensional. Then $\pi(\7F)=\7s,$ and possibly after replacing
3883: $\7s\cong \su(2)$ by the Levi factor contained in $\7F$ we may assume
3884: that $\7s\subset \7F.$ By dimension reasons $\7r_o:=\7g_o\cap
3885: \7r=\7F\cap \7r$. Further, $\7r_o$ is an ideal in $\7F$ and we have
3886: $\7F=\7s\ltimes \7r_o$ as well as $\7H=\7s\ltimes \7r_\7H$ with
3887: $\7r_\7H:=\7H\cap \7r.$ Since $\dim \7r_\7H-\dim \7r_o=2,$ the
3888: ad-representation of $\7s$ on $\qu{\7r_\7H}{\7r_o}$ is trivial, i.e.,
3889: $[\7s,\7r_\7H]\subset \7r_o.$ Since $\7f=\7l_o\oplus \CC \4y$ for some
3890: $\4y\in \7s^{\CC}\cong\7{sl}(2,\CC)$ and $\sigma\7q=\sigma\7f\oplus
3891: \CC\overline{\4x}$ for some $\overline{\4x}\in \7r^{\CC}_\7H,$ we
3892: would have $[\7f,\sigma\7q]=[\7l_o\oplus \CC \4y,\sigma\7f\oplus \CC
3893: \overline{\4x}]\subset \7f+\sigma\7f,$ in contradiction to \25. \nline
3894: $\Bullet\pi(\7l_o)=\pi(\7f)\subsetneq\pi(\7q).$ The possibility
3895: $0=\pi(\7f)$ can be excluded since otherwise $\7F=\7r$ and
3896: $\7f+\sigma\7f$ would be an ideal in $\7l,$ contradicting condition
3897: \25. Next we deal with the case when
3898: $\pi(\7l_o)=\pi(\7f)=\pi(\sigma\7f)$ is a (1-dimensional) toral
3899: subalgebra. By assumption $\7b:=\pi(\7q)$ is then 2-dimensional, i.e.,
3900: a Borel subalgebra, which implies that $\7b+\sigma{\7b}=\7s^{\CC}\cong
3901: \7{sl}(2,\CC),$ or equivalently $\pi(\7H)=\7s.$ Counting dimensions,
3902: $\7r_\7F:=\7F\cap \7r=\7H\cap \7r$. Further, $[\7r_\7F,\7H]\subset
3903: \7r\cap \7H=\7r_\7F$ and since, due to condition \23, $\7H$ generates
3904: $\7g$ as a Lie algebra, we deduce that $\7r_\7F$ is an ideal in $\7g.$
3905: From $\pi(\7l_o)=\pi(\7f)$ follows that $\7f=\7l_o\oplus \CC\4r$ with
3906: $\4r\in \7r^{\CC}_\7F.$ But this implies $[\7f,\sigma\7q]=
3907: [\7l_0\oplus \CC\4r,\sigma \7q]\subset \sigma\7q+\7r^{\CC}_\7F\subset
3908: \sigma\7q\oplus \7f,$ violating \25. It remains one last possibility
3909: for the flag $\7l_o\subset \7f\subset \7q$ in $\7l$:
3910:
3911: \Claim{} If $\pi(\7l_o)\subsetneq\pi(\7f)\subsetneq\pi(\7q)$ then
3912: $\7l_o=0$ and $\7g\cong \su(2)\times\7r$ with $\7r:=\rad(\7g)$ of
3913: dimension 2.
3914:
3915: \ProofC The properness of the inclusions implies $\7l_o\cap
3916: \7r^{\CC}=\7f\cap\7r^{\CC}=\7q\cap \7r^{\CC}=\sigma\7q\cap \7r^{\CC}$.
3917: Since $\7q$ and $\sigma \7q$ generate $\7l$ as a Lie algebra, it
3918: follows that $\7l_o\cap \7r^{\CC}$ is an ideal in $\7l,$ or
3919: equivalently $\7g_o\cap \7r\vartriangleleft \7g.$ By the effectivity
3920: assumption \26$_{1}$, we have $\7g_o\cap \7r=0.$ Consequently, since
3921: $\dim\pi(\7l_o)\le1$ the same estimate holds for $\dim\7l_o$. Next we
3922: show that the case $1=\dim_{\RR}\pi(\7g_o)\,$ ($\,=\dim
3923: \pi(\7l_o)=\dim \7l_o$) cannot happen: Assume on the contrary that
3924: $\pi(\7g_o)=1$. Then $\pi(\7F)=\7s$ and, possibly after replacing the
3925: Levi factor $\7s\subset \7g$ by a conjugate one, we may assume
3926: $\7s\subset \7F.$ Counting dimensions yields then $\7F=\7s\cong
3927: \su(2).$ Since $[\7F,\7H]\subset \7H$ by \22, we have a representation
3928: of $\7s$ on $\qu{\7H}{\7F}$. But this yields the contradiction as
3929: $\dim \qu{\7H}{\7F}=2$, compare \21, implies that this representation
3930: is trivial, i.e. $[\7s,\7H]=[\7F,\7H]\subset \7F$, violating
3931: \25. \nline We have proved that the only possibility for $\7g$ is
3932: $\su(2)\ltimes\7r$ with a 2-dimensional radical $\7r.$ Since $\su(2)$
3933: can act only trivially on such an $\7r$, the above semidirect product
3934: is in fact direct. This proves the claim.\qd
3935:
3936: \Claim{} The case $\7g\cong\su(2)\times \7r$, $\dim \7r=2$, also
3937: cannot occur.
3938:
3939: \ProofC We write $\pi=\pi_\7s$ and $\pi_\7r$ for the projections onto
3940: $\7s^{\CC}\cong \7{sl}(2,\CC)$ and $\7r^{\CC},$ respectively. The
3941: proof analyses various possibilities for $\7f\subset \7q.$ \nline If
3942: $\pi_\7s(\7F)=0$ then $\7F=\7r$ would be an ideal in $\7g$
3943: contradicting \25. It remains the case when $\pi_\7s(\7F)$ as well as
3944: $\pi_\7s(\7f)$ is 1-dimensional. In that situation necessarily
3945: $\pi_\7s(\7q)\ne \pi_\7s (\7f),$ and $\pi(\7q)$ is a Borel subalgebra
3946: in $\7{sl}(2,\CC).$ Further,
3947: $\pi_\7s(\7q)+\sigma(\pi_\7s(\7b))=\7{sl}(2,\CC)$ and
3948: $\pi_\7s(\7q)\cap\pi_\7s(\sigma\7b)$ is a Cartan subalgebra of
3949: $\7{sl}(2,\CC).$ Similar to the situation considered in \ruf{FN} we
3950: also can choose here a {\sl standard triple} $\4e,\4f,\4h\in
3951: \7s^{\CC}$ with $\pi_\7s(\7q)=\CC \4h \oplus \CC \4e $ and
3952: $\sigma(\4h)=-\4h$, but now $\sigma(\4e)=-\4f$ for the Cartan
3953: involution $\sigma$. The radical $\7r$ cannot be abelian: Otherwise,
3954: exactly as in the proof of Claim \ruf{FP}, we obtain a
3955: contradiction. In the remaining non-abelian case we proceed as in the
3956: proof of Claim \ruf{FQ} by investigating the various positions of
3957: $\7q$ in $\7l$: The cases $\dim \7r^{\CC}\cap \7q>0$ are easily ruled
3958: out (same argument as in the proof of \ruf{FQ}, following
3959: \Ruf{FR}). Hence, we may assume that $\7q$ and $\sigma\7q$ are given
3960: by the formula \Ruf{FR}, except that now $\sigma\7q =\CC\cd
3961: (-\4h,\overline\lambda \4x+\overline \mu \4z)\oplus \CC\cd (-\4f,
3962: \overline \nu \4z)$, i.e., the sign in front of $\4f$ has
3963: changed. This slight difference is precisely the reason why in case
3964: $\nu\ne 0,$ contrary to \Ruf{FS}, the CR-germ $(M,o)$ associated to
3965: $(\7g,\7q)$ would be Levi nondegenerate, as shown by a simple
3966: computation. This contradicts our fundamental assumption and concludes
3967: the proof of the claim as well as the proof of Proposition
3968: \ruf{VD}.\qed
3969:
3970:
3971: \KAP{Vierzehn}{Reduction to the case where $\7g$ is solvable and of
3972: dimension 5}
3973:
3974: Striking the balance for the proof of Theorem II obtained so far,
3975: we have shown the following:
3976:
3977: \noindent{\sl Let $(M,o)$ be an arbitrary
3978: locally homogeneous $2$-nondegenerate CR-germ of dimension $5$ and let
3979: $(\7g,\7q)$ be an associated CR-algebra. If the Lie algebra $\7g$ is
3980: not solvable then $M$ is locally
3981: CR-equivalent at $o\in M$ to the tube $\5M$ over the future light cone.}
3982:
3983: For the rest of the proof we therefore assume that every CR-algebra
3984: $(\7g,\7q)$ under consideration satisfies the fundamental assumption
3985: \ruf{FD} and that $\7g$ is solvable. As main result of this section we
3986: show
3987:
3988:
3989: \Lemma{FT} The solvable Lie algebra $\7g$ has dimension $5$, i.e.,
3990: $\7g_o=0$.\Formend
3991:
3992: \noindent The {\sl proof} of this Lemma will be subdivided into several
3993: steps. Recall that by definition the {\sl nilradical}
3994: $\7g^{\nil}$ of $\7g$ is the maximal nilpotent ideal in $\7g$. It is
3995: well-known that $\7g^{\nil}$ contains the commutator subalgebra
3996: $[\7g,\7g]$, and each element $\xi\in\7g^{\nil}$ is ad-nilpotent in
3997: $\7g.$ Similarly, we denote the (complex) nilradical of $\7l$ by
3998: $\7l^{\nil}.$ We retain the notation from \ruf{FA} and \ruf{CA}.
3999:
4000:
4001: \Claim{FU} ${\7g}_o\subset \7g^{\nil}$.
4002:
4003: \ProofC Since $[\7g,\7g]\subset \7g^{\nil},$ the quotient
4004: $\qu{\7g}{\7g^{\nil}}$ is abelian. Let $\pi:\7g\to
4005: \qu{\7g}{\7g^{\nil}}$ be the projection. Assume that $\7g_o\not\subset
4006: \7g^{\nil}$, i.e., $\pi(\7g_o)\ne 0.$ Select a subalgebra $\7r\subset
4007: \qu{\7g}{\7g^{\nil}}$ (possibly 0) which is a complement of
4008: $\pi(\7g_o)$. But then $\7g':=\pi^{-1}(\7r)$ would violate \26$_{2}$.
4009: Consequently we necessarily have $\pi(\7g_o)=0$, i.e., $\7g_o\subset
4010: \7g^{\nil}$ as claimed. \qd
4011: \Parag{XY} Recall that for the CR-algebra $(\7g,\7q)$ under
4012: consideration there exist flags $\7l_o\subset\7f\subset\7q$ and
4013: $\7l_o\subset\sigma\7f\subset\sigma\7q$ of subalgebras in $\7l$, as
4014: defined in \ruf{FA}. For the subsequent considerations we select the
4015: following elements in $\7l$: $\4y\in\7f\setminus \7l_o$ and $\4x\in
4016: \7q\setminus \7f,$ and write $\overline{\4x}:=\sigma(\4x)$ and
4017: $\overline{\4y}:=\sigma(\4y),$ for short. Then
4018:
4019: \noindent \hfil$\7f= \CC\4y \oplus \7l_o,\quad \7q=\CC\4x\oplus
4020: \CC\4y\oplus \7l_o, \quad \sigma\7f= \CC\overline{\4y} \oplus
4021: \7l_o\Steil{and} \sigma\7q=\CC\overline{\4x}\oplus
4022: \CC\overline{\4y}\oplus \7l_o\,.$
4023:
4024: \noindent The inclusion $\7l_o\subset \7l^{\nil}$ is guaranteed by
4025: \ruf{FU}. Hence, $\7l_o$ acts by ad-nilpotent endomorphisms on
4026: $\7l.$ In particular, $[\7l_o,\7f]\subset\7l_o$ and
4027: $[\7l_o,\7q]\subset \7f.$ The condition \23 means
4028: that $[\4x,\overline{\4x}]\not\in \7q +\sigma\7q$, and
4029: \25 is equivalent to $[\overline{\4x},\4y]\not\in
4030: \7f+\sigma\7q.$ \nline Let $\7z\subset \7l^{\nil}$ be the center of
4031: the nilradical (which is nontrivial if $\7l \ne 0$), to which we refer
4032: as to the {\sl nilcenter} of $\7l$. As for every characteristic
4033: ideal, we have $\sigma(\7l^{\nil})=\7l^{\nil}$ and $\sigma(\7z)=\7z.$
4034:
4035:
4036: \Claim{FV} {\rm (i)} The nilcenter $\7z$ of $\7l$ is not contained in
4037: $\7q+\sigma\7q.$ \ite=2 \1 $\7l_o=0$ if $\dim \7z\ge 2$.
4038:
4039: \Proof {\fett ad (i):} Assume that (i) is not true and let
4040: $\4x,\4y,\overline{\4x},\overline{\4y}\in \7l$ be as in \ruf{XY}. Let
4041: $\zeta:=a\4x+\overline a\overline{\4x}+b\4y+\overline
4042: b\overline{\4y}+\gamma_o,$ $\gamma_o\in\7g_o,$ $a,b\in \CC$, be an
4043: arbitrary element in $\7z^\sigma.$ Then
4044: $[\overline{\4x},\zeta]\in\7z\subset \7q+\sigma\7q$ since $\7z$ is an
4045: ideal in $\7l$. On the other hand
4046: $$
4047: [\overline{\4x},a\4x+\overline a\overline{\4x}+b\4y+\overline
4048: b\overline{\4y}+\gamma_o
4049: ]=a[\overline{\4x},\4x] + [\overline{\4x},b\4y+\overline
4050: b\overline{\4y}+\gamma_o
4051: ]\equiv a[\overline{\4x},\4x]\equiv0\;\mod\; \7q{+}\sigma\7q\,,
4052: $$
4053: which implies $a=0.$ This shows that $\7z\subset \FFF.$ Given then
4054: $\zeta=b\4y+\overline b\overline{\4y}+\gamma_o\in \7z^\sigma$, the
4055: inclusion $[\overline{\4x},\zeta]\subset\7z$ holds since $\7z$ is an
4056: ideal. On the other hand
4057: $$
4058: [\overline{\4x},b\4y+\overline
4059: b\overline{\4y}+\gamma_o\;
4060: ]=b[\overline{\4x},\4y]+[\overline{\4x},\overline
4061: b\overline{\4y}+\gamma_o]\equiv b[\overline{\4x},\4y]\equiv0\;\mod \;
4062: \7f{+}\sigma\7q\;.
4063: $$ Hence, $b=0$ as the consequence of the above equation and \25. But
4064: this cannot be true since then $\7z\subset \7l_o$ would be a
4065: nontrivial ideal of $\7l$, violating \26$_{1}$. \nline {\fett ad
4066: (ii):} Assume $\dim \7z\ge 2,$ i.e. $\7z\cap \QQQ \ne 0$. Recall that
4067: $\7l_o\subset \7l^{\nil}$ by \ruf{FU} and consequently
4068: $[\4y,\7l_o]\subset\7l_o,$ $[\overline{\4y},\7l_o]\subset\7l_o.$ Let
4069: an arbitrary $\zeta :=a\4x+\overline a\overline{\4x}+b\4y+\overline
4070: b\overline{\4y}+\gamma_o \in \7z^\sigma\cap \QQQ$ be given. Computing
4071: its bracket with $\7l_o$ we obtain:
4072: $$
4073: [\zeta,\7l_o]=a[\4x,\7l_o]+\overline a[\overline{\4x},\7l_o]\equiv0
4074: \;\mod\;\7l_o \;.
4075: $$ Either $a=0$ for all such $\zeta,$ and then $\7z\cap \QQQ=\7z\cap
4076: \FFF$, or $a\ne 0,$ and then
4077: $[\4x,\7l_o]\subset\7l_o\supset[\overline{\4x},\7l_o].$ In the latter
4078: case it follows that $\7l_o$ is an ideal in $\7l$ (due to
4079: \ruf{FC}.{\ninecmss III}, $\4x,\4y,\overline{\4x},\overline{\4y}$ and
4080: $\7l_o$ generate $\7l$ as a Lie algebra), hence, $\7l_o=0$ as claimed.
4081: The other possibility would be $\7z\cap (\7q+\sigma\7q)=\7z\cap
4082: (\7f+\sigma\7f)$ and we show that this cannot happen: Given an
4083: arbitrary $\zeta=b\4y+\overline b\overline{\4y}+\gamma_o\in
4084: \7z^\sigma\cap (\7f+\sigma\7f)$, note that
4085: $$
4086: [\overline{\4x},\zeta]\in \7z\cap \QQQ= \7z\cap\FFF.
4087: $$
4088: More explicitly, $[\overline{\4x}\,,\,b\4y+\overline
4089: b\overline{\4y}+\gamma_o ]=b[\overline{\4x},\4y]+[\overline{\4x}\,,\,
4090: \overline b\overline{\4y}+\gamma_o]\in
4091: b[\overline{\4x},\4y]+\sigma\7q$. But then the above equation together
4092: with \25 would imply $b=0,$ i.e., $\7z\cap \QQQ=\7z\cap\7l_o\ne 0.$
4093: This is absurd, as
4094: $$
4095: [\7q+\sigma\7q,\7z\cap\7l_o]\subset \7z\cap
4096: (\7q{+}\sigma\7q)=\7z\cap\7l_o \ne 0,
4097: $$
4098: i.e., since $\7q+\sigma\7q$ generates $\7l$ as a Lie algebra, $\7z\cap
4099: \7l_o$ would be a nontrivial ideal in $\7l$. \qd
4100:
4101: \noindent
4102: In remains the case when the nilcenter is 1-dimensional.
4103:
4104: \Claim{FW}Suppose $\dim\7z=1.$ Then
4105: $\7l=\7z\oplus \QQQ\,$ and $\,\7g_o=\7l_o=0.$
4106:
4107: \Proof Since $\7z\not\subset \7q+\sigma\7q$ by \ruf{FV}, the sum $\7z
4108: + \QQQ $ is direct. Recall that the recursively defined subspaces
4109: $C_0(\7l^{\nil}):=0,$ $C_k:=C_k(\7l^{ \nil}):=\{u\in \7l^{
4110: \nil}:[u,\7l^{ \nil}]\subset C_{k-1}(\7l^{ \nil})\}$ for every $k>0$
4111: form the ascending central series of $\7l^{\nil}.$ Clearly, $\7z=C_1$
4112: and $\sigma(C_k)=C_k$ for all $k.$ Either $\7z=C_1=C_2=\7l^{\nil}$ and
4113: consequently $\7l_o=0$ (due to \26$_{1}$ $\7l_o$ must be a proper
4114: subalgebra of the 1-dimensional algebra $\7l^{ \nil}$) or $C_1\ne
4115: C_2.$ In the latter case $C_2\cap \QQQ\ne 0.$ Let $\eta=a\4x+\overline
4116: a \overline{\4x}+b\4y+\overline b\overline{\4y}+\gamma_o\in
4117: C_2^\sigma\cap \QQQ$ be arbitrary. Since $[\eta,\7l_o]\subset \7z\cap
4118: \QQQ=0,$ we have
4119: $$
4120: [\eta,\7l_o]\equiv a[\4x,\7l_o]+\overline
4121: a[\overline{\4x},\7l_o]\equiv0\;\mod\;\7l_{o}\,.
4122: $$
4123: If $a\ne 0$ then $[\4x,\7l_o],[\overline{\4x},\7l_o] \subset \7l_o$
4124: and $\7l_o$ is an ideal in $\7l,$ i.e., $\7l_o=0$ by \26$_{1}$. If
4125: $a=0$ for every choice of $\eta\in C_2^\sigma\cap \QQQ$ as above, then
4126: $C_2\cap \QQQ=C_2\cap \FFF.$ This possibility can be ruled out as
4127: follows: For a nonzero $\eta=b\4y+\overline
4128: b\overline{\4y}+\gamma_o\in C_2^\sigma\cap\FFF$ we have $[\eta,\4x]\in
4129: \QQQ \cap C_2=\FFF\cap C_2,$ i.e.,
4130: $$
4131: [b\4y+\overline b\overline{\4y}+\gamma_o,\4x]=\overline
4132: b[\overline{\4y},\4x] \;+
4133: ([b\4y+\gamma_o,\4x])\equiv0\;\mod\;\FFF\,,
4134: $$
4135: which is only possible if $\overline b=0.$ This would imply $\QQQ\cap
4136: C_2=\7l_o\cap C_2.$ On the other hand, the identity $\QQQ\cap\
4137: C_2=\7l_o\cap C_2$ implies that $\7l_o\cap C_2$ is an ideal of $\7l.$
4138: The effectivity of $(\7g,\7q)$ forces then $\7l_o\cap C_2=0,$
4139: contradicting $\QQQ\cap C_2\ne 0.$ This completes the proof of \ruf{FW}
4140: and, together with \ruf{FU}, \ruf{FV} also the proof
4141: of Lemma \ruf{FT}. \qed
4142:
4143:
4144:
4145: \KAP{Fuenfzehn}{The existence of a 3-dimensional abelian ideal in
4146: $\7g$ suffices}
4147:
4148: The proof of Theorem II has brought us to the point where we may and
4149: do henceforth assume that the Lie algebra $\7g$ in the CR-algebra
4150: $(\7g,\7q)$ satisfying \ruf{FD} is solvable. The
4151: subalgebra $\7q\subset\7l=\7g\oplus i\7g$ then necessarily is of
4152: complex dimension $2$.
4153:
4154: The main result \ruf{GS} of this section states that the CR-germ
4155: associated to $(\7g,\7q)$ is represented by the tube $F+i\RR^3$ over
4156: an affinely homogeneous surface $F\subset \RR^3$ if $\7g$ is
4157: isomorphic to a semidirect product $\7h\ltimes \7r$ with $\7h$ being a
4158: 2-dimensional Lie subalgebra and $\7r\cong \RR^3$ being an abelian
4159: ideal. Once this Main Lemma is proved, our proof of the
4160: classification theorem will be complete as soon as we can show that
4161: every 5-dimensional solvable Lie algebra $\7g$ occurring in $(\7g,\7q)$
4162: indeed is isomorphic to a semidirect product as above. This will be
4163: achieved in the final section \ruf{Sechzehn}. In this section we only
4164: prove the partial result that if $\7q$ is abelian, then also the
4165: commutator $[\7g,\7g]$ is abelian and 3-dimensional. Moreover, there
4166: exists an abelian subalgebra $\7h\subset \7g$ such that
4167: $\7g=\7h\ltimes [\7g,\7g]$.
4168:
4169:
4170:
4171: Since there is no general structure theory for solvable Lie algebras,
4172: we develop ad hoc methods and describe the structure constants in
4173: $\7l=\7g^{\CC}$ with respect to a particularly chosen basis. Every
4174: CR-algebra $(\7g,\7q)$ under consideration gives rise to the
4175: 1-dimensional subalgebras $\7f$ and $\sigma\7f$ of the 2-dimensional
4176: subalgebras $\7q$ and $\sigma\7q,$ respectively. We construct a basis
4177: of $\7l$ which reflects the conditions \21-\25 and investigate the
4178: various possibilities for the values of the corresponding structure
4179: constants. Select a non-zero $\4z\in \7z^{-\sigma}\setminus \QQQ$
4180: (this is possible due to Claim \ruf{FW}) and an $\4x\in \7q\setminus
4181: \7f$ such that for $\overline{\4x}:=\sigma\4x$ the congruence
4182: $[\4x,\overline{\4x}]\equiv \4z\; \mod\;\7q +\sigma\7q$ holds (this is
4183: possible due to \23). By \25 it is further possible to select $\4y\in
4184: \7f\setminus \7l_o$ such that for $\overline{\4y}:=\sigma\4y$ the
4185: structure equations of $\7l$ are of the following form (in particular,
4186: the coefficient in front of $\overline{\4x}$ in the second equation is
4187: $b_2=1$):
4188: $$
4189: \diagram{[\4x,\overline\4x]&\;=\;&\4z+\;& a_1\4x &\;- \;&\overline
4190: a_1\overline \4x&\;+\;& a_2\4y &\;-\;&\overline a_2\overline \4y \cr
4191: [\4x,\overline{\4y}]&\;=\;& & b_1\4x&+& \phantom{\;1\,}\overline\4x
4192: &+&b_3\4y&+&\hfill b_4 \overline{\4y} \cr [\4y,\overline{\4y}]&\;=\;&
4193: & & & & &\hfill c\,\4y & - &\hfill\overline c\,\overline{\4y} \cr
4194: [\4y,\4x]&\;=\;& & d_1\4x & & &+ & d_2\4y & & \cr
4195: [\4z,\eta]&\in&\7z(\7l^{\nil}) \rlap{ \qquad for every $\eta\in \7l\;.$}
4196: }\Leqno{FI}
4197: $$
4198: The brackets $ [\4y,\overline{\4x}]$ and
4199: $[\overline{\4y},\overline{\4x}]$ are completely determined by the
4200: above 5 equations due to the fact that $\sigma:\7l\to \7l$ is an
4201: antilinear Lie algebra automorphism. Of course, not for all values of
4202: the constants the above identities give rise to a Lie {\sl
4203: algebra}. In fact the above structure constants $a_1,...\,,d_2,$ are
4204: subject to further constraints, imposed by (1) the Jacobi identity,
4205: (2) our assumption that $\7l$ is solvable and, (3) our assumption
4206: $\4z\in \7z(\7l^{\nil}).$ \nline Conversely, let a 5-dimensional
4207: solvable complex Lie algebra $\7l=\CC\4z\oplus \CC \4x\oplus\CC
4208: \overline{\4x}\oplus \CC\4y\oplus \CC\overline{\4y}$ be given with
4209: structure equations as in \Ruf{FI} (together with
4210: $[\overline{\4x},\4y]= \4x +\overline b_1\overline\4x + \overline
4211: b_4\4y +\overline b_3\overline{\4y} $ and
4212: $[\overline{\4y},\overline{\4x}]=\overline d_1\overline\4x + \overline
4213: d_2\overline{\4y}\,$) for certain $a_1,..\,,d_2\in\CC$ and
4214: $\4z\in \7z^{-\sigma},$ where $\7z$ is the nilcenter of $\7l$. Define
4215: $$
4216: \7g:=\RR \,i\4z \oplus \RR(\4x+\overline{\4x})\oplus
4217: \RR(i\4x-i\overline{\4x})\oplus \RR(\4y+\overline{\4y})\oplus
4218: \RR(i\4y-i\overline{\4y})\Steil{ and }\7q:=\CC\4x\oplus \CC\4y \;\;.
4219: $$
4220: Then the CR-algebra $(\7g,\7q)$ satisfies the fundamental assumption
4221: \ruf{FD}.
4222:
4223:
4224:
4225: \medskip As already mentioned, besides the geometrically motivated
4226: conditions \ruf{FC}, which already are incorporated in \Ruf{FI},
4227: further conditions affect the particular values of the structure
4228: constants, for example those given by the Jacobi identity. We use
4229: $\Jac{\xi_1}{\xi_2}{\xi_3}:=[[\xi_1,\xi_2],\xi_3]+
4230: [[\xi_2,\xi_3],\xi_1]+[[\xi_3,\xi_1],\xi_2]$ as shorthand. The
4231: identity $\Jac{\4x}{\4y}{\overline{\4y}}=0$ implies
4232: $$
4233: |c|=1, \qquad d_1=\overline c -\overline b_1 \qquad \hbox{and}\quad
4234: \overline c\re2b_1\in\RR\,.\Leqno{FY}
4235: $$
4236:
4237:
4238: \Remark{VJ} From $c\ne 0$, i.e., $[\4y,\overline{\4y}]\ne0$,
4239: follows that the solvable subalgebra $\7f+\sigma\7f,$ and in turn
4240: $\7l,$ cannot be nilpotent. \Formend
4241:
4242: \medskip\noindent Keeping in mind the identities \Ruf{FY} it is
4243: possible to readjust the basis $\4x,...\,,\4z$ of $\7l$ as follows:
4244: Write $c=e^{2}$ for the coefficient $c$ in the third equation of
4245: \Ruf{FI} and replace $\4x$ by $e\4x$ as well as $\4y$ by $c\4y.$ After
4246: this replacement the structure equations \Ruf{FI} keep their form,
4247: only $c$ changes to $c=1$ and $b_{1}$ becomes real.
4248:
4249:
4250: \noindent {\sl Notational agreement:} For the rest of this section we
4251: use $\alpha,\beta,\gamma$ and $\delta$ to denote real numbers, while
4252: $a_j, b_j,c_j$ and $d_j$ stand for complex numbers. In particular, we
4253: write $\beta_1:=b_1$ to underline that the structure
4254: constant $b_1$ is real.
4255:
4256: \noindent The Jacobi identity $\Jac{\4x}{\4y}{\overline{\4y}{}}=0$
4257: implies further relations between the coefficients in the structure
4258: equations \Ruf{FI}:\vskip-25pt
4259: $$
4260: \diagram{c=1,\qquad\qquad \beta_1:=b_1\in \RR \qquad\cr
4261: d_1=1-\beta_1\;,\qquad \overline b_3-d_2=b_4(\beta_1-1)\;,\qquad
4262: -\beta_1(b_3+d_2)=b_4-\overline b_4 \;.}\Leqno{FZ}
4263: $$ \Joker{GQ}{Perfect basis.} Summarizing, there exists a basis
4264: $\4z,\4x,\overline{\4x},\4y,\overline{\4y}$
4265: ($\overline{\4x}:=\sigma(\4x), \;\overline{\4y}:=\sigma(\4y)$) of
4266: $\7l=\7g^{\CC}$ with $\4z\in \7z^{-\sigma}\setminus
4267: (\7q{\oplus}\sigma\7q)$ and $\7q=\CC\4x\oplus \CC\4y$, $\7f=\CC \4y$
4268: such that the corresponding structure constants in \Ruf{FI} satisfy
4269: \Ruf{FZ}. We call each such basis {\sl perfect}.
4270:
4271: \def\FFF{(\7f\mkern-1mu\oplus\mkern-1mu\sigma\7f)}
4272: \def\QQQ{(\7q\mkern-1mu\oplus\mkern-1mu\sigma\7q)}
4273:
4274: There are still more constraints for the structure
4275: constants given by the Jacobi identity for further triples of elements
4276: and also by the fact that $\7z$ is in the center of
4277: $\7l^{\nil}$. Later on, we characterize these additional conditions
4278: more explicitly. For now, we elaborate the particular structure of the
4279: nilcenter:
4280:
4281: \Lemma{GR} The nilcenter $\7z\subset \7g$ has dimension 1 or 3.
4282:
4283: \Proof We closely analyze the conditions in \ruf{FC} in order to get
4284: the dimension estimates. It is clear that $1\le \dim \7z\le 4$, see
4285: Remark \ruf{VJ}. Assume $\dim \7z=4$. Then $\dim\7z\cap \QQQ=3$
4286: follows by Lemma \ruf{FV}. But since $\7f+\sigma\7f$ is not abelian,
4287: i.e., $\dim \7z\cap \FFF=1,$ there exist elements in $\7z$ of the form
4288: $\4x+\eta_1,\overline{\4x}+\eta_2$ with $\eta_j\in \7f+\sigma\7f.$
4289: This leads to a contradiction: Indeed, on the one side
4290: $[\4x+\eta_1,\overline{\4x}+\eta_2]=0$, and on the other side \25
4291: implies $[\4x+\eta_1,\overline{\4x}+\eta_2]
4292: \equiv[\4x,\overline{\4x}]\not\equiv0\;\mod\;{\7q +\sigma\7q}$. Hence,
4293: we have proved that $1\le \dim \7z\le 3.$
4294:
4295: \noindent
4296: The main difficulty is to rule out the possibility $\dim\7z=2$.
4297: Select a perfect basis in $\7l$ as described in
4298: \ruf{GQ} keeping in mind \Ruf{FI} and \Ruf{FZ}. We have to deal with 2
4299: subcases.
4300:
4301: \Claim{VF} If $[\4x,\4y]\ne 0$, i.e., $\7q$ is not abelian, then
4302: $\dim \7z=1$.
4303:
4304: \ProofC We first show that $\7z\cap \QQQ=\7z\cap \FFF$. Select
4305: an arbitrary
4306: $\4z':=\lambda\4x+\overline\lambda\overline{\4x}+\mu\4y+\overline\mu
4307: \overline{\4y} \in \7z^\sigma\cap \QQQ.$ We have to investigate the 2
4308: possibilities `$\beta_1\ne 1$' and `$\beta_1=1$': In the first
4309: case we get
4310:
4311: \noindent \hfil
4312: $[\4z',(1-\beta_1)\4x+ d_2\4y]\equiv\overline\lambda(1-\beta_1)
4313: [\overline{\4x},\4x]\equiv0\;\mod\;\7q +\sigma\7q$
4314:
4315: \noindent since $[\4y,\4x]\in \7l^{\nil}$. Therefore$\lambda=0$ by the
4316: condition \23, i.e., $\4z'\in \7f +\sigma\7f.$ If $\beta_1=1,$ i.e,
4317: $d_1=0$, then by our assumption $d_2\ne 0,$ and in turn $\4y\in
4318: \7l^{\nil}.$ Hence,
4319: $[\4z',\4y]\equiv\overline\lambda[\overline{\4x},\4y]\equiv0\;\mod\;
4320: \7f +\sigma\7f,$ which, together with \25 also forces
4321: $\lambda=0$. This proves $\7z\cap \QQQ=\7z\cap \FFF$. \nline We claim
4322: that this identity can only h old if both sides vanish, i.e., $\dim
4323: \7z=1$ by Lemma \ruf{FV}.i: Assuming to the contrary that $\7z\cap
4324: (\7q\oplus\sigma\7q)\ne 0$, then on the one hand there exists
4325: $\4z'=\mu\4y+\overline\mu\overline{\4y} \in \7z\cap \FFF$ with $\mu\ne
4326: 0.$ On the other hand,
4327: $[\4x,\4z']=[\4x,\mu\4y+\overline\mu\overline{\4y}]\in
4328: \7z\cap\QQQ=\7z\cap \FFF$ which in view of \25 is only possible if
4329: $\mu=0$. This shows $\dim\7z=1$.
4330:
4331: \noindent It remains to rule out the second subcase:
4332:
4333: \Claim{VG} If $[\4y,\4x]=0$ , i.e., $\7q$ is abelian, then the
4334: commutator $[\7l,\7l]$ is a 3-dimensional abelian ideal. Further, if
4335: $\dim \7z\ge 2$ then $[\7l,\7l]=\7l^{\nil}$ and consequently
4336: $\7z=[\7l,\7l]$ is 3-dimensional.
4337:
4338:
4339: \ProofL This is the most involved case. The assumption $[\4y,\4x]=0$,
4340: i.e, $d_1=d_2=0$ implies $\beta_1=1,$ $b_3=0$ and $b_4:=\beta_4\in
4341: \RR$, see the table below:
4342: $$
4343: \diagram{[\4x,\overline\4x]&\;=\;&\4z \;+\;\;\;& a_1\4x &\;- \;
4344: &\overline a_1\overline \4x&\;+\;& a_2\4y &\;-\;&\overline
4345: a_2\overline \4y\cr [\4x,\overline{\4y}]&\;=\;& &\hfill \4x&+&
4346: \hfill\overline\4x && &+ &\beta_4\overline{\4y} \cr
4347: [\4y,\overline{\4y}]&\;=\;& & & & & &\hfill\4y & -
4348: &\hfill\overline{\4y} \cr [\4y,\4x]&\;=\;&0\hfill \cr [\4z,\eta]&\;=
4349: \;&\rlap{$ c_\eta \4z+ \4q_\eta$\quad $\eta\in \7l$, $\;\4q_\eta\in
4350: \7z\cap \QQQ\;.$ }\hfill }\Leqno{VA}
4351: $$
4352: We need to analyze the relations between the structure constants in
4353: more detail. Let\nline
4354: \centerline{$\4q_{\4x}=z_1\4x+z_2\overline{\4x}+z_3\4y+z_4\overline{\4y}
4355: \Steil{and}\4q_{\4y}=w_1\4x+w_2\overline{\4x}+w_3\4y+w_4\overline{\4y}$.}
4356: \nline The fact that the $\4q_\eta$ commute with all elements of
4357: $[\7l,\7l]\subset \7l^{\nil}$ (and in particular with
4358: $\4y-\overline{\4y}$) implies $z_1=z_2$, $w_1=w_2$ and
4359: $z_4=z_1\beta_4-z_3$, $ w_4=w_1\beta_4-w_3$. The Jacobi identity
4360: $\Jac{\4x}{\overline{\4x}}{\4y}=0$ yields
4361: $$
4362: w_j=0,\; c_{\4y}=1,
4363: \Steil{i.e.,}[\4z,\4y]=[\4z,\overline{\4y}]=\4z,\qquad
4364: a_1=i\alpha_1,\qquad a_2=\alpha_2+{i\over 2}\alpha_1\beta_4\;,
4365: \Leqno{}
4366: $$
4367: for some $\alpha_1,\alpha_2\in \RR$, and
4368: $\Jac{\overline{\4x}}{\4y}{\4z}=0$ implies
4369: $$
4370: \Im z_1=0, \qquad \Re z_3=\qu{z_1\beta_4}2 \Steil{ and }\Re
4371: c_{\4x}=-\qu{\beta_4}2\;.\Leqno{}
4372: $$
4373: Summarizing,
4374: $$
4375: [\7l,\7l]=\CC
4376: \4z\;\oplus\;\CC(\4x+\overline{\4x}+\hbox{\Klein$\displaystyle
4377: {\beta_4\over 2}$} \4y+\hbox{\Klein$\displaystyle {\beta_4\over
4378: 2}$}\overline{\4y})\;\oplus\; \CC(\4y-\overline{\4y})\;, \Leqno{VK}
4379: $$
4380: and, using the above relations between the structure constants, a
4381: simple check shows that this ideal is abelian. Moreover, the subset
4382: $$
4383: \7h:=\RR \; i\big(\4x-\overline{\4x}+\hbox{\Klein$\displaystyle
4384: {\beta_4\over 2}$} \4y-\hbox{\Klein$\!\displaystyle {\beta_4\over
4385: 2}$}\overline{\4y}\big) \;\oplus \;\RR (\4y+\overline{\4y}) \Leqno{VL}
4386: $$
4387: is an abelian subalgebra of $\7g$
4388: (and $\7g=\7h\oplus [\7g,\7g]$; in fact $\7g=\7h\ltimes [\7g,\7g]$).
4389: \nline
4390: It should be noted that the nilcenter $\7z$ may be 1-dimensional,
4391: and then $\7l^{\nil}$ properly contains
4392: $[\7l,\7l]$.
4393:
4394: \noindent We next show that the situation `$\dim \7z=2$' does not
4395: occur: Assume to the contrary that $\dim \7z= 2$. Then $\7z\cap \QQQ$
4396: is nonzero. Let $\4z':=\lambda\4x + \overline \lambda\overline{\4x} +
4397: \mu\4y +\overline \mu\overline{\4y}\in \7z^\sigma\cap \QQQ$ be
4398: arbitrary. Either, for all such $\4z'$ the coefficient $\lambda$ is 0,
4399: i.e., $\7z\cap \QQQ=\7z\cap \FFF$, and then this case can be ruled out
4400: by a similar argument as in the proof of the above claim. Or, there
4401: exists $\4z'$ with $\lambda\ne 0.$ In such a situation the identity
4402: $[\4z',\4y-\overline{\4y}]=0$ implies $\lambda\in \RR^*$ and without
4403: loss of generality we may suppose $\4z'=\4x + \overline{\4x} + z_3\4y
4404: +\overline z_3\overline{\4y}$ for some $z_3\in \CC$. The condition
4405: $[\4z,\4z']=0$ gives $\Re z_3=\qu{\beta_4}2$. But then
4406: $\4z'=(\4x+\overline{\4x}+\qu{\beta_4}2\cd
4407: (\4y+\overline{\4y}))+i\,\Im z_3 \cd (\4y-\overline{\4y})\in
4408: [\7l,\7l]$, compare \Ruf{VK}. We claim, that in the situation under
4409: consideration, $\7l^{\nil}=[\7l,\7l]$. To prove this, we simply
4410: compute the ad-action of $\4y+\overline{\4y}$ and $\4x-\overline{\4x}$
4411: on $[\7l,\7l]$. Since $[\4x-\overline{\4x},\4z' ]\in \CC \4z\oplus \CC
4412: \4z'$, the relation $\beta_4^2+4(\Im z_3)^2=4\alpha_1 \Im
4413: z_3+4\alpha_2$ must also be fulfilled. Once again, a simple
4414: computation yields
4415: $$
4416: \ad(\4y+\overline{\4y})\rest {[\7l,\7l]}=-2\cd\id, \qquad
4417: [\4x-\overline{\4x},\4z']=2\4z+2i(\alpha_1-\Im\,z_3)\4z'\;.
4418: $$
4419: The above identities show that for every $\4v:=u_1\cd
4420: (\4x-\overline{\4x})+u_2(\4y+\overline{\4y})+u_3\eta,$ $\,u_j\in \CC$,
4421: $\eta\in [\7l,\7l]$, the condition $[\4v,\4z']=0$ implies
4422: $u_1=u_2=0$, i.e., the centralizer $C_{\7l}(\4z')$ coincides with
4423: $[\7l,\7l]$. This proves $[\7l,\7l]=\7l^{\nil}=C_{\7l}(\4z')$. But
4424: this is absurd, since then the nilcenter $\7z$ would coincide with
4425: the 3-dimensional abelian ideal $[\7l,\7l]$, contrary to our
4426: assumption `$\dim \7z=2$'.
4427:
4428: \noindent Finally, we need to investigate the case $\dim \7z=3.$ We
4429: claim that $\7z=[\7l,\7l]=\7l^{\nil}$: To see this, it enough to show
4430: that $\7l^{\nil}$ is 3-dimensional, as, due to \Ruf{VK}, $[\7l,\7l]$
4431: is 3-dimensional, too. As already mentioned (see the sentence
4432: following \Ruf{FY}) $\7l^{\nil}$ can be at most 4-dimensional. But the
4433: 4-dimensional case can be excluded, otherwise $\7l^{\nil}=\7z\oplus
4434: \CC \4n$ for $\4n\in \7l^{\nil}\setminus [\7l,\7l]$ which would imply
4435: that $\7l^{\nil}$ is abelian. Hence the nilradical is
4436: 3-dimensional. This proves \ruf{VG} and Lemma \ruf{GR}.\qed
4437:
4438:
4439: \medskip The next statement is one of the key points in our
4440: classification of 5-dimensional 2-nondegenerate homogeneous
4441: CR-germs. Before stating it we first fix some notation. Given a vector
4442: space $V$, write $\aff(V)$ for the Lie algebra consisting of affine
4443: maps of $V$. This Lie algebra (as well as the corresponding Lie group
4444: $\Aff(V)$) has the natural semidirect product structure:
4445: $\aff(V)=V\rtimes \gl(V)$ (with $\gl(V)=\{X\in \aff(V):X(0)=0\}$). Let
4446: $\pi:\aff(V)\to \gl(V)$ be the projection homomorphism. We use similar
4447: notation on the Lie group level and write, for instance,
4448: $\pi:\Aff(V)=V\rtimes \GL(V)\to \GL(V)$ for the corresponding group
4449: homomorphism. Sometimes we simply write $\psi^{\lin}:=\pi(\psi)$ for
4450: the linear part of an element in $\aff(V)$ or $\Aff(V).$
4451:
4452: \ifarx\eject\fi
4453:
4454: \MLemma{GS} Let $(\7g,\7q)$ be a CR-algebra satisfying the fundamental
4455: assumption \ruf{FD} and let $\7g$ be solvable and of dimension 5.
4456: Suppose that there exists a 3-dimensional abelian ideal $\7v\subset
4457: \7g$ and a 2-dimensional subalgebra $\7h\subset \7g$ with $\7h\cap
4458: \7v=\{0\}$. Then the associated CR-germ $(M,o)$ is locally
4459: CR-equivalent to a tube $F\times i\RR^3 \subset \CC^3$, where
4460: $F\subset \RR^3$ is an affinely homogeneous surface.\Formend
4461:
4462: \noindent The {\sl proof} is divided in several steps which give more
4463: precise (but also more technical) information concerning the structure
4464: of the Lie groups corresponding to $\7g$, $\7l$ and $\7q$ and a
4465: realization of the CR-germ $(M,o)$:
4466:
4467:
4468: \Claim{} The adjoint representation $\ad:\7h\to \gl(\7v)$ is
4469: faithful. Consequently, identifying $\7h$ with the subalgebra
4470: $\ad(\7h)\subset\gl(\7v)$, the Lie algebras $\7g$ and $\7l$ can be
4471: realized as Lie subalgebras of affine transformations:
4472: $\;\;\7g=\7v\rtimes \7h=\7v\rtimes \ad(\7h)\subset \aff(\7v)\cong
4473: \aff(\RR^3)\;$ {and} $\;\7l=\7v^{\CC}\!\rtimes \7h^{\CC}\subset
4474: \aff(\7v^{\CC})\cong \aff(\CC^3).$
4475:
4476: \ProofC Let $\7n\subset \7h$ be the kernel of the adjoint
4477: representation $\ad:\7h\to \gl(\7v).$ The case $\dim \7n=1$ can be
4478: excluded, otherwise $\7n\oplus\7v=\7g^{\nil}=\7z$ would be
4479: 4-dimensional, contradicting Lemma \ruf{GR}. The case $\dim
4480: \7n=2$ can be also excluded: Otherwise $\7g=\7v\times \7h$ would be
4481: abelian or contain a 4-dimensional abelian nilradical which in both
4482: cases would contradict \23 -- \25.\qd
4483:
4484:
4485: Write $V\cong \RR^3$ for a vector group with Lie algebra $\7v$ and
4486: $E:=V^{\CC}$ for its complexification. Let $H_{\GL}\subset \GL(V)$ and
4487: $H_{\GL}^{\CC}\subset \GL(E)$ be the Lie subgroups, corresponding to
4488: the Lie algebras $\ad(\7h)$ and $\ad(\7h^{\CC})$, respectively. Since
4489: $\GL(V)\cong \GL(3,\RR)$ contains no compact torus of dimension
4490: $\ge2$, each subgroup, in particular $H_{\GL}$, is closed. This is in
4491: general not true for the complex subgroup $H_{\GL}^{\CC}$. Let
4492: $H^{\CC}$ be the simply connected Lie group with Lie algebra
4493: $\7h^{\CC}$, $\pr:H^{\CC}\to H_{\GL}^{\CC}\subset\GL(\7v^{\CC})$ the
4494: homomorphism induced by $\ad:\7h^{\CC}\to
4495: \ad(\7h^{\CC})\subset\gl(\7v^{\CC})$ and $L:=V^{\CC}\rtimes
4496: H^{\CC}$. For simplicity, for each $h\in H^{\CC}$ we also write
4497: $h^{\lin}\subset \GL(E)$ instead of $\pi(h)$. Let $G=V\rtimes H\subset
4498: L$ be the real form. Since every 2-dimensional Lie algebra is
4499: solvable, we deduce that also $\7l=\7v^{\CC}\!\rtimes \7h^{\CC}$ (as
4500: well as $\7g$, $L$ and $G$) is solvable.
4501:
4502:
4503: \Claim{} Let $Q\subset L$ be the subgroup corresponding to the Lie
4504: subalgebra $\7q\subset \7l$. Then $Q$ is closed and $Q\cap
4505: V^{\CC}=\{e\}$. Hence $L=V^{\CC}\!\rtimes Q$ is a semidirect product.
4506:
4507: \ProofC Let $\pi:\7l\to \7h^{\CC}$ be the projection homomorphism. Our
4508: first observation is that $\pi(\7q)=\7h^{\CC}$: The case
4509: `$\pi(\7q)=0$' can clearly be excluded as in such a situation
4510: $\7q\subset \7v^{\CC}$, and in turn $\7q+\sigma\7q\subset \7v^{\CC}$,
4511: which is absurd. \nline The possibility $\dim \pi(\7q)=1$ can be
4512: ruled out as follows: We may assume that $\pi(\7q\oplus
4513: \sigma\7q)=\7h^{\CC}$ (otherwise $\7q\oplus \sigma\7q$ would be a
4514: subalgebra). Either $\pi(\7f)=0,$ i.e., $\7q\cap \7v^{\CC}=\7f$, and
4515: in turn $(\7q\oplus\sigma\7q)\cap \7v^{\CC}=\7f\oplus \sigma\7f$: This
4516: leads to a contradiction since then $[\7F,\7H]\subset \7H\cap
4517: \7v=\7F$, violating \25. Or, $\pi(\7f)=\pi(\7q)$. But then
4518: $\7q=\7f\oplus \CC \4x$ with a nonzero $\4x\in \7q\cap \7v^{\CC}$, and
4519: in turn $[\7q,\sigma\7q]\subset \7q\oplus\sigma\7q$, contradicting
4520: \23. Summarizing, $\pi(\7q)=\7h^{\CC}$. \nline On the group level,
4521: since $L$ is simply connected and solvable, every connected subgroup
4522: is closed. The restriction of $\pi$ to $Q$ induces a surjective
4523: homomorphism $Q\to H^{\CC}$. Since both groups are 2-dimensional,
4524: this homomorphism is a covering. Our assumption that $H^{\CC}$ is
4525: simply connected finally implies that
4526: $\pi\lower1pt\hbox{\mathsurround=0pt$|$}_{Q}:Q\to H^{\CC}$ is an
4527: isomorphism. In particular $Q\cap V^{\CC}=Q\cap \ker\pi=\{e\}$.\qd
4528:
4529:
4530: \Claim{VM}With respect to the identification $Z:=L/Q=V^{\CC}\cong
4531: \CC^3$ the real form $G$ acts on $L/Q$ by affine transformations and
4532: $V\subset G$ by translations.
4533:
4534: \ProofC The existence of the decomposition $L=V^{\CC}\!\rtimes Q$
4535: implies that there are well-defined functions $v\colon L\to V^{\CC}$,
4536: $q\colon L\to Q$ such that $\ell=v(\ell)\cd q(\ell)$ for every
4537: $\ell\in L.$ Let $g=w\cd h\in V\rtimes H=G$ (with $w\in V, h\in H$) be
4538: arbitrary. Then, for any $z\in V^{\CC}$ we have
4539: $$
4540: g\cd zQ=w\cd h \cd zQ =w \cd v(h)\cd q(h)\cd zQ = w \cd v(h)\cd
4541: (q(h)\cd z \cd q(h)^{-1})Q
4542: $$
4543: and $ q(h)\cd z \cd q(h)^{-1}= v(h)^{-1}h \cd z \cd h^{-1}v(h)=
4544: h^{\lin}(z)$. Hence, with respect to the identification $V^{\CC}=L/Q$
4545: (induced by the inclusion $V^{\CC}\into L$ such that $0$ corresponds
4546: to the point $eQ\in L/Q$), the action of $G$ can written as follows:
4547: $$
4548: g\cdot z= h^{\lin}(z) + v(h) +w \qquad g=w\cd h\in L,\;\; z\in V^{\CC}\,.
4549: \Leqno{VN}
4550: $$
4551: In particular, the subgroup $V\subset G$ acts by translations
4552: $z\mapsto z+w$. \qd
4553:
4554: \noindent
4555: Consequently $M:= G\cdot 0=VH\cdot 0= V +
4556: F\subset V\oplus iV$ with $F:=M\cap iV$.
4557: It should be noted, however, that in general $F :=(G\cdot
4558: 0)\cap iV \ne H\cdot 0$. Nevertheless, as we shortly will see, $F$ is
4559: affinely homogeneous under a slightly different subgroup of
4560: $\Aff(iV)$. Clearly, multiplying a tube manifold $F+ iV\subset
4561: V\oplus iV$, $F\subset V$, by the imaginary unit $i$ we get the
4562: CR-equivalent realization $V+ iF=V+ F'$ with $F'=iF\subset
4563: iV$. The latter form of a tube manifold is more suitable in our
4564: particular setup, and we keep this notation until the end of the
4565: proof of the Main Lemma.
4566:
4567: \Claim{} Retaining the previous notation, there exists a subgroup
4568: $B\subset iV\!\rtimes \GL(iV)=\Aff(iV)$ such that $F:= (G\cdot 0)\cap
4569: iV =B\cdot 0$.
4570:
4571: \ProofC Let $\pr^{i}:V\oplus iV\to iV$ be the linear projection. A
4572: glance at \Ruf{VN} shows that $F=\pr^{i}\{v(h): h\in H\}.$ In order to
4573: determine $v(h)$ more explicitly we need to analyze the position of
4574: $Q$ in $V^{\CC}\!\rtimes H^{\CC}$ in greater detail: Since $\7h$ is
4575: 2-dimensional, there exists a basis $\4s,\4n\in \7h$ such that
4576: $[\4s,\4n]=\epsilon \4n$ with $\epsilon\in \{0,1\}$. Recall that the
4577: projection map $\pi:\7q\to \7h^{\CC}$ is an isomorphism. Consequently
4578: there exist $\4w_{\4s},\4w_{\4n}\in \7v^{\CC}=V^{\CC}$ such that the
4579: elements $\4w_{\4s}+\4s$ and $\4w_{\4n}+\4n$ in
4580: $\7l=V^{\CC}\oplus\7h^{\CC}$ generate $\7q$. Since then $\sigma\7q=\CC
4581: (\overline {\4w}_{\4s}+\4s)\oplus \CC (\overline{\4w}_{\4n}+\4n)$, we
4582: must have $\4w_{\4s}\ne \overline {\4w}_{\4s}$ and $\4w_{\4n}\ne
4583: \overline {\4w}_{\4n}$ (otherwise $\7q\cap\sigma\7q \ne 0$). Let
4584: $\exp:\7l\to L$ and $\Exp:\ad(\7h^{\CC})\to \GL(V^{\CC})$ be the
4585: exponential maps (i.e., $\Exp({\ad \4v})=\pi(\exp \4v)$ with $\pi$ as
4586: in the paragraph preceeding \ruf{GS}). Furthermore, let $\Psi$ be the
4587: entire function defined by $$\Psi(z)={e^{z}-1\over
4588: z}=\sum^{\infty}_{k=0}{z^{k}\over(k+1)!}z^{k}\,.$$
4589: Then for $\4S^t:=\Psi(t\ad(\4s))$ and $\4N^u:=\Psi(u\ad(\4n))$
4590: a simple computation shows:
4591: $$
4592: \eqalign{ H&=\{\exp(t\4s)\cd \exp(u\4n):t,u\in \RR\} \cr Q&=\{\exp
4593: t(\4w_{\4s}+\4s)\cdot\exp u(\4w_{\4n}+\4n): t,u\in \CC \}\cr &=\{
4594: \4S^t(t\4w_{\4s})\cd\exp(t\4s)\cdot
4595: \4N^u(u\4w_{\4n})\cd\exp(u\4n):t,u\in \CC \} \cr
4596: &=\{\big(\4S^t(t\4w_{\4s})\cd \Exp(t\ad(\4s))\,(\4N^u(u\4w_{\4n}))\big)
4597: \cdot \exp (t\4s)\cd\exp(u\4n):t,u\in \CC \} \;\;\subset\;\; V^{\CC}
4598: \cdot H^{\CC}=L\,. } \Leqno{VO}
4599: $$
4600: The explicit form of $v(h)$ (compare the proof of Claim \ruf{VM}) can
4601: be read off the last line in \Ruf{VO}:
4602: $$
4603: v(h)=v(\exp (t\4s)\exp(u\4n))
4604: =\big(\4S^t(t\4w_{\4s})\big)^{-1}\cdot\big(\Exp(t\ad(\4s))\,
4605: (\4N^u(u\4w_{\4n}))\big)^{-1}\,.
4606: $$
4607: Since $\ad(\4s), \4N^u$ and $\4S^t$ are real operators, it follows for
4608: $h=\exp(t\4s)\cdot\exp( u\4n)$ as before:
4609: $$
4610: \eqalign{ \pr^{i}(v(h))\;&=\;\big(\4S^t(t\4w_{\4s})\big)^{-1}\!\cdot
4611: \big(\Exp(t\ad(\4s))\, (\4N^u(u\4w_{\4n}))\big)^{-1}\cr&=\;\; \exp
4612: t(-\4w^{i}_{\4s}+\4s)\cd \exp u(-\4w^{i}_{\4n}+\4n)\cdot 0 \;\subset
4613: iV\;.\cr }\Leqno{VP}
4614: $$
4615: (Using additive notation, $\pr^{i}(v(h))=-\4S^t(t\4w^{i}_{\4s})-
4616: \Exp(t\ad(\4s))\,(\4N^u(u\4w^{i}_{\4n}))\;\subset iV\,$). Define
4617:
4618: \hfil $\7b:=\RR(-\4w^{i}_{\4s}+\4s)\oplus
4619: \RR(-\4w^{i}_{\4n}+\4n)\;\;\subset \; \7l=V^{\CC}\!\rtimes
4620: \7h^{\CC}$\hfil
4621:
4622: \noindent and check that this is a Lie algebra. Then $B:=\exp
4623: \RR(-\4w^{i}_{\4s}+\4s)\cd \exp \RR(-\4w^{i}_{\4n}+\4n)$ is the
4624: subgroup of $L$ with Lie algebra $\7b$, and \Ruf{VP} shows that
4625: $F=\pr^{i}\{v(h):h\in H\}=B\cdot 0$. This finishes the proof of the
4626: claim and of the Main Lemma.\qed
4627:
4628: \KAP{Sechzehn}{The final steps}
4629:
4630:
4631: Our final step toward the complete classification of all
4632: 5-dimensional 2-nondegenerate and homogeneous CR-germs is to deduce
4633: that each 5-dimensional solvable Lie algebra $\7g$ which occurs in a
4634: CR-algebra $(\7g,\7q)$ subject to \ruf{FD}, also satisfies the
4635: assumptions of the preceeding Main Lemma \ruf{GS}:
4636:
4637:
4638:
4639: \Lemma{VQ} Let $(\7g,\7q)$ be a CR-algebra satisfying \ruf{FD} and
4640: suppose that $\7g$ is solvable and of dimension 5. Then there exists a
4641: semidirect product decomposition $\7g=\7v\rtimes \7h$ with a
4642: 3-dimensional abelian ideal $\7v\subset \7g$ and a 2-dimensional
4643: subalgebra $\7h$. \Formend
4644:
4645: \Proof We have already observed in Lemma \ruf{GR} that the nilcenter
4646: $\7z$ has dimension 1 or 3. If $\dim \7z=3$ then, due to \Ruf{VF}, we
4647: can apply \ruf{VG}: Consequently, we can choose $\7v=[\7g,\7g]$
4648: (compare \Ruf{VK}) and $\7h$ as defined in \Ruf{VL}.
4649:
4650: \noindent The situation $\dim \7z=1$ requires some more elaborated
4651: work. We classify all CR-algebras $(\7g,\7q)$ under consideration in
4652: terms of the corresponding structure equations with respect to some
4653: perfect basis $\4x,...\,,\4z$ of $\7l$ (see \ruf{GQ}): Given
4654: $(\7g, \7q)$, let the corresponding structure equations be as in
4655: \Ruf{FI}, taking into account \Ruf{FZ}. To handle the various sets of
4656: relations between the structure constants, we divide the class of
4657: CR-algebras under consideration into the subclasses A, B and C, see
4658: below.
4659:
4660: \noindent {\fett Case A: $\beta_1\ne \pm 1$}. In this situation it is
4661: possible to assume $a_1=0$ (simply replace $\4x$ by
4662: $\4x+\lambda\4y$ with $\lambda=u+iv$ defined by $u:=\qu{\Re
4663: a_1}{(1-\beta_1)}$ and $v:=\qu{\Im a_1}{(1+\beta_1)}$). The structure
4664: equations then read
4665: $$
4666: \diagram{[\4x,\overline\4x]&\;=\;&\4z+\;& & & &\;+\;& a_2\4y &\;-\;
4667: &\overline a_2\overline \4y \cr [\4x,\overline{\4y}]&\;=\;& &\kern1em
4668: \beta_1\4x&+& \;\overline\4x &+&b_3\4y&+&b_4\overline{\4y} \cr
4669: [\4y,\overline{\4y}]&\;=\;& & & & & & \;\;\4y & - & \;\;\overline{\4y}
4670: \cr [\4y,\4x]&\;=\;& & \kern-10pt(1-\beta_1)\4x & & &+ &d_2\4y & &
4671: \rlap{\qquad\qquad $d_2=\overline b_3+(1-\beta_1)b_4$} \cr
4672: [\4z,\eta]&\;= \;&c_\eta\4z \rlap{ \qquad for every $\eta\in \7l\;$
4673: ,}}\qquad\qquad\qquad\qquad \Leqno{GT}
4674: $$
4675: and we now work out further constraints imposed on the constants: An
4676: explicit evaluation of the Jacobi identity
4677: $\Jac{\4x}{\overline{\4x}}{\4y}=0$ yields $c_\4y=c_{\overline{\4y}}
4678: =2\beta_1-1.$ Furthermore, we obtain the equations $\overline
4679: b_3(1+\beta_1)=b_4(\beta_1^2-\beta^{}_2)$ and $\overline
4680: b_3(1+\beta_1)=(b_4-\overline b_4)(\beta_1-1)$ which imply
4681: $b_4(1-\beta_1)=\overline b_4.$ \nline In order to investigate most
4682: conveniently further relations between the structure constants, we
4683: deal separately with the following 3 subcases:
4684:
4685: \medskip\centerline
4686: { {\bf AI:} $b_4\in i\RR^{*}$ and $\beta_1=2,$\hfil
4687: {\bf AII:} $b_4\in \RR^{*}$ and $\beta_1=0,$\hfil
4688: {\bf AIII:} $b_4=0.$}
4689:
4690: \noindent {\bf Ad AI:} Put $\beta_{4}:=-ib_{4}$. In this particular
4691: situation the identity $\Jac{\4x}{\overline{\4x}}{\4y}=0$ implies
4692: $b_3=-{2\over 3} i\beta_4,$ $a_2=-{2\over 9}\beta_4^2$ and
4693: $\Jac{\4x}[\4y]{\4z}=0$ implies $c_\4x=-i\beta_4.$ There are no more
4694: conditions imposed by the Jacobi identity and the structure
4695: equations of $\7l$ are now
4696: $$
4697: \diagram{[\4x,\overline\4x]&\;=\;&\4z+\;& & & &\;-\;& 2 \gamma^2\4y
4698: &\;+\;&2 \gamma^2\overline{\4y} \cr [\4x,\overline{\4y}]&\;=\;&
4699: &\kern1em 2\4x&+& \; \overline\4x &-&2i \gamma\4y&+&3i
4700: \gamma\overline{\4y} \cr [\4y, \overline{\4y}]&\;=\;& & & & & &
4701: \;\;\4y & - & \;\;\overline{\4y} \cr [\4y,\4x]&\;=\;& & \kern-3pt-\4x
4702: & & &- &i \gamma \4y & & \cr [\4z,\4x]&\;= \;&\rlap{$-3i \gamma\,\4z$
4703: \qquad\qquad $[\4z,\4y]=-3\4z\;,$}\hfill }\qquad\qquad\qquad\qquad
4704: \Leqno{GU}
4705: $$ where $ \gamma:=\beta_4/3\in \RR^{*}$. Keeping in mind Lemma
4706: \ruf{GS}, the structure of $\7l$, $\7g$ and the position of $\7q$ is
4707: determined by \Ruf{GU}. This can be seen more clearly by decomposing
4708: $\7g$ and $\7l$ into the eigenspaces of $\ad(\4s),$ where
4709: $\4s:=-(\4y+\overline{\4y}/2:$ Define the following elements
4710: $\4n,\4v_1,\4v_2,\4v_3$ from $\7g:$
4711: $$
4712: \diagram{\4n&\;:= \;& i\4x-i\overline{\4x}-\;\gamma\4y-\;
4713: \gamma\overline{\4y} \cr
4714: \4v_1&:=& \qquad\qquad \hbox{\klein$\displaystyle{i\over 2}$}
4715: \4y-\hbox{\klein$\displaystyle{i\over 2}$}\overline{\4y} \cr
4716: \4v_2&:=& \;\4x+\;\overline{\4x}-2i\gamma\4y+2i \gamma\overline{\4y}\cr
4717: \4v_3&:=& 2i\4z\;.\vphantom{{}^{\textstyle|}}\hfill}
4718: $$
4719: It is clear that $\4s,\4n,\4v_1,\4v_2,\4v_3$ form a basis of $\7g.$
4720: The bracket relations are:
4721: $$
4722: [\4s,\4v_k]=k\4v_k,\;[\4s,\4n]=\4n,\quad [\4n,\4v_1]=\4v_2,\;
4723: [\4n,\4v_2]=\4v_3,\;[\4n,\4v_3]=0.
4724: $$ Further, $\7v:=\RR\4v_1\oplus \RR\4v_2\oplus \RR\4v_3\cong \RR^3$
4725: is an abelian ideal in $\7g$ with $[\7g,\7g]=\7g^{\nil}=\RR\4n\oplus
4726: \7v$. Hence, $\7g$ has the structure of the semidirect product
4727: $\7v\rtimes\7h$ with $\7h=\RR\4s\oplus\RR\4n$ and the Main Lemma
4728: applies.
4729:
4730: \noindent {\bf Remark.} A direct verification shows that for every
4731: $\gamma\in\RR^*$ $(\7g,\7q)$ in \Ruf{GU} is associated to Example
4732: \ruf{EV}.
4733:
4734: \noindent We show that in the next case the Lie algebra cannot be
4735: solvable, hence, this case can be discarded.
4736:
4737: \noindent {\bf Ad AII:} Write $\beta_4:=b_4$. The Jacobi identity
4738: implies $a_2=0=b_3$ and $d_2=\beta_4$, and \Ruf{VA} reads \vskip-18pt
4739: $$
4740: \diagram{[\4x,\overline\4x]&\;=\;&\4z\;\;\cr
4741: [\4x,\overline{\4y}]&\;=\;& \;\;& \kern1em & \kern1em \;&
4742: \;\overline\4x & \;& \;&+&\beta_4\overline{\4y} \cr
4743: [\4y,\overline{\4y}]&\;=\;& & & & & & \kern.7em \;\4y & - & \kern.7em
4744: \; \overline{\4y} \cr [\4y,\4x]&\;=\;& & \;\4x & & &+ &\beta_4 \4y & &
4745: \cr [\4z,\4x]&\;= \; &\rlap{$\beta_4\4z$ \qquad\qquad
4746: $[\4z,\4y]=-\4z\;.$}\hfill }\qquad\qquad\qquad\qquad \Leqno{}
4747: $$
4748: But then the linear span of the vectors
4749: $$
4750: \displaylines{ \4e^+:=\4y-\overline{\4y}, \qquad\qquad
4751: \4h:=-\frac1{\beta_4}\big (\4x+\overline{\4x}+(\beta_4-1)\4y \;+
4752: \;(\beta_4+1) \overline{\4y}\big),\kern1em \cr \4e^-:=
4753: \frac1{4\beta_4^2}\big(2\4z+(2-2\beta_4)\4x+(2+2\beta_4)
4754: \overline{\4x}-(1-\beta_4)^2\4y+(1+\beta_4)^2\overline{\4y}\big) }
4755: $$
4756: is a copy of $\7{sl}(2,\CC)$ in $\7l$, that is, $\7l$ is not solvable.
4757:
4758: \noindent{\bf Ad AIII:}. The condition $b_4=0$ together with the Jacobi
4759: identity $\Jac{\4x}{\overline{\4x}}{\4y}=0$ implies $a_2=b_3=0$ and
4760: $c_\4y=2\beta_1-1$. Since $(1-\beta_1)\ne 0$, the identity
4761: $[\4z,[\4y,\4x]]=0$ implies $[\4z,\4x]=0$, see the table below.
4762: $$
4763: \diagram{[\4x,\overline\4x]&\;=\;&\4z\;& \cr
4764: [\4x,\overline{\4y}]&\;=\;& &\kern1em \beta_1\4x&+& \;\overline\4x &
4765: \cr
4766: [\4y,\overline{\4y}]&\;=\;& & & & & & \;\;\4y & - & \;\;\overline{\4y}
4767: \cr
4768: [\4y,\4x]&\;=\;& & \kern-6pt(1-\beta_1)\4x & & & & & &
4769: \cr
4770: [\4z,\4x]&\;= \;&0\hfill
4771: \rlap{\qquad,\qquad $[\4z,\4y] \;=\; (2\beta_1-1)\4z\;.$} \hfill
4772: }\qquad\qquad\qquad\qquad
4773: \Leqno{GV}
4774: $$
4775: Select the following basis of $\7g$:
4776: $$
4777: \diagram{\4n\,&\;:= \;& \hbox{\klein$\displaystyle{i\over
4778: 2}$}(\4x-\overline{\4x}) \qquad,\qquad\quad \4s\,:=
4779: \hbox{\Klein$\displaystyle{1\over 2-2\beta_1}$}(\4y+\overline{\4y})\cr
4780: \4v_1&:=& \qquad\qquad i\4y\;- i\overline{\4y} \hfill\cr \4v_2&:=&
4781: \;\;\4x \; + \;\overline{\4x} \hfill\cr \4v_3&:=&
4782: i\4z\;.\vphantom{\big|}\hfill}
4783: $$
4784: One checks immediately that $\7v:=\RR \4v_1\oplus\RR \4v_2\oplus\RR
4785: \4v_3$ is an abelian ideal and $\7h:= \RR \4s\oplus\RR \4n$ is a
4786: subalgebra with $[\4s,\4n]=\4n$. Further,
4787: $$
4788: [\4s,\4v_j]= \hbox{\klein$\displaystyle{2-j+(j-1)\beta_1\over
4789: \beta_1-1}$}\4v_j\steil{for}j=1,2,3\Steil{and}[\4n,\4v_1]=\4v_2,\;
4790: [\4n,\4v_2]=\4v_3,\;[\4n,\4v_3]=0.
4791: $$
4792: Hence, $\7g=\7v\rtimes \7h$ as claimed, and the Main Lemma applies.
4793: Also in this case for all $\beta_1\ne \pm 1$ the CR-algebra
4794: $(\7g,\7q)$ is associated to Example \ruf{EV}.
4795:
4796:
4797: \medskip It remains to discuss the cases $\beta_1=\pm1.$
4798:
4799: \noindent {\fett Case B: $\beta_1=1$.} Plugging $\beta_1$ into
4800: \Ruf{FZ} gives $d_2=\overline b_3$ and $d_1=0.$ A direct check shows
4801: that $\Jac{\4x}{\overline{\4x}}{\4y}=0$ implies $0=b_3=d_2,$ i.e.,
4802: $[\4x,\4y]=0$. Lemma \ruf{VG} then gives that $\7g$
4803: is isomorphic to the semidirect product $\7v\rtimes\7h$ with
4804: $\7v=[\7g,\7g]$ and the abelian subalgebra $\7h$ as defined in
4805: \Ruf{VL}.
4806:
4807: \medskip\noindent {\fett Case C: $\beta_1=-1$.} We proceed as in the
4808: preceeding cases. Starting from the structure equations \Ruf{FI} with
4809: respect to some perfect basis
4810: $\4x,\overline{\4x},\4y,\overline{\4y},\4z\,$, we first evaluate
4811: \Ruf{FZ} for this particular value of $\beta_1$. We get $d_1=2$ and
4812: $d_2=\overline b_3+2b_4.$ Next, the Jacobi identity
4813: $\Jac{\4x}{\overline{\4x}}{\4y}=0$ implies $b_4:=\beta_4\in \RR$ and
4814: $a_1=\beta_4.$ Further, $\Jac{\4x}{\4y}{\overline{\4y}}=0$ implies
4815: $b_3=-\beta_4 +i\gamma,$ $\gamma\in \RR.$ Next,
4816: $\Jac{\4x}{\overline{\4x}}{\4y}=0$ determines the value of $a_2:$
4817: $a_2=(3\beta_4^2+\gamma^2)/4-i\beta_4\gamma/2.$ Finally
4818: $\Jac{\4x}{\4y}{\4z}=0$ implies
4819: $[\4z,\4x]=(\quu{3\beta_4}2-\quu{3i\gamma}2)\4z,$ see diagram below:
4820: $$
4821: \diagram{[\4x,\overline\4x]&\;=\;&\4z+\;& \beta_4\4x & -
4822: &\beta_4\overline{\4x} &\;+\;& a_2\4y &\;-\;&\overline a_2\overline
4823: \4y \rlap{\qquad
4824: $a_2=\quu{(3\beta_4^2+\gamma^2)}4-\quu{i\gamma\beta_4}2 $} \cr
4825: [\4x,\overline{\4y}]&\;=\;& &\kern1em -\4x&+& \;\overline\4x &
4826: &+b_3\4y&+&\beta_4\overline{\4y} \rlap{\qquad $b_3=-\beta_4+i\gamma$}
4827: \cr [\4y,\overline{\4y}]&\;=\;& & & & & & \;\;\4y & - &
4828: \;\;\overline{\4y} \cr [\4y,\4x]&\;=\; & & \kern-4pt 2\4x & & &-
4829: &b_4\4y & & \cr [\4z,\4x]&\;= \;\rlap{$
4830: (\quu{3\beta_4}2-\quu{3i\gamma}2) \4z$ \qquad $[\4z,\4y]=-3\4z\;\;$.}
4831: }\qquad\qquad\qquad\qquad\qquad\qquad\Leqno{GX}
4832: $$
4833: Select the following basis of $\7g$:
4834: $$
4835: \diagram{\4n\;&\;:= \;& 2i\4x-2i\overline{\4x} \;-ib_3\4y
4836: \;+\;i\overline b_3\overline{\4y} \qquad\qquad \4s\;:=
4837: -\hbox{\klein$\displaystyle{1\over 4}$}(\4y+\overline{\4y})\cr
4838: \4v_1&:=& \qquad\qquad\kern1.3em \hbox{\Klein$\displaystyle{i\over
4839: 2}$}\4y\;\;- \hbox{\Klein$\displaystyle{i\over 2}$}\overline{\4y}
4840: \hfill\cr \4v_2&:=& \;2\4x + 2\overline{\4x} -\;b_3\4y \;- \;\;\;
4841: \overline b_3 \overline{\4y} \hfill\cr \4v_3&:=&
4842: 4i\4z\;.\vphantom{\big|}\hfill}
4843: $$
4844: A direct computation (using \Ruf{GX}) shows that $\7v:=\RR\4v_1\oplus
4845: \RR\4v_2\oplus \RR\4v_3$ is an abelian ideal and
4846: $$
4847: [\4s,\4n]=\4n, \qquad \diagram{[\4s,\4v_1]=
4848: -\hbox{\klein$\displaystyle{1\over 2}$}\4v_1 , \quad[\4s,\4v_2]=
4849: \hbox{\klein$\displaystyle{1\over 2}$}\4v_2 , \quad[\4s,\4v_3]=
4850: \hbox{\klein$\displaystyle{3\over 2}$}\4v_3 , \cr \kern.6em
4851: [\4n,\4v_1]=\;\4v_2 , \kern2.4em [\4n,\4v_2]= \;\4v_3 , \kern1.6em
4852: [\4n,\4v_3]= \;0\;.\qquad }
4853: $$ Again, this shows that $\7g$ is isomorphic to the semidirect
4854: product $\7v\rtimes \7h$ with $\7h=\RR \4n\oplus \RR \4s$. Actually,
4855: an explicit realization of the corresponding CR-manifold $M$ along the
4856: lines of proof of the Main Lemma shows that $(\7g,\7q)$ is associated
4857: to the tube over the future light cone. \qed
4858:
4859: \medskip We close by stating that Lemma \ruf{GS} together with Lemma
4860: \ruf{VQ} finishes the proof of the classification theorem.
4861:
4862:
4863: \Partskip
4864: {\gross\noindent References\write\lst{\def\csname Refa\endcsname{\folio}}}
4865: \bigskip
4866:
4867: {\Klein
4868: \parindent 15pt\advance\parskip-1pt
4869: \def\Masca#1{Math. Scand. {\bf #1}}
4870: \def\Springer{Ber\-lin-Hei\-del\-berg-New York: Sprin\-ger~}
4871: \def\LecNotes#1{Lecture Notes in Mathematics Vol. #1. \Springer}
4872: \def\Manus#1{ma\-nus\-crip\-ta math. {\bf #1}}
4873: \def\PAMS#1{Proc. Amer. Math. Soc. {\bf #1}} \def\TAMS#1{Transactions
4874: Amer. Math. Soc. {\bf #1}} \def\Quaox#1{Quart. J. Math. Ox\-ford {\bf
4875: #1}}
4876:
4877:
4878: \font\klCC=txmia scaled 900
4879:
4880:
4881: \Ref{ANFR}Andreotti, A., Fredricks, G.A.: Embeddability of real analytic Cauchy-Riemann manifolds. Ann. Scuola Norm. Sup. Pisa Cl. Sci. {\bf 6} (1979), 285--304.
4882: \Ref{ANHI}Andreotti, A., Hill, C.D.: Complex characteristic coordinates and tangential Cauchy-Riemann equations. Ann. Scuola Norm. Sup. Pisa Sci. Fis. {\bf 26} (1972), 299-324.
4883: \Ref{BELO}Beloshapka, V.K.: Symmetries of Real Hypersurfaces in Complex 3-Space. Math. Notes {\bf 78} (2005), 156--163.
4884: \Ref{BHUR}Baouendi, M.S., Huang, X., Rothschild, L.P.: Regularity of CR mappings between algebraic hypersurfaces. Invent. Math. {\bf 125} (1996), 13-36.
4885: \Ref{BERO}Baouendi, M.S., Ebenfelt, P., Rothschild, L.P.: {\sl Real Submanifolds in Complex Spaces and Their Mappings}. Princeton Math. Series {\bf 47}, Princeton Univ. Press, 1998.
4886: \Ref{BAHR}Baouendi, M.S., Huang, X., Rothschild, L.P.: Regularity of CR mappings between algebraic hypersurfaces. Invent. Math. {\bf 125} (1996), 13-36.
4887: \Ref{BARZ} Baouendi, M.S., Rothschild, L.P., Zaitsev, D.: Equivalences of real submanifolds in complex space. J. Differential Geom. 59 (2001), 301--351.
4888: \Ref{BOGG}Boggess, A.: {\sl CR manifolds and the tangential Cauchy-Riemann complex.} Studies in Advanced Mathematics. CRC Press. Boca Raton~ Ann Arbor~ Boston~ London 1991.
4889: \Ref{BOUR}Bourbaki, N.: {\sl \'El\'ements de math\'ematique: groupes et alg\`ebres de Lie.} Chapitre I-IX. Herman, Paris 1960.
4890: \Ref{BUSH}Burns, D., Shnider, S.: Spherical hypersurfaces in complex manifolds. Invent. Math. {\bf 33} (1976), 223--246.
4891: \Ref{CART}Cartan, \'E.: Sur la g\'eom\'etrie pseudo-conforme des hypersurfaces de l'espace de deux variables complexes. Annali di Matematica Pura ed Applicata {\bf 11/1} (1933) 17--90.
4892: \Ref{CHMO}Chern, S.S., Moser, J.K.: Real hypersurfaces in complex manifolds. Acta. Math. {\bf 133} (1974), 219-271.
4893: \Ref{DAYA}Dadok, J.,Yang, P.: Automorphisms of tube domains and spherical hypersurfaces. Amer. J. Math. {\bf 107} (1985), 999-1013.
4894: \Ref{DOKR}Doubrov, B., Komrakov, B., Rabinovich, M.: Homogeneous surfaces in the three-dimenaional affine geometry. In: {\sl Geometry and Topology of Submanifolds, VIII,} World Scientific, Singapore 1996, pp. 168-178
4895: \Ref{EAEZ}Eastwood, M., Ezhov, V.: On Affine Normal Forms and a Classification of Homogeneous Surfaces in Affine Three-Space. Geometria Dedicata {\bf 77} (1999), 11-69.
4896: \Ref{EBFT}Ebenfelt, P.: Normal Forms and Biholomorphic Equivalence of Real Hypersurfaces in \hbox{\klCC\char131}$^3$. Indiana J. Math. {\bf 47} (1998), 311-366.
4897: \Ref{EBEN}Ebenfelt, P.: Uniformly Levi degenerate CR manifolds: the 5-dimensional case. Duke Math. J. {\bf 110} (2001), 37--80.
4898: \Ref{FELS}Fels, G.: Locally homogeneous finitely nondegenerate CR-manifolds. To appear. {\tt arXiv:math.CV/0606032}
4899: \Ref{FEKA}Fels, G., Kaup, W.: CR-manifolds of dimension 5: A Lie algebra approach. J. Reine Angew. Math, to appear. {\ninett http://arxiv.org/pdf/math.DS/0508011}
4900: \Ref{GAME}Gaussier, H., Merker, J.: A new example of a uniformly Levi degenerate hypersurface in \hbox{\klCC\char131}$^3$. Ark. Mat. {\bf 41} (2003), 85--94.
4901: \Ref{HOCH}Hochschild, G.: {\sl The structure of Lie groups.} Holden-Day, Inc., San Francisco-London-Amsterdam 1965
4902: \Ref{KNAP}Knapp, A.W.: {\it Lie groups beyond an introduction}. Second edition. Progress in Mathematics {\bf 140}. Birkh\"auser Boston, Inc., Boston, MA, 2002.
4903: \Ref{HART}Hartshorne, R.: {\sl Algebraic Geometry}. Graduate Texts in Mathematics {\bf 52}. \Springer 1977
4904: \Ref{HORM}H\"ormander, L.: {\sl An Introduction to Complex Analysis in Several Variables}. Third edition. North-Holland Publishing Co. Amsterdam - New York, 1990.
4905: \Ref{ISMI}Isaev, A.V., Mishchenko, M.A.: Classification of spherical tube hypersurfaces that have one minus in the Levi signature form. Math. USSR-Izv. {\bf 33} (1989), 441-472.
4906: \Ref{KAZT}Kaup, W., Zaitsev, D.: On local CR-transformations of Levi degenerate group orbits in compact Hermitian symmetric spaces. J. Eur. Math. Soc. {\bf 8} (2006), 465-490.
4907: \Ref{LOBO}Loboda, A.V.: Homogeneous real hypersurfaces in \hbox{\klCC\char131}$^3$ with $2$-dimensional isotropy groups. Tr. Mat. Inst. Steklova {\bf 235} (2001), 114-142 and Proc. Steklov Inst. Math {\bf 235} (2001), 107-135.
4908: \Ref{LOBP}Loboda, A.V.: Homogeneous nondegenerate surfaces in \hbox{\klCC\char131}$^{3}$ with two-dimensional isotropy groups. (Russian) translation in Funct. Anal. Appl. {\bf 36} (2002), 151--153.
4909: \Ref{LOBQ}Loboda, A.V.: On the determination of a homogeneous strictly pseudoconvex hypersurface from the coefficients of its normal equation. (Russian) translation in Math. Notes {\bf 73} (2003), 419--423.
4910: \Ref{MENA}Medori, C., Nacinovich, M.: Algebras of infinitesimal CR automorphisms. J. Algebra {\bf287} (2005), 234--274.
4911: \Ref{PALA}{Palais, R.~S.:} {\sl A global formulation of the Lie theory of transformation groups.} Mem. AMS 1957.
4912: \Ref{SAKA}Sakai, T.: {\sl Riemannian geometry.} Translations of Mathematical Monographs, 149. American Mathematical Society, Providence, RI, 1996.
4913: \Ref{SEVL}Sergeev, A.G., Vladimirov, V.S.: Complex analysis in the future tube. In: {\sl Several Complex Variables II,} Encyclopedia Math. Sci {\bf 8}, \Springer 1994.
4914: \Ref{TANA}Tanaka, N.: On the pseudo-conformal geometry of hypersurfaces of the space of $n$ complex variables. J. Math. Soc. Japan {\bf 14} (1962), 397-429.
4915: \bigskip
4916: }
4917:
4918: \medskip\centerline{Mathematisches Institut, Universit\"at
4919: T\"ubingen, Auf der Morgenstelle 10, 72076 T\"ubingen, Germany}
4920: \smallskip
4921: \centerline{e-mail: {\ninett gfels@uni-tuebingen.de, kaup@uni-tuebingen.de}}
4922:
4923:
4924: \closeout\aux\closeout\lst\bye
4925:
4926:
4927:
4928: %%% Local Variables:
4929: %%% mode: plain-tex
4930: %%% TeX-master: t
4931: %%% End:
4932:
4933:
4934: