1: \documentclass{amsart}
2: %\documentclass{article}
3: %\documentclass{article}
4: \usepackage{amsfonts}
5: \usepackage[all]{xy}
6: \usepackage{amssymb, amscd}
7:
8: \usepackage{amssymb}
9: \usepackage{graphicx}
10: \usepackage{epstopdf}
11:
12:
13: \usepackage[dvips,bookmarks=false]{hyperref}
14:
15: \newcommand{\dsty}{\displaystyle}
16:
17:
18: %\usepackage[dvips]{graphicx}
19: %*********** myfullpagestyle ********************
20: %\textwidth=460pt\textheight=656pt\oddsidemargin=0pt\evensidemargin=0pt%%%\topmargin=0pt\headheight=0pt\headsep=0pt
21:
22: %******************** \newtheorem
23: \newtheorem{theorem}{Theorem}
24: \newtheorem{lemma}{Lemma}
25: \newtheorem{proposition}{Proposition}
26: \newtheorem{corollary}{Corollary}
27: \newtheorem{definition}{Definition}
28:
29: %******************** \Bbb
30: \newcommand{\R}{{\Bbb R}}
31: \newcommand{\C}{{\Bbb C}}
32: \renewcommand{\H}{{\Bbb H}}
33: \renewcommand{\O}{{\Bbb O}}
34: \newcommand{\Z}{{\Bbb Z}}
35: \newcommand{\vo}{\overrightarrow{\omega}}
36:
37:
38: %******************** \mathcal
39: \newcommand{\D}{\mathcal D}
40: \newcommand{\dperp}{\mathcal D ^{2 \perp}}
41:
42: %\newcommand(\dtp}{{\mathcal D}^{2 \perp}}
43: \newcommand{\I}{{\mathcal I}}
44: \newcommand{\F}{{\mathcal F}}
45: \newcommand{\tF}{\widetilde{\mathcal F}}
46:
47:
48: %******************** \mathfrak
49: \renewcommand{\t}{\mathfrak{t}}
50: \newcommand{\h}{\mathfrak{h}}
51: \renewcommand{\l}{\mathfrak{l}}
52: \newcommand{\s}{\mathfrak{s}}
53: \renewcommand{\k}{\mathfrak{K}}
54: \newcommand{\hp}{{\h}^\perp}
55: \newcommand{\g}{\mathfrak{g}}
56: \newcommand{\p}{\mathfrak{p}}
57: \newcommand{\gl}{\mathfrak{gl}}
58: \renewcommand{\sp}{\mathfrak{sp}}
59: \newcommand{\m}{\mathfrak{m}}
60: %\newcommand{\n}{\mathfrak{n}} oopes used twice
61: \newcommand{\so}{\mathfrak{so}}
62: \renewcommand{\sl}{\mathfrak{sl}}
63:
64: %******************** \rm
65: \newcommand{\End}{{\rm End}}
66: \newcommand{\GL}{{\rm GL}}
67: \newcommand{\Hom}{{\rm Hom}}
68: \newcommand{\Iso}{{\rm Iso}}
69: \newcommand{\Aut}{{\rm Aut}}
70: \newcommand{\Sp}{{\rm Sp}}
71: \newcommand{\SO}{{\rm SO}}
72: \newcommand{\SL}{{\rm SL}}
73: \newcommand{\TS}[1]{H^{2,0}(\g_{#1})}
74: \newcommand{\TSS}{H^{2,0}(\g)}
75: \newcommand{\W}{{\rm\bf W}}
76: \newcommand{\V}{{\rm\bf V}}
77: \newcommand{\E}{{\rm\bf E}}
78: \newcommand{\SU}{{\rm SU}}
79: \newcommand{\x}{{\bf x}}
80: \newcommand{\y}{{\bf y}}
81: \renewcommand{\P}{{\bf P}}
82: \newcommand{\bp}{{\bf p}}
83:
84: \renewcommand{\c}{ {\bf c}}
85:
86:
87: \newcommand{\by}{{\bf y}}
88: \newcommand{\Y}{{\bf Y}}
89:
90:
91: %\renewcommand{\S}{{\rm\bf S}}
92:
93: %******************** Misc
94: \renewcommand{\L}{{\mathcal L}}
95: \newcommand{\marco}[1]{\framebox{$\displaystyle #1 $}}
96: \newcommand{\bs}{\bigskip}
97: \newcommand{\ms}{\medskip}
98: \newcommand{\mn}{\medskip\noindent}
99: \newcommand{\n}{\noindent}
100: \newcommand{\bn}{\bs\n}
101: \newcommand{\bnbf}{\bs\n\bf}
102: \newcommand{\tensor}{\otimes}
103: \newcommand{\w}{\wedge}
104: \newcommand{\X}{\times}
105: \renewcommand{\o}{{\omega}}
106: \newcommand{\ov}{\overrightarrow{\omega}}
107: \newcommand{\pf}{\n\textsc{Proof. }}
108: %\newcommand{\qed}{ $\Box$}
109: \newcommand{\e}{ {\bf e}}
110: \renewcommand{\a}{ {\bf a}}
111: \renewcommand{\b}{ {\bf b}}
112: \newcommand{\tQ}{\widetilde {Q}}
113: \newcommand{\tD}{\widetilde {D}}
114: \newcommand{\tC}{\widetilde {C}}
115:
116:
117: %\usepackage{pdfsync}
118: %package which lets you jump back and forth from pdf and tex files easily
119: %getting to the corresponding place within the tex file
120:
121: \title{$G_2$ and the ``rolling distribution''}
122:
123: \author{ Gil Bor and Richard Montgomery}
124:
125: %\input{pdfsync}
126:
127: \begin{document}
128: \maketitle
129:
130: \tableofcontents
131:
132: \section*{0 Introduction}
133:
134:
135: \begin{figure}[h]
136: \centerline{\includegraphics[width=8cm, angle=90]{roll2.eps}}
137: \caption{Rolling a ball on another ball.}\label{roll}
138: \end{figure}
139:
140:
141:
142:
143: Consider two balls of different sizes,
144: rolling on each other, without slipping or spinning.
145: % It may help to think of the balls as covered in Velcro.
146: The
147: configuration space for this system is a 5-dimensional manifold
148: $Q \cong \SO_3\times
149: S^2$ on which the no-slip/no-spin condition defines a rank 2
150: distribution $D\subset TQ$, the ``rolling-distribution''.
151:
152: Now $D$ is a non-integrable distribution (unless the balls are of equal size) which has an ``obvious'' 6-dimensional transitive symmetry group
153: $\SO_3\times\SO_3$ arising from the isometry groups of each ball, but for balls whose radii are in the ratio 3:1, and only for this ratio,
154: something strange happens: the {\it local } symmetry group
155: of the distribution increases from $\SO_3\times\SO_3$ to $G_2$, a
156: 14-dimensional Lie group.
157:
158: More precisely, let $\g_2$ be the real ``split
159: form" of the complex $14$-dimensional exceptional simple Lie algebra $\g_{2 \C}$.
160: There are precisely two connected Lie groups whose Lie algebras
161: are $\g_2$. (See Appendix \ref{twoG2s}.) We choose the one corresponding to the adjoint
162: representation, and call it $G_2$. (The other one is $\tilde G_2$,
163: the universal cover of the one we chose.)
164: $G_2$ is a subgroup of $\SO(3,4)$ and its maximal compact subgroup $K\subset G_2$ is isomorphic to $\SU_2\times \SU_2/\{\pm(1,1)\}$ which double-covers $\SO_3\times\SO_3$.
165: Let $\tQ=S^3\times S^2$ be
166: the universal cover of $Q$ equipped with the distribution $\tD$ induced by the double covering $\tQ \to Q$.
167: Let $\Aut(\tQ,\tD)$ be the group of diffeomorphisms of $\tQ$ leaving $\tD$ invariant.
168: Then we have
169:
170: \begin{theorem} The connected component of the identity in $\Aut(\tQ,\tD)$ for radius ratio 3:1 or 1:3 is isomorphic to $G_2$. The $G_2$ action on $\tQ$
171: does not descend to $Q$, but its restriction to the maximal compact $K \subset G_2$ does, covering the $\SO_3 \times \SO_3$ action on $Q$. For any other radius ratio (other then 1:1) $\Aut(\tQ,\tD)$ is ismorphic to $K$.
172: \end{theorem}
173:
174: This theorem was communicated to us by Robert Bryant, for whom it is but a variation on a theme of E. Cartan's work on the method of equivalence, contained in his notoriously difficult ``Five Variables Paper" \cite{C1} from 1910 . Bryant wrote to us recently:
175: \begin{quote}
176: ``Cartan himself gave a geometric description of the flat
177: $G_2$-structure as the differential system that describes space
178: curves of constant torsion $2$ or $1/2$ in the standard unit 3-sphere.
179: (See the concluding remarks of Section 53 in Paragraph XI in the
180: Five Variables Paper.) One can easily transform the rolling balls
181: problem (for arbitrary ratios of radii) into the problem of curves
182: in the 3-sphere of constant torsion and, in this guise, one can
183: recover the 3:1 or 1:3 ratio as Cartan's torsion 2 or 1/2 with
184: a minimum of fuss. Thus, one could say that Cartan's calculation
185: essentially covers the rolling ball case.''
186: \end{quote}
187:
188: \bn\centerline{$*\qquad*\qquad*$}
189:
190: \bn
191:
192: Our main purpose in this note is to try to explain this beautiful and mysterious theorem
193: in a direct manner which does not appeal to Cartan's method of
194: equivalence. We consider it an expansion of Section 4 in
195: Bryant's lecture notes \cite{Bryant}. Our contribution consists basically of a description of two constructions of $(\tQ, \tD)$ with
196: a built-in $G_2$-invariance. Using these constructions we show here that, for radius ratio 3:1 or 1:3, $G_2$ is {\em contained} in $Aut(\tilde Q, \tilde D)$. But
197: we do not know how to show, without the more sophisticated Cartan's methods
198: (or its variants such as those of Tanaka)
199: that $G_2$ is the {\em full} identity component
200: of $Aut(\tilde Q, \tilde D)$,
201: nor that for radius ratio different from 3:1, 1:3 or 1:1, $Aut(\tilde Q, \tilde D)$ is not larger than $K$.
202:
203:
204: A secondary purpose of this article is to correct an error appearing in the book
205: \cite{Montbook} by one of us. We had mistakenly said there that the symmetry group for the rolling distribution for a ball on a plane (ratio $1:\infty$) was $G_2$.
206:
207: A tertiary purpose is to obtain a bit of a feel for the simplest
208: exceptional Lie algebra $\g_2$ and its Lie groups, and
209: to provide a refresher course on roots and weights.
210:
211:
212: \bn {\bf Structure of Paper.}
213: In the next section (section 1) we describe the background and wider context of the problem,
214: with references to the literature. In section 2 we give a detailed description
215: of the distributions associated with the rolling of balls, noting their $\SO_3\times\SO_3$-symmetries.
216: In section 3 we describe the homogeneous distributions of a Lie group $G$
217: in terms of data $(G, H, W)$, where $H\subset G$ is a closed subgroup and $W\subset\g/\h$ is an $H$-invariant subspace.
218: We then identify this data for the rolling distribution with respect the group $G=\SO_3\times\SO_3$. In section 4 we use the root diagram of $G_2$ to give our first construction of a $G_2$-invariant distribution data $(G_2, P, W)$. The identification of the resulting $G_2$-homogenous distribution on $G_2/P$ with $(\tilde Q, \tilde D)$
219: amounts to the embedding of
220: $\so_3\times\so_3$ in $\g_2$ and is the subject of section 5 (and Appendix B) which forms the heart
221: of this article.
222: In section 6 we give the second $G_2$-invariant construction of $(\tilde Q, \tilde D)$ an
223: explicit construction applying projective geometry to the space of
224: purely imaginary split octonions $V$, the lowest dimensional non-trivial representation space
225: for $G_2$. Appendix C is historical. Following suggestions by Bryant
226: we looked into Cartan's thesis and found that much of content of section 6,
227: and hence of the rolling distribution already appears there.
228:
229:
230:
231:
232:
233: \bn\centerline{$*\qquad*\qquad*$}
234:
235: \bn
236:
237:
238: Despite all our efforts, the ``$3$'' of the ratio $1:3$ remains
239: mysterious. In this article it simply arises out of the structure constants for
240: $G_2$ and appears in the construction
241: of the embedding of $\so_3\times\so_3$ into $\g_2$ (section 5 and Appendix B).
242: Algebraically speaking, this `3' traces back to the 3 edges
243: in $\g_2$'s Dynkin diagram
244: and the consequent relative positions of the long and short
245: roots in the root diagram (see figure 2 below) for $\g_2$ which the Dynkin diagram is encoding.
246:
247: \bn {\bf Open problem.} Find a geometric or dynamical
248: interpretation for the ``$3$'' of the $3:1$ ratio.
249:
250: \bn
251:
252: For work in this direction see Agrachev \cite{Agrachev} and also Kaplan and Levstein
253: \cite{Kaplan}.
254:
255:
256:
257:
258: \bn {\bf Acknowledgements.}
259: Robert
260: Bryant has been crucial, at various key steps along the way,
261: in steering us in the right direction. Marti Weissmann
262: supplied us with key information regarding $G_2$, and the
263: crucial Vogan reference.
264:
265: \section{History and Background}
266:
267: {\bf On distributions.} By a distribution we mean here a linear subbundle of
268: the tangent bundle of a manifold.
269: The distributions first encountered are usually
270: the integrable and the contact distributions and have infinite dimensional symmetry groups. In dimension
271: $5$ we first encounter distributions whose symmetry groups
272: are finite-dimensional. Indeed, the generic distribution of rank $2$ or $3$ in $5$ dimension has {\em no} local symmetries.
273: Cartan \cite{C1} investigated rank $2$ and
274: $3$ distributions in $5$ dimensions in detail.
275: The growth vector of
276: a generic rank $2$ distribution on a $5$-dimensional manifold, at a generic
277: point of that manifold, is $(2, 3, 5)$.
278: This ``$(2,3,5)$ at a point''
279: means that if $X, Y$ are any local vector fields spanning
280: the distribution in a neighbhorhood of the point , then $[X,Y] = Z$
281: is pointwise linearly independent of $X, Y$ (in a neighborhood of the point)
282: and $X, Y, Z, [X, Z], [Y, Z]$ span the tangent bundle in a neighborhood of the point.
283: Cartan worked out the
284: complete local invariants -- analogues of the Riemann curvature
285: tensor -- for these $(2,3,5)$ distributions.
286: For the distribution's symmetry group to act transitively all of
287: Cartan's invariants must be constant. To get the maximal dimensional symmetry group
288: all Cartan's invariants must vanish, in which case we call
289: the distribution ``flat''. Any such distribution
290: is locally diffeomorphic to that of the ``Carnot group'' distribution
291: associated to the unique graded nilpotent Lie group $\mathfrak{n} = \mathfrak{n}_{2,3,5}$ of this same growth, and its local symmetry
292: algebra is $\g_2$.
293: (By the ``local symmetry algebra'' of a distribution we mean the algebra of vector fields
294: $X$ satisfying $[X, \Gamma(D)] \subset \Gamma(D)$ where $\Gamma(D)$ is the sheaf of local sections of vector fields tangent to the distribution.)
295:
296: As mentioned in the above quote from Bryant, Cartan \cite{C1}
297: presented several geometric realizations of the flat case. Bryant and Hsu
298: \cite{BryantHsu} (see section 3.4) pointed out the
299: rolling incarnation of $G_2$. A $(2,3,5)$
300: distribution will arise whenever one rolls one Riemannian surface on another
301: provided their Gaussian curvatures are not equal.
302: The Cartan invariants vanish if and only if the ratio of their curvatures are $1:9$ .
303: Hence the $1:3$ radii for spheres. We could also achieve the maximal
304: local symmetry algebra $\g_2$ by
305: rolling two hyperbolic planes along each other, provided their ``radii''
306: are in the ratio $i: 3i$.
307: More history, and more instances of the flat $G_2$ system
308: are explained in Byrant \cite{Bryant}.
309:
310:
311: Non-integrable rank $2$ distributions in dimension $n$ ($n > 3$) admit special families
312: of integral curves known as ``singular'' or ``abnormal'')
313: (\cite{Montbook}). These are curves which
314: admit no local variations through integral curves and having endpoints fixed. In the case of $(2,3,5)$ distributions
315: there is precisely $1$ singular curve (up to reparameterization)
316: through every point {\it in every direction} tangent to $D$. In the case of
317: rolling one Riemannian surface along another, these singular curves correspond
318: to rolling along geodesics. Using the symplectic geometry associated to
319: variations of singular curves
320: Zelenko and Agrachev have been able to
321: rederive Cartan's (2,3,5) invariants. See \cite{Agrachev} and references therein.
322:
323: Tanaka and his school have established a wonderful
324: generalization of the passage from the flat nilpotent model $\mathfrak{n}_{2,3,5}$
325: to $\g_2$.
326: Associated to each point $p$ of a manifold endowed with a non-integrable
327: distribution there is a graded nilpotent Lie algebra $\m = \m (p)$ called the
328: `nilpotentization' of the distribution, or sometimes the ``symbol algebra''. The dimension of $\m$ is that of the underlying manifold.
329: Call the distribution ``of type $\m$'' if the different algebras $\m(p)$
330: are all isomorphic to the same $\m$, i.e. the isomorphism type does not change
331: from point to point.
332: (Every (2,3,5) distribution is of type $\mathfrak{n}_{2,3,5}$.) Associated to each
333: graded nilpotent $\m$ there is graded Lie algebra $\g \supset \m$, possibly infinite dimensional,
334: called the `prolongation' of $\m$ and built from $\m$ in a purely algebraic manner.
335: This $\g$ represents, roughly speaking,
336: the maximal possible symmetry of a distribution of type $\m$:
337: every symmetry algebra for a type $\m$-distribution, after applying a
338: grading process to it,
339: must be a subalgebra of $\g$. The prolongation of
340: the (2,3,5) algebra is $\g_2$, and this fact can be viewed
341: as the algebraic restatement of Cartan's work on the flat model.
342: This Tanaka prolongation method thus yields a proof that $Aut(\tilde Q, \tilde D)
343: \subset G_2$ in theorem 1, alternative to Cartan's proof.
344: Yamaguchi \cite{Yamaguchi} has classified all $\m$'s whose $\g$'s are simple.
345: To each of these pairs $(\m, \g)$ is associated an intricate differential geometry
346: and most of these have not been explored in any detail.
347:
348:
349:
350:
351:
352: \bn{\bf On $G_2$.} The Lie algebra $\g_2$ is
353: the smallest of the exceptional simple Lie algebras.
354: In 1894 Killing uncovered
355: the existence of the root lattice for $\g_2$'s,
356: but without establishing the existence of the
357: corresponding Lie algebra. Cartan, in his thesis, established the existence of $\g_2$
358: in one page of his thesis \cite{C2}. He did so by constructing
359: the $7$-dimensional representation of $\g_2$, in a way which is closely related to our second ``projective split octonion''
360: model for $\tilde Q$, the universal cover of the rolling space.
361: We have devoted appendix C to this page of his thesis
362: and its connection with this second model.
363: In 1914
364: Cartan \cite{C3} showed that $G_2$ can be realized as the automorphism
365: group of the octonions. For our split $G_2$ he used `split
366: octonions'. The compact form of $G_2$
367: appears in the Berger list of
368: potential holonomy groups of Riemannian metrics. Recently, the compact $G_2$ has been
369: featured in string theories, but perhaps that fad has passed already.
370:
371:
372:
373: \section{Distribution for rolling balls}
374:
375: \subsection{The distribution}
376:
377: \bn
378:
379: Take the first ball to be stationary, of radius $R$, with its center at the
380: origin. Roll a second ball of radius $r$ on the first ball. The position of the
381: second ball is given by an isometry (rigid motion)
382: $\varphi_{(g,\x)}:\R^3\to\R^3$, mapping a point $\P$ to
383: $$\bp = \varphi_{(g,\x)}(\P)=g\P+(R+r)\x,$$ where $(g,\x)\in\SO_3\times
384: S^2$. Here, $R\x$ is the point of contact of the two balls,
385: $(R+r)\x$ is the center of the second ball and $g\in\SO_3$ describes the
386: rotation of the second ball relative to its initial position. See figure 1 in the introduction. Thus the configuration
387: space $Q$ for our rolling problem has been identified with the manifold
388: $\SO_3\times S^2$. (For a visceral account of rolling a sphere on a plane,
389: accessible to upper division undergraduates, we recommend
390: \cite{Hammersley}.)
391:
392:
393:
394: %[PICTURE MISSING]
395:
396: %\includegraphics[width=10cm, angle=90]{roll2.eps}
397: %Figure 1.
398: %\end{picture}
399: %\end{center}
400: \bn
401:
402:
403:
404:
405: Let $(g_t,\x_t)\in Q$ be a differentiable rolling motion. Let $\o_t\in
406: \R^3\cong\so_3$ be the angular velocity of the rolling ball relative to
407: its center, measured with respect to
408: inertial axes. In other words, if $\P$ is a material point fixed
409: on the second ball, $\dot\P=0$, and if
410: we write $\bp_t=g_t\P$, then $\dot \bp=\dot gg^{-1}\bp=\omega\times
411: \bp$. Then we have
412:
413:
414: \begin{proposition} Let $Q=\SO_3\times S^2$ be the configuration space of two rolling balls of
415: radii $R$ and $r$. Let $\rho=R/r$ .
416: Then a curve $(g_t,\x_t)\in Q$ describes a rolling motion without slipping and spinning iff
417: \mn \begin{quote}
418: (1) $(\rho+1)\dot \x=\omega\times \x$ (no-slip condition),
419:
420: (2) $\langle\omega,\x\rangle=0$ (no-spin condition,
421: i.e. $\omega$ need to be tangent to the stationary ball at $R\x$).
422: \end{quote}
423: \end{proposition}
424:
425:
426: \pf (1) The contact point between the two balls is $\bp=R\x$ on the
427: first ball, $\P=-g^{-1}r\x$ with respect to the second ball. For
428: non-slip, their velocities must match: $\dot\bp= g\dot\P.$ Now
429: $\dot\bp=R\dot\x$ and
430: $$\dot\P= [-{d\over dt}g^{-1}]r\x-g^{-1}r\dot\x=g^{-1}\dot gg^{-1}r\x-g^{-1}r\dot\x=g^{-1}r(\o\X\x-\dot\x),$$
431: hence the non-slip condition $\dot\bp= g\dot\P$ is equivalent to $R\dot\x=r(\o\X\x-\dot\x),
432: $
433: from which (1) follows.
434:
435: \mn (2) Let $\bf P$ be a material point {\em fixed} on the second ball
436: ($\dot\P=0$). From the inertial point of view, which is to say,
437: from the point of view of the first ball with origin at its center,
438: the position of this material point is $\bp=g\P+(R+r)\x$, and
439: so its velocity
440: $$\dot\bp=\dot g\P+(R+r)\dot\x=\dot g
441: g^{-1}[\bp-(R+r)\x]+(R+r)\dot\x=\o\X[\bp-(R+r)\x]+(R+r)\dot\x.$$ Using
442: the no-slip equation, $(R+r)\dot \x=r\omega\times \x$, we get
443: $$\dot\bp=\o\X[\bp-(R+r)\x]+r\o\X\x=\o\X(\bp-R\x).$$
444: The equation
445: $\dot\bp=\o\X(\bp-R\x)$ asserts that the instantaneous motion of the second ball is
446: a rotation whose
447: axis of rotation (a line) passes through the point of contact $R\x$,
448: in the direction of $\o$ and with angular velocity of magnitude
449: $\|\o\|$. The no-spin condition is
450: that the second ball does not spin about
451: the point of contact of the two balls, which is to say that $\o$ should have no
452: component orthogonal to the common tangent plane of the two balls,
453: i.e. $\langle\omega,\x\rangle=0$. \qed
454:
455: The two conditions in the last Proposition define together a rank 2 distribution on $Q$. This is the rolling distribution.
456:
457: \vskip .3cm
458:
459:
460: \subsection{The ``obvious'' symmetry}
461:
462: The group $\SO_3\times\SO_3$ acts on $Q$ by
463: $\varphi_{(g,\x)}\mapsto g'\circ\varphi_{(g,\x)}\circ g''^{-1},$ where
464: $g',g''\in\SO_3.$ In terms of $(g,\x)$ this
465: action is
466: $$(g,\x)\mapsto (g' g g''^{-1},g'\x), \quad g',g''\in\SO_3.$$
467: This action is transitive and preserves the rolling distribution $D$ for any value of $\rho=R/r$. The proofs of these assertions are easy and left as exercises.
468:
469:
470: \section{Group theoretic description of the rolling distribution
471: }
472:
473: \bn
474:
475: In the previous section we wrote down a distribution $D$ on $Q=\SO_3\times S^2$, depending on the real parameter $\rho$. We showed
476: that $Q$ admits an $\SO_3\times\SO_3$-transitive action which preserves the distribution. Our aim in this paper is to show that for two specific values of the parameter, $\rho=3$ and $\rho=1/3$, the distribution admits a larger {\em local} group of symmetries, namely the group $G_2$. We do so by defining a $G_2$-homogeneous space $\tQ=G_2/P$, together with a $G_2$-invariant rank 2 distribution $\tD$ on it.
477: We then define a 2:1 covering map $\tQ\to Q$ which maps $\tD$ to $D$. Furthermore, the group $G_2$ contains a maximal compact subgroup $K\subset G_2$ which is a double cover of $\SO_3\times\SO_3$, such that the map $\tQ\to Q$ is $K$-equivariant (with respect to the covering homomorphism $K\to\SO_3\times\SO_3$).
478: The constructions
479: are most easily done on the group level or on the Lie algebra level. We describe in what follows the general set up required for ``working on the group level'' and
480: then calculate the group theoretic data corresponding to the rolling distribution.
481:
482: \bn
483:
484: Let $G$ be a Lie group. A ``$G$-homogeneous distribution'' is a pair $(Q,D)$ where $Q$ is a manifold on which $G$ acts transitively and $D\subset TQ$ is a $G$-invariant distribution. Fixing a base point
485: $q_0\in Q$ with isotropy $H\subset G$ we obtain a $G-$equivariant identification $ Q\cong G/H$ and an $H$-equivariant identification $T_{q_0}Q \cong \g/\h$
486: where $\h \subset \g$ denote the Lie algebras corresponding to $H \subset G$.
487: Then $D_{q_0}\subset T_{q_0}Q$ corresponds to an $H$-invariant subspace $W\subset\g/\h$.
488: In this way every $G$-homogeneous distribution $(Q,D)$ corresponds to data
489: $(G,H,W)$, where
490: $H\subset G$ is a closed subgroup with Lie algebra $\h\subset\g$ and
491: $W\subset\g/\h$ is an $H$-invariant subspace.
492: The adjoint action of $G$ defines an equivalence relation on the set of pairs $(H,W)$ so that
493: different choices of base points on $Q$ correspond to equivalent pairs $(H,W)\sim(H',W')$.
494: Conversely, given the data $(G,H,W)$, we can construct a $G$-homogeneous distribution $(Q,D)$ by letting $G$ act by left translations on the
495: right $H$-coset space $Q:=G/H$, and define a $G$-invariant distribution
496: $D\subset TQ$ using the $G$-action to push $D_{[e]}:=W\subset\g/\h\cong
497: T_{[e]}(G/H)$ around to all other points of $Q$.
498:
499:
500:
501: On the level of Lie algebras, the data $(\g,\h,W)$ determines $(Q,D)$
502: up to a cover. If, as in our case of $\g= \so_3 \oplus \so_3$, the simply
503: connected Lie group $G$ realizing $\g$ is compact,
504: then there are only finitely many homogeneous distributions $(G, H, W)$ which realize the given Lie algebraic data $(\g, \h, W)$.
505:
506:
507:
508: \bn
509:
510: We now determine the data $(G,H,W)$ corresponding to the rolling distribution
511: $(Q,D)$
512: of section 2.1. Here $G=\SO_3\X\SO_3$, $Q=\SO_3\times S^2$, $\dim H=1$, $\dim W =2$.
513: Identify the Lie algebra of $\SO_3\times\SO_3$ with $\R^3\times\R^3$, the set of pairs of angular velocities $(\omega', \omega'')$, with
514: Lie bracket given by the cross product:
515: $$[(\omega', \omega''),(\eta', \eta'')]=(\omega'\X\eta',
516: \omega''\X\eta'').$$ The first factor $\omega '$ corresponds to the
517: first (stationary) sphere, of radius $R$, while the second $\omega ''$ factor corresponds to second (rolling) sphere of radius $r$.
518:
519: Fix a base point, say $(1,\e_3)\in\SO_3\times S^2=Q$. The
520: isotropy at this base point is the circle subgroup $H$ consisting of elements of the form $(h,h)$, where $h$ is a
521: rotation around the $\e_3$ axis, so $\h=\R(\e_3,
522: \e_3)\subset\R^3\X\R^3.$ Using the Killing metric on $\g=\so_3\X\so_3=\R^3\X\R^3$ we can
523: identify $\g/\h\cong\h^\perp$, so that the plane of the distribution at the
524: base point is given by some 2-plane in $\h^\perp.$ Let us
525: determine explicitly this 2-plane.
526:
527:
528:
529: \begin{proposition} The rolling distribution on $\SO_3\times S^2$
530: corresponding to rolling a ball of radius $r$ along one of radius $R$ is given by the 2-plane
531: in $\R^3\times\R^3$ (the Lie
532: algebra of $\SO_3\X\SO_3$) defined by the equations
533: $$\langle\omega',\e_3\rangle=\langle\omega'',\e_3\rangle=0, \quad
534: \rho \omega'+\omega''=0,$$
535: where $\rho = R/r$.
536: \end{proposition}
537:
538: \pf Since $\h\subset\R^3\times\R^3$ is generated by the vector $(\omega', \omega'')=(\e_3,\e_3)$ and the Killing metric corresponds to some multiple of the standard metric on $\R^3\times\R^3$, $\h^\perp\subset \R^3\times\R^3$ is given by the equation $\langle\omega'+\omega'',\e_3\rangle=0.$
539:
540:
541:
542:
543: From the formula for the $\SO_3\X\SO_3$-action in \S2.2 we get the
544: infinitesimal action at the base point $$\omega=\omega'-\omega'',
545: \quad\dot \x=\omega'\X \e_3.$$ Substituting these into the
546: rolling conditions at the base point (see \S2.1),
547: $$\langle\omega,\e_3\rangle=0, \quad (R+r)\dot \x= r\omega\times \e_3,$$
548: we obtain
549: $$\langle\omega'-\omega'',\e_3\rangle=0, \quad [R\omega'+ r\omega'']\times \e_3=0.$$
550: Adding the condition of orthogonality
551: to $\h$, $\langle\omega'+\omega'',\e_3\rangle=0,$
552: we obtain the above equations.$\Box$
553:
554: \bn
555:
556: \subsection{Shrinking the group.}
557: \label{shrinking} The following observation will be key to proving
558: that part of theorem 1 which we are going to prove, namely
559: that $G_2 \subset Aut(\tilde Q, \tilde D)$ for $\rho=3$ or $1/3$.
560: Suppose that $(Q,D)$ is a $G$-homogeneous distribution with $G$-data $(H,W)$.
561: Let $G_1\subset G$ be a subgroup for which the restriction of the $G$-action on $Q$ to $G_1$ is still transitive.
562: Then $(Q,D)$ is also $G_1$-homogeneous distribution and its $G_1$-data is $(H_1,W_1)$ where $H_1=H\cap G_1$ and $W_1\subset \g_1/\h_1$
563: corresponds to
564: $W$ under the linear isomorphism $\g_1/\h_1\to\g/\h$
565: induced by the diffeomorphism $Q = G/H = G_1/ H_1$.
566: Since $(Q, D)$ has not been changed, it follows that
567: the $G$-data $(H, W)$ and the $G_1$-data $(H_1, W_1)$
568: yield diffeomorphic manifolds with distributions.
569: At the Lie algebra level, this discussion asserts that
570: $(\g_1, \h_1, W_1)$ and $(\g, \h, W)$ define manifolds-with-distributions
571: which are diffeomorphic up to a cover. To prove that
572: $G_2 \subset Aut(\tilde Q, \tilde D)$ we will be applying this observation to
573: the case $\g_1 = \so_3 \oplus \so_3 \subset \g_2$.
574:
575:
576:
577:
578: \section{A $G_2$-homogeneous distribution }
579:
580:
581:
582: \bn
583:
584: We now describe the other main actor in this paper,
585: a distribution with Lie algebraic data $(\g_2, \p, W)$.
586: Please see the root diagram of $\g_2$ in figure 2. This diagram will be explained
587: immediately below. The decorations on the diagram are used to indicate
588: the Lie algebraic data and will be explained a bit later.
589:
590: \bn
591:
592: \begin{center}
593: \unitlength=1cm
594: \begin{picture}(5,4)(-2,-2)
595: \put(0,0){\circle*{.4}}
596:
597:
598: \put(.4,.75){$\oplus$}
599: \put(-.5,-.866){\circle*{.25}}
600: \put(.6,1.2){$\sigma_1$}
601: \put(-1.2,-1.3){$-\sigma_1$}
602:
603: \put(-.5,.866){\circle*{.25}}
604: \put(.4,-.95){$\oplus$}
605: \put(-.95,1.2){$\sigma_2$}
606: \put(.3,-1.3){$-\sigma_2$}
607:
608: \put(-1,0){\circle*{.25}}
609: \put(1,0){\circle{.25}}
610: \put(1.3,-0.1){$\sigma_3$}
611: \put(-2,-0.1){$-\sigma_3$}
612:
613:
614: \put(1.4,-.866){\circle{.25}}
615: \put(-1.5,.866){\circle*{.25}}
616: \put(1.6, -1.1){$-\lambda_1$}
617: \put(-2.2, 0.866){$\lambda_1$}
618:
619: \put(1.4,.866){\circle{.25}}
620: \put(-1.5,-.866){\circle*{.25}}
621: \put(1.7,.9){$\lambda_2$}
622: \put(-2.4,-1.1){$-\lambda_2$}
623:
624: \put(0,1.5){\circle*{.25}}
625: \put(0,-1.5){\circle*{.25}}
626: \put(-.2,1.8){$\lambda_3$}
627: \put(-0.5,-2.1){$-\lambda_3$}
628:
629: \put(0,-1.4){\line(0,1){3}}
630: \put(-1.5, -0.866){\line(5,3){2.77}}
631: \put(-1.5, 0.866){\line(5,-3){2.77}}
632:
633: \put( -1, 0){\line(1 , 0 ){1.85}}
634: \put(-.5,-.866){\line(3,5){.96}}
635: \put(-.5,.866){\line(3,-5){.96}}
636: \end{picture}
637:
638: \bn
639:
640: {\small Figure 2: The root diagram of $\g_2$.}
641: \end{center}
642:
643:
644: \bn{\bf A reminder of the meaning of the root diagram.} The plane in which the diagram is drawn is the dual of a Cartan subalgebra $\t\subset\g_2$. A Cartan
645: subalgebra of a semi-simple Lie algebra $\g$ is a maximal abelian
646: subalgebra $\t\subset\g$ of semi-simple elements, i.e. each
647: $ad(T)\in\End(\g)$, $T\in\t$, is diagonalizable. In the case of
648: $\g=\g_2$, $\t$ is 2-dimensional, hence the subscript 2 in $G_2$, the
649: {\em rank} of the group. The root diagram of $\g$ encodes the adjoint
650: action of $\t$ on $\g$, from which one can recover the whole structure
651: of $\g$.
652:
653: The commutativity of the Cartan
654: subalgebra $\t$ implies that the diagonalizable endomorphisms
655: $ad(T)\in\End(\g)$, $T\in\t$, are {\em simultaneously} diagonalizable,
656: resulting in a $\t$-invariant decomposition
657: $$\g=\t\oplus\sum_\alpha\g_\alpha,$$
658: where each $\g_\alpha\subset\g$ is a 1-dimensional subspace of $\t$-common eigenvectors called a {\em root space}.
659: The corresponding eigenvalue depends linearly on
660: the acting element of $\t$, so is given by a linear functional $\alpha\in\t^*$, called {\em root}. Thus
661: $$[T,X]=\alpha(T)X, \quad T\in\t,\quad X\in\g_\alpha.$$
662:
663: When we draw the root diagram in $\t^*$
664: we use the Killing metric in $\g$ to determine the size of the roots
665: and especially the angles between them. The Killing metric in $\g$ is the inner
666: product $\langle X,Y\rangle=-{\rm tr}(ad(X)ad(Y))$. It is
667: non-degenerate (this is equivalent to semi-simplicity) and its
668: restriction to $\t$ is positive definite.
669:
670:
671: \bn{\bf Example of $\g=\sl_3(\R).$} The more familiar example of
672: $\sl_3 (\R)$ is useful to keep
673: in mind before proceeding with $\g_2$. The Lie algebra $\sl_3(\R)$ is the
674: vector space of
675: of 3 by 3 traceless real matrices with Lie bracket the usual matrix Lie bracket.
676: It is Lie algebra of the Lie group $\SL_3(\R)$ of 3 by 3 real matrices with determinant $1$. Like $\g_2$, the Lie
677: algebra $\sl_3(\R)$ is a non-compact split form of its
678: complexification ($\sl_3(\C)$) and has rank 2. We take as a Cartan
679: subalgebra the subspace $\t\subset\sl_3(\R)$ of traceless diagonal
680: matrices,
681: $$\t:=\{\left(\begin{array}{ccc} t_1&0&0\\0& t_2&0\\0&0& t_3\end{array}\right) | t_1+ t_2+ t_3=0, t_i\in\R\}.$$
682:
683: $\sl_3(\R)$ has 6 roots:
684: $$\alpha_{ij}:= t_i- t_j\in\t^*, \quad i \neq j, \quad
685: i,j\in\{1,2,3\},$$ with corresponding root spaces
686: $$\g_{\alpha_{ij}}=\R E_{ij}, $$ where $E_{ij}$ is the matrix whose
687: $ij$ entry is 1 and all of whose other entries are 0. The corresponding root space
688: decomposition $$\sl_3=\t\oplus\sum_{i\neq j}\g_{\alpha_{ij}},$$ is
689: just the decomposition of a matrix as a diagonal matrix plus its off
690: diagonal terms. The metric induced on $\t$ by the Killing metric is
691: some multiple of the standard euclidean metric, so that $\langle T,
692: T'\rangle=c\sum_i t_i t'_i$ for some $c>0$.
693:
694:
695:
696: \bn
697:
698: \begin{center}
699: \unitlength=1cm
700: \begin{picture}(5,4)(-2,-2)
701: \put(0,0){\circle*{.4}}
702:
703: \put(-1.5,.866){\circle*{.25}}
704: \put(1.3,-1){$\oplus$}
705: \put(-2.3,1){$\alpha_{12}$}
706: \put(1.2,-1.3){$-\alpha_{12}$}
707:
708:
709:
710: \put(0,1.5){\circle*{.25}}
711: \put(0,-1.5){\circle{.25}}
712: \put(-.3,1.8){$\alpha_{13}$}
713: \put(-.5,-2){$-\alpha_{13}$}
714:
715:
716:
717: \put(1.4,.866){\circle*{.25}}
718: \put(-1.7,-1){$\oplus$}
719: \put(1.6,1){$\alpha_{23}$}
720: \put(-2.2,-1.3){$-\alpha_{23}$}
721:
722: \put(-1.4, -0.8){\line(5,3){2.9}}
723: \put(0,-1.4){\line(0,1){3}}
724: \put(-1.5, 0.866){\line(5,-3){2.8}}
725: \end{picture}
726:
727: \bn
728:
729: {\small Figure 3: The root diagram of $\sl_3$}
730:
731: \end{center}
732:
733:
734: \bn
735:
736:
737:
738:
739:
740:
741: \bn{\bf Reading the root diagram}. One can read much of the structure
742: of $\g$ from its root diagram in a formula-free manner. Here
743: is the key observation. Let $\alpha, \beta$ be two roots with
744: (non-zero) root vectors $E_\alpha\in\g_\alpha,$ $ E_\beta\in\g_\beta$.
745: That is, $$[T,E_\alpha]=\alpha(T)E_\alpha,\quad T\in\t, $$ and
746: similarly for $\beta$. It then follows immediately from the Jacobi
747: identity that
748: $$[T,[E_\alpha, E_\beta]]=(\alpha+\beta)(T)[E_\alpha, E_\beta].$$ This means that
749: \begin{itemize}
750:
751: \item[(1)] if $\alpha+\beta\neq 0$ and is not a root then $[E_\alpha, E_\beta]=0$;
752:
753: \item[(2)] if $\alpha+\beta\neq 0$ and is a root then $[E_\alpha, E_\beta]\in\g_{\alpha+\beta}$;
754:
755: \item[(3)] if $\alpha+\beta= 0$, i.e. $\beta=-\alpha$, then $[E_\alpha, E_\beta]\in\t$.
756: \end{itemize}
757:
758: This set of 3 conclusions permit us to see at a glance from the
759: diagram a fair amount of the structure of $\g$. In the last two cases
760: one can further show that $[E_\alpha, E_\beta]$ is non-zero and
761: determine, with some calculations, the actual bracket, as will be
762: illustrated in Appendix B.
763:
764: \bn{\bf Example: reading the root diagram of $\sl_3$.} Let us consider
765: the subspace $\p\subset\sl_3$ spanned by $\t$ and the root spaces
766: corresponding to the roots marked with dark dots in figure 3.
767:
768:
769:
770: The diagram shows that $\p$ is a 5-dimensional subalgebra, i.e. it is
771: closed under the Lie bracket (there are 4 dark dots, but remember that
772: the thick dot at the origin stands for the 2-dimensional Cartan
773: subalgebra). Indeed, $\p$ is the subalgebra of upper triangular
774: matrices (including diagonal ones), with corresponding subgroup
775: $P\subset\SL_3$, the subgroup of upper triangular matrices with
776: determinant=1. The quotient space $\SL_3(\R)/P$ can be
777: identified with the space $F$ of full flags in $\R^3$. A full
778: flag is a pairs
779: $(l,\pi)$, where $l$ is a line and $\pi$ is a plane, and
780: $l\subset\pi\subset\R^3$. The ``standard flag'' consisting
781: of the $x$ axis sitting inside the $xy$ plane has isotropy group $P$. The tangent space to $F$ at this base
782: point is naturally identified with $\sl_3/\p$,
783: represented in the root diagram by the remaining three light dots.
784: %(labelled $-\alpha_{12},-\alpha_{13},-\alpha_{23}$).
785: Two of the light dots are marked +. The diagram shows that the root spaces corresponding to these roots
786: %(labelled $-\alpha_{12}$ and $-\alpha_{23}$)
787: span a $\p$-invariant 2-dimensional subspace of $\sl_3/\p$ which Lie
788: generates the root space associated with the third light
789: dot. %($-\alpha_{13}$).
790: This means that we have on $F$ an
791: $SL_3(\R)$-invariant rank 2 contact distribution, i.e. a
792: non-integrable distribution that Lie generates the tangent bundle.
793:
794: This distribution can be geometrically interpreted
795: as the ``tautological"
796: contact distribution on $F$. This distribution is spanned by two vector fields,
797: corresponding to the two +s in figure 3. One vector field generates the flow
798: in which the line $l$ spins within the plane $\pi$, while the plane remains fixed. The other vector field generates the flow in which the plane $\pi$ rotates about
799: the line $l$, while the line remains fixed.
800:
801:
802:
803:
804: \bn{\bf Reading the $\g_2$ diagram.}
805: Now let us draw conclusions in a
806: similar fashion from the $\g_2$ diagram. There are twelve roots
807: in the diagram (figure 2) and so 12 root spaces. The rank of $\g_2$ is $2$
808: and so the dimension of $\g_2$ is $14 = 2 + 12$.
809: Consider the 9-dimensional
810: subspace $\p\subset\g_2$ spanned by $\t$ and the root spaces
811: associated with the roots marked by the dark dots in the diagram of figure 2.
812: Then the diagram shows that
813: \begin{itemize}
814: \item $\p$ is closed under the Lie bracket, i.e. is a subalgebra (a
815: so-called parabolic subalgebra, a subalgebra containing a Cartan
816: subalgebra).
817: \item Let $P\subset G_2$ be
818: the corresponding subgroup. It follows that $G_2$ has a 5-dim
819: homogeneous space $G_2/P$, whose tangent space $\g_2/\p$ at a point is
820: represented by the remaining 5 light dots.
821:
822: \item Two of the light
823: dots are marked with +. The diagram shows that their root spaces generate a 2-dim $\p$-invariant
824: subspace $W_1\subset\g_2/\p$, hence a $G_2$-invariant rank 2 distribution on $G_2/P$.
825:
826: \item This
827: distribution is not integrable, in fact, it is a distribution of type $(2,3,5)$,
828: since the diagram shows that bracketing once gives the light dot marked with
829: $\sigma_3$ and bracketing again gives the remaining two light dots.
830:
831: \end{itemize}
832:
833: To summarize, we have assembled the ingredients for the data
834: $(G_2, P, W_1)$. To prove the theorem
835: is to provide the geometric interpretation
836: of this distribution as $(\tilde Q, \tilde D)$ of theorem 1.
837: The first step in doing so is to embed $\so_3\oplus\so_3$ in $\g_2$.
838:
839:
840:
841:
842:
843:
844:
845:
846:
847: \section{The maximal compact subgroup of $G_2$}
848:
849: \subsection{Algebraic strategy of the proof.}
850:
851: In the previous sections we assembled the Lie algebraic data,
852: $(\so_3 \oplus \so_3, \h, D)$ and $(\g_2, \p, W)$
853: with corresponding group data $(K, H, D)$ and $(G_2, P, W)$.
854: The key to theorem 1 is to embed
855: $\so_3\oplus\so_3$ in $ \g_2$. This embedding is constructed
856: in the next section. Once established, we obtain
857: a diffeomorphism between corresponding distributions
858: by following the observation made in section \ref{shrinking}.
859:
860: We recap that observation, adding a bit of topology.
861: Suppose that $\k \subset \g$
862: and $\p \subset \g$ are Lie subalgebras of the Lie algebra $\g$.
863: Suppose that the natural map $\k/\k \cap \p \to \g/ \p$ is a linear
864: isomorphism. Let $W \subset \g/\p$ be an $ad$- $\p$-invariant subspace
865: and $W_1 \subset \k/\k \cap \p$
866: the corresponding subspace.
867: Then we will say that the Lie algebraic distributional data
868: $(\g, \p, W)$ and
869: $(\k, \k \cap \p, W_1)$ are isomorphic. If the corresponding
870: connected Lie groups are $K \subset G$ and
871: {\it if } $K$ is compact, then we can conclude
872: that the data $(G; P, W)$ and $(K; K \cap P, W_1)$
873: define isomorphic manifolds with distributions.
874: For when $K$ is compact we have that $K/K \cap P$ is a
875: compact and open submanifold of $G/P$ and hence is diffeomorphic
876: to $G/P$. And under this diffeomorphism
877: the distribution corresponding to $W$ is the same as the one represented
878: by $W_1$.
879:
880: The compactness assumption on $K$ is neccessary to conclude
881: that $G/P = K/(K \cap P)$.
882: Think of the case $K = \C^* \subset G = SL(2,\C)$
883: where $G$ acts on the sphere $Q = \C \cup \{ \infty \}$ by Mobius transformations
884: and where $\C^*$ corresponds to the complex scalings $z \mapsto \lambda z$,
885: $\lambda \ne 0$.
886: The fixed points of the $\C^*$-action are $0, \infty$.
887: The $\C^*$ orbit through any point $z_0 \ne 0, \infty$
888: is open, being the whole sphere minus the two fixed points.
889: Thus $G/P \ne K/ (K \cap P)$ where $P$ is the isotropy group of $z_0$.
890: But we still have $\k/(\k \cap \p) = \g/ \p$ since the orbit of $z_0$ is open.
891:
892: Lie algebraic data defines the corresonding Lie group data only up to a covering.
893: We can insure that there are only finitely many such coverings
894: by knowing that $\k$, like $\so_3 \oplus \so_3$, is the compact
895: real form of its corresponding complex Lie algebra.
896: For in this case there are only a finite number of
897: connected Lie groups $K$ with Lie algebra $\k$,
898: all of these being compact and covered by the simply connected $K$.
899: To say this in another way, suppose we are given Lie algebra data
900: $(\g, \p, W)$ and $(\k, \k \cap \p, W_1)$ as above, and
901: suppose that $\k$ is a compact real form. Let
902: $(G, P, W)$ and $(K, H, W_1)$ denote {\it any} Lie-group data realizing these
903: respective Lie algebraic data {\it where we are no longer assuming that $K \subset G$}.
904: Then the two manifolds-with-distribution which they stand for
905: are isomorphic up to a finite cover.
906: By this we mean, there is a third manifold-with-distribution $(X, E)$
907: and covering maps $\pi_G: X \to G/P$,
908: $\pi_K: X \to K/H$ such that $\pi_G ^* W = \pi_K ^* W_1 = E$.
909: Indeed, we can take $X$ to be $\tilde G/ \tilde P$
910: where $\tilde G$ is the unique simply connected Lie group with algebra
911: $\g$.
912:
913: To establish theorem 1 we will apply these considerations
914: to the case $G = G_2$ and $K \subset G_2$
915: its maximal compact subgroup.
916: We will show that $(\k, \k \cap \p)$
917: is isomorphic to $(\so_3\X\so_3, \h = \R(\e_3,\e_3))$.
918: And we show that under this isomorphism
919: $W_1 \subset \k/ \k \cap \p$
920: corresponds to the rolling distribution
921: when the ratios of the rolling spheres are $1:3$.
922:
923:
924: \subsection{ Finding Maximal compacts.} How can we ``see'' a maximal compact subgroup of $G_2$ tangled within its root diagram? Let us look back again
925: at the example of $\SL_3(\R)$. Here the
926: maximal compact subgroup is $\SO_3$, with Lie algebra $\so_3$, the set of 3 by 3 antisymmetric matrices.
927: These are spanned by the vectors $E_{ij}-E_{ji}$, $i>j$. So we see that corresponding to each pair
928: of ``antipodal" roots $\pm\alpha_{ij}$ we have one generator of $\k$, lying in the sum of the two corresponding root spaces.
929:
930:
931: More generally, for the ``split'' real form of any semi-simple Lie
932: algebra (such as our $\g_2$), the situation is similar: we get the Lie
933: algebra $\k$ of a maximal compact subgroup $K\subset G$ by taking the
934: sum of 1-dimensional subspaces, one subspace for each pair of
935: antipodal roots $\pm\alpha$. In fact, there is a certain
936: particulary ``nice" choice of root vectors $E_\alpha\in\g_\alpha$
937: (sometimes called a ``Weyl basis''), so that the sought-for line is
938: given by $\R(E_\alpha-E_{-\alpha}),$ as in the $\sl_3$ case.
939:
940:
941:
942: \bn
943:
944: In the case of $\g_2$ we thus have that \begin{itemize}
945: \item $\k$ is the sum of six 1-dimensional subspaces $\s_{i},\l_{i}$, $i=1,2,3,$ where
946: $\s_{i}$ lies in the sum of the root spaces corresponding to $\pm\sigma_i$, and
947: $\l_{i}$ lies in the sum of the root spaces corresponding to $\pm\lambda_i$.
948:
949: \item The isotropy of the $K$-action, $H=K\cap P\subset K$, is given in the diagram by
950: the vertical segment, $\h=\l_{3}$.
951:
952: \item The distribution plane $W\subset \k/\h$ is generated by $\s_{1},\s_{2}$ (mod $\h$).
953:
954: \end{itemize}
955:
956:
957: We have thus assembled the required ingredients for a ``distribution data" $(\k,\h,W)$.
958:
959:
960:
961:
962: \subsection{$\k\simeq\so_3\oplus\so_3$}
963:
964: Our task here is to define an isomorphism $\k\simeq\so_3\oplus\so_3$
965: that maps $(\k,\h,W)$ to the data of \S4 with $\rho=3$ or $1/3$. This entails the
966: decomposition of $\k$ into the direct sum of two ideals, each
967: isomorphic to $\so_3$. It would have been quite nice and simple if the
968: sought-for decomposition of $\k$ had been the decomposition into
969: ``long" ($\l_{i}$) and ``short" ($\s_{i}$). But this is not the case. For the
970: diagram shows that although the $\l_{i}$ generate an
971: $\so_3$ subalgebra of $\k$, this subalgebra is not an ideal, so is not one
972: of the summands in the decomposition. And the $\s_{i}$
973: do not generate even a subalgebra. So we have to work harder, i.e. write
974: down the precise commutation relations.
975:
976:
977: \begin{proposition} There is
978: a basis $\{S_i,L_i | i=1,2,3\}$ of $\k$, with $S_i\in \s_{i}$ and $L_i\in \l_{i}$,
979: such that
980: $$[L_i,L_j]=\epsilon_{ijk}L_k,\quad [L_i,S_j]=\epsilon_{ijk}S_k,\quad [S_i,S_j]=\epsilon_{ijk}({3\over 4}L_k-S_k),$$
981: where $\epsilon_{ijk}$ is the ``totally antisymmetric tensor on 3 indices" (
982: $\epsilon_{ijk}=1$
983: if $ijk$ is a cyclic permutation of 123, -1 if anticyclic permutation, and
984: 0 otherwise).
985: \end{proposition}
986:
987: The proof of this proposition is relegated to Appendix B.
988: It consists of simple but tedious
989: calculations which we could not ``see'' in the diagram. We tried. We
990: were reduced to picking up as nice as possible basis for $\g_2$ and
991: calculating the corresponding structure constants with the help of Serre \cite{Serre}.
992:
993: \bn Now set $$\e_i':={3L_i+2S_i\over 4}, \quad \e_i'':={L_i-2S_i\over 4}, \quad i=1,2,3.$$
994: These 6 vectors form a new basis for $\k$ and satisfy the standard $\so_3\oplus\so_3$ commutation relations
995: \begin{equation}
996: [\e_i',\e_j']=\epsilon_{ijk}\e_k', \quad
997: [\e''_i,\e''_j]=\epsilon_{ijk}\e''_k, \quad [\e_i',\e''_j]=0,
998: \label{std}
999: \end{equation}
1000: thus
1001: establishing the desired Lie algebra isomorphism $\k\simeq\so_3\oplus\so_3$.
1002:
1003:
1004:
1005:
1006: \begin{corollary} The map $\k\to\so_3 \oplus\so_3$ defined by
1007: $\e_i'\mapsto (\e_i,0),$ $\e_i''\mapsto (0,\e_i),$ $i=1,2,3,$ is a Lie
1008: algebra isomorphism. It maps $\h=\R L_3$ to $\R(\e_3,\e_3)$ and the
1009: 2-plane in $\k$ generated by $S_1,S_2$ to the 2-plane in
1010: $\so_3\X\so_3$ defined in the Proposition of \S4 for $\rho=3$.
1011: Interchanging the summands in $\so_3 \oplus\so_3$, i.e. mapping
1012: $\e_i'\mapsto (0,\e_i),$ $\e_i''\mapsto (\e_i,0),$ correponds to $\rho=1/3$.
1013: \end{corollary}
1014:
1015: The first assertion is eq (\ref{std}). The second assertion
1016: is easily verified using the last Proposition. We have thus
1017: defined a $G_2$-action on some finite cover of the rolling configuration,
1018: one which preserves the pulled-back distribution when the
1019: radii of the two balls are in the ratio $3:1$ or $1:3$. QED
1020:
1021:
1022: \bn{\bf How we came up with the formulae for $\e'_i,
1023: \e''_i$.} The first thing to observe is that since $L_3$ generates the
1024: isotropy $H=P\cap K$ we should have $L_3=\e_3'+\e_3''.$ Since
1025: everything is symmetric in 1,2,3 we conclude that $L_i=\e_i'+\e_i'',$
1026: $i=1,2,3.$ Next since $S_3$ commutes with $L_3$ we should have
1027: $S_3=a\e_3'+b\e_3''$ for some constants $a,b$, and again by symmetry
1028: $S_i=a\e_i'+b\e_i''$, $i=1,2,3.$ Now by using the sought-after commutations
1029: relations for the $\e'_i, \e''_i$ and the known commutations for $L_i, S_i$ we get that $a,b$
1030: are roots of the equation $x^2+x-3/4=0,$ i.e. $a=1/2, b=-3/2.$ Hence,
1031: $$L_i=\e_i'+\e_i'',\quad S_i=(\e_i'-3\e_i'')/2,\quad i=1,2,3.$$
1032: Inverting these equations we obtain the above equations for $\e_i', \e_i''$.
1033:
1034:
1035: \section{Split Octonions and the projective quadric realization of $\tilde Q$}
1036:
1037: To show that $G_2 \subset Aut(\tilde Q, \tilde D)$
1038: (theorem 1), it remains to identify $G_2/P$ with
1039: the $\tilde Q = S^3 \times S^2$ of the theorem
1040: and to show that the covering map $S^3 \times S^2 \to Q$
1041: corresponds to the identification $G_2 /P = K/H$
1042: composed with the projection $K/H \to (\pm 1, \pm 1) \backslash K / H$.
1043: In order to do these things will use the fact, discovered by Cartan \cite{C3}
1044: in 1914, that $G_2$
1045: is the group of automorphisms
1046: of the ``split octonions'' $\tilde \O$. We will
1047: follow the treatment in the book \cite{Harvey}, in the section ``The
1048: Cayley-Dickson process" (p.104). There further consequences and motivation
1049: can be found.
1050:
1051: The split octonions $\tilde \O$ are a real eight-dimensional
1052: algebra with unit and which is neither associative nor commutative.
1053: We identify $\tilde \O$ with $\H^2$, the 2 dimensional quaternionic vector space.
1054: Its multiplication law is
1055: \begin{equation}
1056: (a,b)(c,d)=(ac+\bar d b,da+b\bar c), \hskip .6cm a, b, c , d \in \H
1057: \label{product}
1058: \end{equation}
1059: The unit $1 \in \tilde \O$ is $(1,0) \in \H^2$
1060:
1061: The automorphism group of a real algebra is $A$
1062: is defined to be
1063: the space of nonzero real invertible linear maps $g: A \to A$
1064: satisfying $g(xy) = g(x) g(y)$ for all $x, y \in A$.
1065: $G_2$ is the automorphism group of $\tilde \O$.
1066:
1067: The unit $1$
1068: is automatically invariant under any automorphism of $\tilde \O$,
1069: so that $\R = \R 1 \subset \tilde \O$ is a $G_2$-invariant subspace.
1070: This subspace has a $G_2$-invariant complement:
1071: $$ \tilde \O = \R 1 \oplus V = Re(\tilde \O) \oplus Im(\tilde \O) $$
1072: In quaternionic terms:
1073: \begin{equation}
1074: V=Im\tilde\O=Im\H\oplus\H \subset \H ^2 .
1075: \label{V}
1076: \end{equation}
1077: To see the invariant nature of $V$, we
1078: use the split-octonion conjugation $x \mapsto \bar x$
1079: defined by $x=(a,b)\in\tilde\O \mapsto \bar x=(\bar a, -b)$
1080: for $x \in \tilde\O$.
1081: Then $x=Re(x)+Im(x)$,
1082: $Re(x)=(x+\bar x)/2 \in \R 1$, and $Im(x)=(x-\bar x)/2$.
1083: Also $x \bar x = - \langle x, x \rangle 1$
1084: where $\langle x,y\rangle=Re(x\bar y)$ is an inner product of signature $4,4$ on
1085: $\tilde\O$ which is invariant under the action of $G_2$.
1086: $V$ is the orthogonal complement of
1087: $1 \in \tilde \O$ relative to this inner product.
1088: Alternatively, it can be shown that $x \in V$ if and only if $x^2 = \langle x, x \rangle 1$
1089: (see \cite{Harvey}, lemma 6.67), proving the
1090: $G_2$-invariance of $V$.
1091: $V$ forms a 7-dimensional inner product
1092: space of signature $(3,4)$ relative to the restriction of
1093: $\langle \cdot , \cdot \rangle$. The $G_2$ action on $V$
1094: leaves this inner product invariant, so that
1095: $G_2$ is realized as a subgroup of $SO(3,4)$ through its representation
1096: on $V$.
1097:
1098: The maximal compact of $G_2$
1099: is $K = SO(4) = ( SU(2) \X SU(2) )/ \pm(1,1)$.
1100: See Appendix B and \cite{Vogan}. Upon restricting from $G_2$
1101: to $K$, the representation $V$ decomposes into irreducibles according
1102: to (\ref{V}). In other words, thinking of $SU(2)$ as unit quaternions,
1103: $(q_1, q_2) \in SU(2) \times SU(2) = \tilde K$ (the universal cover of $K$)
1104: and $(a,b) \in Im(\H) \oplus \H = V$ we
1105: have $(q_1, q_2) \cdot (a,b) = (q_1 a \bar q_1, q_1 b \bar q_2)$.
1106:
1107: In quaternionic terms (\ref{V}) the quadratic form associated to
1108: our $(3,4)$ inner product on $V$ is
1109: $$\langle (v, q), (v, q) \rangle = -|v|^2 + |q|^2.$$
1110: Since $K$ acts transitively on the product of spheres
1111: $S^2 \times S^3 \subset Im (\H) \oplus \H = \tilde \O$
1112: we have that
1113: $G_2$ acts transitively on the null cone
1114: $\{ x = (v, h): \langle x,x\rangle=0, x\neq 0 \}$.
1115: (To see that we can change the `length'
1116: of an $x$ in the null cone using $G_2$,
1117: use the fact that each such null vector is a nonzero weight vector relative
1118: to some choice of maximal Cartan $T \subset G_2$.
1119: This maximal Cartan then acts on $x$ by scaling. See the description
1120: following eq. (\ref{basis}) below.)
1121: Thus $G_2$ acts
1122: transitively on the space of null rays
1123: $$ C=\{\R^+ x\subset V|\langle x,x\rangle=0, x\neq
1124: 0\}\subset P^+ (V) := \hbox{ rays in } V $$.
1125: This $C$ is a nondegenerate 5-dimensional quadric
1126: sitting in the 6-dimensional real ray space
1127: $P^+(V)$ (diffeomorphic to $S^6$). We can describe points
1128: of the ray space $P^+ (V)$ using homogeneous coordinates
1129: $[x] = [v,h] = [\lambda v, \lambda h]$,
1130: $\lambda \in \R^+$ with $(v,h)\in Im\H\oplus\H =V$.
1131: $C$ is
1132: defined by the homogeneous equation $\|v\|^2=\|h\|^2.$
1133: Using the $\R^+$ action, we normalize $\|v \| = 1$,
1134: proving that $C$ is diffeomorphic to
1135: the product of spheres $S^2 \times S^3 \cong S^3 \times S^2 =\tilde Q$
1136: which appears in theorem 1.
1137:
1138:
1139: Given a point $\R^+ x = [x] \in C,$ set $$x^\perp=\{y\in V|\langle x,y\rangle=0\}, \quad
1140: x^0=\{y\in V| xy=0\}.$$ Then
1141: \begin{proposition}
1142: $$\R x\subset x^0\subset (x^0)^\perp\subset x^\perp\subset V,$$
1143: and the dimensions are $1,3,4,6,7.$
1144: \end{proposition}
1145:
1146: \pf Use the definitions of the split octonion product (eq (\ref{product})) and inner product.\qed
1147:
1148: When we projectivize, $x^\perp$ maps to the tangent plane $T_{[x]} C$ to $C$ at $[x]$, and $x^0$
1149: maps to a 2-dimensional subspace $D_{[x]}\subset T_{[x]} C$.
1150: Letting $[x]$ vary over $C$ we have
1151: defined a rank 2 distribution $D\subset TC$.
1152: This construction of $(C, D)$ depends only on
1153: on the algebraic structure of $\tilde \O$, so that $G_2=Aut(\tilde\O)$
1154: acts on $C$ preserving $D$.
1155:
1156:
1157: \begin{proposition}
1158: \label{projective}
1159: The (ray) projective quadric $C$ is a homogeneous
1160: space for $G_2$. $C$ is diffeomorphic to $\tilde Q =S^3 \times S^2$
1161: of theorem 1, and is naturally
1162: endowed with a $G_2$-invariant distribution $D$ of rank $2$.
1163: Viewed as a $G_2$-homogeneous space, the data for $(C,D)$
1164: coincides with the data $(G_2, P, W)$ of section 3.3. Viewed
1165: as a $K$-homogeneous space, its Lie algebraic data
1166: coincides with that of the rolling distribution $(\k, \h, D)$ for
1167: the ratio $1:3$. The distribution on $C$
1168: pushes down to the rolling distribution for ratios $3:1$
1169: under the two-to-one cover $C = S^3 \to S^2 \to Q = SO_3 \times S^2$
1170: \end{proposition}
1171:
1172: This proposition immediately implies
1173: that part of the theorem we are going to prove:
1174: that $G_2 \subset Aut(\tilde Q, \tilde D)$.
1175:
1176: {\bf Steps of the proof.}
1177: In the paragraph preceding the
1178: proposition we proved that $C$ is a homogeneous
1179: space for $G_2$, that $D$ is invariant under this $G_2$ action,
1180: and that $C$ is diffeomorphic to $\tilde Q$.
1181: Next, we will prove that the coincidence of the $\g_2$-data for $(C,D)$
1182: and the data $(\p, W)$ of the previous section.
1183: For this we will use the weights for the
1184: $G_2$-representation space $V = Im(\tilde \O)$.
1185:
1186:
1187: {\bf Weights for the $7$-dimensional representation.}
1188:
1189: Here is the weight diagram for this representation.
1190:
1191: \begin{center}
1192: \unitlength=1cm
1193: \begin{picture}(5,4)(-2,-2)
1194:
1195:
1196: \put(0,0){\circle{.3}}
1197:
1198: \put(0,0){\circle{.18}}
1199:
1200:
1201:
1202: \put(-.5,-.866){\circle{.25}}
1203: \put(-.8,-1){X}
1204: \put(.5,.866){\circle{.25}}
1205: \put(-.5,.866){\circle{.25}}
1206: \put(-.68,.866){X}
1207: \put(.5,-.866){\circle{.25}}
1208:
1209: \put(-.5,-.866){\circle{.1}}
1210: \put(.5,-.866){\circle{.1}}
1211:
1212: \put(.5,.866){\circle{.1}}
1213: \put(-.5,.866){\circle{.1}}
1214:
1215:
1216:
1217:
1218: \put(-1,0){\circle*{.25}}
1219: \put(1,0){\circle{.25}}
1220: \put(-1,0){\circle{.1}}
1221: \put(1,0){\circle{.1}}
1222:
1223: %this marks the weights for V : one circle inside the other
1224:
1225: %\put(1.3,-0.1){$\sigma_3$}
1226: %\put(-2,-0.1){$-\sigma_3$}
1227: \put(-1.4,-0.1){X}
1228:
1229:
1230: \put(1.4,-.866){\circle{.25}}
1231: \put(-1.5,.866){\circle{.25}}
1232: %\put(-1.5,.866){X}
1233: %\put(1.6, -1.1){$-\lambda_1$}
1234: %\put(-2.2, 0.866){$\lambda_1$}
1235: \put(-1.9, .8){ X }
1236:
1237: \put(1.4,.866){\circle{.25}}
1238: \put(-1.5,-.866){\circle{.25}}
1239: \put(-1.9, -1){X}
1240: %\put(1.7,.9){$\lambda_2$}
1241: %\put(-2.4,-1.1){$-\lambda_2$}
1242:
1243: \put(0,1.5){\circle{.25}}
1244: \put(-.2,1.6){X}
1245: \put(0,-1.5){\circle{.25}}
1246: \put(-0.3,-1.7){X}
1247:
1248: \put(0,-1.4){\line(0,1){3}}
1249: \put(-1.5, -0.866){\line(5,3){2.77}}
1250: \put(-1.5, 0.866){\line(5,-3){2.77}}
1251:
1252: \put( -1, 0){\line(1 , 0 ){1.85}}
1253: \put(-.5,-.866){\line(3,5){.96}}
1254: \put(-.5,.866){\line(3,-5){.96}}
1255:
1256: \end{picture}
1257:
1258: Figure 4: Weights and roots associated with the representation $V$.
1259:
1260: \end{center}
1261:
1262:
1263: The weights of the representation $V$ form a subset of the roots of $\g_2$.
1264: In figure 4 we redrew the root diagram of $\g_2$, marking
1265: those roots which are weights for $V$ with bullseye's.
1266: They are the six short roots and one zero root.
1267: The corresponding weight spaces $V_w$ are all one-dimensional.
1268: The black dot is a selected weight vector and corresponds to
1269: a `choice of base point' for $C$. The meaning of the X's will be given below.
1270:
1271:
1272: \bn{\bf A reminder of the meaning of the weight diagram.}
1273:
1274: We recall the general case of a representation $V$ of a
1275: semi-simple Lie algebra $\g$
1276: with Cartan subalgebra $\t$.
1277: A {\it weight} for the representation $V$ of $\g$
1278: is an element $w \in \t^*$ such that there is a nonzero
1279: vector $v \in V$ with the property that
1280: $\zeta \cdot v = w(\zeta) v$ for all $\zeta \in \t$. The space of $v$'s for a given weight $w$
1281: is called the {\it weight space} for $w$ and is denoted $V_w$.
1282: If, for given $w \in \t^*$ there is no such nonzero $v$
1283: then we set $V_w = 0$. For a finite-dimensional representation
1284: the set of weights is finite. We have
1285: $$V = \bigoplus_{w\in\t^*} V_w.$$
1286: The roots of $\g$ are the non-zero weights of the adjoint representation.
1287:
1288: If, as in our situation, the roots for $\t$ are real, then its `torus' $T = exp(\t)$ is noncompact and acts on the weight spaces by scaling, as follows.
1289: If $\lambda = exp(\xi) \in T$, with $\xi \in \t$,
1290: then $\lambda e_w = exp(w(\xi))e_w$ for $w \in V_w$.
1291:
1292: From $\zeta \cdot \xi \cdot v = \xi \cdot \zeta \cdot v + [\zeta, \xi ] \cdot v$
1293: it follows that if $v$ is in the weight space for $w$
1294: and $\xi \in \g_{\alpha}$ is in the root space for $\alpha$ then $\xi v$
1295: is in the weight space for $w + \alpha$ (which, as above, could be zero).
1296: In other words:
1297: $\g_{\alpha} V_w \subset V_{w + \alpha}$. This inclusion is half of the rule:
1298: \begin{equation}
1299: w \hbox{ a weight}, \alpha \hbox{ a root} \Rightarrow \g_{\alpha} V_w = V_{w + \alpha}
1300: \label{rule}
1301: \end{equation}
1302: which is true for $V$.
1303: It follows in particular that if $v \in V_w$
1304: and $\xi \in \g_{\alpha}$ and if $w + \alpha$ is not a weight for the representation,
1305: then $\xi (v) = 0$. We will use this fact momentarily.
1306:
1307:
1308: We now construct the weight spaces and the action of
1309: the torus for our $G_2$-representation
1310: $V = Im(\tilde \O)$.
1311: Let $n$ be an imaginary quaternion.
1312: Then $(n, n)$ and $(n, -n)$ are both null vectors in $V$.
1313: Take as basis for $V$:
1314: \begin{equation}
1315: e_1 = \frac{1}{2}(i, i), e_2 = \frac{1}{2}(j, j), e_3 = \frac{1}{2}(k,k) ; f_1 = \frac{1}{2}(i, -i), f_2 = \frac{1}{2}(j, -j), f_3 = \frac{1}{2}(k, -k)
1316: \label{basis}
1317: \end{equation}
1318: and
1319: $$U = (0,1).$$ Then we have
1320: the multiplication table:
1321: $$e_i ^2 = f_i ^2 = 0 $$
1322: $$e_i f_j = f_j e_i = 0, \hskip .2cm \hbox{ if } i \ne j$$
1323: $$e_i e_j = f_k; i, j, k \hbox{ a cyclic permutation of } 1,2,3 $$
1324: $$f_i f_j = e_k ; i, j, k \hbox{ a cyclic permutation of } 1,2,3 $$
1325: $$e_i f_i = -\frac{1}{2} + \frac{1}{2}U$$
1326: $$f_i e_i = -\frac{1}{2} - \frac{1}{2}U$$
1327: $$e_i U = e_i$$
1328: $$f_i U = - f_i $$
1329: To complete the multiplication table, use that the conjugate
1330: of $xy$ is $\bar x \bar y$,
1331: so that if $x, y \in V = Im(\tilde \O)$ we have
1332: $y x = \bar z$ where
1333: $z = xy$.
1334: Thus, for example since $\bar f_k = -f_k$
1335: we see that $e_j e_i =-f_k$, for $i,j,k$ a cyclic permutation of $1,2,3$.
1336: Now let $\lambda_1, \lambda_2, \lambda_3$
1337: be nonzero reals with $\lambda_1 \lambda_2 \lambda_3 = 1$.
1338: Let $\alpha_i, \beta_i, \gamma_i$
1339: and $\tilde \alpha_i, \tilde \beta_i, \tilde \gamma_i$ be real exponents for $i = 1, 2, 3$
1340: satisfying $\alpha_i + \beta_i + \gamma_i = 0$
1341: Then the scaling transformation
1342: $$e_i \mapsto \lambda_1 ^{\alpha_i} \lambda_2 ^{\beta_i} \lambda_3 ^{\gamma_i} e_i$$
1343: $$f_i \mapsto \lambda_1 ^{\tilde \alpha_i} \lambda_2 ^{\tilde \beta_i} \lambda_3 ^{\tilde\gamma_i} f_i$$
1344: together with $z \mapsto z$
1345: preserves the multiplication table, and hence defines an element of $G_2$,
1346: provided
1347: $$\tilde \alpha_i = - \alpha_i, \tilde \beta_i = - \beta_i, \tilde \gamma_i = - \gamma_i$$
1348: and provided that
1349: $(\alpha_i, \beta_i, \gamma_i)$ are multiples of the values from the following
1350: weight table
1351:
1352: $$\begin{array}{|c||c|c|cl}
1353: \hline
1354: &\alpha_i&\beta_i&\gamma_i\\
1355: \hline\hline
1356: i = 1 &2 &-1& -1\\
1357: \hline
1358: i=2& -1 &2 &-1 \\
1359: \hline
1360: i=3 &-1 & -1& 2 \\
1361: \hline
1362: \end{array}
1363: $$
1364: These scaling transformations generate the Cartan $T$ of $G_2$,
1365: and the table gives the corresponding weights. Thus for example
1366: $e_1$ is a weight vector with corresponding weight, relative to
1367: the basis for $\t$, being $(2,-1,-1)$. Here we view $\t$
1368: as being the collection of vectors $(a, b, c)$ with $a + b + c = 0$.
1369: Looking at the inner products of these vectors we see that they are arranged
1370: on the weight diagram according to:
1371:
1372: \begin{center}
1373: \unitlength=1cm
1374: \begin{picture}(5,4)(-2,-2)
1375:
1376:
1377: \put(0,0){\circle{.3}}
1378:
1379: \put(0,0){\circle{.18}}
1380:
1381:
1382:
1383: \put(-.5,-.866){\circle{.25}}
1384: \put(.5,.866){\circle{.25}}
1385: \put(-.5,.866){\circle{.25}}
1386:
1387: \put(.5,-.866){\circle{.25}}
1388: \put(.8, 1){$e_2$}
1389: \put(.76, -1){$e_3$}
1390:
1391: \put(-.5,-.866){\circle{.1}}
1392: \put(.5,-.866){\circle{.1}}
1393: \put(-.8, -1.18){$f_2$}
1394: \put(-.8, 1.1){$f_3$}
1395:
1396: \put(.5,.866){\circle{.1}}
1397: \put(-.5,.866){\circle{.1}}
1398:
1399:
1400:
1401:
1402: \put(-1,0){\circle*{.25}}
1403: \put(1,0){\circle{.25}}
1404: \put(-1,0){\circle{.1}}
1405: \put(1,0){\circle{.1}}
1406:
1407: \put(-1.45, 0){$e_1$}
1408: \put(1.2, 0){$f_1$}
1409:
1410: %this marks the weights for V : one circle inside the other
1411:
1412:
1413: \put(.2, .2){$z$}
1414:
1415: \put( -1, 0){\line(1 , 0 ){1.85}}
1416: \put(-.5,-.866){\line(3,5){.96}}
1417: \put(-.5,.866){\line(3,-5){.96}}
1418:
1419: \end{picture}
1420:
1421: Figure 5: The weight space basis.
1422:
1423: \end{center}
1424:
1425: We are now in a position to compute the $\g_2$-data
1426: associated to $(C, D)$ from the proposition.
1427:
1428: {\bf Weight vectors are null vectors: } Because the inner product is $G_2$-invariant,
1429: the $\g_2$ action on $V$ satisfies $\langle \xi x, x \rangle = 0$
1430: for any $\xi \in \g_2$, $x \in V$. Take $x$ a weight vector with nonzero weight $w$,
1431: and
1432: take $\xi \in \t$ with $w(\xi) \ne 0$. From
1433: $\langle \xi x, x \rangle = w(\xi) \langle x, x \rangle$
1434: we have that $x$ is a null-vector.
1435:
1436:
1437: {\bf Computing the isotropy data.}
1438: Set $c_0 = [e_1]$, the ray through $e_1$.
1439: We must show that the isotropy group of $c_0$
1440: is $P$.
1441:
1442: We begin by showing that the isotropy algebra
1443: $\g_{c_0}$ of $c_0 = [e_1] $ is
1444: $\p$. We have that $\g_{c_0} = \{ \xi \in \g_2 : \xi e_1 = \lambda e_1 \hbox{ for some
1445: real number } \lambda \}$.
1446: The black dot in figure 4 indicates
1447: the weight space spanned by $e_1$,
1448: with corresponding weight
1449: by $w_1$. (This weight is the root
1450: marked $- \sigma_3$ in Figure 3.)
1451: According to the addition rule,
1452: (\ref{rule}) if $\alpha \in \t^*$ is a root
1453: and $w_1 + \alpha$ {\it is not a weight} for $V$,
1454: then $\xi_{\alpha} x_0 = 0$. Those
1455: roots for $w_0 + \alpha$ {\it is not a weight} are marked by X's in figure 4.
1456: The sum of the corresponding $\g_{\alpha} \subset \g_2$ is a vector
1457: space of elements $\xi$ satisfying $\xi (e_1) = 0$.
1458: Now the isotropy algebra $\g_{c_0}$of the ray through $e_1$
1459: consists of all those $\xi$ such that $\xi e_1 = \lambda e_1$
1460: for some real scalar $\lambda$.
1461: The elements $ H \in \t$ act on $e_1$ by
1462: scalar multiplication by $\lambda = w_1 (H)$.
1463: Referring to the diagram then, we see that
1464: $\p \subset \g_{c_0}$. But there is no
1465: subalgebra of $\g_2$ lying between $\p$ and all of $\g_2$.
1466: It follows that the isotropy algebra for the ray {\it is} $\p$.
1467:
1468: It follows from this Lie algebra computation
1469: that the isotropy subgroup $G_{c_0} $
1470: {\it contains} $P$ and has Lie algebra equalling the Lie
1471: algebra $\p$ of $P$. Now $P$
1472: was defined to be the {\it connected} Lie subgroup of $G_2$
1473: whose Lie algebra is $\p$, thus to show
1474: $G_{c_0} = P$ it suffices to show that $G_{c_0}$ is connected.
1475: We demonstrate connectivity
1476: by applying the homotopy exact sequence
1477: to the fiber bundle $G_{c_0} \to G_2 \to C = G_2 / G_{[x]}$.
1478: This exact sequence
1479: is $\ldots \to \pi_1(C) \to \pi_0 ( G_{c_0}) = \pi_0 (G_2) \to \pi_0 (C)$.
1480: Since $C$ is simply connected and connected we
1481: get that $\pi_0 ( G_{c_0} = \pi_0 (G_2)$
1482: and since $\pi_0 (G_2) = 0$ we have our connectivity:
1483: $\pi_0 (G_{c_0}) = 0$.
1484:
1485: We have established the isotropy ($P$ part) of the data for $(C, D)$.
1486:
1487:
1488:
1489: {\bf Computing the distribution data}.
1490: The distribution plane $D(c_0)$ at $c_0$
1491: corresponds to $e_1 ^0$ -- the
1492: subspace $S \subset V$ consisting of
1493: those vectors $y \in V$ for which
1494: $e_1 y = 0$. From the multiplication table
1495: following the description of our basis (\ref{basis}) we see that
1496: $S = span \{ e_1, f_2, f_3 \}$.
1497: From Figure 4, we see that weights
1498: corresponding to $f_2, f_3$, , say $w_2, w_3$,
1499: are given by $w_2 = w_1 + \sigma_1$,
1500: $w_3 = w_1 + (-\sigma_2)$. Compare Figure 3.
1501: Let $x_1, y_1 \in \g_2$ be the corresponding
1502: nonzero root vectors for $\sigma_1, - \sigma_2$.
1503: (We follow the $x, y$ notation from Figure 5, Appendix 2.)
1504: It follows from rule (\ref{rule})
1505: that $f_2$ is a multiple of $x_1 (c_0)$
1506: and $f_2$ is a multiple of $y_2 (c_0)$.
1507: In other words, $S = W(e_1)$ mod $\p (e_1)$
1508: where $W$ is the space spanned by the roots indicated
1509: by the pluses in Figure 2. We have proved
1510: that the Lie algebraic data for
1511: $(C, D)$ is $(\g_2, \p, W)$.
1512:
1513:
1514:
1515: {\bf The covering map.}
1516: On the Lie algebra level we have shown that the data for
1517: $(C, D)$ is $(\g_2, \p, W)$. As computed in section 3.3, Cor. 1, upon
1518: restricting the action of $G_2$ to $K$ this Lie algebraic data
1519: $(\g_2, \p, W)$ corresponds to
1520: the data $(\so_3 \oplus \so_3, \h, D(3;1))$.
1521: Thus, up to a finite cover, $(C,D)$ is the rolling distribution.
1522: Now $C$ is simply connected, and $2:1$ covers $Q$.
1523: This covering map $C = S^3 \times S^2 \to Q = SO_3 \times S^2$
1524: is realized by forming the quotient of
1525: $C$ by the $\Z_2$ subgroup
1526: generated by image of $\sigma = (\pm 1, 1) \in K = SU_1 \X_{\pm(1,1)} SU_2$.
1527: Being an element of the symmetry group $\sigma$ preserves
1528: the distribution $D$ on $C$, and so $D$ does push down to
1529: the rolling space $Q$. The $\k$ data of the pushed-down distribution
1530: remains $(\so_3 \oplus \so_3, \h, D(3;1))$. Thus the pushed-down distribution
1531: is isomorphic to the rolling distribution on $Q$.
1532:
1533: QED
1534:
1535:
1536:
1537: \section{Summary. Lack of action on the rolling space. The theorem is done. }
1538:
1539: We have proved that $G_2 \subset Aut( \tilde Q, \tilde D)$
1540: and that is all that we are going to prove of
1541: theorem 1, with the exception of the fact that the
1542: $G_2$-action does {\it not} descend to $Q$. (Recall from the introduction
1543: we are not going to prove that $G_2 = Aut( \tilde Q, \tilde D)$.)
1544: To prove that the $G_2$ action does not descend to $Q$, we
1545: realize as above that $Q = \Z_2 \backslash C$
1546: where the $\Z_2 \subset K \subset G_2$ is generated by $\sigma = (\pm 1, 1)$.
1547: Now we use the following fact about group actions. Suppose
1548: that a group $G$ (here $G_2$) acts effectively on a set $C$
1549: and that $\Gamma \subset G$. (``Effectively'' means that the only
1550: group element acting as the identity on $C$ is the identity.)
1551: Then the action of an element $g \in G$ descends to the quotient space $\Gamma \backslash C$ if and only if $g \Gamma g^{-1} = \Gamma$.
1552: In particular, if $\Gamma$ is not normal in $G$ then
1553: the action of all of $G$ does not descend to the quotient $\Gamma \backslash C$.
1554: Returning to our situation, we see that
1555: if the $G_2$ action were to descend then this
1556: $\Z_2$ generated by $\sigma$
1557: would have to be normal. But a discrete normal subgroup
1558: of a connected Lie group is central, and $G_2$ has
1559: no center. See Appendix A, or \cite{Vogan}. So our $\Z_2$ is not normal,
1560: and the $G_2$ action does not descend.
1561:
1562:
1563: {\bf Remark.} Had we used lines instead of rays
1564: when constructing $C = \tilde Q$, we would have
1565: arrived at a quadric $Q_f$ in the standard real projective space $P(V)$
1566: which is double covered by $C = \tilde Q$.
1567: (The subscript `f' is for `false'.)
1568: $Q_f \subset P(V)$ is diffeomorphic to $S^3 \times_{\Z_2} S^2 = {\pm I} \backslash C$ where
1569: the notation $\times_{\Z_2}$
1570: indicates that we divide out by the action of the involution $(v, h) \mapsto (-v, -h)$.
1571: (This involution does not lie in $G_2$.)
1572: $C = \tilde Q$ double-covers both $Q_f$ and
1573: $Q$, and the distribution $\tilde D$ pushes down to both covered spaces.
1574: But $Q_f$ is topologically distinct from $Q$. Both
1575: $Q$ and $Q_f$ are $\SO_3$-bundles over $S^2$. $Q$ is the trivial $SO(3)$-bundle.
1576: $Q_f$ is the other one. (Since $\pi_1 (\SO_3) = \Z_2$ there are precisely
1577: two topologically distinct $\SO_3$ bundles over $S^2$.)
1578: Because $-I \in GL(V)$ commutes with the $G_2$ action on $V$
1579: the $G_2$-action on $\tilde Q$ does descend to $Q_f$.
1580: We find it curious that the action of $G_2$
1581: on $\tilde Q$ does descend to this `false' rolling configuration space $Q_f$, but not to the real one $Q$.
1582:
1583:
1584:
1585: %\section{Appendices.}
1586: \appendix{}
1587: \section{Covers. Two $G_2$'s. }
1588: \label{twoG2s}
1589:
1590: To understand our results, it helps to understand that
1591: up to isomorphism, there are precisely two connected $G_2$'s:
1592: the adjoint one which is the one we have been using,
1593: and the simply connected one, which is the universal cover
1594: of the adjoint one.
1595: For a general semi-simple Lie algebra $\g$ we can always form
1596: the simply connected Lie group $\tilde G$ having $\g$ as its Lie algebra.
1597: If $Z$ is the center of $\tilde G$, then $Ad(\tilde G) = \tilde G/Z$
1598: where $Ad(\tilde G)$ is the image of $G$ under the adjoint map
1599: from $\tilde G$ to $Hom(\g)$. If $Z \ne I$ then $G \ne Ad(G)$.
1600: There are as many distinct connected Lie groups with
1601: algebra $\g$ as there are distinct subgroups of $Z$, these being
1602: the connected topological groups covered by $\tilde G$ and covering $Ad(G)$.
1603: So, when $Z = \Z_2$ there are precisely two such Lie groups, $\tilde G$, the simply connected one, and $G = Ad(\tilde G)$, the adjoint one.
1604:
1605: We find on p. 3 of Vogan \cite{Vogan} that the center of the simply connected
1606: $G_2$ is indeed $\Z_2$, and hence we have precisely two $G_2$'s.
1607: It will be useful to explain a few details of this computation of $Z(G_2)$.
1608: The universal cover of any $G$ contracts onto its maximal compact.
1609: Thus, if the maximal compact of $Ad(G)$ has finite fundamental group,
1610: then the universal cover $\tilde G \to Ad(G) = \tilde G /Z$
1611: is a finite cover, and so the center $Z$ must be finite.
1612: (At the other extreme, the maximal compact of
1613: $SL(2, \R)$ is a circle group, corresonding to the fact that its
1614: universal cover has infinite center $\Z$.)
1615:
1616: We saw above that the Lie algebra of
1617: the maximal compact of any $G_2$ realizing $\g_2$
1618: is $\k = \so_3 \times \so_3$. The connected Lie groups $K$
1619: having $\k$ as Lie algebra have fundamental groups
1620: consisting of
1621: either $1$, $2$ or $4$ elements.
1622: It follows that the center $Z(G_2)$ of any $G_2$ is finite, and
1623: hence compact. Being compact and central, this center lies
1624: in every maximal compact: $Z(G_2) \subset K \subset G_2$.
1625: If we take the simply connected $G_2$, call
1626: it $\tilde G_2$, then its maximal compact is
1627: $\tilde K = \SU_2 \times \SU_2$. The center of $\tilde K$ is the group of the four elements $(\pm 1, \pm 1)$.
1628: The center of $\tilde K$ need not be the center of $G_2$
1629: but it must
1630: contain it: $Z(\tilde G_2) \subset Z(\tilde K)$.
1631: To see what the actual center of $\tilde G_2$ is, it suffices to see how
1632: $Z(\tilde K) \subset \tilde K$ acts on
1633: the Lie algebra $\g_2$ under the adjoint action.
1634: This can be done using roots.
1635: The center of $\tilde G_2$ is that part of $Z(\tilde K)$ which
1636: acts trivially on $\g_2$. A computation using roots and
1637: the restriction of the adjoint representation to $\tilde K$ shows that
1638: this part is $(1,1)$ and $-(1,1)$.
1639:
1640:
1641: %\appendix{}
1642: \section{The isomorphism of $\k$ and $\so_3 \oplus \so_3$ from Proposition 3. }
1643: \label{maxcpt}
1644:
1645: We complete the proposition 3 from section 5, in which the
1646: explicit identification of $\so_3 \oplus \so_3$
1647: as the Lie algebra $\k$ of the maximal compact in $\g_2$. We follow Serre \cite{Serre},
1648: page VI-11: $\g_2$ is Lie-generated by the elements
1649: $x,y,h,X,Y,H$, subject to the following relations, which one can read
1650: off the root diagram.
1651:
1652: $$\begin{array}{llll}
1653: [x,y]=h,& [h,x]=2x,& [h,y]=-2y,& \\
1654: \left[X,Y\right]=H,& [H,X]=2X,&[H,Y]=-2Y;&\\
1655: \left[h,X\right]=-3X, &[h,Y]=3Y;&[H,x]=-x,&[H,y]=y;\\
1656: \left[x,Y\right]=[X,y]=[h,H]=0;&&&\\
1657: \left[ad(x)\right]^4 X=0; &\left[ad(X)\right]^2x=0;&&\\
1658: \left[ad(y)\right]^4Y=0;& \left[ad(Y)\right]^2x=0.&&\\
1659: \end{array}
1660: $$
1661:
1662:
1663:
1664: Taking Lie brackets of the vectors $x,y,h,X,Y,H$ we generate a complete set $\{x_i,X_i, y_i,Y_i|i=1,2,3\}$ of root vectors for
1665: $\g_2$, which, together with the basis $h,H$ for the Cartan subalgebra form a basis for $\g_2$ as follows:
1666: %
1667: $$\begin{array}{llllll}
1668: x_3=x,&X_1=X, &x_2=[x,X_1],&x_1=[x,x_2],&X_2=[x,x_1],&X_3=[X_1,X_2];\\
1669: y_3=y,& Y_1=Y,&y_2=-[y,Y_1],&y_1=-[y,y_2],&Y_2=-[y,y_1],&Y_3=-[Y_1,Y_2].
1670: \end{array}
1671: $$
1672: We label each root in the diagram
1673: with the corresponding root vector.
1674:
1675: \bn
1676:
1677: \begin{center}
1678:
1679: \unitlength=1cm
1680: \begin{picture}(5,4)(-2,-2)
1681: \put(0,0){\circle*{.4}}
1682: %
1683: \put(.5,.866){\circle*{.25}}
1684: \put(.65,1.05){$x_1$}
1685: \put(-.5,-.866){\circle*{.25}}
1686: \put(-.75,-1.25){$y_1$}
1687: %
1688: \put(-.5,.866){\circle*{.25}}
1689: \put(-.95,1.05){$x_2$}
1690: \put(.5,-.866){\circle*{.25}}
1691: \put(.65,-1.25){$y_2$}
1692: %
1693: \put(1,0){\circle*{.25}}
1694: \put(1.2,-0.1){$x_3=x$}
1695: \put(-1,0){\circle*{.25}}
1696: \put(-2.4,-0.1){$y=y_3$}
1697: %
1698: \put(0,1.7){\circle*{.25}}
1699: \put(-.2,2){$X_3$}
1700: \put(0,-1.732){\circle*{.25}}
1701: \put(-.2,-2.3){$Y_3$}
1702: %
1703: \put(-1.5,.866){\circle*{.25}}
1704: \put(-3.1, 1.05){$X=X_1$}
1705: \put(1.4,-.866){\circle*{.25}}
1706: \put(1.6, -1.25){$Y_1=Y$}
1707: %
1708: \put(1.4,.866){\circle*{.25}}
1709: \put(1.7, 1.05){$X_2$}
1710: \put(-1.5,-.866){\circle*{.25}}
1711: \put(-2.2,-1.25){$Y_2$}
1712:
1713:
1714:
1715: \put(0,-1.6){\line(0,1){3.3}}
1716: \put(-1.4, -0.8){\line(5,3){2.77}}
1717: \put(-1.5, 0.866){\line(5,-3){2.77}}
1718: \put( -.9, 0){\line(1 , 0 ){1.85}}
1719: \put(-.46,-.75){\line(3,5){.96}}
1720: \put(-.5,.866){\line(3,-5){.96}}
1721:
1722: \end{picture}
1723:
1724: \vskip .8cm
1725: Figure 5. A basis for the Lie algebra.
1726: \end{center}
1727:
1728:
1729: \bn
1730:
1731: We end up with a ``nice'' basis wrt which the structure constants are
1732: particulary pleasant; they are integers and have symmetry properties
1733: which facilitate greatly the work involved in their determination; you
1734: can also apply some elementary $\sl_2$ representation theory that
1735: further facilitate the calculation; it helps to work with the root
1736: diagram nearby.
1737:
1738:
1739: \bn{\bf Symmetry properties of the structure constants.} Suppose $\alpha,\beta$ are two roots such
1740: that $\alpha+\beta$ is also a root. Let $E_\alpha, E_\beta$
1741: be the corresponding root vectors, as chosen above. Then
1742: $[E_\alpha,E_\beta]=c_{\alpha,\beta}E_{\alpha+\beta}$, for some
1743: non-zero constant $c_{\alpha,\beta}\in\Z$. The nice feature of
1744: our base is that the structure constants satisfy
1745: $$c_{-\alpha,-\beta}=-c_{\alpha,\beta}.$$ This cuts in half the amount
1746: of work involved, since you need only consider say $\alpha>0$ (the positive roots are the six dots in the last root diagram marked
1747: with $x$'s and $X$'s).
1748: Combining this with the obvious $c_{\alpha,\beta}=-c_{\beta,\alpha}$
1749: (antisymmetry of Lie bracket) you obtain
1750: $$c_{\alpha,-\beta}=c_{\beta,-\alpha}.$$
1751: This cuts in half again the amount of work.
1752:
1753: \begin{proposition} The structure constants of $\g_2$, with respect to the basis of root vectors $\{x_i,X_i,y_i, Y_i | i=1,2,3\}$
1754: and the Cartan algebra elements $\{h,H\}$ are given as follows. The
1755: basis elements are grouped in three sets: positive (three $x$'s and
1756: three $X$'s), negative (three $y$'s and three $Y$'s), and Cartan
1757: subalgebra elements ($h$ and $H$).
1758: \begin{itemize}
1759: \item $[Positive, positive]$: other then the ones given above, and
1760: those which are zero for obvious reasons from the root diagram (sum of
1761: roots which is not a root): $$[x_1, x_2]= X_3.$$
1762:
1763: \item $[ Positive, negative ]$:
1764:
1765: $$\begin{array}{|c||c|c|c|c|c|c|}
1766: \hline
1767: c_{\alpha,\beta}&y_1&y_2&y_3&Y_1&Y_2&Y_3\\
1768: \hline\hline
1769: x_1&1&4&-4&0&12&-12\\
1770: \hline
1771: x_2& 4 &1 &-3 &1 &0 &3 \\
1772: \hline
1773: x_3 &-4 & -3& 1& 0&-3&0 \\
1774: \hline
1775: X_1 &0 &1 &0 &1 &0&-1 \\
1776: \hline
1777: X_2 & 12& 0& -3& 0& 1&36\\
1778: \hline
1779: X_ 3& -12& 3& 0&-1 &36&1\\
1780: \hline
1781: \end{array}
1782: $$
1783:
1784: \mn The 1's on the diagonal stand for the relations $[x_i,y_i]=h_i,$
1785: $[X_i,Y_i]=H_i,$ where, in terms of our basis $\{h,H\}$ for the
1786: Cartan subalgebra, $$h_1=8h+12H,\quad h_2=h+3H, \quad h_3=h,$$
1787: $$ H_1=H,\quad H_2=36(h+H), \quad H_3=36(h+2H).$$
1788: \item $[Cartan, anything]$: this is coded directly by the root diagram:
1789:
1790: - $ad(x)$ has eigenvalues and eigenvectors
1791: $$\begin{array}{l||c|c|c|c|c|c|c}
1792: \hbox{eigenvalue}&3&2&1&0&-1&-2&-3\\
1793: \hline
1794: \hbox{eigenvectors}&X_2,Y_1&x_3& x_1, y_2&X_3,Y_3, h,H&x_2,y_1&y_3&X_1,Y_2
1795: \end{array}$$
1796:
1797: - $ad(X)$ has
1798: eigenvalues and eigenvectors
1799: $$\begin{array}{l||c|c|c|c|c}
1800: \hbox{eigenvalue}&2&1&0&-1&-2\\
1801: \hline
1802: \hbox{eigenvectors}&X_1&X_3,x_2,y_3,Y_2&x_1,y_1,h,H&X_2,x_3,y_2,Y_3&Y_1
1803: \end{array}$$
1804:
1805:
1806: \end{itemize}
1807: \end{proposition}
1808:
1809:
1810:
1811:
1812: \pf This is elementary, using only the Jacobi identity, but takes time. We will give as a typical example
1813: the calculation of $[x_1, x_2]$:
1814:
1815: $$\begin{array}{rcll}
1816: [x_1,x_2]&=&[x_1,[x,X]]&\quad\hbox{(by definition of $x_2$)}\\
1817: &=&[x,[x_1,X]]+[X, [x,x_1]]&\quad\hbox{(Jacobi identity)}\\
1818: &=&[X, [x,x_1]]&\quad\hbox{(since $[x_1,X]=0$)}\\
1819: &=&[X, X_2]=X_3&\quad\hbox{(by definitions of $X_2, X_3$).}
1820: \end{array}
1821: $$
1822: The rest of the relations are derived in a similar fashion. \qed
1823:
1824: \bn
1825:
1826:
1827: Now we are ready to define the generators of the Lie algebra of a maximal compact subgroup $K\subset G_2$. Let
1828: $$L_1=X_1-Y_1, \quad L_2={X_2-Y_2\over 6}, \quad L_3={X_3-Y_3\over 6},$$
1829: $$S_1= {x_1-y_1\over 4}, \quad S_2= {x_2-y_2\over 2}, \quad S_3= {x_3-y_3\over 2}.$$
1830:
1831:
1832: Using the commutation relations of the last Proposition one checks easily that
1833: $$[L_i,L_j]=\epsilon_{ijk}L_k,\quad [L_i,S_j]=\epsilon_{ijk}S_k,\quad [S_i,S_j]=\epsilon_{ijk}({3\over 4}L_k-S_k).$$
1834:
1835: \bn Note: the strange-looking coefficients 2,4,6 in the definition of the $L_i, S_i$ are chosen precisely so that we get
1836: these pleasing commutation relations.
1837:
1838:
1839: \section{The rolling distribution in Cartan's thesis}
1840:
1841: \subsection{Cartan's constructions and claims.}
1842: \label{claims}
1843:
1844: In E. Cartan's thesis \cite{C2}, p.146, we find the following constructions:
1845: consider $V=\R^7=\R^3\X \R^3\X \R$ with coordinates $(\x,\y,z)$, where $\x,\y\in\R^3$, $z\in\R$,
1846: and the following 15 linear vector fields (hence linear operators) on $V $:
1847:
1848:
1849:
1850: \begin{itemize}
1851: \item $X_{ii}=-x_i \partial_{x_i} + y_i\partial_{y_i}+
1852: {1\over 3}\sum_{j=1}^3 (x_j\partial_{x_j}-y_j\partial_{y_j})$, $i=1,2,3$.
1853:
1854: \item $X_{i0}=2z\partial_{x_i}- y_i \partial_z - x_j \partial_{y_k} + x_k\partial_{y_j},$ $(ijk)\in A_3=\{(123),(231),(312)\}.$
1855:
1856: \item $X_{0i}=-2z\partial_{y_i}+ x_i \partial_z +y_j \partial_{x_k} - y_k\partial_{x_j},$ $(ijk)\in A_3.$
1857:
1858: \item $X_{ij}=- x_j \partial_{x_i} + y_i\partial_{y_j},$ $i\neq j,$ $i,j=1,2,3.$
1859:
1860: \end{itemize}
1861:
1862: Cartan makes the following claims without proof:
1863:
1864: \begin{enumerate}
1865:
1866: \item The linear span of these 15 operators is a 14 dimensional Lie subalgebra $\g\subset\End(V)$ isomorphic to $\g_2$.
1867:
1868: \item $\g$ preserves the quadratic form on $V$ given by
1869: $$J=z^2+\x\cdot\y.$$
1870:
1871: \item The linear group $G\subset\GL(V)$ generated by $\g$ acts transitively on the projectivized null cone of $J$.
1872:
1873:
1874: \item $G$ preserves the system of 6 Pfaffian equations on $V,$
1875: given by the 6 components of
1876:
1877: $$
1878: \left\{
1879: \begin{array}{rcl}
1880: \alpha&:=&zd\x-\x dz+\y\X d\y=0,\\
1881: \beta &:=&zd\y-\y dz+\x\X d\x=0,
1882: \end{array}
1883: \right.
1884: $$
1885: which have as a consequence
1886: $$
1887: \left\{
1888: \begin{array}{rcl}
1889: \gamma_1&:=&zdz+\x\cdot d\y=0,\\
1890: \gamma_2&:=&zdz+\y\cdot d\x=0.
1891: \end{array}
1892: \right.
1893: $$
1894:
1895: $$$$
1896: \item $G$ preserves a 5 parameter family of
1897: 3 dimensional linear subspaces of $V,$ contained in the null cone of $J$,
1898:
1899: $$
1900: \left\{
1901: \begin{array}{l}
1902: \x-z\a +\b\X \y=0,\\
1903: \y-z\b +\a\X \x=0,
1904: \end{array}
1905: \right.,
1906: $$
1907: where
1908: $$\a\cdot\b+1=0.$$
1909:
1910:
1911: \end{enumerate}
1912:
1913: Our goal in this appendix is to sketch proofs of these claims,
1914: provide a minor correction in one place,
1915: relate Cartan's construction to the octonions,
1916: and show how they contain, in essence, the
1917: construction of the rolling distribution $\tilde Q$
1918: via projective geometry, as in the proposition \ref{projective}
1919: from section 5.
1920:
1921:
1922: \subsection{Relation with Octonions}.
1923: \label{connectOctonions}
1924:
1925: Recall the basis $e_i , f_i , U $ of section \ref{basis} for $V$ (imaginary split octonions)
1926: with its consequent multiplication table.
1927: Make the change of basis $e_i \mapsto -e_i$,
1928: keeping $f_i, U$ as they were, thus changing the
1929: signs of some entries of the multiplication table.
1930: Use this new basis $E_i = -e_i , f_i , U$
1931: to identify $V$ with $\R^3 \times \R^3 \times \R$
1932: by setting $(\x, \y, z) = \Sigma x_i E_i + \Sigma y_i f_i + z U \in V$.
1933: Referring to the multiplication
1934: table we compute
1935: \begin{eqnarray*}
1936: (\x, \y, z) (\x', \y', z') &= &(-\y \times \y' -z \x' + z' \x, \x \times \x' + z \y' - z' \y, \frac{1}{2}( \x \cdot \y' - \x' \cdot \y))
1937: \\
1938: &&+ 1 \{ z z' + \frac{1}{2} ( \x \cdot \y' - \x' \cdot \y) \}.
1939: \end{eqnarray*}
1940: The last term is in the real part of the split octonions, and not in $V$.
1941: It follows from this formula that $(\x, \y, z)^2 =J$, of Cartan's claim
1942: 2 in the preceding paragraph.
1943: Multiplying out $(\x, \y, z) (d \x, d \y, dz)$ we find that
1944: $$(\x, \y, z) (d \x, d \y, dz)
1945: = (\alpha, \beta, \frac{1}{2} (\gamma_1 - \gamma_2)) +
1946: 1 \{ \frac{1}{2} (\gamma_1 + \gamma_2) \},$$
1947: where $\alpha, \beta, \gamma_1, \gamma_2$
1948: are as in Cartan's claim 4 of the previous paragraph.
1949: It follows that the any element of $G_2 = Aut(\tilde \O)$
1950: preserves $J$ and preserves the Pfaffian system of Cartan's claim 4.
1951: The distribution $D$ defined by this system is, upon restriction
1952: to the null cone $\{J =0 \} \setminus \{ 0 \}$, precisely the distribution $D$ which we defined in the final section of
1953: our paper: $D(\x, \y, z) : =
1954: \{ (\a, \b , c) : (\x, \y, z) (\a, \b, c) = 0 \}$.
1955: It follows that Cartan's construction,
1956: pushed down to the space of rays using the
1957: $\R^+$-action, yields
1958: precisely our $\tilde Q$.
1959:
1960: \subsection{Commentary and proofs of Cartan's claims.}
1961:
1962:
1963: \subsubsection{Definition of $\g_2$}
1964:
1965: {\bf The Cartan subalgebra.} The first 3 operators of claim 1 are linearly dependent since
1966: $\sum_i X_{ii}=0$. This is the only linear relation (proof below) and explains why $\g$ is 14 dimensional and not 15 dimensional. The flows
1967: of $3 X_{ii}$ generate the scalings
1968: $x_i \mapsto \lambda_1 ^{\alpha_i} \lambda_2 ^{\beta_i} \lambda_3 ^{\gamma_i}x_i$
1969: $y_i \mapsto \lambda_1 ^{-\alpha_i} \lambda_2 ^{-\beta_i} \lambda_3 ^{-\gamma_i}y_i$
1970: $z \mapsto z$ as described in section 5. Hence these operators should span the
1971: Cartan $\t$ of $\g = \g_2$.
1972:
1973:
1974:
1975:
1976: \begin{proposition} $\g$ is a 14 dimensional Lie subalgebra of $\End(V)$, isomorphic to $\g_2$, with Cartan subalgebra as just described.
1977: \end{proposition}
1978:
1979: \pf It is convenient to put $\g$ in block matrix form. For each ${\bf u}\in\R^3$ let $R_{\bf u}\in\End(\R^3)$ be given by $ {\bf v}\mapsto{\bf u}\X {\bf v}$; i.e.
1980: $$R_{\bf u}=\left(\begin{array}{ccc}
1981: 0&-u_3&u_2\\
1982: u_3&0&-u_1\\
1983: -u_2&u_1&0
1984: \end{array}\right).$$
1985:
1986: \mn Define the linear map $\rho:\sl_3(\R)\X\R^3\X\R^3\to\End(V)$ by
1987:
1988: $$
1989: \rho(A,\b,\c)=\left(\begin{array}{ccc}
1990: A&R_{\c} &2\b\\
1991: -R_\b&-A^t&-2\c\\
1992: \c^t&-\b^t&0
1993: \end{array}\right).
1994: $$
1995: %
1996: Now $\rho$ is clearly injective, hence its image is a 14 dimensional linear subspace of $\End(V)$.
1997: Denote the components of $A,\b,\c$ by $a_{ij},b_i,c_i$ (resp.), then it is easy to check that
1998: $$\rho(A,\b,\c)=-\sum_{i,j}a_{ij}X_{ij}+\sum_i b_iX_{i0}+\sum_i c_iX_{0i}.$$
1999: %
2000: This shows that $\g$ is the image of $\rho$ and hence a 14 dimensional subspace of $\End(V)$.
2001:
2002: \bn To show that $\g$ is a lie algebra one calculates that $$[\rho(A,\b,\c),\rho(A',\b',\c')]=\rho(A'',\b'',\c''),$$
2003: where
2004: \begin{eqnarray*}
2005: A''&=&[A,A']+3(\b\c'^t-\b'\c^t)- [\b\cdot \c'-\b'\cdot\c]I,\\
2006: \b''&=&A\b'-A'\b-2\c\X \c',\\
2007: \c''&=&-A^t\c'+A'^t\c+2\b\X \b'.
2008: \end{eqnarray*}\\
2009:
2010:
2011: These formulae show that $\{\rho(A,0,0)|A\in\sl_3(\R)\}$ forms a lie subalgebra of $\g$ isomorphic to $\sl_3(\R)$. This
2012: subalgebra corresponds to the sum of the long root
2013: spaces in the root diagram, and the Cartan subalgebra
2014: (the sum of the $X_ii$) as identified earliter.
2015: The formulae also show that
2016: the images of the $\rho(0,\b,0)$ and $\rho(0,0,\c)$ are stable under the adjoint action of the Cartan, hence they must correspond to the remaining short roots.
2017:
2018: A tedious computation now yields the root diagram and the structure
2019: constants of $\g_2$.
2020: \bn
2021:
2022: \subsubsection{Invariance of $J$}
2023: Let $G_2\subset\GL_7(\R)$ be the subgroup generated by $\g$.
2024:
2025:
2026: \begin{proposition} $J$ is $G_2$-invariant.
2027: \end{proposition}
2028:
2029: \pf This is equivalent to showing that every $X\in\g$ is $J$-antisymmetric, i.e. that $X$ anti-commutes with
2030: $$
2031: \left(\begin{array}{ccc}
2032: 0&I/2&0\\
2033: I/2&0&0\\
2034: 0&0&1
2035: \end{array}\right).
2036: $$
2037:
2038:
2039: One now checks easily that the set of $J$-antisymmetric matrices consists of the matrices of the form
2040: $$\left(\begin{array}{ccc}
2041: A&R_\c&2\tilde\b\\
2042: -R_\b&-A^t&-2\tilde\c\\
2043: \tilde\c^t&-\tilde\b^t&0
2044: \end{array}\right),
2045: $$
2046: where $A\in\End(\R^3)$ and $\b,\tilde\b,\c,\tilde\c\in\R^3$. Looking at the formula for
2047: $\rho(A,\b,\c)$ we see that $\g$ is the subset of the $J$-antisymmetric matrices
2048: satisfying $\rm tr A=0, \b=\tilde\b,\c=\tilde\c$ (a codimension 7 condition).\qed
2049:
2050:
2051: \subsubsection{Invariance of the Pfaffian system.}
2052:
2053: \mn
2054:
2055:
2056: \n {\bf Generalities.} A ``Pfaffian system''on a manifold $M$ is given locally by the common kernels of a finite set of 1-forms,
2057: $$\alpha_1=\ldots=\alpha_m=0.$$
2058: Two sets of 1-forms $$\{\alpha_1,\ldots,\alpha_m\},\quad \{\beta_1,\ldots,\beta_n\},$$
2059: give equivalent systems if one can express each element of one set as a linear combination
2060: (with coefficients in $C^{\infty}(M)$) of the elements of the other set. We write this as
2061: $$\alpha_i\equiv 0 \quad \mod \beta_1,\ldots,\beta_n, \quad i=1,\dots, m,$$
2062: and similarly for the $\beta$'s.
2063:
2064:
2065:
2066:
2067: Consequently, if we want to prove that a system is preserved by some diffeomorphism $f:M\to M$
2068: we must show that $$f^*\alpha_i\equiv 0\quad \mod \alpha_1,\ldots,\alpha_m, \quad i=1,\dots, m,$$
2069: and if we want to show that the flow of some vector field $X$ on $M$ preserves the system
2070: we must show that
2071: $$\L_X\alpha_i\equiv 0\quad \mod \alpha_1,\ldots,\alpha_m, \quad i=1,\dots, m.$$
2072:
2073: Given such a system we can consider the common kernels $D_x\subset T_xM$ of the 1-forms at each point $x\in M$.
2074: This is well defined independently of the 1-forms chosen to represent the system. If $\dim D_x$ (the rank of the system)
2075: is constant we obtain a distribution $D\subset TM$ (a subbundle of the tangent bundle). But the rank may vary.
2076: For example, the system on $\R$ given by $xdx=0$ has rank 1 at $x=0$ and rank 0 for $x\neq 0$. However, if $G$ acts on $M$ preserving a Pfaffian system, then the rank must clearly be constant along the $G$-orbits.
2077:
2078:
2079: \bn {\bf Cartan's Pfaffian system. Rank jumps. A correction.} Due to jumping of rank,
2080: as discussed in the last remark, the Pfaffian system
2081: which Cartan defined by
2082: the vanishing of the 6 components of $\alpha,\beta$
2083: cannot be $G_2$ invariant, even when restricted to $\tC$, the $J$ null cone.
2084: For at $(\e_1,0,0)$ the system reduces
2085: to $dx_2=dx_3=dz=0$ and so has rank $4$. On the other hand, at
2086: the point $(\e_1,\e_2,0)$ the system is equivalent to $dy_1=dx_2=dz-dy_3=dz+dx_3=0,$ and so has rank $3$.
2087: And both points lie in $\tC\setminus\{0\}$,
2088: which is a single $G_2$-orbit, contradicting $G_2$ invariance.
2089: A related problem with Cartan's claim 4 of subsection (\ref{claims}) is his claim that $\gamma_1=\gamma_2=0$ is a consequence of $\alpha=\beta$. But this is true only on the $z\neq 0$ part of $\tC$.
2090:
2091: Both errors are fixed by imposing the extra equation $\gamma:=\gamma_1-\gamma_2=0.$ Then, as in section (\ref{connectOctonions}),
2092: we do obtain a $G_2$-invariant system on $V$. Furthermore,
2093: as proved immediately below, the two equations $\gamma_1=\gamma_2=0$ are indeed a consequence of $\alpha=\beta=0, \gamma=0$ on $\tC$, and are a consequence $\alpha=\beta=0$ on the subset $z\neq 0$ of $\tC$.
2094: So Cartan's claim is correct on the open dense set $z\neq 0$ of the null cone $\tC\subset V$.
2095: (See also page 11 of Bryant's paper on Geometric Duality \cite{Bryant}, where he adds the equation $\gamma=0$ to
2096: $\alpha=\beta=0$.)
2097:
2098:
2099:
2100: \begin{proposition} The Pfaffian system on $V$ given by $\alpha=\beta=0, \gamma=0$ is $G_2$-invariant.
2101: On $\tC$ the system is equivalent to $\alpha=\beta=0, \gamma_1=\gamma_2=0$. On the subset $z\neq 0$ of $\tC$
2102: it is equivalent to $\alpha=\beta=0.$
2103: \end{proposition}
2104:
2105: \pf We prove the claims of the last two sentences first.
2106: Note that $\gamma_1+\gamma_2=dJ$. It follows that on $\tC$, where $J=0$,
2107: we have that $\gamma_1=\gamma_2=0$
2108: is a consequence of $\gamma:=\gamma_1-\gamma_2=0.$ Thus, restricted to $\tC$, the system $\alpha=\beta=0, \gamma=0$
2109: is equivalent to $\alpha=\beta=0,\gamma_1=\gamma_2=0$.
2110: Next, note that $\x\cdot\beta-\y\cdot\alpha=z\gamma.$ It follows that on $z\neq 0 $ the equation $\gamma=0$ is a consequence of $\alpha=\beta=0$.\qed
2111:
2112: It remains to establish invariance.
2113: We need to show that $$\L_X\alpha_i\equiv\L_X\beta_j \equiv\L_X\gamma\equiv 0 \mod \alpha_i,\beta_j,\gamma,$$
2114: for all $X=\rho(A,\b,\c)\in\g.$
2115: Divide into 3 cases, corresponding to $(A, 0,0)$ , $(0, \a, 0)$ and $(0, 0, \b)$
2116: in our coordinatization of $\g$.
2117: \begin{itemize}
2118: \item case 1: $X=\rho(A,0,0),$ $A\in\sl_3(\R)$.
2119:
2120: \begin{lemma} If $A\in\End(\R^3)$ and ${\bf u},{\bf v}\in\R^3$, then
2121: $$A({\bf u}\X {\bf v})+A^t{\bf u}\X {\bf v}+ {\bf u}\X A^t{\bf v}={\rm tr }A({\bf u}\X {\bf v}).$$
2122: \end{lemma}
2123: \pf Sketch: divide in 2 cases. If $A^t=-A$ then ${\rm tr }A=0$ and the identity is a consequence of the fact the
2124: $\SO_3$ preserves de cross product and that $\so_3$ are the antisymmetric matrices. If $A^t=A$ then can assume w.l.o.g. that $A$ is diagonal and do an explicit easy calculation. \qed
2125:
2126:
2127: Now since $$X(\x,\y,z)=(A\x,-A^t\y,0), \quad \alpha=zd\x-\x dz+\y\X d\y,$$
2128: we get, using the lemma and ${\rm tr }A=0$, that
2129: \begin{eqnarray*}
2130: \L_X\alpha&=&zAd\x-A\x dz-A^t\y\X d\y-\y\X A^td\y=\\
2131: &=&A(zd\x-\x dz+\y\X d\y)=A\alpha\equiv 0\quad \mod \alpha.
2132: \end{eqnarray*}
2133: Similarly, $\L_X\beta=-A^t\beta\equiv 0\quad (\mod \beta).$
2134:
2135: Finally, $\L_X\gamma=(A\x)\cdot d\y- \x\cdot (A^t d\y)=0.$
2136:
2137: \item case 2: $X=\rho(0,\b,0),$ $\b\in\R^3$.
2138:
2139: Here $$X(\x,\y,z)=(2\b z, -\b\X \x, -\b\cdot\y), $$
2140: and one calculates that
2141: $$\L_X\alpha=\b\gamma,\quad \L_X\beta=\b\X\beta,\quad\L_X\gamma=\b\cdot\beta.$$
2142:
2143:
2144: \item case 3: $X=\rho(0,0,\c), \c\in\R^3.$ The proof for this case is very similar to the previous case. Just interchange $\x$ and $\y$, and $\b$ and $\c$.
2145: \end{itemize}
2146: This completes the proof of invariance,
2147: and hence the proof of the proposition.
2148:
2149:
2150:
2151: %{\bf Caveat. Rank jumps. The $\gamma_i$.}
2152: %Cartan's Pfaffian system from claim 4, defined
2153: %by the vanishing of $\alpha$ and $\beta$ does not agree with our
2154: %$D$ on the locus $z =0$ , due to the fact that its rank jumps along this locus. This discrepancy is corrected by imposing $\gamma_1 = \gamma_2 = 0$.
2155: %Contrary to Cartan's assertion, these last equations are not a consequence of $\alpha =\beta = 0$ when$z = 0$.
2156:
2157:
2158:
2159: \begin{thebibliography}{9}
2160:
2161: \bibitem{Agrachev}
2162: Andrei Agrachev, {\em Rolling balls and Octonions},
2163: math arXivs: math.OC/0611812.
2164:
2165: \bibitem{Serre} J.P.~Serre,
2166: {\em Complex semi-simple Lie algebras},
2167: Springer, reprinted and translated from
2168: {\em Alg\`ebre de Lie semi-simple complexes}, W. A. Benjamin
2169: (1966).
2170:
2171: \bibitem{BryantHsu}
2172: {R. Bryant and L. Hsu},
2173: {Rigidity of Integral Curves of Rank Two Distributions},
2174: {Invent.\ Math.},
2175: vol 114,
2176: p. {435--461},
2177: 1993.
2178:
2179:
2180: \bibitem{Bryant} Robert L. Bryant,
2181: Elie Cartan and Geometric Duality,
2182: lecture notes from a lecture given at the Institut dÕHenri Poincare , on 19 June 1998;
2183: available on Bryant's website.
2184:
2185:
2186: \bibitem{C1} E. Cartan, {Les syst{\`e}mes de {P}faff {\`a} cinque variables et l{\`e}s
2187: {\'e}quations aux d{\'e}riv{\'e}es partielles du second ordre},
2188: {Ann.\ Sci.\ {\`E}cole Normale},
2189: vol. 27, no. 3, p. {109--192}, 1910. (Reprinted in Oeuvres completes, Partie III, vol 2. 137-288.)
2190:
2191:
2192: \bibitem{C2} E. Cartan, {Sur la structure des groupes de transformations finis et continus}, These, Paris, 1894. (Reprinted in Oeuvres completes, Partie I, vol 1.)
2193:
2194:
2195: \bibitem{C3} E. Cartan. Le groupse reels simples finis et continus, Ann. Ec. Norm. v. 31, pp 263-355, (1914). (Reprinted in Oeuvres completes, Partie I, vol 1.)
2196:
2197:
2198: \bibitem{Chern} S.S Chern and C. Chevalley, {Elie Cartan and his mathematical work}
2199: (Reprinted in Oeuvres, partie III vol. 2. )
2200:
2201:
2202: \bibitem{Hammersley} J. M. Hammersley, {\em Oxford commemoration ball},
2203: in {\bf Probability, Statistics, and Analysis}, London Math. Society, Lecture Note
2204: Series, vol. 79, Cambridge U. Press, [1983].
2205:
2206:
2207: \bibitem{Harvey} Reese
2208: Harvey, {\bf Spinors and Calibration}, Acad. Press, 1990.
2209:
2210: \bibitem{Kaplan} Aroldo Kaplan and F. Levstein, {\em A split Fano plane}
2211: in preparation.
2212:
2213:
2214: \bibitem{Montbook}{R. Montgomery} {\bf A tour of sub-Riemannian Geometry},
2215: AMS, 2001.
2216:
2217: \bibitem{Serre} J-P. Serre, (translated from the French by G.A. Jones)
2218: {\bf Complex Semisimple Lie algebras}, Springer-Verlag.
2219: 1987.
2220:
2221: \bibitem{Tanaka} N. Tanaka, {\em On the differential systems, graded Lie
2222: algebras and pseudo-groups}, J. Math. Kyoto Univ. , v. 10, 1-82, 1970.
2223:
2224: \bibitem{Tanaka2}N. Tanaka, {\em On the equivalence problem associated
2225: with simple graded Lie algebras}, Hokkaido Math. Journal,
2226: v. 8, 23-84, 1979.
2227:
2228: \bibitem{Vogan} D. Vogan, {\em The unitary dual of $G_2$}, Inventiones, vol. 116, 677-791,
2229: (esp. p. 679), 1994.
2230:
2231: \bibitem{Yamaguchi} K. Yamaguchi, {\em Differential Systems Associated with Simple graded Lie Algebras}, in {\bf Progress in Differential Geometry}, Advanced Studies in Math, v. 22, 413-494, 1993.
2232:
2233:
2234: \end{thebibliography}
2235:
2236:
2237: \end{document}
2238:
2239:
2240: