1:
2: %prepared in LaTeX
3:
4: \input{epsf}
5:
6: \hfuzz = 9pt
7:
8: \newcommand\Z{{\mathbb Z}}
9: \newcommand\R{{\mathbb R}}
10: \newcommand\C{{\mathbb C}}
11: \renewcommand\H{{\mathbb H}}
12: \newcommand\K{{\mathbb K}}
13: \renewcommand\O{{\mathbb O}}
14: \renewcommand\S{{\mathbb S}}
15:
16: \newcommand\RP{{\mathbb {RP}}}
17: \newcommand\CP{{\mathbb {CP}}}
18: \newcommand\HP{{\mathbb {HP}}}
19: \newcommand\KP{{\mathbb {KP}}}
20: \newcommand\OP{{\mathbb {OP}}}
21: \renewcommand\P{{\mathbb P}}
22:
23: \newcommand{\Cliff}{{\rm Cliff}}
24: \newcommand{\J}{{\rm J}}
25: \newcommand{\M}{{\rm M}}
26:
27: \newcommand{\OO}{{\rm O}}
28: \newcommand{\SO}{{\rm SO}}
29: \newcommand{\SL}{{\rm SL}}
30: \newcommand{\PSL}{{\rm PSL}}
31: \newcommand{\SU}{{\rm SU}}
32: \newcommand{\Sp}{{\rm Sp}}
33: \newcommand{\Spin}{{\rm Spin}}
34: \newcommand{\Pin}{{\rm Pin}}
35: \newcommand{\U}{{\rm U}}
36: \newcommand{\E}{{\rm E}}
37: \newcommand{\F}{{\rm F}}
38: \newcommand{\G}{{\rm G}}
39:
40: \newcommand{\so}{{\mathfrak {so}}}
41: \newcommand{\Sl}{{\mathfrak {sl}}}
42: \newcommand{\symp}{{\mathfrak {sp}}}
43: \renewcommand{\u}{{\mathfrak {u}}}
44: \newcommand{\e}{{\mathfrak {e}}}
45: \newcommand{\f}{{\mathfrak {f}}}
46: \newcommand{\g}{{\mathfrak {g}}}
47: \newcommand{\su}{{\mathfrak {su}}}
48: \renewcommand{\a}{{\mathfrak {a}}}
49: \newcommand{\sa}{{\mathfrak {sa}}}
50: \newcommand{\h}{{\mathfrak {h}}}
51: \newcommand{\sh}{{\mathfrak {sh}}}
52: \newcommand{\coll}{{\mathfrak {coll}}}
53: \newcommand{\isom}{{\mathfrak {isom}}}
54: \newcommand{\Isom}{{\rm Isom}}
55: \newcommand{\Der}{{\mathfrak {der}}}
56: \newcommand{\Tri}{{\mathfrak {tri}}}
57:
58:
59: \newcommand{\et}{\hspace{-0.08in}{\bf .}\hspace{0.1in}}
60: \newcommand{\DOT}{\hspace{-0.08in}{\bf .}\hspace{0.1in}}
61: \newcommand{\BOX}{\hbox {$\sqcap$ \kern -1em $\sqcup$}}
62: \newcommand{\qed}{\hskip 3em \hbox{\BOX} \vskip 2ex}
63:
64: \renewcommand{\Re}{{\rm Re}}
65: \renewcommand{\Im}{{\rm Im}}
66: \newcommand{\Hom}{{\rm hom}}
67: \newcommand{\Aut}{{\rm Aut}}
68: \newcommand{\Inv}{{\rm Inv}}
69: \newcommand{\tensor}{\otimes}
70: \newcommand{\iso}{\cong}
71: \newcommand{\implies}{\Longrightarrow}
72: \newcommand{\impliedby}{\Longleftarrow}
73:
74: \newcommand{\be}{\begin{equation}}
75: \newcommand{\ee}{\end{equation}}
76: \newcommand{\ba}{\begin{eqnarray}}
77: \newcommand{\ea}{\end{eqnarray}}
78: \newcommand{\ban}{\begin{eqnarray*}}
79: \newcommand{\ean}{\end{eqnarray*}}
80:
81: \newcommand{\ad}{{\rm ad}}
82: \newcommand{\Cob}{{\rm Cob}}
83: \newcommand{\Vect}{{\rm Vect}}
84: \newcommand{\Hilb}{{\rm Hilb}}
85: \newcommand{\maps}{\colon}
86: \newcommand{\tr}{{\rm tr}}
87: \newcommand{\rank}{{\rm rank}}
88: \newcommand{\id}{{\rm id}}
89: \newcommand{\Fun}{{\rm Fun}}
90: \newcommand{\Ad}{{\rm Ad}}
91: \newcommand{\om}{\omega}
92: \newcommand{\eps}{\epsilon}
93:
94: \newtheorem{thm}{Theorem}
95: \newtheorem{prop}{Proposition}
96: \newtheorem{lemma}{Lemma}
97: \newtheorem{corollary}{Corollary}
98: \newtheorem{ex}{Example}
99:
100: \documentclass {article}
101: \usepackage {amsfonts}
102:
103: \textwidth 6in
104: \textheight 8.5in \evensidemargin .25in
105: \oddsidemargin .25in
106: \topmargin .25in
107: \headsep 0in
108: \headheight 0in
109: \footskip .5in
110: %If you want single spaced copy, delete the next two lines.
111: % \parskip 1.75\parskip plus 3pt minus 1pt
112: % \renewcommand{\baselinestretch}{1.5}
113: \pagestyle{plain}
114: \pagenumbering{arabic}
115: \begin{document}
116:
117: \begin{center}
118: {\bf The Octonions \\}
119: {\em John\ C.\ Baez\\}
120: \vspace{0.3cm}
121: {\small Department of Mathematics \\
122: University of California\\
123: Riverside CA 92521\\}
124: {\small email: baez@math.ucr.edu\\}
125: \vspace{0.3cm}
126: {May 16, 2001 \\}
127: \vspace{0.3cm}
128: \end{center}
129:
130: \begin{abstract}
131: \noindent
132: The octonions are the largest of the four normed division algebras.
133: While somewhat neglected due to their nonassociativity, they stand at
134: the crossroads of many interesting fields of mathematics. Here we
135: describe them and their relation to Clifford algebras and spinors, Bott
136: periodicity, projective and Lorentzian geometry, Jordan algebras, and
137: the exceptional Lie groups. We also touch upon their applications in
138: quantum logic, special relativity and supersymmetry.
139: \end{abstract}
140:
141: \section{Introduction}
142:
143: There are exactly four normed division algebras: the real numbers
144: ($\R$), complex numbers ($\C$), quaternions ($\H$), and octonions
145: ($\O$). The real numbers are the dependable breadwinner of the family,
146: the complete ordered field we all rely on. The complex numbers are a
147: slightly flashier but still respectable younger brother: not ordered,
148: but algebraically complete. The quaternions, being noncommutative, are
149: the eccentric cousin who is shunned at important family gatherings. But
150: the octonions are the crazy old uncle nobody lets out of the attic: they
151: are {\it nonassociative}.
152:
153: Most mathematicians have heard the story of how Hamilton invented the
154: quaternions. In 1835, at the age of 30, he had discovered how to treat
155: complex numbers as pairs of real numbers. Fascinated by the relation
156: between $\C$ and 2-dimensional geometry, he tried for many years to
157: invent a bigger algebra that would play a similar role in 3-dimensional
158: geometry. In modern language, it seems he was looking for a 3-dimensional
159: normed division algebra. His quest built to its climax in October 1843.
160: He later wrote to his son, ``Every morning in the early part of the
161: above-cited month, on my coming down to breakfast, your (then) little
162: brother William Edwin, and yourself, used to ask me: `Well, Papa, can you
163: {\it multiply} triplets?' Whereto I was always obliged to reply, with a sad
164: shake of the head: `No, I can only {\it add} and subtract them'.''
165: The problem, of course, was that there exists no 3-dimensional normed
166: division algebra. He really needed a 4-dimensional algebra.
167:
168: Finally, on the 16th of October, 1843, while walking with his wife along
169: the Royal Canal to a meeting of the Royal Irish Academy in Dublin, he made
170: his momentous discovery. ``That is to say, I then and there felt the
171: galvanic circuit of thought {\it close}; and the sparks which fell from it
172: were the {\it fundamental equations between $i,j,k$; exactly such} as I have
173: used them ever since.'' And in a famous act of mathematical vandalism, he
174: carved these equations into the stone of the Brougham Bridge:
175: \[ i^2 = j^2 = k^2 = ijk = -1 .\]
176:
177: One reason this story is so well-known is that Hamilton spent the rest
178: of his life obsessed with the quaternions and their applications to
179: geometry \cite{Graves,Hankins}. And for a while, quaternions were
180: fashionable. They were made a mandatory examination topic in Dublin,
181: and in some American universities they were the only advanced
182: mathematics taught. Much of what we now do with scalars and vectors in
183: $\R^3$ was then done using real and imaginary quaternions. A school
184: of `quaternionists' developed, which was led after Hamilton's death by
185: Peter Tait of Edinburgh and Benjamin Peirce of Harvard. Tait wrote 8
186: books on the quaternions, emphasizing their applications to physics.
187: When Gibbs invented the modern notation for the dot product and cross
188: product, Tait condemned it as a ``hermaphrodite monstrosity''. A war of
189: polemics ensued, with such luminaries as Heaviside weighing
190: in on the side of vectors. Ultimately the quaternions lost, and
191: acquired a slight taint of disgrace from which they have never fully
192: recovered \cite{Crowe}.
193:
194: Less well-known is the discovery of the octonions by Hamilton's friend
195: from college, John T.\ Graves. It was Graves' interest in algebra that
196: got Hamilton thinking about complex numbers and triplets in the first
197: place. The very day after his fateful walk, Hamilton sent an 8-page
198: letter describing the quaternions to Graves. Graves replied on October
199: 26th, complimenting Hamilton on the boldness of the idea, but adding
200: ``There is still something in the system which gravels me. I have not
201: yet any clear views as to the extent to which we are at liberty
202: arbitrarily to create imaginaries, and to endow them with supernatural
203: properties.'' And he asked: ``If with your alchemy you can make three
204: pounds of gold, why should you stop there?''
205:
206: Graves then set to work on some gold of his own! On December 26th,
207: he wrote to Hamilton describing a new 8-dimensional algebra, which he
208: called the `octaves'. He showed that they were a normed division
209: algebra, and used this to express the product of two sums of eight
210: perfect squares as another sum of eight perfect squares: the `eight
211: squares theorem' \cite{Hamilton}.
212:
213: In January 1844, Graves sent three letters to Hamilton expanding on his
214: discovery. He considered the idea of a general theory of
215: `$2^m$-ions', and tried to construct a 16-dimensional normed division
216: algebra, but he ``met with an unexpected hitch'' and came to doubt that
217: this was possible. Hamilton offered to publicize Graves' discovery, but
218: being busy with work on quaternions, he kept putting it off. In July he
219: wrote to Graves pointing out that the octonions were nonassociative:
220: ``$A \cdot BC = AB \cdot C = ABC$, if $A,B,C$ be quaternions, but not
221: so, generally, with your octaves.'' In fact, Hamilton first invented
222: the term `associative' at about this time, so the octonions may have
223: played a role in clarifying the importance of this concept.
224:
225: Meanwhile the young Arthur Cayley, fresh out of Cambridge, had been
226: thinking about the quaternions ever since Hamilton announced their
227: existence. He seemed to be seeking relationships between the
228: quaternions and hyperelliptic functions. In March of 1845, he published
229: a paper in the {\it Philosophical Magazine} entitled `On Jacobi's
230: Elliptic Functions, in Reply to the Rev.\ B.\ Bronwin; and on Quaternions'
231: \cite{Cayley}. The bulk of this paper was an attempt to rebut an
232: article pointing out mistakes in Cayley's work on elliptic functions.
233: Apparently as an afterthought, he tacked on a brief description of the
234: octonions. In fact, this paper was so full of errors that it was
235: omitted from his collected works --- except for the part about octonions
236: \cite{Cayley2}.
237:
238: Upset at being beaten to publication, Graves attached a postscript to a
239: paper of his own which was to appear in the following issue of the same
240: journal, saying that he had known of the octonions ever since Christmas,
241: 1843. On June 14th, 1847, Hamilton contributed a short note to the
242: Transactions of the Royal Irish Academy, vouching for Graves' priority.
243: But it was too late: the octonions became known as `Cayley numbers'.
244: Still worse, Graves later found that his eight squares theorem had
245: already been discovered by C.\ F.\ Degen in 1818 \cite{Curtis,Dickson}.
246:
247: Why have the octonions languished in such obscurity compared to the
248: quaternions? Besides their rather inglorious birth, one reason is that
249: they lacked a tireless defender such as Hamilton. But surely the reason
250: for {\it this} is that they lacked any clear application to geometry and
251: physics. The unit quaternions form the group $\SU(2)$, which is the
252: double cover of the rotation group $\SO(3)$. This makes them nicely
253: suited to the study of rotations and angular momentum, particularly in
254: the context of quantum mechanics. These days we regard this phenomenon
255: as a special case of the theory of Clifford algebras. Most of us no
256: longer attribute to the quaternions the cosmic significance that
257: Hamilton claimed for them, but they fit nicely into our understanding of
258: the scheme of things.
259:
260: The octonions, on the other hand, do not. Their relevance to geometry
261: was quite obscure until 1925, when \'Elie Cartan described `triality'
262: --- the symmetry between vectors and spinors in 8-dimensional Euclidean
263: space \cite{Cartan3}. Their potential relevance to physics was noticed
264: in a 1934 paper by Jordan, von Neumann and Wigner on the foundations of
265: quantum mechanics \cite{JNW}. However, attempts by Jordan and others to
266: apply octonionic quantum mechanics to nuclear and particle physics met
267: with little success. Work along these lines continued quite slowly
268: until the 1980s, when it was realized that the octonions explain some
269: curious features of string theory \cite{KT}. The Lagrangian for
270: the classical superstring involves a relationship between vectors and
271: spinors in Minkowski spacetime which holds only in 3, 4, 6, and 10
272: dimensions. Note that these numbers are 2 more than the dimensions of
273: $\R,\C,\H$ and $\O$. As we shall see, this is no coincidence: briefly,
274: the isomorphisms
275: \[ \begin{array}{lcl}
276: \Sl(2,\R) &\iso& \so(2,1) \\
277: \Sl(2,\C) &\iso& \so(3,1) \\
278: \Sl(2,\H) &\iso& \so(5,1) \\
279: \Sl(2,\O) &\iso& \so(9,1)
280: \end{array}
281: \]
282: allow us to treat a spinor in one of these dimensions as a pair of
283: elements of the corresponding division algebra. It is fascinating
284: that of these superstring Lagrangians, it is the 10-dimensional
285: octonionic one that gives the most promising candidate for a realistic
286: theory of fundamental physics! However, there is still no {\it proof}
287: that the octonions are useful for understanding the real world. We
288: can only hope that eventually this question will be settled one way or
289: another.
290:
291: Besides their possible role in physics, the octonions are important
292: because they tie together some algebraic structures that otherwise
293: appear as isolated and inexplicable exceptions. As we shall explain,
294: the concept of an octonionic projective space $\OP^n$ only makes sense
295: for $n \le 2$, due to the nonassociativity of $\O$. This means that
296: various structures associated to real, complex and quaternionic
297: projective spaces have octonionic analogues only for $n \le 2$.
298:
299: Simple Lie algebras are a nice example of this phenomenon. There are
300: 3 infinite families of `classical' simple Lie algebras, which come from
301: the isometry groups of the projective spaces $\RP^n$, $\CP^n$ and
302: $\HP^n$. There are also 5 `exceptional' simple Lie algebras. These
303: were discovered by Killing and Cartan in the late 1800s. At the time,
304: the significance of these exceptions was shrouded in mystery: they did
305: not arise as symmetry groups of known structures. Only later did their
306: connection to the octonions become clear. It turns out that 4 of them
307: come from the isometry groups of the projective planes over $\O$, $\O
308: \tensor \C$, $\O \tensor \H$ and $\O \tensor \O$. The remaining one is
309: the automorphism group of the octonions!
310:
311: Another good example is the classification of simple formally real
312: Jordan algebras. Besides several infinite families of these, there
313: is the `exceptional' Jordan algebra, which consists of $3 \times 3$
314: hermitian octonionic matrices. Minimal projections in this Jordan
315: algebra correspond to points of $\OP^2$, and the automorphism group of
316: this algebra is the same as the isometry group of $\OP^2$.
317:
318: The octonions also have fascinating connections to topology. In 1957,
319: Raoul Bott computed the homotopy groups of the topological group
320: $\OO(\infty)$, which is the inductive limit of the orthogonal groups
321: $\OO(n)$ as $n \to \infty$. He proved that they repeat with period
322: 8:
323: \[ \pi_{i+8}(\OO(\infty)) \iso \pi_i(\OO(\infty)). \]
324: This is known as `Bott periodicity'. He also computed the first 8:
325: \ban
326: \pi_0(\OO(\infty)) &\iso& \Z_2 \\
327: \pi_1(\OO(\infty)) &\iso& \Z_2 \\
328: \pi_2(\OO(\infty)) &\iso& 0 \\
329: \pi_3(\OO(\infty)) &\iso& \Z \\
330: \pi_4(\OO(\infty)) &\iso& 0 \\
331: \pi_5(\OO(\infty)) &\iso& 0 \\
332: \pi_6(\OO(\infty)) &\iso& 0 \\
333: \pi_7(\OO(\infty)) &\iso& \Z
334: \ean
335: Note that the nonvanishing homotopy groups here occur in dimensions one
336: less than the dimensions of $\R,\C,\H$, and $\O$. This is no coincidence!
337: In a normed division algebra, left multiplication by an element of norm
338: one defines an orthogonal transformation of the algebra, and thus an
339: element of $\OO(\infty)$. This gives us maps from the spheres $S^0,
340: S^1, S^3$ and $S^7$ to $\OO(\infty)$, and these maps generate the
341: homotopy groups in those dimensions.
342:
343: Given this, one might naturally guess that the period-8 repetition in
344: the homotopy groups of $\OO(\infty)$ is in some sense `caused' by the
345: octonions. As we shall see, this is true. Conversely, Bott
346: periodicity is closely connected to the problem of how many pointwise
347: linearly independent smooth vector fields can be found on the
348: $n$-sphere \cite{Husemoller}. There exist $n$ such vector fields only
349: when $n+1 = 1, 2, 4,$ or $8$, and this can be used to show that
350: division algebras over the reals can only occur in these dimensions.
351:
352: In what follows we shall try to explain the octonions and their role in
353: algebra, geometry, and topology. In Section \ref{constructing} we give
354: four constructions of the octonions: first via their multiplication
355: table, then using the Fano plane, then using the Cayley--Dickson
356: construction and finally using Clifford algebras, spinors, and a
357: generalized concept of `triality' advocated by Frank Adams \cite{Adams}.
358: Each approach has its own merits. In Section \ref{proj} we discuss
359: the projective lines and planes over the normed division algebras ---
360: especially $\O$ --- and describe their relation to Bott periodicity,
361: the exceptional Jordan algebra, and the Lie algebra isomorphisms listed
362: above. Finally, in Section \ref{lie} we discuss octonionic
363: constructions of the exceptional Lie groups, especially the `magic
364: square'.
365:
366: \subsection{Preliminaries} \label{preliminaries}
367:
368: Before our tour begins, let us settle on some definitions. For us a
369: {\bf vector space} will always be a finite-dimensional module over the
370: field of real numbers. An {\bf algebra} $A$ will be a vector space that
371: is equipped with a bilinear map $m \maps A \times A \to A$ called
372: `multiplication' and a nonzero element $1 \in A$ called the `unit' such
373: that $m(1,a) = m(a,1) = a$. As usual, we abbreviate $m(a,b)$ as $ab$.
374: We do not assume our algebras are associative! Given an algebra, we
375: will freely think of real numbers as elements of this algebra via the
376: map $\alpha \mapsto \alpha 1$.
377:
378: An algebra $A$ is a {\bf division algebra} if given $a,b \in A$ with $ab
379: = 0$, then either $a = 0$ or $b = 0$. Equivalently, $A$ is a division
380: algebra if the operations of left and right multiplication by any
381: nonzero element are invertible. A {\bf normed division algebra} is an
382: algebra $A$ that is also a normed vector space with $\|ab\| = \|a\|
383: \|b\|$. This implies that $A$ is a division algebra and that $\|1\| =
384: 1$.
385:
386: We should warn the reader of some subtleties. We say an algebra $A$ has
387: {\bf multiplicative inverses} if for any nonzero $a \in A$ there is an
388: element $a^{-1} \in A$ with $aa^{-1} = a^{-1}a = 1$. An associative
389: algebra has multiplicative inverses iff it is a division
390: algebra. However, this fails for nonassociative algebras! In Section
391: \ref{cayley-dickson} we shall construct algebras that have
392: multiplicative inverses, but are not division algebras. On the other
393: hand, we can construct a division algebra without multiplicative
394: inverses by taking the quaternions and modifying the product slightly,
395: setting $i^2 = -1 + \epsilon j$ for some small nonzero real number
396: $\epsilon$ while leaving the rest of the multiplication table unchanged.
397: The element $i$ then has both right and left inverses, but they are not
398: equal. (We thank David Rusin for this example.)
399:
400: There are three levels of associativity. An algebra is {\bf
401: power-associative} if the subalgebra generated by any one element is
402: associative. It is {\bf alternative} if the subalgebra generated by any
403: two elements is associative. Finally, if the subalgebra generated by any
404: three elements is associative, the algebra is associative.
405:
406: As we shall see, the octonions are not associative, but they are alternative.
407: How does one check a thing like this? By a theorem of Emil Artin
408: \cite{Schafer}, an algebra $A$ is alternative iff for all $a,b \in A$ we have
409: \ba (aa)b = a(ab), \qquad (ab)a = a(ba), \qquad (ba)a = b(aa)
410: \label{alternative} \ea
411: In fact, any two of these equations implies the remaining one, so people
412: usually take the first and last as the definition of `alternative'.
413: To see this fact, note that any algebra has a trilinear map
414: \[ [\cdot,\cdot,\cdot] \maps A^3 \to A \]
415: called the {\bf associator}, given by
416: \[ [a,b,c] = (ab)c - a(bc) .\]
417: The associator measures the failure of associativity just as the
418: commutator $[a,b] = ab - ba$ measures the failure of commutativity.
419: Now, the commutator is an alternating bilinear map, meaning that it
420: switches sign whenever the two arguments are exchanged:
421: \[ [a,b] = -[b,a] \]
422: or equivalently, that it vanishes when they are equal:
423: \[ [a,a] = 0 .\]
424: This raises the question of whether the associator is alternating too.
425: In fact, this holds precisely when $A$ is alternative! The reason is
426: that each equation in (\ref{alternative}) says that the associator
427: vanishes when a certain pair of arguments are equal, or equivalently,
428: that it changes sign when that pair of arguments is switched. Note,
429: however, that if the associator changes sign when we switch the $i$th
430: and $j$th arguments, and also when we switch the $j$th and $k$th
431: arguments, it must change sign when we switch the $i$th and $k$th.
432: Thus any two of equations (\ref{alternative}) imply the third.
433:
434: Now we can say what is so great about $\R,\C,\H,$ and $\O$:
435:
436: \begin{thm} \et \label{hurwitz}
437: $\R,\C,\H$, and $\O$ are the only normed division algebras.
438: \end{thm}
439:
440: \begin{thm} \et \label{zorn}
441: $\R,\C,\H$, and $\O$ are the only alternative division algebras.
442: \end{thm}
443:
444: The first theorem goes back to an 1898 paper by Hurwitz \cite{Hurwitz}.
445: It was subsequently generalized in many directions, for example, to
446: algebras over other fields. A version of the second theorem appears in
447: an 1930 paper by Zorn \cite{Zorn} --- the guy with the lemma. For
448: modern proofs of both these theorems, see Schafer's excellent book on
449: nonassociative algebras \cite{Schafer}. We sketch a couple proofs of
450: Hurwitz's theorem in Section \ref{clifford}.
451:
452: Note that we did {\it not} state that $\R,\C,\H$ and $\O$ are the only
453: division algebras. This is not true. For example, we have already
454: described a way to get 4-dimensional division algebras that do not have
455: multiplicative inverses. However, we do have this fact:
456:
457: \begin{thm} \et \label{bott-milnor} All division algebras have dimension
458: $1, 2, 4,$ or $8$.
459: \end{thm}
460:
461: \noindent
462: This was independently proved by Kervaire \cite{Kervaire} and
463: Bott--Milnor \cite{BM} in 1958. We will say a bit about the proof in
464: Section \ref{OP1}. However, in what follows our main focus will not be
465: on general results about division algebras. Instead, we concentrate on
466: special features of the octonions. Let us begin by constructing them.
467:
468: \section{Constructing the Octonions} \label{constructing}
469:
470: The most elementary way to construct the octonions is to give their
471: multiplication table. The octonions are an 8-dimensional algebra
472: with basis $1, e_1,e_2,e_3,e_4,e_5,e_6,e_7$,
473: and their multiplication is given in this table, which describes
474: the result of multiplying the element in the $i$th row by the
475: element in the $j$th column:
476:
477: \vskip 2em
478: {\vbox{
479: \begin{center}
480: {\small
481: \begin{tabular}{|c|c|c|c|c|c|c|c|c|} \hline
482: & $e_1$ & $e_2$ & $e_3$ & $e_4$ & $e_5$ & $e_6$ & $e_7$ \\ \hline
483: $e_1$ & $-1$ & $e_4$ & $e_7$ & $-e_2$ & $e_6$ & $-e_5$ & $-e_3$ \\ \hline
484: $e_2$ & $-e_4$ & $-1$ & $e_5$ & $e_1$ & $-e_3$ & $e_7$ & $-e_6$ \\ \hline
485: $e_3$ & $-e_7$ & $-e_5$ & $-1$ & $e_6$ & $e_2$ & $-e_4$ & $e_1$ \\ \hline
486: $e_4$ & $e_2$ & $-e_1$ & $-e_6$ & $-1$ & $e_7$ & $e_3$ & $-e_5$ \\ \hline
487: $e_5$ & $-e_6$ & $e_3$ & $-e_2$ & $-e_7$ & $-1$ & $e_1$ & $e_4$ \\ \hline
488: $e_6$ & $e_5$ & $-e_7$ & $e_4$ & $-e_3$ & $-e_1$ & $-1$ & $e_2$ \\ \hline
489: $e_7$ & $e_3$ & $e_6$ & $-e_1$ & $e_5$ & $-e_4$ & $-e_2$ & $-1$ \\ \hline
490: \end{tabular}}
491: \end{center}
492: \vskip 1em
493: \centerline{Table 1 --- Octonion Multiplication Table}
494: }}
495: \vskip 1em
496:
497: \noindent
498: Unfortunately, this table is almost completely unenlightening! About the only
499: interesting things one can easily learn from it are:
500: \begin{itemize}
501: \item $e_1,\dots,e_7$ are square roots of -1,
502: \item $e_i$ and $e_j$ anticommute when $i \ne j$:
503: \[ e_i e_j = -e_j e_i \]
504: \item the {\bf index cycling} identity holds:
505: \[ e_i e_j = e_k \; \implies\; e_{i+1} e_{j+1} = e_{k+1} \]
506: where we think of the indices as living in $\Z_7$, and
507: \item the {\bf index doubling} identity holds:
508: \[ e_i e_j = e_k \; \implies \; e_{2i} e_{2j} = e_{2k} . \]
509: \end{itemize}
510: Together with a single nontrivial product like $e_1 e_2 = e_4$, these
511: facts are enough to recover the whole multiplication table. However, we
512: really want a better way to remember the octonion product. We should
513: become as comfortable with multiplying octonions as we are with
514: multiplying matrices! And ultimately, we want a more conceptual
515: approach to the octonions, which explains their special properties and
516: how they fit in with other mathematical ideas. In what follows, we give
517: some more descriptions of octonion multiplication, starting with a nice
518: mnemonic, and working up to some deeper, more conceptual ones.
519:
520: \subsection{The Fano plane} \label{fano}
521:
522: The quaternions, $\H$, are a 4-dimensional algebra with basis $1,i,j,k$.
523: To describe the product we could give a multiplication
524: table, but it is easier to remember that:
525: \begin{itemize}
526: \item $1$ is the multiplicative identity,
527: \item $i,j,$ and $k$ are square roots of -1,
528: \item we have $ij = k$, $ji = -k$, and all identities obtained
529: from these by cyclic permutations of $(i,j,k)$.
530: \end{itemize}
531: We can summarize the last rule in a picture:
532:
533: \centerline{\epsfysize=1.5in\epsfbox{triangle.eps}}
534: \label{triangle}
535:
536: \noindent
537: When we multiply two elements going clockwise around the circle we get
538: the next one: for example, $ij = k$. But when we multiply two
539: going around counterclockwise, we get {\it minus} the next one:
540: for example, $ji = -k$.
541:
542: We can use the same sort of picture to remember how to multiply
543: octonions:
544: \medskip
545:
546: \centerline{\epsfysize=1.5in\epsfbox{fano.eps}}
547: \label{Fano}
548: \medskip
549:
550: \noindent
551: This is the {\bf Fano plane}, a little gadget
552: with 7 points and 7 lines. The `lines' are the sides of the triangle,
553: its altitudes, and the circle containing all the midpoints of the sides.
554: Each pair of distinct points lies on a unique line. Each line contains
555: three points, and each of these triples has has a cyclic ordering
556: shown by the arrows. If $e_i, e_j,$ and $e_k$ are cyclically ordered
557: in this way then
558: \[ e_i e_j = e_k, \qquad e_j e_i = -e_k . \]
559: Together with these rules:
560: \begin{itemize}
561: \item $1$ is the multiplicative identity,
562: \item $e_1, \dots, e_7$ are square roots of -1,
563: \end{itemize}
564: the Fano plane completely describes the algebra structure of the
565: octonions. Index-doubling corresponds to rotating the picture
566: a third of a turn.
567:
568: This is certainly a neat mnemonic, but is there anything deeper lurking
569: behind it? Yes! The Fano plane is the projective plane over the 2-element
570: field $\Z_2$. In other words, it consists of lines through the origin
571: in the vector space $\Z_2^3$. Since every such line contains a single
572: nonzero element, we can also think of the Fano plane as consisting of the
573: seven nonzero elements of $\Z_2^3$. If we think of the origin in $\Z_2^3$
574: as corresponding to $1 \in \O$, we get the following picture of the
575: octonions:
576:
577: \medskip
578: \centerline{\epsfysize=1.5in\epsfbox{cube.eps}}
579: \label{cube}
580: \medskip
581:
582: \noindent
583: Note that planes through the origin of this 3-dimensional vector space
584: give subalgebras of $\O$ isomorphic to the quaternions, lines through
585: the origin give subalgebras isomorphic to the complex numbers, and
586: the origin itself gives a subalgebra isomorphic to the real numbers.
587:
588: What we really have here is a description of the octonions as a
589: `twisted group algebra'. Given any group $G$, the group algebra
590: $\R[G]$ consists of all finite formal linear combinations of elements
591: of $G$ with real coefficients. This is an associative algebra with
592: the product coming from that of $G$. We can use any function
593: \[ \alpha \maps G^2 \to \{ \pm 1 \} \]
594: to `twist' this product, defining a new product
595: \[ \star \maps \R[G] \times \R[G] \to \R[G] \]
596: by:
597: \[ g \star h = \alpha(g,h) \; gh, \]
598: where $g,h \in G \subset \R[G]$. One can figure out an equation
599: involving $\alpha$ that guarantees this new product will be associative.
600: In this case we call $\alpha$ a `2-cocycle'. If $\alpha$ satisfies a
601: certain extra equation, the product $\star$ will also be commutative,
602: and we call $\alpha$ a `stable 2-cocycle'. For example, the group
603: algebra $\R[\Z_2]$ is isomorphic to a product of 2 copies of $\R$,
604: but we can twist it by a stable 2-cocyle to obtain the complex numbers.
605: The group algebra $\R[\Z_2^2]$ is isomorphic to a product of 4 copies
606: of $\R$, but we can twist it by a 2-cocycle to obtain the quaternions.
607: Similarly, the group algebra $\R[\Z_2^3]$ is a product of 8 copies of $\R$,
608: and what we have really done in this section is describe a function
609: $\alpha$ that allows us to twist this group algebra to obtain the
610: octonions. Since the octonions are nonassociative, this function is
611: not a 2-cocycle. However, its coboundary is a `stable 3-cocycle', which
612: allows one to define a new associator and braiding for the category of
613: $\Z_2^3$-graded vector spaces, making it into a symmetric monoidal
614: category \cite{AM}. In this symmetric monoidal category, the octonions
615: are a commutative monoid object. In less technical terms: this category
616: provides a context in which the octonions {\it are} commutative and
617: associative! So far this idea has just begun to be exploited.
618:
619: \subsection{The Cayley--Dickson construction} \label{cayley-dickson}
620:
621: It would be nice to have a construction of the normed division algebras
622: $\R,\C,\H,\O$ that explained why each one fits neatly inside the next.
623: It would be nice if this construction made it clear why $\H$ is
624: noncommutative and $\O$ is nonassociative. It would be even better if
625: this construction gave an infinite sequence of algebras, doubling in
626: dimension each time, with the normed division algebras as the first
627: four. In fact, there is such a construction: it's called the
628: Cayley--Dickson construction.
629:
630: As Hamilton noted, the complex number $a+bi$ can be thought of as a pair
631: $(a,b)$ of real numbers. Addition is done component-wise, and
632: multiplication goes like this:
633: \[ (a,b)(c,d) = (ac - db,ad + cb) \]
634: We can also define the conjugate of a complex number by
635: \[ (a,b)^* = (a,-b). \]
636:
637: Now that we have the complex numbers, we can define the
638: quaternions in a similar way. A quaternion can be thought of
639: as a pair of complex numbers. Addition is done component-wise,
640: and multiplication goes like this:
641: \be (a,b)(c,d) = (ac - db^*, a^* d + cb) \label{cd1} \ee
642: This is just like our formula for multiplication of complex numbers, but
643: with a couple of conjugates thrown in. If we included them in
644: the previous formula nothing would change, since the conjugate of a
645: real number is just itself. We can also define the conjugate of a
646: quaternion by
647: \be (a,b)^* = (a^*,-b). \label{cd2} \ee
648:
649: The game continues! Now we can define an octonion to be a pair of
650: quaternions. As before, we add and multiply them using formulas
651: (\ref{cd1}) and (\ref{cd2}). This trick for getting new algebras from
652: old is called the {\bf Cayley--Dickson construction}.
653:
654: Why do the real numbers, complex numbers, quaternions
655: and octonions have multiplicative inverses? I take it as
656: obvious for the real numbers. For the complex numbers,
657: one can check that
658: \[ (a,b) (a,b)^* = (a,b)^* (a,b) = k (1,0) \]
659: where $k$ is a real number, the square of the norm of $(a,b)$.
660: This means that whenever $(a,b)$ is nonzero, its multiplicative
661: inverse is $(a,b)^*/k$. One can check that the same holds for the
662: quaternions and octonions.
663:
664: But this, of course, raises the question: why isn't there an {\it
665: infinite} sequence of division algebras, each one obtained from the
666: preceding one by the Cayley--Dickson construction? The answer is that
667: each time we apply the construction, our algebra gets a bit worse.
668: First we lose the fact that every element is its own conjugate, then we
669: lose commutativity, then we lose associativity, and finally we lose the
670: division algebra property.
671:
672: To see this clearly, it helps to be a bit more formal. Define a {\bf
673: $\ast$-algebra} to be an algebra $A$ equipped with a {\bf conjugation},
674: that is, a real-linear map $\ast \maps A \to A$ with
675: \[ a^{**} = a, \quad \quad (ab)^* = b^* a^* \]
676: for all $a,b \in A$. We say a $\ast$-algebra is {\bf real} if $a =
677: a^*$ for every element $a$ of the algebra. We say the $\ast$-algebra
678: $A$ is {\bf nicely normed} if $a + a^* \in \R$ and $aa^* = a^* a > 0$
679: for all nonzero $a \in A$. If $A$ is nicely normed we set
680: \[ \Re(a) = (a + a^\ast)/2 \in \R, \qquad
681: \Im(a) = (a - a^\ast)/2 , \]
682: and define a norm on $A$ by
683: \[ \|a\|^2 = aa^\ast . \]
684: If $A$ is nicely normed, it has multiplicative inverses given by
685: \[ a^{-1} = a^\ast / \|a\|^2 .\]
686: If $A$ is nicely normed and alternative, $A$ is a normed division
687: algebra. To see this, note that for any $a,b \in A$, all 4 elements
688: $a,b,a^\ast,b^\ast$ lie in the associative algebra generated by $\Im(a)$
689: and $\Im(b)$, so that
690: \[ \|ab\|^2 = (ab)(ab)^\ast = ab(b^\ast a^\ast) =
691: a(bb^\ast)a^\ast = \|a\|^2 \|b\|^2 . \]
692:
693: Starting from any $\ast$-algebra $A$, the Cayley--Dickson construction
694: gives a new $\ast$-algebra $A'$. Elements of $A'$ are pairs $(a,b) \in
695: A^2$, and multiplication and conjugation are defined using equations
696: (\ref{cd1}) and (\ref{cd2}). The following propositions show the
697: effect of repeatedly applying the Cayley--Dickson construction:
698:
699: \begin{prop} \et \label{CD1}
700: $A'$ is never real.
701: \end{prop}
702:
703: \begin{prop} \et \label{CD2}
704: $A$ is real (and thus commutative) $\iff$ $A'$ is commutative.
705: \end{prop}
706:
707: \begin{prop} \et \label{CD3}
708: $A$ is commutative and associative $\iff$ $A'$ is associative.
709: \end{prop}
710:
711: \begin{prop} \et \label{CD4}
712: $A$ is associative and nicely normed $\iff$
713: $A'$ is alternative and nicely normed.
714: \end{prop}
715:
716: \begin{prop} \et\label{CD5}
717: $A$ is nicely normed $\iff$ $A'$ is nicely normed.
718: \end{prop}
719:
720: \noindent All of these follow from straightforward calculations; to
721: prove them here would merely deprive the reader of the pleasure of doing
722: so. It follows from these propositions that:
723: \begin{center}
724: $\R$ is a real commutative associative nicely normed
725: $\ast$-algebra $\implies$
726:
727: $\C$ is a commutative associative nicely normed $\ast$-algebra $\implies$
728:
729: $\H$ is an associative nicely normed $\ast$-algebra $\implies$
730:
731: $\O$ is an alternative nicely normed $\ast$-algebra
732: \end{center}
733: and therefore that $\R,\C,\H,$ and $\O$ are normed division algebras.
734: It also follows that the octonions are neither real, nor commutative, nor
735: associative.
736:
737: If we keep applying the Cayley--Dickson process to the octonions we get a
738: sequence of $\ast$-algebras of dimension 16, 32, 64, and so on. The
739: first of these is called the {\bf sedenions}, presumably alluding to the
740: fact that it is 16-dimensional \cite{LPS}. It follows from the above
741: results that all the $\ast$-algebras in this sequence are nicely normed
742: but neither real, nor commutative, nor alternative. They all have
743: multiplicative inverses, since they are nicely normed. But they are not
744: division algebras, since an explicit calculation demonstrates that the
745: sedenions, and thus all the rest, have zero divisors. In fact
746: \cite{Cohen,Moreno}, the zero divisors of norm one in the sedenions
747: form a subspace that is homeomorphic to the exceptional Lie group $\G_2$.
748:
749: The Cayley--Dickson construction provides a nice way to obtain the
750: sequence $\R,\H,\C,\O$ and the basic properties of these algebras. But
751: what is the meaning of this construction? To answer this, it is better
752: to define $A'$ as the algebra formed by adjoining to $A$ an element $i$
753: satisfying $i^2 = -1$ together with the following relations:
754: \be a(ib) = i(a^* b) , \qquad
755: (ai)b = (ab^*)i, \qquad
756: (ia)(bi^{-1}) = (ab)^* \label{cd3} \ee
757: for all $a,b \in A$. We make $A'$ into a $\ast$-algebra using the
758: original conjugation on elements of $A$ and setting $i^* = -i$. It is
759: easy to check that every element of $A'$ can be uniquely written as $a +
760: ib$ for some $a,b \in A$, and that this description of the
761: Cayley--Dickson construction becomes equivalent to our previous one
762: if we set $(a,b) = a + ib$.
763:
764: What is the significance of the relations in (\ref{cd3})? Simply
765: this: {\sl they express conjugation in terms of conjugation!} This is a pun
766: on the double meaning of the word `conjugation'. What I really mean is
767: that they express the $\ast$ operation in $A$ as conjugation by $i$. In
768: particular, we have
769: \[ a^\ast = (ia)i^{-1} = i(ai^{-1}) \]
770: for all $a \in A$. Note that when $A'$ is associative, any one of the
771: relations in (\ref{cd3}) implies the other two. It is when $A'$ is
772: nonassociative that we really need all three relations.
773:
774: This interpretation of the Cayley--Dickson construction makes it easier
775: to see what happens as we repeatedly apply the construction starting with
776: $\R$. In $\R$ the $\ast$ operation does nothing, so when we do the
777: Cayley--Dickson construction, conjugation by $i$ must have no effect on
778: elements of $\R$. Since $\R$ is commutative, this means that $\C = \R'$
779: is commutative. But $\C$ is no longer real, since $i^* = -i$.
780:
781: Next let us apply the Cayley--Dickson construction to $\C$. Since $\C$
782: is commutative, the $\ast$ operation in $\C$ is an automorphism. Whenever
783: we have an associative algebra $A$ equipped with an automorphism $\alpha$,
784: we can always extend $A$ to a larger associative algebra by adjoining an
785: invertible element $x$ with
786: \[ \alpha(a) = xax^{-1} \]
787: for all $a \in A$. Since $\C$ is associative, this means that $\C' =
788: \H$ is associative. But since $\C$ is not real, $\H$ cannot be
789: commutative, since conjugation by the newly adjoined element $i$ must
790: have a nontrivial effect.
791:
792: Finally, let us apply the Cayley--Dickson construction to $\H$. Since
793: $\H$ is noncommutative, the $\ast$ operation in $\H$ is not an
794: automorphism; it is merely an antiautomorphism. This means we cannot
795: express it as conjugation by some element of a larger associative
796: algebra. Thus $\H' = \O$ must be nonassociative.
797:
798: \subsection{Clifford Algebras} \label{clifford}
799:
800: William Clifford invented his algebras in 1876 as an attempt to
801: generalize the quaternions to higher dimensions, and he published a
802: paper about them two years later \cite{Clifford}. Given a real inner
803: product space $V$, the {\bf Clifford algebra} $\Cliff(V)$ is the
804: associative algebra freely generated by $V$ modulo the relations
805: \[ v^2 = -\|v\|^2 \]
806: for all $v \in V$. Equivalently, it is the associative algebra
807: freely generated by $V$ modulo the relations
808: \[ vw + wv = -2\langle v,w\rangle \]
809: for all $v, w \in V$. If $V = \R^n$ with its usual inner product, we
810: call this Clifford algebra $\Cliff(n)$. Concretely, this is the
811: associative algebra freely generated by $n$ anticommuting square roots
812: of $-1$. From this we easily see that
813: \[ \Cliff(0) = \R, \qquad\qquad \Cliff(1) = \C, \qquad\qquad
814: \Cliff(2) = \H .\]
815: So far this sequence resembles the iterated Cayley-Dickson construction
816: --- but the octonions are {\it not} a Clifford algebra, since they are
817: nonassociative. Nonetheless, there is a profound relation between
818: Clifford algebras and normed division algebras. This relationship gives
819: a nice way to prove that $\R, \C, \H$ and $\O$ are the only normed
820: dvivision algebras. It is also crucial for understanding the
821: geometrical meaning of the octonions.
822:
823: To see this relation, first suppose $\K$ is a normed division algebra.
824: Left multiplication by any element $a \in \K$ gives an operator
825: \[ \begin{array}{ccccc}
826: L_a &\maps& \K & \to & \K \\
827: & & x &\mapsto& ax .
828: \end{array} \]
829: If $\|a\| = 1$, the operator $L_a$ is norm-preserving, so it maps the
830: unit sphere of $\K$ to itself. Since $\K$ is a division algebra, we can
831: find an operator of this form mapping any point on the unit sphere to
832: any other point. The only way the unit sphere in $\K$ can have this much
833: symmetry is if the norm on $\K$ comes from an inner product. Even better,
834: this inner product is unique, since we can use the polarization identity
835: \[ \langle x, y\rangle = {1\over 2}(\|x+y\|^2 - \|x\|^2 - \|y\|^2) \]
836: to recover it from the norm.
837:
838: Using this inner product, we say an element $a \in \K$ is {\bf imaginary}
839: if it is orthogonal to the element $1$, and we let $\Im(\K)$ be the space
840: of imaginary elements of $\K$. We can also think of $\Im(\K)$ as the
841: tangent space of the unit sphere in $\K$ at the point $1$. This has a
842: nice consequence: since $a \mapsto L_a$ maps the unit sphere in $\K$ to
843: the Lie group of orthogonal transformations of $\K$, it must send
844: $\Im(\K)$ to the Lie algebra of skew-adjoint transformations of $\K$.
845: In short, $L_a$ is skew-adjoint whenever $a$ is imaginary.
846:
847: The relation to Clifford algebras shows up when we compute the square of
848: $L_a$ for $a \in \Im(\K)$. We can do this most easily when $a$ has norm
849: $1$. Then $L_a$ is both orthogonal and skew-adjoint. For any
850: orthogonal transformation, we can find some orthonormal basis in which
851: its matrix is block diagonal, built from $2 \times 2$ blocks that look
852: like this:
853: \[ \left( \begin{array}{cc} \cos \theta & \sin \theta \\
854: -\sin \theta & \cos \theta
855: \end{array} \right) \]
856: and possibly a $1 \times 1$ block like this: $\left( 1 \right)$.
857: Such a transformation can only be skew-adjoint if it consists solely of
858: $2 \times 2$ blocks of this form:
859: \[ \pm \left( \begin{array}{cc} 0 & 1 \\
860: -1 & 0
861: \end{array} \right). \]
862: In this case, its square is $-1$. We thus have $L_a^2 = -1$ when
863: $a \in \Im(\K)$ has norm 1. It follows that
864: \[ L_a^2 = -\| a \|^2 \]
865: for all $a \in \Im(\K)$. We thus obtain a representation of the Clifford
866: algebra $\Cliff(\Im(\K))$ on $\K$. Any $n$-dimensional normed division
867: algebra thus gives an $n$-dimensional representation of $\Cliff(n-1)$.
868: As we shall see, this is very constraining.
869:
870: We have already described the Clifford algebras up to $\Cliff(2)$.
871: Further calculations \cite{Harvey,Porteous} give the following table,
872: where we use $A[n]$ to stand for $n\times n$ matrices with entries in
873: the algebra $A$:
874:
875: \medskip
876: {\vbox{
877: \begin{center}
878: {\small
879: \begin{tabular}{|c|c|c|c|c|c|c|c|c|} \hline
880: $n$ & $\Cliff(n)$ \\ \hline
881: $0$ & $\R$ \\ \hline
882: $1$ & $\C$ \\ \hline
883: $2$ & $\H$ \\ \hline
884: $3$ & $\H \oplus \H$ \\ \hline
885: $4$ & $\H[2]$ \\ \hline
886: $5$ & $\C[4]$ \\ \hline
887: $6$ & $\R[8]$ \\ \hline
888: $7$ & $\R[8] \oplus \R[8]$ \\ \hline
889: \end{tabular}}
890: \vskip 1em
891: \centerline{Table 2 --- Clifford Algebras}
892: \end{center}
893: }}
894: \medskip
895:
896: \noindent
897: Starting at dimension 8, something marvelous happens: the table continues
898: in the following fashion:
899: \[ \Cliff(n+8) \iso \Cliff(n) \tensor \R[16] .\]
900: In other words, $\Cliff(n+8)$ consists of $16 \times 16$ matrices
901: with entries in $\Cliff(n)$. This `period-8' behavior was discovered
902: by Cartan in 1908 \cite{Cartan2}, but we will take the liberty of
903: calling it {\bf Bott periodicity}, since it has a far-ranging set of
904: applications to topology, some of which were discovered by Bott.
905:
906: Since Clifford algebras are built from matrix algebras over $\R,\C$ and
907: $\H$, it is easy to determine their representations. Every
908: representation is a direct sum of irreducible ones. The only
909: irreducible representation of $\R[n]$ is its obvious one via matrix
910: multiplication on $\R^n$. Similarly, the only irreducible
911: representation of $\C[n]$ is the obvious one on $\C^n$, and the only
912: irreducible representation of $\H[n]$ is the obvious one on $\H^n$.
913:
914: Glancing at the above table, we see that unless $n$ equals $3$ or $7$
915: modulo $8$, $\Cliff(n)$ is a real, complex or quaternionic matrix
916: algebra, so it has a unique irreducible representation. For reasons to
917: be explained later, this irreducible representation is known as the
918: space of {\bf pinors} and denoted $P_n$. When $n$ is $3$ or $7$ modulo
919: $8$, the algebra $\Cliff(n)$ is a direct sum of two real or quaternionic
920: matrix algebras, so it has two irreducible representations, which we
921: call the {\bf positive pinors} $P_n^+$ and {\bf negative pinors}
922: $P_n^-$. We summarize these results in the following table:
923:
924: \medskip
925: {\vbox{
926: \begin{center}
927: {\small
928: \begin{tabular}{|c|c|l|} \hline
929: $n$ & $\Cliff(n)$ & irreducible representations \\ \hline
930: $0$ & $\R$ & $P_0 = \R$ \\ \hline
931: $1$ & $\C$ & $P_1 = \C$ \\ \hline
932: $2$ & $\H$ & $P_2 = \H$ \\ \hline
933: $3$ & $\H \oplus \H$ & $P^+_3 = \H,\, P^-_3 =\H$ \\ \hline
934: $4$ & $\H[2]$ & $P_4 = \H^2$ \\ \hline
935: $5$ & $\C[4]$ & $P_5 = \C^4$ \\ \hline
936: $6$ & $\R[8]$ & $P_6 =\R^8$ \\ \hline
937: $7$ & $\R[8] \oplus \R[8]$& $P_7^+ = \R^8,\, P_7^- =\R^8$ \\ \hline
938: \end{tabular}}
939: \vskip 1em
940: \centerline{Table 3 --- Pinor Representations}
941: \end{center}
942: }}
943: \medskip
944:
945: \noindent
946: Examining this table, we see that in the range of dimensions listed
947: there is an $n$-dimensional representation of $\Cliff(n-1)$ only for $n
948: = 1,2,4,$ and $8$. What about higher dimensions? By Bott periodicity,
949: the irreducible representations of $\Cliff(n+8)$ are obtained by
950: tensoring those of $\Cliff(n)$ by $\R^{16}$. This multiplies the
951: dimension by 16, so one can easily check that for $n > 8$, the
952: irreducible representations of $\Cliff(n-1)$ always have dimension
953: greater than $n$.
954:
955: It follows that normed division algebras are only possible in dimensions
956: $1,2,4,$ and $8$. Having constructed $\R,\C,\H$ and $\O$, we also know
957: that normed division algebras {\it exist} in these dimensions. The only
958: remaining question is whether they are {\it unique}. For this it helps
959: to investigate more deeply the relation between normed division algebras
960: and the Cayley-Dickson construction. In what follows, we outline an
961: approach based on ideas in the book by Springer and Veldkamp \cite{SV}.
962:
963: First, suppose $\K$ is a normed division algebra. Then there is a unique
964: linear operator $\ast \maps \K \to \K$ such that $1^\ast = 1$ and $a^\ast
965: = -a$ for $a \in \Im(\K)$. With some calculation one can prove this
966: makes $\K$ into a nicely normed $\ast$-algebra.
967:
968: Next, suppose that $\K_0$ is any subalgebra of the normed division algebra
969: $\K$. It is easy to check that $\K_0$ is a nicely normed $\ast$-algebra in
970: its own right. If $\K_0$ is not all of $\K$, we can find an element $i \in
971: \K$ that is orthogonal to every element of $\K_0$. Without loss of
972: generality we shall assume this element has norm 1. Since this element
973: $i$ is orthogonal to $1 \in \K_0$, it is imaginary. From the definition
974: of the $\ast$ operator it follows that $i^\ast = -i$, and from results
975: earlier in this section we have $i^2 = -1$. With further calculation
976: one can show that for all $a,a' \in \K_0$ we have
977: \[ a(ia') = i(a^* a') , \qquad
978: (ai)a' = (aa'^*)i, \qquad
979: (ia)(a'i^{-1}) = (aa')^* \]
980: A glance at equation (\ref{cd3}) reveals that these are exactly the
981: relations defining the Cayley-Dickson construction! With a little
982: thought, it follows that the subalgebra of $\K$ generated by $\K_0$ and $i$
983: is isomorphic as a $\ast$-algebra to $\K'_0$, the $\ast$-algebra obtained
984: from $\K_0$ by the Cayley-Dickson construction.
985:
986: Thus, whenever we have a normed division algebra $\K$ we can find a
987: chain of subalgebras $\R = \K_0 \subset \K_1 \subset \cdots \subset
988: \K_n = \K$ such that $\K_{i+1} \iso \K_i'$. To construct $\K_{i+1}$, we
989: simply need to choose a norm-one element of $\K$ that is orthogonal to
990: every element of $\K_i$. It follows that the only normed division
991: algebras of dimension 1, 2, 4 and 8 are $\R,\C,\H$ and $\O$. This also
992: gives an alternate proof that there are no normed division algebras of
993: other dimensions: if there were any, there would have to be a
994: 16-dimensional one, namely $\O'$ --- the sedenions. But as mentioned
995: in Section \ref{cayley-dickson}, one can check explicitly that the
996: sedenions are not a division algebra.
997:
998: \subsection{Spinors and Trialities} \label{triality}
999:
1000: A nonassociative division algebra may seem like a strange thing to
1001: bother with, but the notion of triality makes it seem a bit more
1002: natural. The concept of duality is important throughout linear algebra.
1003: The concept of triality is similar, but considerably subtler. Given
1004: vector spaces $V_1$ and $V_2$, we may define a {\bf duality} to be a
1005: bilinear map
1006: \[ f \maps V_1 \times V_2 \to \R \]
1007: that is nondegenerate, meaning that if we fix either argument
1008: to any nonzero value, the linear functional induced on the other vector
1009: space is nonzero. Similarly, given vector spaces $V_1,V_2,$ and $V_3$,
1010: a {\bf triality} is a trilinear map
1011: \[ t \maps V_1 \times V_2 \times V_3 \to \R \]
1012: that is nondegenerate in the sense that if we fix any two arguments to
1013: any nonzero values, the linear functional induced on the third vector
1014: space is nonzero.
1015:
1016: Dualities are easy to come by. Trialities are much rarer. For suppose
1017: we have a triality
1018: \[ t \maps V_1 \times V_2 \times V_3 \to \R . \]
1019: By dualizing, we can turn this into a bilinear map
1020: \[ m \maps V_1 \times V_2 \to V_3^\ast \]
1021: which we call `multiplication'. By the nondegeneracy of our triality,
1022: left multiplication by any nonzero element of $V_1$ defines an
1023: isomorphism from $V_2$ to $V_3^\ast$. Similarly, right multiplication
1024: by any nonzero element of $V_2$ defines an isomorphism from $V_1$ to
1025: $V_3^\ast$. If we choose nonzero elements $e_1 \in V_1$ and $e_2 \in
1026: V_2$, we can thereby identify the spaces $V_1$, $V_2$ and $V_3^\ast$
1027: with a single vector space, say $V$. Note that this identifies
1028: all three vectors $e_1 \in V_1$, $e_2 \in V_2$, and $e_1e_2 \in V_3^\ast$
1029: with the same vector $e \in V$. We thus obtain a product
1030: \[ m \maps V \times V \to V \]
1031: for which $e$ is the left and right unit. Since left or right
1032: multiplication by any nonzero element is an isomorphism, $V$ is
1033: actually a division algebra! Conversely, any division algebra
1034: gives a triality.
1035:
1036: It follows from Theorem \ref{bott-milnor} that trialities only occur in
1037: dimensions 1, 2, 4, or 8. This theorem is quite deep. By comparison,
1038: Hurwitz's classification of {\it normed} division algebras is easy to
1039: prove. Not surprisingly, these correspond to a special sort of
1040: triality, which we call a `normed' triality.
1041:
1042: To be precise, a {\bf normed triality} consists of inner product
1043: spaces $V_1, V_2, V_3$ equipped with a trilinear map
1044: $t \maps V_1 \times V_2 \times V_3 \to \R$ with
1045: \[ |t(v_1, v_2, v_3)| \le \|v_1\| \, \|v_2\| \, \|v_3 \|, \]
1046: and such that for all $v_1, v_2$ there exists $v_3 \ne 0$ for which this
1047: bound is attained --- and similarly for cyclic permutations of $1,2,3$.
1048: Given a normed triality, picking unit vectors in any two of the spaces
1049: $V_i$ allows us to identify all three spaces and get a normed division
1050: algebra. Conversely, any normed division algebra gives a normed triality.
1051:
1052: But where do normed trialities come from? They come from the theory of
1053: spinors! From Section \ref{clifford}, we already know that any
1054: $n$-dimensional normed division algebra is a representation of
1055: $\Cliff(n-1)$, so it makes sense to look for normed trialities here.
1056: In fact, representations of $\Cliff(n-1)$ give certain representations
1057: of $\Spin(n)$, the double cover of the rotation group in $n$ dimensions.
1058: These are called `spinors'. As we shall see, the relation between
1059: spinors and vectors gives a nice way to construct normed trialities in
1060: dimensions 1, 2, 4 and 8.
1061:
1062: To see how this works, first let $\Pin(n)$ be the group sitting inside
1063: $\Cliff(n)$ that consists of all products of unit vectors in $\R^n$.
1064: This group is a double cover of the orthogonal group $\OO(n)$, where
1065: given any unit vector $v \in \R^n$, we map both $\pm v \in \Pin(n)$ to
1066: the element of $\OO(n)$ that reflects across the hyperplane perpendicular to
1067: $v$. Since every element of $\OO(n)$ is a product of reflections, this
1068: homomorphism is indeed onto.
1069:
1070: Next, let $\Spin(n) \subset \Pin(n)$ be the subgroup consisting of all
1071: elements that are a product of an even number of unit vectors in
1072: $\R^n$. An element of $\OO(n)$ has determinant 1 iff it is the product
1073: of an even number of reflections, so just as $\Pin(n)$ is a double cover
1074: of $\OO(n)$, $\Spin(n)$ is a double cover of $\SO(n)$. Together with a
1075: French dirty joke which we shall not explain, this analogy is the origin
1076: of the terms `$\Pin$' and `pinor'.
1077:
1078: Since $\Pin(n)$ sits inside $\Cliff(n)$, the irreducible representations
1079: of $\Cliff(n)$ restrict to representations of $\Pin(n)$, which turn out
1080: to be still irreducible. These are again called {\bf pinors}, and we
1081: know what they are from Table 3. Similarly, $\Spin(n)$ sits inside the
1082: subalgebra
1083: \[ \Cliff_0(n) \subseteq \Cliff(n) \]
1084: consisting of all linear combinations of products of an even number of
1085: vectors in $\R^n$. Thus the irreducible representations of
1086: $\Cliff_0(n)$ restrict to representations of $\Spin(n)$, which turn out
1087: to be still irreducible. These are called {\bf spinors} --- but we warn
1088: the reader that this term is also used for many slight variations on
1089: this concept.
1090:
1091: In fact, there is an isomorphism
1092: \[ \phi \maps \Cliff(n-1) \to \Cliff_0(n) \]
1093: given as follows:
1094: \[ \phi(e_i) = e_i e_n , \qquad \qquad 1 \le i \le n-1 ,\]
1095: where $\{e_i\}$ is an orthonormal basis for $\R^n$. Thus spinors in $n$
1096: dimensions are the same as pinors in $n-1$ dimensions! Table 3
1097: therefore yields the following table, where we use similar notation but
1098: with `$S$' instead of `$P$':
1099:
1100: \medskip
1101: {\vbox{
1102: \begin{center}
1103: {\small
1104: \begin{tabular}{|c|c|l|} \hline
1105: $n$ & $\Cliff_0(n)$ & irreducible representations \\ \hline
1106: $1$ & $\R$ & $S_1 = \R$ \\ \hline
1107: $2$ & $\C$ & $S_2 = \C$ \\ \hline
1108: $3$ & $\H$ & $S_3 = \H$ \\ \hline
1109: $4$ & $\H \oplus \H$ & $S_4^+ = \H, \, S_4^- = \H$ \\ \hline
1110: $5$ & $\H[2]$ & $S_5 = \H^2$ \\ \hline
1111: $6$ & $\C[4]$ & $S_6 = \C^4$ \\ \hline
1112: $7$ & $\R[8]$ & $S_7 = \R^8$ \\ \hline
1113: $8$ & $\R[8] \oplus \R[8]$ & $S_8^+ = \R^8,\, S_8^- = \R^8$ \\ \hline
1114: \end{tabular}}
1115: \vskip 1em
1116: \centerline{Table 4 --- Spinor Representations}
1117: \end{center}
1118: }}
1119: \medskip
1120:
1121: \noindent
1122: We call $S_n^+$ and $S_n^-$ the {\bf right-handed} and {\bf left-handed}
1123: spinor representations. For $n > 8$ we can work out the spinor
1124: representations using Bott periodicity:
1125: \[ S_{n+8} \iso S_n \tensor \R^{16} \]
1126: and similarly for right-handed and left-handed spinors.
1127:
1128:
1129: Now, besides its pinor representation(s), the group $\Pin(n)$ also has
1130: an irreducible representation where we first apply the 2--1 homomorphism
1131: $\Pin(n) \to \OO(n)$ and then use the obvious representation of $\OO(n)$
1132: on $\R^n$. We call this the {\bf vector} representation, $V_n$. As a
1133: vector space $V_n$ is just $\R^n$, and $\Cliff(n)$ is generated by
1134: $\R^n$, so we have an inclusion
1135: \[ V_n \hookrightarrow \Cliff(n) .\]
1136: Using this, we can restrict the action of the Clifford algebra on pinors
1137: to a map
1138: \[
1139: \begin{array}{lcc}
1140: m_n \maps& V_n \times P_n^\pm \to P_n^\pm &n \equiv 3,7 \bmod 8 \\
1141: m_n \maps& V_n \times P_n \to P_n & {\rm otherwise.}
1142: \end{array}
1143: \]
1144: This map is actually an intertwining operator between
1145: representations of $\Pin(n)$. If we restrict the vector representation
1146: to the subgroup $\Spin(n)$, it remains irreducible. This is not always
1147: true for the pinor representations, but we can always decompose them as
1148: a direct sum of spinor representations. Applying this decomposition to
1149: the map $m_n$, we get a map
1150: \[
1151: \begin{array}{lc}
1152: m_n \maps V_n \times S_n^\pm \to S_n^\mp &n \equiv 0,4 \bmod 8 \\
1153: m_n \maps V_n \times S_n \to S_n & {\rm otherwise.}
1154: \end{array}
1155: \]
1156: All the spinor representations appearing here are self-dual, so we can
1157: dualize the above maps and reinterpret them as trilinear maps
1158: \[
1159: \begin{array}{lc}
1160: t_n \maps V_n \times S_n^+ \times S_n^- \to \R & n \equiv 0,4 \bmod 8 \\
1161: t_n \maps V_n \times S_n \times S_n \to \R & {\rm otherwise.}
1162: \end{array}
1163: \]
1164:
1165: These trilinear maps are candidates for trialities! However, they can
1166: only be trialities when the dimension of the vector representation
1167: matches that of the relevant spinor representations. In the range of
1168: the above table this happens only for $n = 1,2,4,8$. In these cases we
1169: actually do get normed trialities, which in turn give normed division algebras:
1170: \[
1171: \begin{array}{ll}
1172: t_1 \maps V_1 \times S_1 \times S_1 \to \R &
1173: {\rm \; gives \;} \R. \\
1174: t_2 \maps V_2 \times S_2 \times S_2 \to \R &
1175: {\rm \; gives \;} \C. \\
1176: t_4 \maps V_4 \times S_4^+ \times S_4^- \to \R &
1177: {\rm \; gives \;} \H. \\
1178: t_8 \maps V_8 \times S_8^+ \times S_8^- \to \R &
1179: {\rm \; gives \;} \O .
1180: \end{array}
1181: \]
1182: In higher dimensions, the spinor representations become bigger than the
1183: vector representation, so we get no more trialities this way --- and of
1184: course, none exist.
1185:
1186: Of the four normed trialities, the one that gives the octonions
1187: has an interesting property that the rest lack. To see this property,
1188: one must pay careful attention to the difference between a normed triality
1189: and a normed division algebra. To construct a normed division $\K$
1190: algebra from the normed triality $t \maps V_1 \times V_2 \times V_3 \to \R$,
1191: we must arbitrarily choose unit vectors in two of the three spaces, so
1192: the symmetry group of $\K$ is smaller than that of $t$. More precisely,
1193: let us define a {\bf automorphism} of the normed triality $t \maps V_1 \times
1194: V_2 \times V_3 \to \R$ to be a triple of norm--preserving maps
1195: $f_i \maps V_i \to V_i$ such that
1196: \[ t(f_1(v_1), f_2(v_2), f_3(v_3)) = t(v_1,v_2,v_3) \]
1197: for all $v_i \in V_i$. These automorphisms form a group we call
1198: $\Aut(t)$. If we construct a normed division algebra $\K$ from $t$
1199: by choosing unit vectors $e_1 \in V_1, e_2 \in V_2$, we have
1200: \[
1201: \Aut(\K) \iso \{(f_1,f_2,f_3) \in \Aut(t)\; \colon \; f_1(e_1) = e_1,
1202: \; f_2(e_2) = e_2 \} .
1203: \]
1204:
1205: In particular, it turns out that:
1206: \be
1207: \begin{array}{lclclcl}
1208: 1 &\iso& \Aut(\R) & \subseteq& \Aut(t_1) & \iso &
1209: \{ (g_1,g_2,g_3) \in \OO(1)^3 \colon \; g_1g_2g_3 = 1 \} \\
1210: \Z_2 &\iso& \Aut(\C) & \subseteq& \Aut(t_2)& \iso &
1211: \{ (g_1,g_2,g_3) \in \U(1)^3 \colon \; g_1g_2g_3 = 1 \} \times \Z_2 \\
1212: \SO(3) &\iso& \Aut(\H) &\subseteq &\Aut(t_4)& \iso &
1213: \Sp(1)^3 / \{ \pm(1,1,1) \} \\
1214: \G_2 &\iso& \Aut(\O) &\subseteq& \Aut(t_8)& \iso& \Spin(8)
1215: \end{array}
1216: \label{Aut(t)}
1217: \ee
1218: where
1219: \[ \OO(1) \iso \Z_2, \qquad \U(1) \iso \SO(2) , \qquad \Sp(1) \iso \SU(2) \]
1220: are the unit spheres in $\R$, $\C$ and $\H$, respectively ---
1221: the only spheres that are Lie groups.
1222: $\G_2$ is just another name for the automorphism group of
1223: the octonions; we shall study this group in Section \ref{G2}.
1224: The bigger group $\Spin(8)$ acts as automorphisms of the triality
1225: that gives the octonions, and it does so in an interesting way.
1226: Given any element $g \in \Spin(8)$, there exist unique elements
1227: $g_\pm \in \Spin(8)$ such that
1228: \[ t(g(v_1), g_+(v_2), g_-(v_3)) = t(v_1,v_2,v_3) \]
1229: for all $v_1 \in V_8, v_2 \in S^+_8,$ and $v_3 \in S^-_8$.
1230: Moreover, the maps
1231: \[ \alpha_\pm \maps g \to g_\pm \]
1232: are outer automorphisms of $\Spin(8)$. In fact ${\rm Out(\Spin(8))}$
1233: is the permutation group on 3 letters, and there exist outer
1234: automorphisms that have the effect of permuting the vector, left-handed
1235: spinor, and right-handed spinor representations any way one likes;
1236: $\alpha_+$ and $\alpha_-$ are among these.
1237:
1238: In general, outer automorphisms of simple Lie groups come from
1239: symmetries of their Dynkin diagrams. Of all the simple Lie groups,
1240: $\Spin(8)$ has the most symmetrical Dynkin diagram! It looks like this:
1241:
1242: \medskip
1243: \centerline{\epsfysize=1.0in\epsfbox{triality.eps}}
1244: \label{triality.figure}
1245: \medskip
1246:
1247: \noindent
1248: Here the three outer nodes correspond to the vector, left-handed spinor
1249: and right-handed spinor representations of $\Spin(8)$, while the central
1250: node corresponds to the adjoint representation --- that is, the
1251: representation of $\Spin(8)$ on its own Lie algebra, better known as
1252: $\so(8)$. The outer automorphisms corresponding to the symmetries of
1253: this diagram were discovered in 1925 by Cartan \cite{Cartan3}, who
1254: called these symmetries {\bf triality}. The more general notion of
1255: `triality' we have been discussing here came later, and is apparently
1256: due to Adams \cite{Adams}.
1257:
1258: The construction of division algebras from trialities has tantalizing
1259: links to physics. In the Standard Model of particle physics, all
1260: particles other than the Higgs boson transform either as vectors or
1261: spinors. The vector particles are also called `gauge bosons', and they
1262: serve to carry the {\it forces} in the Standard Model. The spinor
1263: particles are also called `fermions', and they correspond to the basic
1264: forms of {\it matter}: quarks and leptons. The interaction between
1265: matter and the forces is described by a trilinear map involving two
1266: spinors and one vector. This map is often drawn as a Feynman diagram:
1267:
1268: \medskip
1269: \centerline{\epsfysize=1.0in\epsfbox{feynman.eps}}
1270: \label{feynman}
1271: \medskip
1272: \noindent
1273: where the straight lines denote spinors and the wiggly one denotes a
1274: vector. The most familiar example is the process whereby an electron
1275: emits or absorbs a photon.
1276:
1277: It is fascinating that the same sort of mathematics can be used both to
1278: construct the normed division algebras and to describe the interaction
1279: between matter and forces. Could this be important for physics? One
1280: {\it prima facie} problem with this speculation is that physics uses
1281: spinors associated to Lorentz groups rather than rotation groups, due to
1282: the fact that spacetime has a Lorentzian rather than Euclidean metric.
1283: However, in Section \ref{lorentz} we describe a way around this problem.
1284: Just as octonions give the spinor representations of $\Spin(8)$, pairs
1285: of octonions give the spinor representations of $\Spin(9,1)$. This is
1286: one reason so many theories of physics work best when spacetime is
1287: 10-dimensional! Examples include superstring theory \cite{Deligne,GSW},
1288: supersymmetric gauge theories \cite{Evans,KT,Schray}, and Geoffrey
1289: Dixon's extension of the Standard Model based on the algebra $\C \tensor
1290: \H \tensor \O$, in which the 3 forces arise naturally from the three
1291: factors in this tensor product \cite{Dixon}.
1292:
1293: \section{Octonionic Projective Geometry} \label{proj}
1294:
1295: Projective geometry is a venerable subject that has its origins in the
1296: study of perspective by Renaissance painters. As seen by the eye,
1297: parallel lines --- e.g., train tracks --- appear to meet at a `point at
1298: infinity'. When one changes ones viewpoint, distances and angles appear
1299: to change, but points remain points and lines remain lines. These facts
1300: suggest a modification of Euclidean plane geometry, based on a set of
1301: points, a set of lines, and relation whereby a point `lies on' a line,
1302: satisfying the following axioms:
1303: \begin{itemize}
1304: \item For any two distinct points, there is a unique line on which they
1305: both lie.
1306: \item For any two distinct lines, there is a unique point which lies on
1307: both of them.
1308: \item There exist four points, no three of which lie on the same line.
1309: \item There exist four lines, no three of which have the same point lying
1310: on them.
1311: \end{itemize}
1312: A structure satisfying these axioms is called a {\bf projective plane}.
1313: Part of the charm of this definition is that it is `self-dual': if we
1314: switch the words `point' and `line' and switch who lies on whom, it
1315: stays the same.
1316:
1317: We have already met one example of a projective plane in Section
1318: \ref{fano}: the smallest one of all, the Fano plane. The example
1319: relevant to perspective is the real projective plane, $\RP^2$. Here the
1320: points are lines through the origin in $\R^3$, the lines are planes
1321: through the origin in $\R^3$, and the relation of `lying on' is taken to
1322: be inclusion. Each point $(x,y) \in \R^2$ determines a point in
1323: $\RP^2$, namely the line in $\R^3$ containing the origin and the point
1324: $(x,y,-1)$:
1325:
1326: \centerline{\epsfysize=2in\epsfbox{plane.eps}}
1327: \label{plane}
1328:
1329: \noindent
1330: There are also other points in $\RP^2$, the `points at infinity',
1331: corresponding to lines through the origin in $\R^3$ that do not
1332: intersect the plane $\{z = -1\}$. For example, any point on the
1333: horizon in the above picture determines a point at infinity.
1334:
1335: Projective geometry is also interesting in higher dimensions.
1336: One can define a {\bf projective space} by the following axioms:
1337: \begin{itemize}
1338: \item For any two distinct points $p,q$, there is a unique line
1339: $pq$ on which they both lie.
1340: \item For any line, there are at least three points lying on this line.
1341: \item If $a,b,c,d$ are distinct points and there is a point lying
1342: on both $ab$ and $cd$, then there is a point lying on both $ac$ and $bd$.
1343: \end{itemize}
1344: Given a projective space and a set $S$ of points in this space, we
1345: define the {\bf span} of $S$ to be the smallest set $T$ of points
1346: containing $S$ such that if $a$ and $b$ lie in $T$, so do all points
1347: on the line $ab$. The {\bf dimension}
1348: of a projective space is defined to be one less than the minimal
1349: cardinality of a set that spans the whole space. The reader may enjoy
1350: showing that a 2-dimensional projective space is the same thing as a
1351: projective plane \cite{Garner}.
1352:
1353: If $\K$ is any field, there is an $n$-dimensional projective space
1354: called $\KP^n$ where the points are lines through the origin in
1355: $\K^{n+1}$, the lines are planes through the origin in $\K^{n+1}$, and
1356: the relation of `lying on' is inclusion. In fact, this construction
1357: works even when $\K$ is a mere {\bf skew field}: a ring such that every
1358: nonzero element has a left and right multiplicative inverse. We just
1359: need to be a bit careful about defining lines and planes through the
1360: origin in $\K^{n+1}$. To do this, we use the fact that $\K^{n+1}$ is a
1361: $\K$-bimodule in an obvious way. We take a line through the origin to
1362: be any set
1363: \[ L = \{ \alpha x \; \colon\; \alpha \in \K \} \]
1364: where $x \in \K^{n+1}$ is nonzero, and take a plane through the
1365: origin to be any set
1366: \[ P = \{ \alpha x + \beta y \; \colon \; \alpha,\beta \in \K \} \]
1367: where $x,y \in \K^{n+1}$ are elements such that
1368: $\alpha x + \beta y = 0$ implies $\alpha,\beta = 0$.
1369:
1370: Given this example, the question naturally arises whether {\it every}
1371: projective $n$-space is of the form $\KP^n$ for some skew field $\K$.
1372: The answer is quite surprising: yes, but only if $n > 2$. Projective
1373: planes are more subtle \cite{Stevenson}. A projective plane comes
1374: from a skew field if and only if it satisfies an extra axiom, the
1375: `axiom of Desargues', which goes as follows. Define a {\bf triangle}
1376: to be a triple of points that don't all lie on the same line. Now,
1377: suppose we have two triangles $xyz$ and $x'y'z'$. The sides of each
1378: triangle determine three lines, say $LMN$ and $L'M'N'$. Sometimes
1379: the line through $x$ and $x'$, the line through $y$ and $y'$, and
1380: the line through $z$ and $z'$ will all intersect at the same point:
1381:
1382: \vskip 1em
1383: {\hskip 15em}{\epsfysize=2in\epsfbox{desargues1.eps}}
1384: \label{desargues1}
1385: \vskip -3em
1386:
1387: \noindent
1388: The {\bf axiom of Desargues} says that whenever this happens, something
1389: else happens: the intersection of $L$ and $L'$, the intersection of $M$
1390: and $M'$, and the intersection of $N$ and $N'$ all lie on the same line:
1391:
1392: {\hskip 11em}{\epsfysize=2in\epsfbox{desargues2.eps}}
1393: \label{desargues2}
1394:
1395:
1396: \noindent
1397: This axiom holds automatically for projective spaces of dimension 3
1398: or more, but not for projective planes. A projective plane satisfying
1399: this axiom is called {\bf Desarguesian}.
1400:
1401: The axiom of Desargues is pretty, but what is its connection to skew
1402: fields? Suppose we start with a projective plane $P$ and try to
1403: reconstruct a skew field from it. We can choose any line $L$, choose
1404: three distinct points on this line, call them $0, 1$, and $\infty$, and
1405: set $\K = L - \{\infty\}$. Copying geometric constructions that work
1406: when $P = \RP^2$, we can define addition and multiplication of points in
1407: $\K$. In general the resulting structure $(\K,+,0,\cdot,1)$ will not
1408: be a skew field. Even worse, it will depend in a nontrivial way on the
1409: choices made. However, if we assume the axiom of Desargues, these
1410: problems go away. We thus obtain a one-to-one correspondence between
1411: isomorphism classes of skew fields and isomorphism classes of
1412: Desarguesian projective planes.
1413:
1414: Projective geometry was very fashionable in the 1800s, with such
1415: worthies as Poncelet, Brianchon, Steiner and von Staudt making important
1416: contributions. Later it was overshadowed by other forms of geometry.
1417: However, work on the subject continued, and in 1933 Ruth Moufang
1418: constructed a remarkable example of a non-Desarguesian projective plane
1419: using the octonions \cite{Moufang}. As we shall see, this projective
1420: plane deserves the name $\OP^2$.
1421:
1422: The 1930s also saw the rise of another reason for interest in projective
1423: geometry: quantum mechanics! Quantum theory is distressingly different
1424: from the classical Newtonian physics we have learnt to love. In
1425: classical mechanics, observables are described by real-valued functions.
1426: In quantum mechanics, they are often described by hermitian $n \times
1427: n$ complex matrices. In both cases, observables are closed under
1428: addition and multiplication by real scalars. However, in quantum
1429: mechanics, observables do not form an associative algebra. Still,
1430: one can raise an observable to a power, and from squaring one
1431: can construct a commutative but nonassociative product:
1432: \[ a \circ b = {1\over 2}((a+b)^2 - a^2 - b^2)
1433: = {1\over 2}(ab + ba) . \]
1434: In 1932, Pascual Jordan attempted to understand this situation better by
1435: isolating the bare minimum axioms that an `algebra of observables'
1436: should satisfy \cite{Jordan}. He invented the definition of what is now
1437: called a {\bf formally real Jordan algebra}: a commutative and
1438: power-associative algebra satisfying
1439: \[ a_1^2 + \cdots + a_n^2 = 0 \quad \implies \quad a_1 = \cdots = a_n = 0 \]
1440: for all $n$. The last condition gives the algebra a partial
1441: ordering: if we write $a \le b$ when the element $b - a$ is a sum of
1442: squares, it says that $a \le b$ and $b \le a$ imply $a = b$. Though it
1443: is not obvious, any formally real Jordan algebra satisfies the identity
1444: \[ a \circ (b \circ a^2) = (a \circ b) \circ a^2 \]
1445: for all elements $a$ and $b$. Any commutative algebra satisfying
1446: this identity is called a {\bf Jordan algebra}. Jordan algebras are
1447: automatically power-associative.
1448:
1449: In 1934, Jordan published a paper with von Neumann and Wigner
1450: classifying all formally real Jordan algebras \cite{JNW}. The
1451: classification is nice and succinct. An {\bf ideal} in the Jordan algebra
1452: $A$ is a subspace $B \subseteq A$ such that $b \in B$ implies $a \circ b
1453: \in B$ for all $a \in A$. A Jordan algebra $A$ is {\bf simple} if its
1454: only ideals are $\{0\}$ and $A$ itself. Every formally real Jordan
1455: algebra is a direct sum of simple ones. The simple formally real Jordan
1456: algebras consist of 4 infinite families and one exception.
1457: \begin{enumerate}
1458: \item The algebra $\h_n(\R)$
1459: with the product $a \circ b = {1\over 2}(ab + ba)$.
1460: \item The algebra $\h_n(\C)$
1461: with the product $a \circ b = {1\over 2}(ab + ba)$.
1462: \item The algebra $\h_n(\H)$
1463: with the product $a \circ b = {1\over 2}(ab + ba)$.
1464: \item The algebra $\R^n \oplus \R$ with the product
1465: \[ (v,\alpha) \circ (w, \beta) =
1466: (\alpha w + \beta v, \langle v,w\rangle + \alpha \beta). \]
1467: \item The algebra $\h_3(\O)$
1468: with the product $a \circ b = {1\over 2}(ab + ba)$.
1469: \end{enumerate}
1470: Here we say a square matrix with entries in the $\ast$-algebra $A$ is
1471: {\bf hermitian} if it equals its conjugate transpose, and we let
1472: $\h_n(A)$ stand for the hermitian $n \times n$ matrices with entries in
1473: $A$. Jordan algebras in the fourth family are called {\bf spin
1474: factors}, while $\h_3(\O)$ is called the {\bf exceptional Jordan
1475: algebra}. This classification raises some obvious questions. Why does
1476: nature prefer the Jordan algebras $\h_n(\C)$ over all the rest? Or does
1477: it? Could the other Jordan algebras --- even the exceptional one ---
1478: have some role to play in quantum physics? Despite much research, these
1479: questions remain unanswered to this day.
1480:
1481: The paper by Jordan, von Neumann and Wigner appears to have been
1482: uninfluenced by Moufang's discovery of $\OP^2$, but in fact they are
1483: related. A {\bf projection} in a formally real Jordan algebra is
1484: defined to be an element $p$ with $p^2 = p$. In the familiar case of
1485: $\h_n(\C)$, these correspond to hermitian matrices with eigenvalues $0$
1486: and $1$, so they are used to describe observables that assume only two
1487: values --- e.g., `true' and `false'. This suggests treating projections
1488: in a formally real Jordan algebra as propositions in a kind of `quantum
1489: logic'. The partial order helps us do this: given projections $p$ and
1490: $q$, we say that $p$ `implies' $q$ if $p \le q$.
1491:
1492: The relation between Jordan algebras and quantum logic is already
1493: interesting \cite{Emch}, but the real fun starts when we note
1494: that projections in $\h_n(\C)$ correspond to subspaces of $\C^n$. This
1495: sets up a relationship to projective geometry \cite{Varadarajan}, since
1496: the projections onto 1-dimensional subspaces correspond to points in
1497: $\CP^n$, while the projections onto 2-dimensional subspaces correspond
1498: to lines. Even better, we can work out the dimension of a subspace $V
1499: \subseteq \C^n$ from the corresponding projection $p \maps \C^n \to V$
1500: using only the partial order on projections: $V$ has dimension $d$ iff
1501: the longest chain of distinct projections
1502: \[ 0 = p_0 < \cdots < p_i = p \]
1503: has length $i = d$. In fact, we can use this to define the {\bf rank}
1504: of a projection in any formally real Jordan algebra. We can then try to
1505: construct a projective space whose points are the rank-1 projections and
1506: whose lines are the rank-2 projections, with the relation of `lying on'
1507: given by the partial order $\le$.
1508:
1509: If we try this starting with $\h_n(\R)$, $\h_n(\C)$ or $\h_n(\H)$, we
1510: succeed when $n \ge 2$, and we obtain the projective spaces $\RP^n$,
1511: $\CP^n$ and $\HP^n$, respectively. If we try this starting with the
1512: spin factor $\R^n \oplus \R$ we succeed when $n \ge 2$, and obtain a
1513: series of 1-dimensional projective spaces related to Lorentzian
1514: geometry. Finally, in 1949 Jordan \cite{Jordan2} discovered that if we
1515: try this construction starting with the exceptional Jordan algebra, we
1516: get the projective plane discovered by Moufang: $\OP^2$.
1517:
1518: In what follows we describe the octonionic projective plane
1519: and exceptional Jordan algebra in more detail. But first let us
1520: consider the octonionic projective line, and the Jordan algebra
1521: $\h_2(\O)$.
1522:
1523: \subsection{Projective Lines} \label{OP1}
1524:
1525: A one-dimensional projective space is called a {\bf projective line}.
1526: Projective lines are not very interesting from the viewpoint of
1527: axiomatic projective geometry, since they have only one line on which
1528: all the points lie. Nonetheless, they can be geometrically and
1529: topologically interesting. This is especially true of the octonionic
1530: projective line. As we shall see, this space has a deep connection to
1531: Bott periodicity, and also to the Lorentzian geometry of 10-dimensional
1532: spacetime.
1533:
1534: Suppose $\K$ is a normed division algebra. We have already defined
1535: $\KP^1$ when $\K$ is associative, but this definition does not work well
1536: for the octonions: it is wiser to take a detour through Jordan
1537: algebras. Let $\h_2(\K)$ be the space of $2 \times 2$ hermitian
1538: matrices with entries in $\K$. It is easy to check that this becomes a
1539: Jordan algebra with the product $a \circ b = {1\over 2}(ab + ba)$. We
1540: can try to build a projective space from this Jordan algebra using the
1541: construction in the previous section. To see if this
1542: succeeds, we need to ponder the projections in $\h_2(\K)$. A little
1543: calculation shows that besides the trivial projections 0 and 1, they
1544: are all of the form
1545: \[
1546: \left( \begin{array}{c} x^* \\ y^* \end{array} \right)
1547: \left( \begin{array}{cc} \! x & y \! \end{array} \right)
1548: =
1549: \left( \begin{array}{cc}
1550: x^* x & x^* y \\
1551: y^* x & y^* y
1552: \end{array} \right)
1553: \]
1554: where $(x,y) \in \K^2$ has
1555: \[ \|x\|^2 + \|y\|^2 = 1. \]
1556: These nontrivial projections all have rank 1, so they are the points of
1557: our would--be projective space. Our would--be projective space has just
1558: one line, corresponding to the projection 1, and all the points lie on
1559: this line. It is easy to check that the axioms for a projective space
1560: hold. Since this projective space is 1-dimensional, we have succeeded
1561: in creating the {\bf projective line over} $\K$. We call the set of
1562: points of this projective line $\KP^1$.
1563:
1564: Given any nonzero element $(x,y) \in \K^2$, we can normalize it and then
1565: use the above formula to get a point in $\KP^1$, which we call
1566: $[(x,y)]$. This allows us to describe $\KP^1$ in terms
1567: of lines through the origin, as follows. Define an equivalence relation
1568: on nonzero elements of $\K^2$ by
1569: \[ (x,y) \sim (x',y') \; \iff \; [(x,y)] = [(x',y')] .\]
1570: We call an equivalence class for this relation a {\bf line through the
1571: origin} in $\K^2$. We can then identify points in $\KP^1$ with lines
1572: through the origin in $\K^2$.
1573:
1574: Be careful: when $\K$ is the octonions, the line through the
1575: origin containing $(x,y)$ is not always equal to
1576: \[ \{(\alpha x, \alpha y)\; \colon \; \alpha \in \K\}. \]
1577: This is only true when $\K$ is associative, or when $x$ or $y$ is
1578: $1$. Luckily, we have $(x,y) \sim (y^{-1}x,1)$ when $y \ne 0$ and
1579: $(x,y) \sim (1,x^{-1}y)$ when $x \ne 0$. Thus in either case we get a
1580: concrete description of the line through the origin containing $(x,y)$:
1581: when $x \ne 0$ it equals
1582: \[ \{(\alpha(y^{-1}x), \alpha)\; \colon \; \alpha \in \K\} , \]
1583: and when $y \ne 0$ it equals
1584: \[ \{(\alpha, \alpha(x^{-1}y)\; \colon \; \alpha \in \K\} . \]
1585: In particular, the line through the origin containing $(x,y)$ is
1586: always a real vector space isomorphic to $\K$.
1587:
1588: We can make $\KP^1$ into a manifold as follows. By the above
1589: observations, we can cover it with two coordinate charts: one containing
1590: all points of the form $[(x,1)]$, the other containing all points of the
1591: form $[(1,y)]$. It is easy to check that $[(x,1)] = [(1,y)]$ iff $y =
1592: x^{-1}$, so the transition function from the first chart to the second
1593: is the map $x \mapsto x^{-1}$. Since this transition function and its
1594: inverse are smooth on the intersection of the two charts, $\KP^1$
1595: becomes a smooth manifold.
1596:
1597: When pondering the geometry of projective lines it is handy to
1598: visualize the complex case, since $\CP^1$ is just the familiar
1599: `Riemann sphere'. In this case, the map
1600: \[ x \mapsto [(x,1)] \]
1601: is given by stereographic projection:
1602:
1603: \begin{figure}[h]
1604: \centerline{\epsfysize=1.5in\epsfbox{stereo.eps}}
1605: \label{stereo}
1606: \end{figure}
1607:
1608: \noindent
1609: where we choose the sphere to have diameter 1. This map from $\C$ to
1610: $\CP^1$ is one-to-one and almost onto, missing only the point at
1611: infinity, or `north pole'. Similarly, the map
1612: \[ y \mapsto [(1,y)] \]
1613: misses only the south pole. Composing the first map with the inverse of
1614: the second, we get the map $x \mapsto x^{-1}$, which goes by the name
1615: of `conformal inversion'. The southern hemisphere of the Riemann
1616: sphere consists of all points $[(x,1)]$ with $\|x\| \le 1$, while the
1617: northern hemisphere consists of all $[(1,y)]$ with $\|y\| \le 1$. Unit
1618: complex numbers $x$ give points $[(x,1)] = [(1,x^{-1})]$ on the equator.
1619:
1620: All these ideas painlessly generalize to $\KP^1$ for any normed division
1621: algebra $\K$. First of all, as a smooth manifold $\KP^1$ is just a
1622: sphere with dimension equal to that of $\K$:
1623: \[
1624: \begin{array}{ccl}
1625: \RP^1 &\iso& S^1 \\
1626: \CP^1 &\iso& S^2 \\
1627: \HP^1 &\iso& S^4 \\
1628: \OP^1 &\iso& S^8.
1629: \end{array}
1630: \]
1631: We can think of it as the one-point compactification of $\K$. The
1632: `southern hemisphere', `northern hemisphere', and `equator' of $\K$ have
1633: descriptions exactly like those given above for the complex case. Also,
1634: as in the complex case, the maps $x \mapsto [(x,1)]$ and $y \mapsto [(1,y)]$
1635: are angle-preserving with respect to the usual Euclidean metric on $\K$
1636: and the round metric on the sphere.
1637:
1638: One of the nice things about $\KP^1$ is that it comes equipped with a
1639: vector bundle whose fiber over the point $[(x,y)]$ is the line
1640: through the origin corresponding to this point. This bundle is called
1641: the {\bf canonical line bundle}, $L_\K$. Of course, when we are working
1642: with a particular division algebra, `line' means a copy of this division
1643: algebra, so if we think of them as real vector bundles, $L_\R, L_\C,
1644: L_\H$ and $L_\O$ have dimensions 1,2,4, and 8, respectively.
1645:
1646: These bundles play an important role in topology, so it is good to
1647: understand them in a number of ways. In general, any $k$-dimensional
1648: real vector bundle over $S^n$ can be formed by taking trivial bundles
1649: over the northern and southern hemispheres and gluing them together
1650: along the equator via a map $f \maps S^{n-1} \to \OO(k)$. We must
1651: therefore be able to build the canonical line bundles $L_\R, L_\C,
1652: L_\H$ and $L_\O$ using maps
1653: \[
1654: \begin{array}{cccl}
1655: f_\R \maps &S^0& \to & \OO(1) \\
1656: f_\C \maps &S^1& \to & \OO(2) \\
1657: f_\H \maps &S^3& \to & \OO(4) \\
1658: f_\O \maps &S^7& \to & \OO(8).
1659: \end{array}
1660: \]
1661: What are these maps? We can describe them all simultaneously. Suppose
1662: $\K$ is a normed division algebra of dimension $n$. In the southern
1663: hemisphere of $\KP^1$, we can identify any fiber of $L_\K$ with $\K$ by
1664: mapping the point $(\alpha x, \alpha)$ in the line $[(x,1)]$ to the
1665: element $\alpha \in \K$. This trivializes the canonical line bundle
1666: over the southern hemisphere. Similarly, we can trivialize this bundle
1667: over the northern hemisphere by mapping the point $(\beta,\beta y)$ in
1668: the line $[(1,y)]$ to the element $\beta \in \K$. If $x \in \K$ has norm
1669: one, $[(x,1)] = [(1,x^{-1})]$ is a point on the equator, so we get two
1670: trivializations of the fiber over this point. These are related as
1671: follows: if $(\alpha x, \alpha) = (\beta, \beta x^{-1})$ then $\beta =
1672: \alpha x$. The map $\alpha \mapsto \beta$ is thus right multiplication
1673: by $x$. In short,
1674: \[ f_\K \maps S^{n-1} \to \OO(n) \]
1675: is just the map sending any norm-one element $x \in \K$ to the operation
1676: of right multiplication by $x$.
1677:
1678: The importance of the map $f_\K$ becomes clearest if we form the
1679: inductive limit of the groups $\OO(n)$ using the obvious inclusions
1680: $\OO(n) \hookrightarrow \OO(n+1)$, obtaining a topological group
1681: called $\OO(\infty)$. Since $\OO(n)$ is included in $\OO(\infty)$,
1682: we can think of $f_\K$ as a map from $S^{n-1}$ to $\OO(\infty)$.
1683: Its homotopy class $[f_\K]$ has the following marvelous property,
1684: mentioned in the Introduction:
1685:
1686: \vbox{
1687: \begin{itemize}
1688: \item $[f_\R]$ generates $\pi_0(\OO(\infty)) \iso \Z_2$.
1689: \item $[f_\C]$ generates $\pi_1(\OO(\infty)) \iso \Z$.
1690: \item $[f_\H]$ generates $\pi_3(\OO(\infty)) \iso \Z$.
1691: \item $[f_\O]$ generates $\pi_7(\OO(\infty)) \iso \Z$.
1692: \end{itemize}
1693: }
1694:
1695: Another nice perspective on the canonical line bundles $L_\K$ comes from
1696: looking at their unit sphere bundles. Any fiber of $L_\K$ is naturally
1697: an inner product space, since it is a line through the origin in $\K^2$.
1698: If we take the unit sphere in each fiber, we get a bundle of
1699: $(n-1)$-spheres over $\KP^1$ called the {\bf Hopf bundle}:
1700: \[ p_\K \maps E_\K \to \KP^1 \]
1701: The projection $p_\K$ is called the {\bf Hopf map}. The total space
1702: $E_\K$ consists of all the unit vectors in $\K^2$, so it is a sphere
1703: of dimension $2n-1$. In short, the Hopf bundles look like this:
1704: \[
1705: \begin{array}{crccl}
1706: \K = \R: \qquad &S^0 \hookrightarrow &S^1& \to &S^1 \\
1707: \K = \C: \qquad &S^1 \hookrightarrow &S^3& \to &S^2 \\
1708: \K = \H: \qquad &S^3 \hookrightarrow &S^7& \to &S^4 \\
1709: \K = \O: \qquad &S^7 \hookrightarrow &S^{15}& \to &S^8
1710: \end{array}
1711: \]
1712:
1713: We can understand the Hopf maps better by thinking about inverse images
1714: of points. The inverse image $p_\K^{-1}(x)$ of any point $x \in S^n$
1715: is a $(n-1)$-sphere in $S^{2n-1}$, and the inverse image of any pair of
1716: distinct points is a pair of linked spheres of this sort. When $\K =
1717: \C$ we get linked circles in $S^3$, which form the famous {\bf Hopf link}:
1718:
1719: \medskip
1720: \centerline{\epsfysize=1.0in\epsfbox{hopf.eps}}
1721: \label{hopf}
1722: \medskip
1723:
1724: \noindent
1725: When $\K = \O$, we get a pair of linked 7-spheres in $S^{15}$.
1726:
1727: To quantify this notion of linking, we can use the `Hopf invariant'.
1728: Suppose for a moment that $n$ is any natural number greater than one,
1729: and let $f \maps S^{2n-1} \to S^n$ be any smooth map. If $\omega$ is
1730: the normalized volume form on $S^n$, then $f^* \omega$ is a closed
1731: $n$-form on $S^{2n-1}$. Since the $n$th cohomology of $S^{2n-1}$
1732: vanishes, $f^\ast \omega = d \alpha$ for some $(n-1)$-form $\alpha$.
1733: We define the {\bf Hopf invariant} of $f$ to be the number
1734: \[ H(f) = \int_{S^{2n-1}} \alpha \wedge d\alpha .\]
1735: This is easily seen to be invariant under smooth homotopies of the map $f$.
1736:
1737: To see how the Hopf invariant is related to linking, we can compute it
1738: using homology rather than cohomology. If we take any two regular
1739: values of $f$, say $x$ and $y$, the inverse images of these points are
1740: compact oriented $(n-1)$-dimensional submanifolds of $S^{2n-1}$. We
1741: can always find an oriented $n$-dimensional submanifold $X \subset
1742: S^{2n-1}$ that has boundary equal to $f^{-1}(x)$ and that intersects
1743: $f^{-1}(y)$ transversely. The dimensions of $X$ and $f^{-1}(y)$ add up
1744: to $2n-1$, so their intersection number is well-defined. By the
1745: duality between homology and cohomology, this number equals the Hopf
1746: invariant $H(f)$. This shows that the Hopf invariant is an integer.
1747: Moreover, it shows that when the Hopf invariant is nonzero, the inverse
1748: images of $x$ and $y$ are linked.
1749:
1750: Using either of these approaches we can compute the Hopf invariant of
1751: $p_\C$, $p_\H$ and $p_\O$. They all turn out to have Hopf invariant 1.
1752: This implies, for example, that the inverse images of distinct points
1753: under $p_\O$ are nontrivially linked 7-spheres in $S^{15}$. It
1754: also implies that $p_\C$, $p_\H$ and $p_\O$ give nontrivial elements of
1755: $\pi_{2n-1}(S^n)$ for $n = 2, 4$, and $8$. In fact, these elements
1756: generate the torsion-free part of $\pi_{2n-1}(S^n)$.
1757:
1758: A deep study of the Hopf invariant is one way to prove that any division
1759: algebra must have dimension 1, 2, 4 or 8. One can show that if there
1760: exists an $n$-dimensional division algebra, then $S^{n-1}$ must be {\bf
1761: parallelizable}: it must admit $n - 1$ pointwise linearly independent
1762: smooth vector fields. One can also show that for $n > 1$, $S^{n-1}$ is
1763: parallelizable iff there exists a map $f \maps S^{2n-1} \to S^n$ with
1764: $H(f) = 1$ \cite{AH,BM,Kervaire}. The hard part is showing that a map from
1765: $S^{2n-1}$ to $S^n$ can have Hopf invariant 1 only if $n = 2, 4$, or
1766: $8$. This was proved by Adams sometime about 1958 \cite{Adams0}.
1767:
1768: \subsection{$\OP^1$ and Bott Periodicity} \label{bott}
1769:
1770: We already touched upon Bott periodicity when we mentioned that the
1771: Clifford algebra $\Cliff_{n+8}$ is isomorphic to the algebra of $16
1772: \times 16$ matrices with entries lying in $\Cliff_n$. This is but one
1773: of many related `period-8' phenomena that go by the name of Bott
1774: periodicity. The appearance of the number 8 here is no coincidence: all
1775: these phenomena are related to the octonions! Since this marvelous
1776: fact is somewhat under-appreciated, it seems worthwhile to say a bit
1777: about it. Here we shall focus on those aspects that are related to
1778: $\OP^1$ and the canonical octonionic line bundle over this space.
1779:
1780: Let us start with K-theory. This is a way of gaining information about
1781: a topological space by studying the vector bundles over it. If the
1782: space has holes in it, there will be nontrivial vector bundles that
1783: have `twists' as we go around these holes. The simplest example is
1784: the `M\"obius strip' bundle over $S^1$, a 1-dimensional real vector
1785: bundle which has a $180^\circ$ twist as we go around the circle. In
1786: fact, this is just the canonical line bundle $L_\R$. The canonical
1787: line bundles $L_\C, L_\H$ and $L_\O$ provide higher-dimensional
1788: analogues of this example.
1789:
1790: K-theory tells us to study the vector bundles over a topological space
1791: $X$ by constructing an abelian group as follows. First, take the set
1792: consisting of all isomorphism classes of real vector bundles over $X$.
1793: Our ability to take direct sums of vector bundles gives this set an
1794: `addition' operation making it into a commutative monoid. Next, adjoin
1795: formal `additive inverses' for all the elements of this set, obtaining
1796: an abelian group. This group is called $KO(X)$, the {\bf real K-theory}
1797: of $X$. Alternatively we could start with complex vector bundles and
1798: get a group called $K(X)$, but here we will be interested in real vector
1799: bundles.
1800:
1801: Any real vector bundle $E$ over $X$ gives an element $[E] \in KO(X)$, and
1802: these elements generate this group. If we pick a point in $X$, there is an
1803: obvious homomorphism $\dim \maps KO(X) \to \Z$ sending $[E]$ to the
1804: dimension of the fiber of $E$ at this point. Since the dimension is a
1805: rather obvious and boring invariant of vector bundles, it is nice to
1806: work with the kernel of this homomorphism, which is called the {\bf
1807: reduced} real K-theory of $X$ and denoted $\widetilde{KO}(X)$. This is
1808: an invariant of pointed spaces, i.e.\ spaces equipped with a designated
1809: point or {\bf basepoint}.
1810:
1811: Any sphere becomes a pointed space if we take the north pole as
1812: basepoint. The reduced real K-theory of the first eight spheres
1813: looks like this:
1814: \ban
1815: \widetilde{KO}(S^1) &\iso& \Z_2 \\
1816: \widetilde{KO}(S^2) &\iso& \Z_2 \\
1817: \widetilde{KO}(S^3) &\iso& 0 \\
1818: \widetilde{KO}(S^4) &\iso& \Z \\
1819: \widetilde{KO}(S^5) &\iso& 0 \\
1820: \widetilde{KO}(S^6) &\iso& 0 \\
1821: \widetilde{KO}(S^7) &\iso& 0 \\
1822: \widetilde{KO}(S^8) &\iso& \Z
1823: \ean
1824: where, as one might guess,
1825: \begin{itemize}
1826: \item $[L_\R]$ generates $\widetilde{KO}(S^1)$.
1827: \item $[L_\C]$ generates $\widetilde{KO}(S^2)$.
1828: \item $[L_\H]$ generates $\widetilde{KO}(S^4)$.
1829: \item $[L_\O]$ generates $\widetilde{KO}(S^8)$.
1830: \end{itemize}
1831: As mentioned in the previous section, one can build any $k$-dimensional
1832: real vector bundle over $S^n$ using a map $f \maps S^{n-1} \to \OO(k)$.
1833: In fact, isomorphism classes of such bundles are in one-to-one correspondence
1834: with homotopy classes of such maps. Moreover, two such bundles determine
1835: the same element of $\widetilde{KO}(X)$ if and only if the corresponding
1836: maps become homotopy equivalent after we compose them with the
1837: inclusion $\OO(k) \hookrightarrow \OO(\infty)$, where
1838: $\OO(\infty)$ is the direct limit of the groups $\OO(k)$. It follows that
1839: \[ \widetilde{KO}(S^n) \iso \pi_{n-1}(\OO(\infty)) .\]
1840: This fact gives us the list of homotopy groups of $\OO(\infty)$ which
1841: appears in the Introduction. It also means that to prove
1842: Bott periodicity for these homotopy groups:
1843: \[ \pi_{i+8}(\OO(\infty)) \iso \pi_i(\OO(\infty)), \]
1844: it suffices to prove Bott periodicity for real K-theory:
1845: \[ \widetilde{KO}(S^{n+8}) \iso \widetilde{KO}(S^n) . \]
1846:
1847: Why do we have Bott periodicity in real K-theory? It turns out
1848: that there is a graded ring $KO$ with
1849: \[ KO_n = \widetilde{KO}(S^n) .\]
1850: The product in this ring comes from our ability to take
1851: `smash products' of spheres and also of real vector bundles over these
1852: spheres. Multiplying by $[L_\O]$ gives an isomorphism
1853: \[
1854: \begin{array}{ccc}
1855: \widetilde{KO}(S^n) &\to& \widetilde{KO}(S^{n+8}) \\
1856: x &\mapsto& [L_\O] \,x
1857: \end{array}
1858: \]
1859: In other words, the canonical octonionic line bundle over $\OP^1$
1860: generates Bott periodicity!
1861:
1862: There is much more to say about this fact and how it relates to Bott
1863: periodicity for Clifford algebras, but alas, this would take us too far
1864: afield. We recommend that the interested reader turn to some
1865: introductory texts on K-theory, for example the one by Dale Husemoller
1866: \cite{Husemoller}. Unfortunately, all the books I know downplay the
1867: role of the octonions. To spot it, one must bear in mind the relation
1868: between the octonions and Clifford algebras, discussed in Section
1869: \ref{clifford} above.
1870:
1871: \subsection{$\OP^1$ and Lorentzian Geometry} \label{lorentz}
1872:
1873: In Section \ref{OP1} we sketched a systematic approach to projective
1874: lines over the normed division algebras. The most famous example is the
1875: Riemann sphere, $\CP^1$. As emphasized by Penrose \cite{PR}, this space
1876: has a fascinating connection to Lorentzian geometry --- or in other words,
1877: special relativity. All conformal transformations of the Riemann
1878: sphere come from fractional linear transformations
1879: \[ z \mapsto {az + b\over cz + d}, \qquad \qquad a,b,c,d \in \C. \]
1880: It is easy to see that the group of such transformations is isomorphic
1881: to $\PSL(2,\C)$: $2 \times 2$ complex matrices with determinant 1,
1882: modulo scalar multiples of the identity. Less obviously, it is also
1883: isomorphic to the Lorentz group $\SO_0(3,1)$: the identity component of
1884: the group of linear transformations of $\R^4$ that preserve the
1885: Minkowski metric
1886: \[ x \cdot y = x_1 y_1 + x_2 y_2 + x_3 y_3 - x_4 y_4 .\]
1887: This fact has a nice explanation in terms of the `heavenly sphere'.
1888: Mathematically, this is the 2-sphere consisting of all lines of the form
1889: $\{\alpha x\}$ where $x \in \R^4$ has $x \cdot x = 0$. In special
1890: relativity such lines represent light rays, so the heavenly sphere is
1891: the sphere on which the stars appear to lie when you look at the night
1892: sky. This sphere inherits a conformal structure from the Minkowski
1893: metric on $\R^4$. This allows us to identify the heavenly sphere with
1894: $\CP^1$, and it implies that the Lorentz group acts as conformal
1895: transformations of $\CP^1$. In concrete terms, what this means is that
1896: if you shoot past the earth at nearly the speed of light, the
1897: constellations in the sky will appear distorted, but all {\it angles}
1898: will be preserved.
1899:
1900: In fact, these results are not special to the complex case: the same
1901: ideas work for the other normed division algebras as well! The algebras
1902: $\R, \C, \H$ and $\O$ are related to Lorentzian geometry in 3, 4, 6, and
1903: 10 dimensions, respectively \cite{MD,MS,MS2,Schray,Sudbery}. Even
1904: better, a full explanation of this fact brings out new relationships
1905: between the normed division algebras and spinors. In what follows we
1906: explain how this works for all 4 normed division algebras, with special
1907: attention to the peculiarities of the octonionic case.
1908:
1909: To set the stage, we first recall the most mysterious of the four
1910: infinite series of Jordan algebras listed at the beginning of Section
1911: \ref{proj}: the spin factors. We described these quite concretely, but
1912: a more abstract approach brings out their kinship to Clifford algebras.
1913: Given an $n$-dimensional real inner product space $V$, let the {\bf spin
1914: factor} $\J(V)$ be the Jordan algebra freely generated by $V$ modulo
1915: relations
1916: \[ v^2 = \|v\|^2 .\]
1917: Polarizing and applying the commutative law, we obtain
1918: \[ v\circ w = \langle v, w \rangle, \]
1919: so $\J(V)$ is isomorphic to $V \oplus \R$ with the product
1920: \[ (v,\alpha) \circ (w, \beta) =
1921: (\alpha w + \beta v, \langle v,w\rangle + \alpha \beta). \]
1922:
1923: Though Jordan algebras were invented to study quantum mechanics, the
1924: spin factors are also deeply related to special relativity. We can
1925: think of $\J(V) \iso V \oplus \R$ as {\bf Minkowksi spacetime}, with
1926: $V$ as space and $\R$ as time. The reason is that $\J(V)$ is naturally
1927: equipped with a symmetric bilinear form of signature $(n,1)$, the {\bf
1928: Minkowski metric}:
1929: \[ (v,\alpha)\cdot (w,\beta) = \langle v,w\rangle - \alpha \beta. \]
1930: The group of linear transformations preserving the Minkowski metric is
1931: called $\OO(n,1)$, and the identity component of this is called the {\bf
1932: Lorentz group}, $\SO_0(n,1)$. We define the {\bf lightcone} ${\rm
1933: C}(V)$ to consist of all nonzero $x \in \J(V)$ with $x \cdot x = 0$. A
1934: 1-dimensional subspace of $\J(V)$ spanned by an element of the
1935: lightcone is called a {\bf light ray}, and the space of all light rays is
1936: called the {\bf heavenly sphere} ${\rm S}(V)$. We can identify the
1937: heavenly sphere with the unit sphere in $V$, since every light ray is
1938: spanned by an element of the form $(v,1)$ where $v \in V$ has norm one.
1939: Here is a picture of the lightcone and the heavenly sphere when $V$ is
1940: 2-dimensional:
1941:
1942: \medskip
1943: \centerline{\epsfysize=1.5in\epsfbox{heavenly.eps}}
1944: \label{heavenly}
1945: \medskip
1946:
1947: When $V$ is at least 2-dimensional, we can build a projective space from
1948: the Jordan algebra $\J(V)$. The result is none other than the heavenly
1949: sphere! To see this, note that aside from the elements 0 and 1, all
1950: projections in $\J(V)$ are of the form $p = \textstyle{1\over 2}(v, 1)$
1951: where $v \in V$ has norm one. These are the points of our projective
1952: space, but as we have seen, they also correspond to points of the
1953: heavenly sphere. Our projective space has just one line, corresponding
1954: to the projection $1 \in \J(V)$. We can visualize this line as the
1955: heavenly sphere itself.
1956:
1957: What does all this have to do with normed division algebras? To answer
1958: this, let $\K$ be a normed division algebra of dimension $n$. Then
1959: the Jordan algebra $h_2(\K)$ is secretly a spin factor! There is an
1960: isomorphism
1961: \[ \phi \maps \h_2(\K) \to J(\K \oplus \R) \iso \K \oplus \R \oplus \R \]
1962: given by
1963: \be \phi \left( \begin{array}{cc} \alpha + \beta & x \\
1964: x^\ast & \alpha - \beta \\
1965: \end{array} \right) = (x, \beta, \alpha) ,
1966: \qquad \qquad x \in \K, \; \alpha, \beta \in \R . \label{h2} \ee
1967: Furthermore, the determinant of matrices in $\h_2(\K)$ is well-defined
1968: even when $\K$ is noncommutative or nonassociative:
1969: \[ \det \left( \begin{array}{cc} \alpha + \beta & x \\
1970: x^\ast & \alpha - \beta \\
1971: \end{array} \right) = \alpha^2 - \beta^2 - \|x\|^2 , \]
1972: and clearly we have
1973: \[ \det(a) = -\phi(a) \cdot \phi(a) \]
1974: for all $a \in \h_2(\K)$.
1975:
1976: These facts have a number of nice consequences. First of all, since the
1977: Jordan algebras $\J(\K \oplus \R)$ and $\h_2(\K)$ are isomorphic, so are
1978: their associated projective spaces. We have seen that the former space
1979: is the heavenly sphere ${\rm S}(\K \oplus \R)$, and that the latter is
1980: $\KP^1$. It follows that
1981: \[ \KP^1 \iso {\rm S}(\K \oplus \R) . \]
1982: This gives another proof of something we already saw in Section
1983: \ref{OP1}: $\KP^1$ is an $n$-sphere. But it shows more. The Lorentz
1984: group $\SO_0(n+1,1)$ has an obvious action on the heavenly sphere, and
1985: the usual conformal structure on the sphere is invariant under this
1986: action. Using the above isomorphism we can transfer this group action
1987: and invariant conformal structure to $\KP^1$ in a natural way.
1988:
1989: Secondly, it follows that the determinant-preserving linear
1990: transformations of $\h_2(\K)$ form a group isomorphic to $\OO(n+1,1)$.
1991: How can we find some transformations of this sort? If $\K = \R$, this
1992: is easy: when $g \in \SL(2,\R)$ and $x \in \h_2(\R)$, we again have
1993: $gxg^* \in \h_2(\R)$, and
1994: \[ \det(gxg^*) = \det(x). \]
1995: This gives a homomorphism from $\SL(2,\R)$ to ${\rm O}(2,1)$. This
1996: homomorphism is two--to--one, since both $g = 1$ and $g = -1$ act
1997: trivially, and it maps $\SL(2,\R)$ onto the identity component of ${\rm
1998: O}(2,1)$. It follows that $\SL(2,\R)$ is a double cover of
1999: $\SO_0(2,1)$. The exact same construction works for $\K = \C$, so
2000: $\SL(2,\C)$ is a double cover of $\SO_0(3,1)$.
2001:
2002: For the other two normed division algebras the above calculation
2003: involving determinants breaks down, and it even becomes tricky to define
2004: the group $\SL(2,\K)$, so we start by working at the Lie algebra level.
2005: We say a $m \times m$ matrix with entries in the normed division algebra
2006: $\K$ is {\bf traceless} if the sum of its diagonal entries is zero. Any
2007: such traceless matrix acts as a real--linear operator on $\K^m$. When
2008: $\K$ is commutative and associative, the space of operators coming from
2009: $m \times m$ traceless matrices with entries in $\K$ is closed under
2010: commutators, but otherwise it is not, so we define $\Sl(m,\K)$ to be the
2011: Lie algebra of operators on $\K^m$ {\it generated} by operators of this
2012: form. This Lie algebra in turn generates a Lie group of real-linear
2013: operators on $\K^m$, which we call $\SL(m,\K)$. Note that
2014: multiplication in this group is given by composition of real-linear
2015: operators, which is associative even for $\K = \O$.
2016:
2017: The Lie algebra $\Sl(m,\K)$ comes born with a representation:
2018: its {\bf fundamental representation} as real-linear operators on $\K^m$,
2019: given by
2020: \[ a \maps x \mapsto ax , \qquad \qquad x \in \K^m \]
2021: whenever $a \in \Sl(m,\K)$ actually corresponds to a traceless $m \times
2022: m$ matrix with entries in $\K$. Tensoring the fundamental representation
2023: with its dual, we get a representation of $\Sl(m,\K)$ on the space
2024: of matrices $\K[m]$, given by
2025: \[ a \maps x \mapsto ax + xa^*, \qquad \qquad x \in \K[m] \]
2026: whenever $a$ is a traceless matrix with entries in $\K$. Since $ax +
2027: xa^*$ is hermitian whenever $x$ is, this representation restricts to a
2028: representation of $\Sl(m,\K)$ on $\h_m(\K)$. This in turn can be
2029: exponentiated to obtain a representation of the group $\SL(m,\K)$ on
2030: $\h_m(\K)$.
2031:
2032: Now let us return to the case $m = 2$. One can prove that the
2033: representation of $\SL(2,\K)$ on $\h_2(\K)$ is determinant-preserving
2034: simply by checking that
2035: \[ {d \over dt} \det(x + t(ax + xa^*))\, \Bigr|_{t = 0} = 0 \]
2036: when $x$ lies in $\h_2(\K)$ and $a \in \K[2]$ is traceless. Here the
2037: crucial thing is to make sure that the calculation is not spoiled by
2038: noncommutativity or nonassociativity. It follows that we have a
2039: homomorphism
2040: \[ \alpha_\K \maps \SL(2,\K) \to \SO_0(n+1,1) \]
2041: One can check that this is onto, and that its kernel consists of the
2042: matrices $\pm 1$. Thus if we define
2043: \[ \PSL(2,\K) = \SL(2,\K) / \{\pm 1\} , \]
2044: we get isomorphisms
2045: \[
2046: \begin{array}{ccl}
2047: \PSL(2,\R)& \iso &\SO_0(2,1) \\
2048: \PSL(2,\C)& \iso &\SO_0(3,1) \\
2049: \PSL(2,\H)& \iso &\SO_0(6,1) \\
2050: \PSL(2,\O)& \iso &\SO_0(9,1) .
2051: \end{array}
2052: \]
2053: Putting this together with our earlier observations, it follows that
2054: $\PSL(2,\K)$ acts as conformal transformations of $\KP^1$.
2055:
2056: We conclude with some words about how all this relates to spinors. The
2057: machinery of Clifford algebras and spinors extends effortlessly from the
2058: case of inner product spaces to vector spaces equipped with an
2059: indefinite metric. In particular, the Lorentz group $\SO_0(n+1,1)$ has
2060: a double cover called $\Spin(n+1,1)$, and this group has certain
2061: representations called spinor representations. When $n = 1,2,4$ or $8$,
2062: we actually have
2063: \[ \Spin(n+1,1) \iso \SL(2,\K) \]
2064: where $\K$ is the normed division algebra of dimension $n$. The
2065: fundamental representation of $\SL(2,\K)$ on $\K^2$ is the left-handed
2066: spinor representation of $\Spin(n+1,1)$. Its dual is the right-handed
2067: spinor representation. Moreover, the interaction between vectors and
2068: spinors that serves as the basis of supersymmetric theories of physics
2069: in spacetimes of dimension 3, 4, 6 and 10 is just the action of
2070: $\h_2(\K)$ on $\K^2$ by matrix multiplication. In a
2071: Feynman diagram, this is represented as follows:
2072:
2073: \medskip
2074: \hskip 25em \raise 1ex \hbox{$\K^2$}
2075:
2076: \centerline{\raise7.5ex \hbox{$\h_2(\K)$} \kern -.5em
2077: \epsfysize=1.0in\epsfbox{feynman.eps}}
2078:
2079: \hskip 25em $\K^2$
2080: \label{feynman2}
2081:
2082: In the case $\K = \C$, Penrose \cite{PR} has described a nice trick for
2083: getting points on the heavenly sphere from spinors.
2084: In fact, it also works for other normed division algebras:
2085: if $(x,y) \in \K^2$ is nonzero, the hermitian matrix
2086: \[
2087: \left( \begin{array}{c} x \\ y \end{array} \right)
2088: \left( \begin{array}{cc} \! x^\ast & y^\ast \! \end{array} \right)
2089: =
2090: \left( \begin{array}{cc}
2091: x x^\ast & x y^\ast \\
2092: y x^\ast & y y^\ast
2093: \end{array} \right)
2094: \]
2095: is nonzero but has determinant zero, so it defines a point on the
2096: heavenly sphere. If we restrict to spinors of norm one, this trick
2097: reduces to the Hopf map. Moreover, it clarifies the curious double role
2098: of $\KP^1$ as both the heavenly sphere in special relativity and a space
2099: of propositions in the quantum logic associated to the Jordan algebra
2100: $\h_2(\K)$: any point on the heavenly sphere corresponds to a
2101: proposition specifying the state of a spinor!
2102:
2103: \subsection{$\OP^2$ and the Exceptional Jordan Algebra} \label{OP2}
2104:
2105: The octonions are fascinating in themselves, but the magic really starts
2106: when we use them to construct the exceptional Jordan algebra $\h_3(\O)$
2107: and its associated projective space, the octonionic projective plane.
2108: The symmetry groups of these structures turn out to be exceptional Lie
2109: groups, and triality gains an eerie pervasive influence over the
2110: proceedings, since an element of $\h_3(\O)$ consists of 3 octonions and
2111: 3 real numbers. Using the relation between normed division algebras and
2112: trialities, we get an isomorphism
2113: \be
2114: \begin{array} {ccc}
2115: \h_3(\O) & \iso & \R^3 \oplus V_8 \oplus S_8^+ \oplus S_8^- \\
2116: {} & {} & {} \\
2117: \left( \begin{array}{ccc}
2118: \alpha & z^* & y^* \\
2119: z & \beta & x \\
2120: y & x^* & \gamma
2121: \end{array} \right) & \mapsto & ((\alpha,\beta,\gamma),x,y,z)
2122: \end{array}
2123: \label{jordan.triality} \ee
2124: where $\alpha,\beta,\gamma \in \R$ and $x,y,z \in \O$. Examining the
2125: Jordan product in $\h_3(\O)$ then reveals a wonderful fact: while
2126: superficially this product is defined using the $\ast$-algebra structure
2127: of $\O$, it can actually be defined using only the natural maps
2128: \[ V_8 \times S_8^+ \to S_8^- , \qquad
2129: V_8 \times S_8^- \to S_8^+ , \qquad
2130: S_8^+ \times S_8^- \to V_8
2131: \]
2132: together with the inner products on these 3 spaces. All this
2133: information is contained in the normed triality
2134: \[ t_8 \maps V_8 \times S_8^+ \times S_8^- \to \R ,\]
2135: so any automorphism of this triality gives a automorphism of $\h_3(\O)$.
2136: In Section \ref{triality} we saw that $\Aut(t_8) \iso \Spin(8)$. With
2137: a little thought, it follows that
2138: \[ \Spin(8) \subseteq \Aut(\h_3(\O)) .\]
2139:
2140: However, this picture of $\h_3(\O)$ in terms of 8-dimensional Euclidean
2141: geometry is just part of a bigger picture --- a picture set in
2142: 10-dimensional Minkowski spacetime! If we regard $\h_2(\O)$ as sitting
2143: in the lower right-hand corner of $\h_3(\O)$, we get an isomorphism
2144: \be
2145: \begin{array} {ccc}
2146: \h_3(\O) & \iso & \R \oplus \h_2(\O) \oplus \O^2 \\
2147: {} & {} & {} \\
2148: \left( \begin{array}{cc}
2149: \alpha & \psi^* \\
2150: \psi & a \\
2151: \end{array} \right) & \mapsto & (\alpha,a,\psi)
2152: \end{array}
2153: \label{jordan.10d} \ee
2154: We saw in Section \ref{lorentz} that $a \in \h_2(\O)$ and $\psi \in
2155: \O^2$ can be identified with a vector and a spinor in 10-dimensional
2156: Minkowski spacetime, respectively. Similarly, $\alpha$ is a scalar.
2157:
2158: This picture gives a representation of $\Spin(9,1)$ as linear
2159: transformations of $\h_3(\O)$. Unfortunately, most of these
2160: transformations do not preserve the Jordan product on $\h_3(\O)$. As we
2161: shall see, they only preserve a lesser structure on $\h_3(\O)$: the {\it
2162: determinant}. However, the transformations coming from the subgroup
2163: $\Spin(9) \subset \Spin(9,1)$ do preserve the Jordan product. We can
2164: see this as follows. As a representation of $\Spin(9)$, $\h_2(\O)$
2165: splits into `space' and `time':
2166: \[ \h_2(\O) \iso V_9 \oplus \R \]
2167: with the two pieces corresponding to the traceless elements of
2168: $\h_2(\O)$ and the real multiples of the identity, respectively.
2169: On the other hand, the spinor representation of $\so(9)$ splits
2170: as $S_8^+ \oplus S_8^-$ when we restrict it to $\so(8)$, so we
2171: have
2172: \[ \O^2 \iso S_9 .\]
2173: We thus obtain an isomorphism
2174: \be
2175: \begin{array} {ccc}
2176: \h_3(\O) & \iso & \R^2 \oplus V_9 \oplus S_9 \\
2177: {} & {} & {} \\
2178: \left( \begin{array}{cc}
2179: \alpha & \psi^* \\
2180: \psi & a + \beta \\
2181: \end{array} \right) & \mapsto & ((\alpha,\beta),a,\psi)
2182: \end{array}
2183: \label{jordan.9d} \ee
2184: where $a \in \h_2(\O)$ has vanishing trace and $\beta$ is a real
2185: multiple of the identity. In these terms, one can easily check that the
2186: Jordan product in $\h_3(\O)$ is built from invariant operations on
2187: scalars, vectors and spinors in 9 dimensions. It follows that
2188: \[ \Spin(9) \subseteq \Aut(\h_3(\O)) .\]
2189: For more details on this, see Harvey's book \cite{Harvey}.
2190:
2191: This does not exhaust all the symmetries of $\h_3(\O)$, since there are
2192: other automorphisms coming from the permutation group on 3 letters,
2193: which acts on $(\alpha,\beta,\gamma) \in \R^3$ and $(x,y,z) \in \O^3$ in
2194: an obvious way. Also, any matrix $g \in \OO(3)$ acts by conjugation as
2195: an automorphism of $\h_3(\O)$; since the entries of $g$ are real, there
2196: is no problem with nonassociativity here. The group $\Spin(9)$ is
2197: 36-dimensional, but the full automorphism group $\h_3(\O)$ is much
2198: bigger: it is 52-dimensional. As we explain in Section \ref{F4}, it
2199: goes by the name of $\F_4$.
2200:
2201: However, we can already do something interesting with the automorphisms
2202: we have: we can use them to diagonalize any element of $\h_3(\O)$. To
2203: see this, first note that the rotation group, and thus $\Spin(9)$, acts
2204: transitively on the unit sphere in $V_9$. This means we can use an
2205: automorphism in our $\Spin(9)$ subgroup to bring any element of $\h_3(\O)$
2206: to the form
2207: \[
2208: \left( \begin{array}{ccc}
2209: \alpha & z^* & y^* \\
2210: z & \beta & x \\
2211: y & x^* & \gamma \end{array} \right)
2212: \]
2213: where $x$ is real. The next step is to apply an automorphism
2214: that makes $y$ and $z$ real while leaving $x$ alone. To do this, note
2215: that the subgroup of $\Spin(9)$ fixing any nonzero vector in $V_9$ is
2216: isomorphic to $\Spin(8)$. When we restrict the representation $S_9$ to
2217: this subgroup, it splits as $S_8^+ \oplus S_8^-$, and with some work
2218: \cite{Harvey} one can show that $\Spin(8)$ acts on $S_8^+ \oplus S_8^-
2219: \iso \O^2$ in such a way that any element $(y,z) \in \O^2$ can be
2220: carried to an element with both components real. The final step is to
2221: take our element of $\h_3(\O)$ with all real entries and use an
2222: automorphism to diagonalize it. We can do this by conjugating it with a
2223: suitable matrix in $\OO(3)$.
2224:
2225: To understand $\OP^2$, we need to understand projections in $\h_3(\O)$.
2226: Here is where our ability to diagonalize matrices in $\h_3(\O)$ via
2227: automorphisms comes in handy. Up to automorphism, every projection in
2228: $\h_3(\O)$ looks like one of these four:
2229: \[
2230: p_0 =
2231: \left( \begin{array}{ccc}
2232: 0 & 0 & 0 \\
2233: 0 & 0 & 0 \\
2234: 0 & 0 & 0 \end{array} \right) , \; \;
2235: p_1 =
2236: \left( \begin{array}{ccc}
2237: 1 & 0 & 0 \\
2238: 0 & 0 & 0 \\
2239: 0 & 0 & 0 \end{array} \right) , \; \;
2240: p_2 =
2241: \left( \begin{array}{ccc}
2242: 1 & 0 & 0 \\
2243: 0 & 1 & 0 \\
2244: 0 & 0 & 0 \end{array} \right) , \; \;
2245: p_3 =
2246: \left( \begin{array}{ccc}
2247: 1 & 0 & 0 \\
2248: 0 & 1 & 0 \\
2249: 0 & 0 & 1 \end{array} \right) . \]
2250: Now, the trace of a matrix in $\h_3(\O)$ is invariant under
2251: automorphisms, because we can define it using only the Jordan algebra
2252: structure:
2253: \[ \tr(a) = {1\over 9} \tr(L_a) , \qquad \qquad a \in \h_3(\O) \]
2254: where $L_a$ is left multiplication by $a$. It follows that the trace of
2255: any projection in $\h_3(\O)$ is 0,1,2, or 3. Furthermore, the rank of
2256: any projection $p \in \h_3(\O)$ equals its trace. To see this, first
2257: note that $\tr(p) \ge \rank(p)$, since $p < q$ implies $\tr(p) <
2258: \tr(q)$, and the trace goes up by integer steps. Thus we only need
2259: show $\tr(p) \le \rank(p)$. For this it suffices to consider the four
2260: projections shown above, as both trace and rank are invariant under
2261: automorphisms. Since $p_0 < p_1 < p_2 < p_3$, it is clear that for
2262: these projections we indeed have $\tr(p) \le \rank(p)$.
2263:
2264: It follows that the points of the octonionic projective plane are
2265: projections with trace 1 in $\h_3(\O)$, while the lines are projections
2266: with trace 2. A calculation \cite{Harvey} shows that any projection
2267: with trace 1 has the form
2268: \be p =
2269: \left( \begin{array}{c} x \\ y \\ z \end{array} \right)
2270: \left( \begin{array}{ccc} \! x^\ast & y^\ast & z^\ast \! \end{array} \right)
2271: =
2272: \left( \begin{array}{ccc}
2273: x x^* & x y^* & x z^* \\
2274: y x^* & y y^* & y z^* \\
2275: z x^* & z y^* & z z^*
2276: \end{array} \right)
2277: \label{projection}
2278: \ee
2279: where $(x,y,z) \in \O^3$ has
2280: \[ (xy)z = x(yz), \qquad \qquad \|x\|^2 + \|y\|^2 + \|z\|^2 = 1 .\]
2281: On the other hand, any projection with trace 2 is of the form $1 - p$
2282: where $p$ has trace 1. This sets up a one-to-one correspondence between
2283: points and lines in the octonionic projective plane. If we use this
2284: correspondence to think of both as trace-1 projections, the point $p$
2285: lies on the line $p'$ if and only if $p < 1 - p'$. Of course, $p < 1 - p'$
2286: iff $p' < 1 - p$. The symmetry of this relation means the octonionic
2287: projective plane is self-dual! This is also true of the real, complex
2288: and quaternionic projective planes. In all cases, the operation that
2289: switches points and lines corresponds in quantum logic to the `negation'
2290: of propositions \cite{Varadarajan}.
2291:
2292: We use $\OP^2$ to stand for the set of points in the octonionic
2293: projective plane. Given any nonzero element $(x,y,z) \in \O^3$ with
2294: $(xy)z = x(yz)$, we can normalize it and then use equation
2295: (\ref{projection}) to obtain a point $[(x,y,z)] \in \OP^2$. Copying the
2296: strategy that worked for $\OP^1$, we can make $\OP^2$ into a smooth manifold
2297: by covering it with three coordinate charts:
2298: \begin{itemize}
2299: \item one chart containing all points of the form $[(x,y,1)]$,
2300: \item one chart containing all points of the form $[(x,1,z)]$,
2301: \item one chart containing all points of the form $[(1,y,z)]$.
2302: \end{itemize}
2303: Checking that this works is a simple calculation. The only interesting
2304: part is to make sure that whenever the associative law might appear
2305: necessary, we can either use the alternativity of the octonions or the
2306: fact that only triples with $(xy)z = x(yz)$ give points $[(x,y,z)] \in
2307: \OP^2$.
2308:
2309: We thus obtain the following picture of the octonionic projective plane.
2310: As a manifold, $\OP^2$ is 16-dimensional. The lines in $\OP^2$ are
2311: copies of $\OP^1$, and thus 8-spheres. For any two distinct points in
2312: $\OP^2$, there is a unique line on which they both lie. For any two
2313: distinct lines, there is a unique point lying on both of them. There is
2314: a `duality' transformation that maps points to lines and vice versa
2315: while preserving this incidence relation. In particular, since the
2316: space of all points lying on any given line is a copy of $\OP^1$, so
2317: is the space of all lines containing a given point!
2318:
2319: To dig more deeply into the geometry of $\OP^2$ one needs another
2320: important structure on the exceptional Jordan algebra: the determinant.
2321: We saw in Section \ref{lorentz} that despite noncommutativity and
2322: nonassociativity, the determinant of a matrix in $\h_2(\O)$ is a
2323: well-defined and useful concept. The same holds for $\h_3(\O)$! We
2324: can define the {\bf determinant} of a matrix in $\h_3(\O)$ by
2325: \[
2326: \det \left( \begin{array}{ccc}
2327: \alpha & z^* & y^* \\
2328: z & \beta & x \\
2329: y & x^* & \gamma \end{array} \right) =
2330: \alpha \beta \gamma - (\alpha \|x\|^2 + \beta \|y\|^2 + \gamma \|z\|^2)
2331: + 2 \Re(xyz) .
2332: \]
2333: We can express this in terms of the trace and product via
2334: \[ \det(a) = {1\over 3} \tr(a^3) - {1\over 2} \tr(a^2) \tr(a) +
2335: {1\over 6} {\tr(a)}^3 .\]
2336: This shows that the determinant is invariant under all automorphisms of
2337: $\h_3(\O)$. However, the determinant is invariant under an even bigger
2338: group of linear transformations. As we shall see in Section \ref{E6},
2339: this group is 78-dimensional: it is a noncompact real form of the
2340: exceptional Lie group $\E_6$. This extra symmetry makes it worth
2341: seeing how much geometry we can do starting with just the determinant
2342: and the vector space structure of $\h_3(\O)$.
2343:
2344: The determinant is a cubic form on $\h_3(\O)$, so there is a unique
2345: symmetric trilinear form
2346: \[ (\cdot,\cdot,\cdot) \maps \h_3(\O) \times \h_3(\O) \times \h_3(\O)
2347: \to \R \]
2348: such that
2349: \[ (a,a,a) = \det(a) .\]
2350: By dualizing this, we obtain the so-called {\bf cross product}
2351: \[ \times \maps \h_3(\O) \times \h_3(\O) \to \h_3(\O)^*. \]
2352: Explicitly, this is given by
2353: \[ (a \times b)(c) = (a,b,c) .\]
2354: Despite its name, this product is commutative.
2355:
2356: We have already seen that points of $\OP^2$ correspond to trace-1
2357: projections in $\h_3(\O)$. Freudenthal \cite{Freudenthal} noticed that
2358: these are the same as elements $p \in \h_3(\O)$ with $\tr(p) = 1$ and $p
2359: \times p = 0$. Even better, we can drop the equation $\tr(p) = 1$ as
2360: long as we promise to work with {\it equivalence classes} of nonzero
2361: elements satisfying $p \times p = 0$, where two such elements are
2362: equivalent when one is a nonzero real multiple of the other. Each
2363: such equivalence class $[p]$ corresponds to a unique point of $\OP^2$,
2364: and we get all the points this way.
2365:
2366: Given two points $[p]$ and $[q]$, their cross product $p \times q$ is
2367: well-defined up to a nonzero real multiple. This suggests that we define
2368: a `line' to be an equivalence class of elements $p \times q \in
2369: \h_3(\O)^*$, where again two such elements are deemed equivalent if one
2370: is a nonzero real multiple of the other. Freudenthal showed that we
2371: get a projective plane isomorphic to $\OP^2$ if we take these as our
2372: definitions of points and lines and decree that the point $[p]$ lies on
2373: the line $[L]$ if and only if $L(p) = 0$. Note that this equation
2374: makes sense even though $L$ and $p$ are only well-defined up to
2375: nonzero real multiples.
2376:
2377: One consequence of all this is that one can recover the structure of
2378: $\OP^2$ as a projective plane starting from just the determinant on
2379: $\h_3(\O)$: we did not need the Jordan algebra structure! However, to
2380: get a `duality' map switching points and lines while preserving the
2381: incidence relation, we need a bit more: we need the nondegenerate
2382: pairing
2383: \[ \langle a,b \rangle = \tr(ab) \]
2384: on $\h_3(\O)$. This sets up an isomorphism
2385: \[ \h_3(\O) \iso \h_3(\O)^* . \]
2386: This isomorphism turns out to map points to lines, and in fact, it sets
2387: up a one-to-one correspondence between points and lines. We can use
2388: this correspondence to think of both points and lines in $\OP^2$ as
2389: equivalence classes of elements of $\h_3(\O)$. In these terms, the
2390: point $p$ lies on the line $\ell$ iff $\langle \ell,p \rangle = 0$. This
2391: relationship is symmetrical! It follows that if we switch points and
2392: lines using this correspondence, the incidence relation is preserved.
2393:
2394: We thus obtain a very pretty setup for working with $\OP^2$. If we
2395: use the isomorphism between $\h_3(\O)$ and its dual to reinterpret
2396: the cross product as a map
2397: \[ \times \maps \h_3(\O) \times \h_3(\O) \to \h_3(\O) ,\]
2398: then not only is the line through distinct points $[p]$ and $[q]$ given
2399: by $[p \times q]$, but also the point in which two distinct lines
2400: $[\ell]$ and $[m]$ meet is given by $[\ell \times m]$. A
2401: triple of points $[p], [q]$ and $[r]$ is collinear iff $(p,q,r) = 0$,
2402: and a triple of lines $[\ell]$, $[m]$, $[n]$ meets at a point iff
2403: $(\ell,m,n) = 0$. In addition, there is a delightful bunch of
2404: identities relating the Jordan product, the determinant, the cross
2405: product and the inner product in $\h_3(\O)$.
2406:
2407: For more on octonionic geometry, the reader is urged to consult the
2408: original papers by Freudenthal
2409: \cite{Freudenthal4,Freudenthal,Freudenthal2,Freudenthal3}, Jacques Tits
2410: \cite{Tits,Tits2} and Tonny Springer
2411: \cite{Springer,Springer2,Springer3}. The book by Helmut Salzmann {\it
2412: et al} is also good \cite{Salzmann}. Unfortunately, we must now bid
2413: goodbye to this subject and begin our trip through the exceptional
2414: groups. However, we shall return to study the symmetries of $\OP^2$ and
2415: the exceptional Jordan algebra in Sections \ref{F4} and \ref{E6}.
2416:
2417: \section{Exceptional Lie Algebras} \label{lie}
2418:
2419: On October 18th, 1887, Wilhelm Killing wrote a letter to Friedrich Engel
2420: saying that he had classified the simple Lie algebras. In the next
2421: three years, this revolutionary work was published in a series of papers
2422: \cite{Killing}. Besides what we now call the `classical' simple Lie
2423: algebras, he claimed to have found 6 `exceptional' ones --- new
2424: mathematical objects whose existence had never before been suspected. In
2425: fact he gave a rigorous construction of only the smallest of these. In
2426: his 1894 thesis, Cartan \cite{Cartan0} constructed all of them and
2427: noticed that the two 52-dimensional exceptional Lie algebras discovered
2428: by Killing were isomorphic, so that there are really only 5.
2429:
2430: The Killing-Cartan classification of simple Lie algebras introduced much
2431: of the technology that is now covered in any introductory course on the
2432: subject, such as roots and weights. In what follows we shall avoid this
2433: technology, since we wish instead to see the exceptional Lie algebras as
2434: octonionic relatives of the classical ones --- slightly eccentric relatives,
2435: but still having a close connection to {\it geometry}, in particular the
2436: Riemannian geometry of projective planes. It is also for this reason
2437: that we shall focus on the compact real forms of the simple Lie algebras.
2438:
2439: The classical simple Lie algebras can be organized in three infinite
2440: families:
2441: \[
2442: \begin{array}{lcl}
2443: \so(n) &=& \{ x \in \R[n] \colon \; x^* = -x, \; \tr(x) = 0\}, \\
2444: \su(n) &=& \{ x \in \C[n] \colon \; x^* = -x, \; \tr(x) = 0\}, \\
2445: \symp(n) &=& \{ x \in \H[n] \colon \; x^* = -x \}.
2446: \end{array}
2447: \]
2448: The corresponding Lie groups are
2449: \[
2450: \begin{array}{lcl}
2451: \SO(n) &=& \{ x \in \R[n] \colon \; xx^* = 1, \; \det(x) = 1\}, \\
2452: \SU(n) &=& \{ x \in \C[n] \colon \; xx^* = 1, \; \det(x) = 1\}, \\
2453: \Sp(n) &=& \{ x \in \H[n] \colon \; xx^* = 1 \}.
2454: \end{array}
2455: \]
2456: These arise naturally as symmetry groups of projective spaces over
2457: $\R$, $\C$, and $\H$, respectively. More precisely, they arise as
2458: groups of {\bf isometries}: transformations that preserve a specified
2459: Riemannian metric. Let us sketch how this works, as a warmup for the
2460: exceptional groups.
2461:
2462: First consider the projective space $\RP^n$. We can think of this as
2463: the unit sphere in $\R^{n+1}$ with antipodal points $x$ and $-x$
2464: identified. It thus inherits a Riemannian metric from the sphere,
2465: and the obvious action of the rotation group $\OO(n+1)$ as isometries
2466: of the sphere yields an action of this group as isometries of $\RP^n$
2467: with this metric. In fact, with this metric, the group of all isometries
2468: of $\RP^n$ is just
2469: \[ \Isom(\RP^n) \iso \OO(n+1)/\OO(1) \]
2470: where $\OO(1) = \{\pm 1\}$ is the subgroup of $\OO(n+1)$ that acts trivially
2471: on $\RP^n$. The Lie algebra of this isometry group is
2472: \[ \isom(\RP^n) \iso \so(n+1) .\]
2473:
2474: The case of $\CP^n$ is very similar. We can think of this as the unit
2475: sphere in $\C^{n+1}$ with points $x$ and $\alpha x$ identified whenever
2476: $\alpha$ is a unit complex number. It thus inherits a Riemannian metric
2477: from this sphere, and the unitary group $\U(n+1)$ acts as isometries.
2478: If we consider only the connected component of the isometry group and
2479: ignore the orientation-reversing isometries coming from complex
2480: conjugation, we have
2481: \[ \Isom_0(\CP^n) \iso \U(n+1)/\U(1) \]
2482: where $\U(1)$ is the subgroup that acts trivially on $\CP^n$.
2483: The Lie algebra of this isometry group is
2484: \[ \isom(\CP^n) \iso \su(n+1) .\]
2485:
2486: The case of $\HP^n$ is subtler, since we must take the noncommutativity
2487: of the quaternions into account. We can think of $\HP^n$ as the unit
2488: sphere in $\H^{n+1}$ with points $x$ and $\alpha x$ identified whenever
2489: $\alpha$ is a unit quaternion, and as before, $\HP^n$ inherits a
2490: Riemannian metric. The group $\Sp(n+1)$ acts as isometries of $\HP^n$,
2491: but this action comes from {\it right} multiplication, so
2492: \[ \Isom(\HP^n) \iso \Sp(n+1)/\{ \pm 1 \}, \]
2493: since not $\Sp(1)$ but only its center $\{\pm 1\}$ acts trivially
2494: on $\HP^n$ by right multiplication. At the Lie algebra level, this
2495: gives
2496: \[ \isom(\HP^n) \iso \symp(n+1) . \]
2497:
2498: For lovers of the octonions, it is tempting to try a similar
2499: construction starting with $\OP^2$. While nonassociativity makes things
2500: a bit tricky, we show in Section \ref{F4} that it can in fact be done.
2501: It turns out that $\Isom(\OP^2)$ is one of the exceptional Lie groups,
2502: namely $\F_4$. Similarly, the exceptional Lie groups $\E_6$, $\E_7$
2503: and $\E_8$ are in a certain subtle sense the isometry groups of
2504: projective planes over the algebras $\C \tensor \O$, $\H \tensor \O$ and
2505: $\O \tensor \O$. Together with $\F_4$, these groups can all be
2506: defined by the so-called `magic square' construction, which makes use of
2507: much of the algebra we have described so far. We explain three versions
2508: of this construction in Section \ref{magic}. We then treat the groups
2509: $\E_6, \E_7$ and $\E_8$ individually in the following sections. But
2510: first, we must introduce $\G_2$: the smallest of the exceptional Lie
2511: groups, and none other than the automorphism group of the octonions.
2512:
2513: \subsection{$\G_2$} \label{G2}
2514:
2515: In 1914, \'Elie Cartan noted that the smallest of the exceptional Lie
2516: groups, $\G_2$, is the automorphism group of the octonions
2517: \cite{Cartan}. Its Lie algebra $\g_2$ is therefore $\Der(\O)$, the
2518: derivations of the octonions. Let us take these facts as definitions of
2519: $\G_2$ and its Lie algebra, and work out some of the consequences.
2520:
2521: What are automorphisms of the octonions like? One way to analyze this
2522: involves subalgebras of the octonions. Any octonion $e_1$ whose square
2523: is $-1$ generates a subalgebra of $\O$ isomorphic to $\C$. If we then
2524: pick any octonion $e_2$ with square equal to $-1$ that anticommutes with
2525: $e_1$, the elements $e_1,e_2$ generate a subalgebra isomorphic to $\H$.
2526: Finally, if we pick any octonion $e_3$ with square equal to $-1$ that
2527: anticommutes with $e_1,e_2,$ and $e_1e_2$, the elements $e_1,e_2,e_3$
2528: generate all of $\O$. We call such a triple of octonions a {\bf basic
2529: triple}. Given any basic triple, there exists a unique way to define
2530: $e_4, \dots , e_7$ so that the whole multiplication table in Section
2531: \ref{constructing} holds. In fact, this follows from the remarks on the
2532: Cayley--Dickson construction at the end of Section \ref{clifford}.
2533:
2534: It follows that given any two basic triples, there exists a unique
2535: automorphism of $\O$ mapping the first to the second. Conversely, it is
2536: obvious that any automorphism maps basic triples to basic triples.
2537: This gives a nice description of the group $\G_2$, as follows.
2538:
2539: Fix a basic triple $e_1,e_2,e_3$. There is a unique automorphism
2540: of the octonions mapping this to any other basic triple, say
2541: $e'_1,e'_2,e'_3$. Now our description of basic triples so far has
2542: been purely algebraic, but we can also view them more geometrically
2543: as follows: a basic triple is any triple of unit imaginary
2544: octonions (i.e.\ imaginary octonions of norm one) such that each is
2545: orthogonal to the algebra generated by the other two. This means that
2546: our automorphism can map $e_1$ to any point $e'_1$ on the 6-sphere of
2547: unit imaginary octonions, then map $e_2$ to any point $e'_2$ on the
2548: 5-sphere of unit imaginary octonions that are orthogonal to $e'_1$, and
2549: then map $e_3$ to any point $e'_3$ on the 3-sphere of unit imaginary
2550: octonions that are orthogonal to $e'_1,e'_2$ and $e'_1 e'_2$. It follows
2551: that
2552: \[ \dim \G_2 = \dim S^6 + \dim S^5 + \dim S^3 = 14 .\]
2553:
2554: The triality description of the octonions in Section \ref{triality}
2555: gives another picture of $\G_2$. First, recall that $\Spin(8)$ is the
2556: automorphism group of the triality $t_8 \maps V_8 \times S^+_8 \times
2557: S^-_8 \to \R$. To construct the octonions from this triality we need to
2558: pick unit vectors in any two of these spaces, so we can think of $\G_2$
2559: as the subgroup of $\Spin(8)$ fixing unit vectors in $V_8$ and $S^+_8$.
2560: The subgroup of $\Spin(8)$ fixing a unit vector in $V_8$ is just
2561: $\Spin(7)$, and when we restrict the representation $S^+_8$ to
2562: $\Spin(7)$, we get the spinor representation $S_7$. Thus $\G_2$ is the
2563: subgroup of $\Spin(7)$ fixing a unit vector in $S_7$. Since $\Spin(7)$
2564: acts transitively on the unit sphere $S^7$ in this spinor representation
2565: \cite{Adams}, we have
2566: \[ \Spin(7)/\G_2 = S^7 . \]
2567: It follows that
2568: \[ \dim \G_2 = \dim (\Spin(7)) - \dim S^7 = 21 - 7 = 14 .\]
2569:
2570: This picture becomes a bit more vivid if we remember that after choosing
2571: unit vectors in $V_8$ and $S^+_8$, we can identify both these
2572: representations with the octonions, with both unit vectors corresponding
2573: to $1 \in \O$. Thus what we are really saying is this: the subgroup of
2574: $\Spin(8)$ that fixes $1$ in the vector representation on $\O$ is
2575: $\Spin(7)$; the subgroup that fixes $1$ in both the vector and
2576: right-handed spinor representations is $\G_2$. This subgroup also fixes
2577: the element $1$ in the left-handed spinor representation of $\Spin(8)$
2578: on $\O$.
2579:
2580: Now, using the vector representation of $\Spin(8)$ on $\O$, we
2581: get homomorphisms
2582: \[ \G_2 \hookrightarrow \Spin(8) \to \SO(\O) \]
2583: where $\SO(\O) \iso \SO(8)$ is the rotation group of the octonions,
2584: viewed as a real vector space with the inner product $\langle x,y\rangle
2585: = \Re(x^\ast y)$. The map from $\Spin(8)$ to $\SO(\O)$ is two-to-one,
2586: but when we restrict it to $\G_2$ we get a one-to-one map
2587: \[ \G_2 \hookrightarrow \SO(\O) . \]
2588:
2589: At the Lie algebra level, this construction gives an inclusion
2590: \[ \g_2 \hookrightarrow \so(\O) \]
2591: where $\so(\O) \iso \so(8)$ is the Lie algebra of skew-adjoint
2592: real-linear transformations of the octonions. Since $\g_2$ is
2593: 14-dimensional and $\so(\O)$ is 28-dimensional, it is nice to see
2594: exactly where the extra 14 dimensions come from. In fact, they come
2595: from two copies of $\Im(\O)$, the 7-dimensional space consisting of all
2596: imaginary octonions.
2597:
2598: More precisely, we have:
2599: \be \so(\O) = \g_2 \oplus L_{\Im(\O)} \oplus R_{\Im(\O)} \label{so(O)} \ee
2600: (a direct sum of vector spaces, not Lie algebras), where $L_{\Im(\O)}$
2601: is the space of linear transformations of $\O$ given by left multiplication
2602: by imaginary octonions and $R_{\Im(\O)}$ is the space of linear
2603: transformations of $\O$ given by right multiplication by imaginary
2604: octonions \cite{Schafer}. To see this, we first check that left
2605: multiplication by an imaginary octonion is skew-adjoint. Using
2606: polarization, it suffices to note that
2607: \[ \langle x,ax \rangle = \Re(x^*(ax)) = \Re((x^*a)x) =
2608: \Re((a^*x)^*x) = -\Re((ax)^* x) = -\langle ax,x \rangle \]
2609: for all $a \in \Im(\O)$ and $x \in \O$. Note that this calculation only
2610: uses the alternative law, not the associative law, since $x, x^\ast$
2611: and $a$ all lie in the algebra generated by the two elements $a$ and
2612: $\Im(x)$. A similar argument shows that right multiplication by
2613: an imaginary octonion is skew-adjoint. It follows that
2614: $\g_2$, $L_{\Im(\O)}$ and $R_{\Im(\O)}$ all naturally lie in $\so(8)$.
2615: Next, with some easy calculations we can check that
2616: \[ L_{\Im(\O)} \cap R_{\Im(\O)} = \{0\} \]
2617: and
2618: \[ \g_2 \cap (L_{\Im(\O)} + R_{\Im(\O)}) = \{0\} .\]
2619: Using the fact that the dimensions of the 3 pieces adds to 28,
2620: equation (\ref{so(O)}) follows.
2621:
2622: We have seen that $\G_2$ sits inside $\SO(8)$, but we can do better: it
2623: actually sits inside $\SO(7)$. After all, every automorphism of the
2624: octonions fixes the identity, and thus preserves the space of octonions
2625: orthogonal to the identity. This space is just $\Im(\O)$, so
2626: we have an inclusion
2627: \[ \G_2 \hookrightarrow \SO(\Im(\O)) \]
2628: where $\SO(\Im(\O)) \iso \SO(7)$ is the rotation group of the imaginary
2629: octonions. At the Lie algebra level this gives an inclusion
2630: \[ \g_2 \hookrightarrow \so(\Im(\O)) . \]
2631:
2632: Since $\g_2$ is 14-dimensional and $\so(\Im(\O))$ is 21-dimensional, it
2633: is nice to see where the 7 extra dimensions come from. Examining
2634: equation (\ref{so(O)}), it is clear that these extra dimensions must
2635: come from the transformations in $ L_{\Im(\O)} \oplus R_{\Im(\O)}$ that
2636: annihilate the identity $1 \in \O$. The transformations that do this are
2637: precisely those of the form
2638: \[ \ad_a = L_a - R_a \]
2639: for $a \in \Im(\O)$. We thus have
2640: \be \so(\Im(\O)) \iso \Der(\O) \oplus \ad_{\Im(\O)} \label{so(Im(O))} \ee
2641: where $\ad_{\Im(\O)}$ is the 7-dimensional space of such transformations.
2642:
2643: We may summarize some of the above results as follows:
2644:
2645: \begin{thm} \et \label{g2-description} The compact real form
2646: of the Lie algebra $\g_2$ is given by
2647: \[ \g_2 = \Der(\O) \subset \so(\Im(\O)) \subset \so(\O) \]
2648: and we have
2649: \ban \so(\Im(\O)) &=& \Der(\O) \oplus \ad_{\Im(\O)} \\
2650: \so(\O) &=& \Der(\O) \oplus L_{\Im(\O)} \oplus R_{\Im(\O)}
2651: \ean
2652: where the Lie brackets in $\so(\Im(\O))$ and $\so(\O)$ are built
2653: from natural bilinear operations on the summands.
2654: \end{thm}
2655:
2656: As we have seen, $\G_2$ has a 7-dimensional representation $\Im(\O)$.
2657: In fact, this is the smallest nontrivial representation of $\G_2$,
2658: so it is worth understanding in as many ways as possible. The space
2659: $\Im(\O)$ has at least three natural structures that are preserved by
2660: the transformations in $\G_2$. These give more descriptions of
2661: $\G_2$ as a symmetry group, and they also shed some new light on the
2662: octonions. The first two of the structures we describe are analogous to
2663: more familiar ones that exist on the 3-dimensional space of imaginary
2664: quaternions, $\Im(\H)$. The third makes explicit use of the
2665: nonassociativity of the octonions.
2666:
2667: First, both $\Im(\H)$ and $\Im(\O)$ are closed under the commutator. In
2668: the case of $\Im(\H)$, the commutator divided by 2 is the familiar {\bf
2669: cross product} in 3 dimensions:
2670: \[ a \times b = {1\over 2}[a,b] .\]
2671: We can make the same definition for $\Im(\O)$, obtaining a 7-dimensional
2672: analog of the cross product. For both $\Im(\H)$ and $\Im(\O)$ the
2673: cross product is bilinear and anticommutative. The cross product
2674: makes $\Im(\H)$ into a Lie algebra, but not $\Im(\O)$. For both
2675: $\Im(\H)$ and $\Im(\O)$, the cross product has two nice
2676: geometrical properties. On the one hand, its norm is determined by the
2677: formula
2678: \[ \|a \times b\|^2 + \langle a,b\rangle^2 = \|a\|^2 \, \|b\|^2 , \]
2679: or equivalently,
2680: \[ \|a \times b\| = | {\sin \theta}| \, \|a\| \, \|b\| , \]
2681: where $\theta$ is the angle between $a$ and $b$. On the other hand, $a
2682: \times b$ is orthogonal to $a$ and $b$. Both these properties follow
2683: from easy calculations. For $\Im(\H)$, these two properties are enough
2684: to determine $x \times y$ up to a sign. For $\Im(\O)$ they are not ---
2685: but they become so if we also use the fact that $x \times y$ lies inside
2686: a copy of $\Im(\H)$ that contains $x$ and $y$.
2687:
2688: It is clear that the group of all real-linear transformations of
2689: $\Im(\H)$ preserving the cross product is just $\SO(3)$, which is also
2690: the automorphism group of the quaternions. One can similarly show that
2691: the group of real-linear transformations of $\Im(\O)$ preserving the
2692: cross product is exactly $\G_2$. To see this, start by noting that any
2693: element of $\G_2$ preserves the cross product on $\Im(\O)$, since the
2694: cross product is defined using octonion multiplication. To show that
2695: conversely any transformation preserving the cross product lies in
2696: $\G_2$, it suffices to express the multiplication of imaginary octonions
2697: in terms of their cross product. Using this identity:
2698: \[ a \times b = ab + \langle a, b\rangle , \]
2699: it actually suffices to express the inner product on $\Im(\O)$ in terms
2700: of the cross product. Here the following identity does the job:
2701: \be \langle a,b\rangle =
2702: -\textstyle{1\over 6}\tr(a \times (b \times \cdot\,))
2703: \label{inner} \ee
2704: where the right-hand side refers to the trace of the map
2705: \[ a \times (b \times \cdot\,) \maps \Im(\O) \to \Im(\O) .\]
2706:
2707: Second, both $\Im(\H)$ and $\Im(\O)$ are equipped with a natural 3-form,
2708: or in other words, an alternating trilinear functional. This is given
2709: by
2710: \[ \phi(x,y,z) = \langle x,yz \rangle .\]
2711: In the case of $\Im(\H)$ this is just the usual volume form, and the
2712: group of real-linear transformations preserving it is $\SL(3, \R)$. In
2713: the case of $\Im(\O)$, the real-linear transformations preserving $\phi$
2714: are exactly those in the group $\G_2$. A proof of this by Robert
2715: Bryant can be found in Reese Harvey's book \cite{Harvey}. The 3-form
2716: $\phi$ is important in the theory of `Joyce manifolds' \cite{Joyce},
2717: which are 7-dimensional Riemannian manifolds with holonomy group equal
2718: to $\G_2$.
2719:
2720: Third, both $\Im(\H)$ and $\Im(\O)$ are closed under the associator.
2721: For $\Im(\H)$ this is boring, since the associator vanishes. On the
2722: other hand, for $\Im(\O)$ the associator is interesting. In fact, it
2723: follows from results of Harvey \cite{Harvey} that a real-linear
2724: transformation $T \maps \Im(\O) \to \Im(\O)$ preserves the associator
2725: if and only if $\pm T$ lies in $\G_2$. Thus the symmetry group of the
2726: associator is slightly bigger than $\G_2$: it is $\G_2 \times \Z_2$.
2727:
2728: Now we must make an embarrassing admission: these three structures on
2729: $\Im(\O)$ are all almost the same thing! Starting with the cross
2730: product
2731: \[ \times \maps \Im(\O) \times \Im(\O) \to \Im(\O) \]
2732: we can recover the usual inner product on $\Im(\O)$ by equation
2733: (\ref{inner}). This inner product allows us to dualize the cross
2734: product and obtain a trilinear functional, which is, up to a constant,
2735: just the 3-form
2736: \[ \phi \maps \Im(\O) \times \Im(\O) \times \Im(\O) \to \R .\]
2737: The cross product also determines an orientation on $\Im(\O)$ (we leave
2738: this as an exercise for the reader). This allows us to take the Hodge
2739: dual of $\phi$, obtaining a 4-form $\psi$, i.e.\ an alternating
2740: tetralinear functional
2741: \[ \psi \maps \Im(\O) \times \Im(\O) \times \Im(\O) \times \Im(\O)
2742: \to \R .\]
2743: Dualizing yet again, this gives a ternary operation which, up
2744: to a constant multiple, is the associator:
2745: \[ [\cdot, \cdot, \cdot] \maps \Im(\O) \times \Im(\O) \times \Im(\O)
2746: \to \Im(\O) .\]
2747:
2748: We conclude this section with a handy explicit formula for all the
2749: derivations of the octonions. In an associative algebra $A$, any
2750: element $x$ defines an {\bf inner derivation} $\ad_x \maps A \to A$ by
2751: \[ \ad_x (a) = [x,a] \]
2752: where the bracket denotes the commutator $xa - ax$.
2753: In a nonassociative algebra, this formula usually does not define
2754: a derivation. However, if $A$ is alternative, any pair of elements
2755: $x,y \in A$ define a derivation $D_{x,y} \maps A \to A$ by
2756: \be D_{x,y} a = [[x,y],a] - 3[x,y,a] \label{D} \ee
2757: where $[a,b,x]$ denotes the associator $(ab)x - a(bx)$. Moreover,
2758: when $A$ is a normed division algebra, every derivation is a linear
2759: combination of derivations of this form. Unfortunately, proving these
2760: facts seems to require some brutal calculations \cite{Schafer}.
2761:
2762: \subsection{$\F_4$} \label{F4}
2763:
2764: The second smallest of the exceptional Lie groups is the 52-dimensional
2765: group $\F_4$. The geometric meaning of this group became clear in a
2766: number of nearly simultaneous papers by various mathematicians. In
2767: 1949, Jordan constructed the octonionic projective plane using
2768: projections in $\h_3(\O)$. One year later, Armand Borel \cite{Borel}
2769: noted that $\F_4$ is the isometry group of a 16-dimensional projective
2770: plane. In fact, this plane is none other than than $\OP^2$. Also
2771: in 1950, Claude Chevalley and Richard Schafer \cite{CS} showed that
2772: $\F_4$ is the automorphism group of $\h_3(\O)$. In 1951, Freudenthal
2773: \cite{Freudenthal4} embarked upon a long series of papers in which he
2774: described not only $\F_4$ but also the other exceptional Lie groups
2775: using octonionic projective geometry. To survey these developments, one
2776: still cannot do better than to read his classic 1964 paper on Lie groups
2777: and the foundations of geometry \cite{Freudenthal3}.
2778:
2779: Let us take Chevalley and Schafer's result as the definition of $\F_4$:
2780: \[ \F_4 = \Aut(\h_3(\O)) .\]
2781: Its Lie algebra is thus
2782: \[ \f_4 = \Der(\h_3(\O)). \label{f4.1} \]
2783: As we saw in Section \ref{OP2}, points of $\OP^2$ correspond to trace-1
2784: projections in the exceptional Jordan algebra. It follows that $\F_4$
2785: acts as transformations of $\OP^2$. In fact, we can equip $\OP^2$ with
2786: a Riemannian metric for which $\F_4$ is the isometry group. To get a
2787: sense of how this works, let us describe $\OP^2$ as a quotient space of
2788: $\F_4$.
2789:
2790: In Section \ref{OP2} we saw that the exceptional Jordan algebra can
2791: be built using natural operations on the scalar, vector and spinor
2792: representations of $\Spin(9)$. This implies that $\Spin(9)$ is a
2793: subgroup of $\F_4$. Equation (\ref{jordan.9d}) makes it clear that
2794: $\Spin(9)$ is precisely the subgroup fixing the element
2795: \[ \left( \begin{array}{ccc} 1 & 0 & 0 \\
2796: 0 & 0 & 0 \\
2797: 0 & 0 & 0 \\
2798: \end{array} \right). \]
2799: Since this element is a trace-one projection, it corresponds to a point
2800: of $\OP^2$. We have already seen that $\F_4$ acts transitively on
2801: $\OP^2$. It follows that
2802: \be \OP^2 \iso \F_4 /\Spin(9) . \label{F4/Spin(9)} \ee
2803:
2804: This fact has various nice spinoffs. First, it gives an easy way to
2805: compute the dimension of $\F_4$:
2806: \[ \dim(\F_4) = \dim(\Spin(9)) + \dim(\OP^2) =
2807: 36 + 16 = 52.\]
2808: Second, since $\F_4$ is compact, we can take any Riemannian metric on $\OP^2$
2809: and average it with respect to the action of this group. The isometry
2810: group of the resulting metric will automatically include $\F_4$ as a
2811: subgroup. With more work \cite{Besse}, one can show that actually
2812: \[ \F_4 = \Isom(\OP^2) \]
2813: and thus
2814: \[ \f_4 = \isom(\OP^2). \label{f4.2} \]
2815:
2816: Equation (\ref{F4/Spin(9)}) also implies that the tangent space of our
2817: chosen point in $\OP^2$ is isomorphic to $\f_4/\so(9)$. But we already
2818: know that this tangent space is just $\O^2$, or in other words, the
2819: spinor representation of $\so(9)$. We thus have
2820: \be \f_4 \iso \so(9) \oplus S_9 \label{f4.3} \ee
2821: as vector spaces, where $\so(9)$ is a Lie subalgebra. The bracket in
2822: $\f_4$ is built from the bracket in $\so(9)$, the action $\so(9) \tensor
2823: S_9 \to S_9$, and the map $S_9 \tensor S_9 \to \so(9)$ obtained by
2824: dualizing this action. We can also rewrite this description
2825: of $\f_4$ in terms of the octonions, as follows:
2826: \[ \f_4 \iso \so(\O \oplus \R) \oplus \O^2 \label{f4.4} \]
2827:
2828: This last formula suggests that we decompose $\f_4$ further using the
2829: splitting of $\O \oplus \R$ into $\O$ and $\R$.
2830: It is easily seen by looking at matrices that for all $n,m$ we have
2831: \be
2832: \so(n+m) \iso \so(n) \oplus \so(m) \, \oplus\, V_n \tensor V_m.
2833: \label{so(n+m)} \ee
2834: Moreover, when we restrict the representation
2835: $S_9$ to $\so(8)$, it splits as a direct sum $S_8^+ \oplus S_8^-$.
2836: Using these facts and equation (\ref{f4.3}), we see
2837: \be
2838: \f_4 \iso \so(8) \oplus V_8 \oplus S_8^+ \oplus S_8^-
2839: \label{f4.5}
2840: \ee
2841: This formula emphasizes the close relation between $\f_4$ and triality:
2842: the Lie bracket in $\f_4$ is completely built out of maps involving
2843: $\so(8)$ and its three 8-dimensional irreducible representations!
2844: We can rewrite this in a way that brings out the role of the octonions:
2845: \[
2846: \f_4 \iso \so(\O) \oplus \O^3
2847: \label{f4.6}
2848: \]
2849:
2850: While elegant, none of these descriptions of $\f_4$ gives a convenient
2851: picture of all the derivations of the exceptional Jordan algebra. In
2852: fact, there is a nice picture of this sort for $\h_3(\K)$ whenever $\K$
2853: is a normed division algebra. One way to get a derivation of the
2854: Jordan algebra $\h_3(\K)$ is to take a derivation of $\K$ and let it
2855: act on each entry of the matrices in $\h_3(\K)$. Another way uses
2856: elements of
2857: \[ \sa_3(\K) = \{ x \in \K[3] \colon \; x^* = -x,\; \tr(x) = 0 \} .\]
2858: Given $x \in \sa_3(\K)$, there is a derivation $\ad_x$ of $\h_3(\K)$ given
2859: by
2860: \[ \ad_x (a) = [x,a] .\]
2861: In fact \cite{BS}, every derivation of $\h_3(\K)$ can be uniquely
2862: expressed as a linear combination of derivations of these two sorts,
2863: so we have
2864: \be \Der(\h_3(\K)) \iso \Der(\K) \oplus \sa_3(\K) \label{der(h3K)} \ee
2865: as vector spaces. In the case of the octonions, this decomposition
2866: says that
2867: \[ \f_4 \iso \g_2 \oplus \sa_3(\O) . \label{f4.7} \]
2868:
2869: In equation (\ref{der(h3K)}), the subspace $\Der(\K)$ is always a Lie
2870: subalgebra, but $\sa_3(\K)$ is not unless $\K$ is commutative and
2871: associative --- in which case $\Der(\K)$ vanishes. Nonetheless, there
2872: is a formula for the brackets in $\Der(\h_3(\K))$ which applies in every
2873: case \cite{OV}. Given $D,D' \in \Der(\K)$ and $x,y \in \sa_3(\K)$, we
2874: have
2875: \be
2876: \begin{array}{lcl}
2877: [D,D'] &=& DD' - D'D \cr
2878: [D,\ad_x] &=& \ad_{Dx} \cr
2879: [\ad_x,\ad_y] &=& \ad_{[x,y]_0} +
2880: \displaystyle{{1\over 3} \sum_{i,j = 1}^3 D_{x_{ij},y_{ij}} }
2881: \end{array}
2882: \label{der(h3K)-bracket}
2883: \ee
2884: where $D$ acts on $x$ componentwise, $[x,y]_0$ is the trace-free
2885: part of the commutator $[x,y]$, and $D_{x_{ij},y_{ij}}$ is the
2886: derivation of $\K$ defined using equation (\ref{D}).
2887:
2888: Summarizing these different descriptions of $\f_4$, we have:
2889: \begin{thm} \et \label{f4-description} The compact real form of
2890: $\f_4$ is given by
2891: \ban \f_4 &\iso& \isom(\OP^2) \\
2892: &\iso& \Der(\h_3(\O)) \\
2893: &\iso& \Der(\O) \oplus \sa_3(\O) \\
2894: &\iso& \so(\O \oplus \R) \oplus \O^2 \\
2895: &\iso& \so(\O) \oplus \O^3
2896: \ean
2897: where in each case the Lie bracket is built from
2898: natural bilinear operations on the summands.
2899: \end{thm}
2900:
2901: \subsection{The Magic Square} \label{magic}
2902:
2903: Around 1956, Boris Rosenfeld \cite{Rosenfeld1} had the remarkable idea
2904: that just as $\F_4$ is the isometry group of the projective plane over
2905: the octonions, the exceptional Lie groups $\E_6$, $\E_7$ and $\E_8$ are
2906: the isometry groups of projective planes over the following three
2907: algebras, respectively:
2908: \begin{itemize}
2909: \item the {\bf bioctonions}, $\C \tensor \O$,
2910: \item the {\bf quateroctonions}, $\H \tensor \O$,
2911: \item the {\bf octooctonions}, $\O \tensor \O$.
2912: \end{itemize}
2913: There is definitely something right about this idea, because one would
2914: expect these projective planes to have dimensions 32, 64, and 128, and
2915: there indeed do exist compact Riemannian manifolds with these
2916: dimensions having $\E_6$, $\E_7$ and $\E_8$ as their isometry groups.
2917: The problem is that the bioctonions, quateroctonions and and
2918: octooctonions are not division algebras, so it is a nontrivial matter to
2919: define projective planes over them!
2920:
2921: The situation is not so bad for the bioctonions: $\h_3(\C \tensor \O)$
2922: is a simple Jordan algebra, though not a formally real one, and one can
2923: use this to define $(\C \tensor \O)\P^2$ in a manner modeled after one
2924: of the constructions of $\OP^2$. Rosenfeld claimed that a similar
2925: construction worked for the quateroctonions and octooctonions, but this
2926: appears to be false. Among other problems, $\h_3(\H \tensor \O)$ and
2927: $\h_3(\O \tensor \O)$ do not become Jordan algebras under the product
2928: $a \circ b = {1\over 2}(ab + ba)$. Scattered throughout the literature
2929: \cite{Besse,Freudenthal3,Freudenthal5} one can find frustrated comments
2930: about the lack of a really nice construction of $(\H \tensor \O)\P^2$
2931: and $(\O \tensor \O)\P^2$. One problem is that these spaces do {\it
2932: not} satisfy the usual axioms for a projective plane. Tits addressed
2933: this problem in his theory of `buildings', which allows one to construct
2934: a geometry having any desired algebraic group as symmetries
2935: \cite{Tits4}. But alas, it still seems that the quickest way to get our
2936: hands on the quateroctonionic and octooctonionic `projective planes' is
2937: by {\it starting} with the Lie groups $\E_7$ and $\E_8$ and then taking
2938: quotients by suitable subgroups.
2939:
2940: In short, more work must be done before we can claim to fully understand
2941: the geometrical meaning of the Lie groups $\E_6, \E_7$ and $\E_8$.
2942: Luckily, Rosenfeld's ideas can be used to motivate a nice construction
2943: of their Lie algebras. This goes by the name of the `magic square'.
2944: Tits \cite{Tits3} and Freudenthal \cite{Freudenthal2} found two very
2945: different versions of this construction in about 1958, but we shall
2946: start by presenting a simplified version published by E.\ B.\ Vinberg
2947: \cite{Vinberg} in 1966.
2948:
2949: First consider the projective plane $\KP^2$ when $\K$ is a normed
2950: division algebra $\K$. The points of this plane are the rank-1
2951: projections in the Jordan algebra $\h_3(\K)$, and this plane admits a
2952: Riemannian metric such that
2953: \[ \isom(\KP^2) \iso \Der(\h_3(\K)). \]
2954: Moreover, we have seen in equation (\ref{der(h3K)}) that
2955: \[ \Der(\h_3(\K)) \iso \Der(\K) \oplus \sa_3(\K) . \]
2956: Combined with Rosenfeld's observations, these facts might lead one to
2957: hope that whenever we have a pair of normed division algebras $\K$ and
2958: $\K'$, there is a Riemannian manifold $(\K \tensor \K')\P^2$ with
2959: \[ \isom((\K\tensor \K')\P^2) \iso
2960: \Der(\K) \oplus \Der(\K') \oplus \sa_3(\K \tensor \K') \]
2961: where for any $\ast$-algebra $A$ we define
2962: \[
2963: \begin{array}{lcl}
2964: \sh_n(A) & =& \{ x \in A[n] \colon \; x^* = x, \; \tr(x) = 0\} \\
2965: \sa_n(A) & =& \{ x \in A[n] \colon \; x^* = -x, \; \tr(x) = 0\}.
2966: \end{array}
2967: \]
2968:
2969: This motivated Vinberg's definition of the {\bf magic square} Lie
2970: algebras:
2971: \[
2972: \M(\K,\K') = \Der(\K) \oplus \Der(\K') \oplus \sa_3(\K \tensor \K').
2973: \]
2974: Now, when $\K \tensor \K'$ is commutative and associative, $\sa_3(\K
2975: \tensor \K')$ is a Lie algebra with the commutator as its Lie bracket,
2976: but in the really interesting cases it is not. Thus to make
2977: $\M(\K,\K')$ into a Lie algebra we must give it a rather subtle bracket.
2978: We have already seen the special case $\K' = \R$ in equation
2979: (\ref{der(h3K)-bracket}). In general, the Lie bracket in $\M(\K,\K')$
2980: is given as follows:
2981: \begin{enumerate}
2982: \item $\Der(\K)$ and $\Der(\K')$ are commuting Lie subalgebras of $\M(\K,\K')$.
2983: \item The bracket of $D \in \Der(K) \oplus \Der(\K')$
2984: with $x \in \sa_3(\K \tensor \K')$ is given by applying
2985: $D$ to every entry of the matrix $x$, using the natural action of
2986: $\Der(K) \oplus \Der(\K')$ as derivations of $\K \tensor \K'$.
2987: \item Given $X,Y \in \sa_3(\K \tensor \K')$,
2988: \[ [X,Y] = [X,Y]_0 +
2989: {1\over 3}\displaystyle{\sum_{i,j = 1}^3 D_{X_{ij},Y_{ij}} } .\]
2990: Here $[X,Y]_0$ is the traceless part of the $3 \times 3$ matrix $[X,Y]$,
2991: and given $x,y \in \K \tensor \K'$ we define $D_{x,y} \in
2992: \Der(\K) \oplus \Der(\K')$ in the following way: $D_{x,y}$ is
2993: real-bilinear in $x$ and $y$, and
2994: \[ D_{a \tensor a',b \tensor b'} = \langle a',b' \rangle D_{a,b} +
2995: \langle a,b \rangle D_{a',b'} \]
2996: where $a,b \in \K$, $a',b' \in \K'$, and $D_{a,b},D_{a',b'}$ are defined
2997: as in equation (\ref{D}).
2998: \end{enumerate}
2999: With this construction we magically obtain the following square of Lie
3000: algebras:
3001: \vskip 1em
3002: {\vbox{
3003: \begin{center}
3004: \renewcommand{\baselinestretch}{1.3}
3005: {\small
3006: \begin{tabular}{|c|c|c|c|c|} \hline
3007: & $\K' = \R$ & $\K' = \C$ & $\K' = \H$ & $\K' = \O$ \\ \hline
3008: $\K = \R$ & $\so(3)$ & $\su(3)$ & $\symp(3)$ & $\f_4$ \\ \hline
3009: $\K = \C$ & $\su(3)$ & $\su(3) \oplus \su(3)$ & $\su(6)$ & $\e_6$ \\ \hline
3010: $\K = \H$ & $\symp(3)$ & $\su(6)$ & $\so(12)$ & $\e_7$ \\ \hline
3011: $\K = \O$ & $\f_4$ & $\e_6$ & $\e_7$ & $\e_8$ \\ \hline
3012: \end{tabular}} \vskip 1em
3013: Table 5 --- Magic Square Lie Algebras $\M(\K,\K')$
3014: \end{center}
3015: } }
3016: \vskip 0.5em
3017: \noindent
3018: We will mainly be interested in the last row (or column), which is the
3019: one involving the octonions. In this case we can take the magic square
3020: construction as {\it defining} the Lie algebras $\f_4$, $\e_6$, $\e_7$
3021: and $\e_8$. This definition turns out to be consistent with our earlier
3022: definition of $\f_4$.
3023:
3024: Starting from Vinberg's definition of the magic square Lie algebras, we
3025: can easily recover Tits' original definition. To do so, we need two
3026: facts. First,
3027: \[
3028: \sa_3(\K \otimes \K')
3029: \iso \sa_3(\K') \, \oplus \, (\Im(\K) \! \tensor \! \sh_3(\K')).
3030: \]
3031: This is easily seen by direct examination of the relevant matrices.
3032: Second,
3033: \[ \Der(\h_3(\K)) \iso \Der(\K) \oplus \sa_3(\K) \]
3034: as vector spaces. This is just equation (\ref{der(h3K)}). Starting with
3035: Vinberg's definition and applying these two facts, we obtain
3036: \[
3037: \begin{array}{lcl}
3038: \M(\K,\K') &=& \Der(\K) \oplus \Der(\K') \oplus \sa_3(\K \tensor \K') \\
3039: &\iso& \Der(\K) \oplus \Der(\K')
3040: \oplus \sa_3(\K') \, \oplus \, (\Im(\K) \tensor \sh_3(\K')) \\
3041: &\iso&
3042: \Der(\K) \oplus \Der(\h_3(\K')) \, \oplus \, (\Im(\K) \tensor \sh_3(\K')) .
3043: \end{array}
3044: \]
3045: The last line is Tits' definition of the magic square Lie algebras.
3046: Unlike Vinberg's, it is not manifestly symmetrical in $\K$ and $\K'$.
3047: This unhappy feature is somewhat made up for by the fact that $\Der(\K)
3048: \oplus \Der(\h_3(\K'))$ is a nice big Lie subalgebra. This subalgebra
3049: acts on $\Im(\K) \tensor \sh_3(\K')$ in an obvious way, using the fact
3050: that any derivation of $\K$ maps $\Im(\K)$ to itself, and any derivation
3051: of $\h_3(\K')$ maps $\sh_3(\K')$ to itself. However, the bracket of two
3052: elements of $(\Im(\K) \tensor \sh_3(\K'))$ is a bit of a mess.
3053:
3054: Yet another description of the magic square was recently given by
3055: Barton and Sudbery \cite{BS}. This one emphasizes the role of
3056: trialities. Let $\Tri(\K)$ be the Lie algebra of the group $\Aut(t)$,
3057: where $t$ is the normed triality giving the normed division algebra
3058: $\K$. From equation (\ref{Aut(t)}) we have
3059: \be
3060: \begin{array}{lcl}
3061: \Tri(\R) &\iso& \{0\} \\
3062: \Tri(\C) &\iso& \u(1)^2 \\
3063: \Tri(\H) &\iso& \symp(1)^3 \\
3064: \Tri(\O) &\iso& \so(8) .
3065: \end{array}
3066: \label{tri}
3067: \ee
3068: To express the magic square in terms of these Lie algebras, we need
3069: three facts. First, it is easy to see that
3070: \[ \sh_3(\K) \iso \K^3 \oplus \R^2 .\]
3071: Second, Barton and Sudbery show that as vector spaces,
3072: \[ \Der(\h_3(\K)) \iso \Tri(\K) \oplus \K^3 . \]
3073: This follows in a case--by--case way from equation
3074: (\ref{tri}), but they give a unified proof that covers all cases.
3075: Third, they show that as vector spaces,
3076: \[ \Tri(\K) \iso \Der(\K) \oplus \Im(\K)^2 . \]
3077: Now starting with Tits' definition of the magic square,
3078: applying the first two facts, regrouping terms, and applying
3079: the third fact, we obtain Barton and Sudbery's version of the
3080: magic square:
3081: \[
3082: \begin{array}{lcl}
3083: \M(\K,\K')
3084: &\iso& \Der(\K) \oplus \Der(\h_3(\K')) \, \oplus \,
3085: (\Im(\K) \tensor \sh_3(\K')) \\
3086: &\iso & \Der(\K) \oplus \Tri(\K') \oplus \K'^3 \oplus
3087: \Im(\K) \! \tensor \! (\K'^3 \oplus \R^2) \\
3088: &\iso & \Der(\K) \oplus \Im(\K)^2 \oplus \Tri(\K') \oplus (\K \tensor \K')^3 \\
3089: &\iso & \Tri(\K) \oplus \Tri(\K') \oplus (\K \tensor \K')^3 .
3090: \end{array}
3091: \]
3092:
3093: In the next three sections we use all these different versions of the
3094: magic square to give lots of octonionic descriptions of $\e_6$, $\e_7$
3095: and $\e_8$. To save space, we usually omit the formulas for the Lie
3096: bracket in these descriptions. However, the patient reader can
3097: reconstruct these with the help of Barton and Sudbery's paper, which
3098: is packed with useful formulas.
3099:
3100: As we continue our tour through the exceptional Lie algebras, we shall
3101: make contact with Adams' work \cite{Adams} constructing
3102: $\f_4,\e_6,\e_7,$ and $\e_8$ with the help of spinors and rotation group
3103: Lie algebras:
3104: \[
3105: \begin{array}{lcl}
3106: \f_4 &\iso& \so(9) \oplus S_9 \\
3107: \e_6 &\iso& \so(10) \oplus \u(1) \oplus S_{10} \\
3108: \e_7 &\iso& \so(12) \oplus \symp(1) \oplus S_{12}^+ \\
3109: \e_8 &\iso& \so(16) \oplus S_{16}^+
3110: \end{array}
3111: \]
3112: as vector spaces. Note that the numbers 9, 10, 12 and 16 are 8 more
3113: than the dimensions of $\R,\C,\H$ and $\O$. As usual, this is no
3114: coincidence! In terms of the octonions, Bott periodicity implies that
3115: \[ S_{n+8} \iso S_n \tensor \O^2 .\]
3116: This gives the following description of spinors in dimensions $\le 16$:
3117:
3118: \medskip
3119: {\vbox{
3120: \begin{center}
3121: \renewcommand{\baselinestretch}{1.3}
3122: {\small
3123: \begin{tabular}{|l|l|} \hline
3124: $S_1 = \R$ & $S_9 = \O^2$ \\ \hline
3125: $S_2 = \C$ & $S_{10} = (\C \tensor \O)^2$ \\ \hline
3126: $S_3 = \H$ & $S_{11} = (\H \tensor \O)^2$ \\ \hline
3127: $S_4^\pm = \H$ & $S_{12}^\pm = (\H \tensor \O)^2$ \\ \hline
3128: $S_5 = \H^2$ & $S_{13} = (\H^2 \tensor \O)^2$ \\ \hline
3129: $S_6 = \C^4$ & $S_{14} = (\C^4 \tensor \O)^2$ \\ \hline
3130: $S_7 = \O$ & $S_{15} = (\O \tensor \O)^2$ \\ \hline
3131: $S_8^\pm = \O$ & $S_{16}^\pm = (\O \tensor \O)^2$ \\ \hline
3132: \end{tabular}}
3133: \vskip 1em
3134: \centerline{Table 6 --- Spinor Representations Revisited}
3135: \end{center}
3136: }}
3137: \medskip
3138:
3139: \noindent
3140: Since spinors in dimensions 1,2,4 and 8 are isomorphic to the division
3141: algebras $\R,\C,\H$ and $\O$, spinors in dimensions 8 higher are
3142: isomorphic to the `planes' $\O^2, (\C \tensor \O)^2, (\H \tensor \O)^2$
3143: and $(\O \tensor \O)^2$ --- and are thus closely linked to $\f_4$, $\e_6$,
3144: $\e_7$ and $\e_8$, thanks to the magic square.
3145:
3146: \subsection{$\E_6$} \label{E6}
3147:
3148: We begin with the 78-dimensional exceptional Lie group $\E_6$.
3149: As we mentioned in Section \ref{OP2}, there is a nice description of a
3150: certain noncompact real form of $\E_6$ as the group of collineations
3151: of $\OP^2$, or equivalently, the group of determinant-preserving
3152: linear transformations of $\h_3(\O)$. But before going into these,
3153: we consider the magic square constructions of the Lie algebra $\e_6$.
3154: Vinberg's construction gives
3155: \[
3156: \e_6 = \Der(\O) \oplus \sa_3(\C \tensor \O) .
3157: \label{e6.1}
3158: \]
3159: Tits' construction, which is asymmetrical, gives
3160: \[
3161: \e_6 \iso \Der(\h_3(\O)) \oplus \sh_3(\O)
3162: \label{e6.2}
3163: \]
3164: and also
3165: \[
3166: \e_6 \iso \Der(\O) \oplus \Der(\h_3(\C)) \oplus
3167: (\Im(\O) \!\tensor \! \sh_3(\C)) .
3168: \label{e6.3}
3169: \]
3170: The Barton-Sudbery construction gives
3171: \[
3172: \e_6 \iso \Tri(\O) \oplus \Tri(\C) \oplus (\C \tensor \O)^3 .
3173: \label{e6.4}
3174: \]
3175: We can use any of these to determine the dimension of $\e_6$. For
3176: example, we have
3177: \[ \dim(\e_6) = \dim(\Der(\h_3(\O))) + \dim(\sh_3(\O)) = 52 + 26 = 78. \]
3178:
3179: Starting from the Barton-Sudbery construction and using the concrete
3180: descriptions of $\Tri(\O)$ and $\Tri(\C)$ from equation
3181: (\ref{tri}), we obtain
3182: \[
3183: \e_6 \iso \so(\O) \oplus \so(\C) \oplus \Im(\C) \oplus (\C \tensor \O)^3
3184: \label{e6.5}
3185: \]
3186: Using equation (\ref{so(n+m)}), we may rewrite this as
3187: \[
3188: \e_6 \iso \so(\O \oplus \C) \oplus \Im(\C) \oplus (\C \tensor \O)^2
3189: \label{e6.6}
3190: \]
3191: and it turns out that the summand $\so(\O \oplus \C) \oplus \Im(\C)$ is
3192: actually a Lie subalgebra of $\e_6$. This result can also be found in
3193: Adams' book \cite{Adams}, phrased as follows:
3194: \[
3195: \e_6 \iso \so(10) \oplus \u(1) \oplus S_{10}
3196: \label{e6.7}
3197: \]
3198: In fact, he describes the bracket in $\e_6$ in terms of natural
3199: operations involving $\so(10)$ and its spinor representation $S_{10}$.
3200: The funny-looking factor of $\u(1)$ comes from the fact that this spinor
3201: representation is complex. The bracket of an element of $\u(1)$ and
3202: an element of $S_{10}$ is another element of $S_{10}$, defined using
3203: the obvious action of $\u(1)$ on this complex vector space.
3204:
3205: If we define $\E_6$ to be the simply connected group with Lie algebra
3206: $\e_6$, it follows from results of Adams that the subgroup generated by
3207: the Lie subalgebra $\so(10) \oplus \u(1)$ is isomorphic to $(\Spin(10)
3208: \times \U(1))/\Z_4$. This lets us define the {\bf bioctonionic
3209: projective plane} by
3210: \[ (\C \tensor \O)\P^2 = \E_6\, / \, ((\Spin(10) \times \U(1))/\Z_4) \]
3211: and conclude that the tangent space at any point of this manifold is
3212: isomorphic to $S_{10} \iso (\C \tensor \O)^2$.
3213:
3214: Since $\E_6$ is compact, we can put an $\E_6$-invariant Riemannian metric on the bioctonionic
3215: projective plane by averaging any metric with respect to the action
3216: of this group. It turns out \cite{Besse} that the isometry group of this
3217: metric is exactly $\E_6$, so we have
3218: \[ \E_6 \iso \Isom((\C \tensor \O)\P^2).
3219: \]
3220: It follows that
3221: \[
3222: \e_6 \iso \isom((\C \tensor \O)\P^2) .
3223: \label{e6.8}
3224: \]
3225:
3226: Summarizing, we have 6 octonionic descriptions of $\e_6$:
3227: \begin{thm} \et \label{e6-description} The compact real form of
3228: $\e_6$ is given by
3229: \ban
3230: \e_6 &\iso& \isom((\C \tensor \O)\P^2) \\
3231: &\iso& \Der(\O) \oplus \Der(\h_3(\C)) \oplus
3232: (\Im(\O) \! \tensor\! \sh_3(\C)) \\
3233: &\iso& \Der(\h_3(\O)) \oplus \sh_3(\O) \\
3234: &\iso& \Der(\O) \oplus \sa_3(\C \tensor \O) \\
3235: &\iso& \so(\O \oplus \C) \oplus \Im(\C) \oplus (\C \tensor \O)^2 \\
3236: &\iso& \so(\O) \oplus \so(\C) \oplus \Im(\C) \oplus (\C \tensor \O)^3
3237: \ean
3238: where in each case the Lie bracket of $\e_6$ is built from
3239: natural bilinear operations on the summands.
3240: \end{thm}
3241:
3242: The smallest nontrivial representations of $\E_6$ are 27-dimensional:
3243: in fact it has two inequivalent representations of this dimension, which
3244: are dual to one another. Now, the exceptional Jordan algebra is also
3245: 27-dimensional, and in 1950 this clue led Chevalley and Schafer \cite{CS}
3246: to give a nice description of $\E_6$ as symmetries of this algebra.
3247: These symmetries do not preserve the product, but only the determinant.
3248:
3249: More precisely, the group of determinant-preserving linear
3250: transformations of $\h_3(\O)$ turns out to be a noncompact real form of
3251: $\E_6$. This real form is sometimes called $\E_{6(-26)}$, because its
3252: Killing form has signature $-26$. To
3253: see this, note that any automorphism of $\h_3(\O)$ preserves the
3254: determinant, so we get an inclusion
3255: \[ \F_4 \hookrightarrow \E_{6(-26)} .\]
3256: This means that $\F_4$ is a compact subgroup of $\E_{6(-26)}$. In fact
3257: it is a maximal compact subgroup, since if there were a larger one, we
3258: could average a Riemannian metric group on $\OP^2$ with respect to this
3259: group and get a metric with an isometry group larger than $\F_4$, but no
3260: such metric exists. It follows that the Killing form on the Lie algebra
3261: $\e_{6(-26)}$ is negative definite on its 52-dimensional maximal compact
3262: Lie algebra, $\f_4$, and positive definite on the complementary
3263: 26-dimensional subspace, giving a signature of $26 - 52 = -26$.
3264:
3265: We saw in Section \ref{OP2} that the projective plane structure of
3266: $\OP^2$ can be constructed starting only with the determinant function
3267: on the vector space $\h_3(\O)$. It follows that $\E_{6(-26)}$
3268: acts as {\bf collineations} on $\OP^2$, that is, line-preserving
3269: transformations. In fact, the group of collineations of $\OP^2$ is
3270: precisely $\E_{6(-26)}$:
3271: \[ \E_{6(-26)} \iso {\rm Coll}(\OP^2). \]
3272: Moreover, just as the group of isometries of $\OP^2$ fixing a specific
3273: point is a copy of $\Spin(9)$, the group of collineations fixing a
3274: specific point is $\Spin(9,1)$. This fact follows with some work
3275: starting from equation (\ref{jordan.10d}), and it gives us a commutative
3276: square of inclusions:
3277: \[
3278: \begin{array}{ccl}
3279: \Spin(9) &\longrightarrow& \Isom(\OP^2) \iso \F_4 \\
3280: \downarrow & & \;\;\;\;\; \downarrow \\
3281: \Spin(9,1) & \longrightarrow & {\rm Coll}(\OP^2) \iso \E_{6(-26)}
3282: \end{array}
3283: \]
3284: where the groups on the top are maximal compact subgroups of those on
3285: the bottom. Thus in a very real sense, $\F_4$ is to 9-dimensional
3286: Euclidean geometry as $\E_{6(-26)}$ is to 10-dimensional Lorentzian
3287: geometry.
3288:
3289: \subsection{$\E_7$} \label{E7}
3290:
3291: Next we turn to the 133-dimensional exceptional Lie group $\E_7$.
3292: In 1954, Freudenthal \cite{Freudenthal2} described this group as
3293: the automorphism group of a 56-dimensional octonionic structure now
3294: called a `Freudenthal triple system'. We sketch this idea
3295: below, but first we give some magic square constructions.
3296: Vinberg's version of the magic square gives
3297: \[
3298: \e_7 = \Der(\H) \oplus \Der(\O) \oplus \sa_3(\H \tensor \O) .
3299: \label{e7.1}
3300: \]
3301: Tits' version gives
3302: \be
3303: \e_7 \iso \Der(\H) \oplus \Der(\h_3(\O)) \oplus
3304: (\Im(\H) \!\tensor \! \sh_3(\O))
3305: \label{e7.2}
3306: \ee
3307: and also
3308: \[
3309: \e_7 \iso \Der(\O) \oplus \Der(\h_3(\H)) \oplus
3310: (\Im(\O) \!\tensor \! \sh_3(\H))
3311: \label{e7.3}
3312: \]
3313: The Barton-Sudbery version gives
3314: \be
3315: \e_7 \iso \Tri(\O) \oplus \Tri(\H) \oplus (\H \tensor \O)^3
3316: \label{e7.4}
3317: \ee
3318:
3319: Starting from equation (\ref{e7.2}) and using the fact that
3320: $\Der(\H) \iso \Im(\H)$ is 3-dimensional, we obtain the elegant
3321: formula
3322: \[
3323: \e_7 \iso \Der(\h_3(\O)) \, \oplus \, \h_3(\O)^3 .
3324: \label{e7.5}
3325: \]
3326: This gives an illuminating way to compute the dimension of $\e_7$:
3327: \[ \dim(\e_7) = \dim(\Der(\h_3(\O))) + 3 \dim(\h_3(\O)) = 52 + 3 \cdot 27 =
3328: 133 .\]
3329: Starting from equation (\ref{e7.4}) and using the concrete
3330: descriptions of $\Tri(\H)$ and $\Tri(\O)$ from equation
3331: (\ref{tri}), we obtain
3332: \[
3333: \e_7 \iso \so(\O) \oplus \so(\H) \oplus \Im(\H) \oplus (\H \tensor \O)^3
3334: \label{e7.6}
3335: \]
3336: Using equation (\ref{so(n+m)}), we may rewrite this as
3337: \[
3338: \e_7 \iso \so(\O \oplus \H) \oplus \Im(\H) \oplus (\H \tensor \O)^2.
3339: \label{e7.7}
3340: \]
3341: Though not obvious from what we have done, the direct summand $\so(\O
3342: \oplus \H) \oplus \Im(\H)$ here is really a Lie subalgebra of $\e_7$. In
3343: less octonionic language, this result can also be found in Adams' book
3344: \cite{Adams}:
3345: \[
3346: \e_7 \iso \so(12) \oplus \symp(1) \oplus S_{12}^+
3347: \label{e7.8}
3348: \]
3349: He describes the bracket in $\e_7$ in terms of natural operations
3350: involving $\so(12)$ and its spinor representation $S_{12}^+$. The
3351: funny-looking factor of $\symp(1)$ comes from the fact that this
3352: representation is quaternionic. The bracket of an element of $\symp(1)$ and
3353: an element of $S_{12}^+$ is the element of $S_{12}^+$ defined using
3354: the natural action of $\symp(1)$ on this space.
3355:
3356: If we let $\E_7$ be the simply connected group with Lie
3357: algebra $\e_7$, it follows from results of Adams \cite{Adams} that the
3358: subgroup generated by the Lie subalgebra $\so(12) \oplus \symp(1)$ is
3359: isomorphic to $(\Spin(12) \times \Sp(1))/\Z_2$.
3360: This lets us define the {\bf quateroctonionic projective plane} by
3361: \[ (\H \tensor \O)\P^2 = \E_7\, / \,((\Spin(12) \times \Sp(1))/\Z_2) \]
3362: and conclude that the tangent space at any point of this manifold is
3363: isomorphic to $S_{12}^+ \iso (\H \tensor \O)^2$. We can put
3364: an $\E_7$-invariant Riemannian metric on this manifold by the technique
3365: of averaging over the group action. It then turns out \cite{Besse} that
3366: \[ \E_7 \iso \Isom((\H \tensor \O)\P^2)
3367: \]
3368: and thus
3369: \[
3370: \e_7 \iso \isom((\H \tensor \O)\P^2) .
3371: \label{e7.9}
3372: \]
3373:
3374: Summarizing, we have the following 7 octonionic descriptions of $\e_7$:
3375:
3376: \begin{thm} \et \label{e7-description} The compact real form of
3377: $\e_7$ is given by
3378: \ban
3379: \e_7 &\iso& \isom((\H \tensor \O)\P^2) \\
3380: &\iso& \Der(\h_3(\O)) \oplus \h_3(\O)^3 \\
3381: &\iso& \Der(\O) \oplus \Der(\h_3(\H)) \oplus
3382: (\Im(\O) \!\tensor \! \sh_3(\H)) \\
3383: &\iso& \Der(\H) \oplus \Der(\h_3(\O)) \oplus
3384: (\Im(\H) \! \tensor\! \sh_3(\O)) \\
3385: &\iso& \Der(\O) \oplus \Der(\H) \oplus \sa_3(\H \tensor \O) \\
3386: &\iso& \so(\O \oplus \H) \oplus \Im(\H) \oplus (\H \tensor \O)^2 \\
3387: &\iso& \so(\O) \oplus \so(\H) \oplus \Im(\H) \oplus (\H \tensor \O)^3
3388: \ean
3389: where in each case the Lie bracket of $\e_7$ is built from
3390: natural bilinear operations on the summands.
3391: \end{thm}
3392:
3393: Before the magic square was developed, Freudenthal \cite{Freudenthal2}
3394: used another octonionic construction to study $\E_7$. The smallest
3395: nontrivial representation of this group is 56-dimensional. Freudenthal
3396: realized we can define a 56-dimensional space
3397: \[ F = \{ \left( \begin{array}{cc}
3398: \alpha & x \\
3399: y & \beta \\
3400: \end{array} \right) : \;
3401: x,y \in \h_3(\O) , \; \alpha , \beta \in \R \}
3402: \]
3403: and equip this space with a symplectic structure
3404: \[ \omega \maps F \times F \to \R \]
3405: and trilinear product
3406: \[ \tau \maps F \times F \times F \to F \]
3407: such that the group of linear transformations preserving both these
3408: structures is a certain noncompact real form of $\E_7$, namely
3409: $\E_{7(-25)}$. The symplectic structure and trilinear product on
3410: $F$ satisfy some relations, and algebraists have made these into the
3411: definition of a `Freudenthal triple system' \cite{Brown,Faulkner,Meyberg}.
3412: The geometrical significance of this rather complicated sort of structure
3413: has recently been clarified by some physicists working on string theory.
3414: At the end of the previous section, we
3415: mentioned a relation between 9-dimensional Euclidean geometry and
3416: $\F_4$, and a corresponding relation between 10-dimensional Lorentzian
3417: geometry and $\E_{6(-26)}$. Murat G\"unaydin \cite{Gunaydin} has
3418: extended this to a relation between 10-dimensional {\sl conformal}
3419: geometry and $\E_{7(-25)}$, and in work with Kilian Koepsell
3420: and Hermann Nikolai \cite{GKN} has explicated how this is connected
3421: to Freudenthal triple systems.
3422:
3423: \subsection{$\E_8$} \label{E8}
3424:
3425: With 248 dimensions, $\E_8$ is the biggest of the exceptional Lie
3426: groups, and in some ways the most mysterious. The easiest way to
3427: understand a group is to realize it as as symmetries of a structure one
3428: already understands. Of all the simple Lie groups, $\E_8$ is the only
3429: one whose smallest nontrivial representation is the adjoint
3430: representation. This means that in the context of linear algebra,
3431: $\E_8$ is most simply described as the group of symmetries of its own
3432: Lie algebra! One way out of this vicious circle would be to describe
3433: $\E_8$ as isometries of a Riemannian manifold. As already mentioned,
3434: $\E_8$ is the isometry group of a 128-dimensional manifold called $(\O
3435: \tensor \O)\P^2$. But alas, nobody seems to know how to define $(\O
3436: \tensor \O)\P^2$ without first defining $\E_8$. Thus this group remains
3437: a bit enigmatic.
3438:
3439: At present, to get our hands on $\E_8$ we must start with its Lie
3440: algebra. We can define this using any of the three equivalent magic
3441: square constructions explained in Section \ref{magic}. Vinberg's
3442: construction gives
3443: \[
3444: \e_8 = \Der(\O) \oplus \Der(\O) \oplus \sa_3(\O \tensor \O) .
3445: \label{e8.1}
3446: \]
3447: Tits' construction gives
3448: \[
3449: \e_8 \iso \Der(\O) \oplus \Der(\h_3(\O)) \oplus
3450: (\Im(\O) \!\tensor \! \sh_3(\O)) .
3451: \label{e8.2}
3452: \]
3453: The Barton-Sudbery construction gives
3454: \be
3455: \begin{array}{lcl}
3456: \e_8 &\iso& \Tri(\O) \oplus \Tri(\O) \oplus (\O \tensor \O)^3 \\
3457: &\iso& \so(\O) \oplus \so(\O) \oplus (\O \tensor \O)^3
3458: \end{array}
3459: \label{e8.3}
3460: \ee
3461: We can use any of these to count the dimension of $\e_8$; for example,
3462: the last one gives
3463: \[ \dim \e_8 = 28 + 28 + 3 \cdot 8^2 = 248.\]
3464:
3465: To emphasize the importance of triality, we can rewrite equation
3466: (\ref{e8.3}) as:
3467: \be
3468: \e_8 \iso \so(8) \oplus \so(8) \oplus (V_8 \tensor V_8) \oplus
3469: (S_8^+ \tensor S_8^+) \oplus (S_8^- \tensor S_8^-).
3470: \label{e8.4}
3471: \ee
3472: Here the Lie bracket is built from natural maps relating $\so(8)$
3473: and its three 8-dimensional irreducible representations. In particular,
3474: $\so(8) \oplus \so(8)$ is a Lie subalgebra, and the first copy of
3475: $\so(8)$ acts on the first factor in $V_8 \tensor V_8$, $S_8^+ \tensor
3476: S_8^+$, and $S_8^- \tensor S_8^-$, while the second copy acts on the
3477: second factor in each of these. The reader should compare this to
3478: the description of $\f_4$ in equation (\ref{f4.5}).
3479:
3480: Now, equation (\ref{so(n+m)}) implies that
3481: \[ \so(16) \iso \so(8) \oplus \so(8) \oplus (V_8 \tensor V_8) .\]
3482: Together with equation (\ref{e8.4}), this suggests that $\e_8$ contains
3483: $\so(16)$ as a Lie subalgebra. In fact this is true! Even better, if we
3484: restrict the right-handed spinor representation of $\so(16)$ to $\so(8)
3485: \oplus \so(8)$, it decomposes as
3486: \[ S^+_{16} \iso (S_8^+ \tensor S_8^+) \oplus (S_8^- \tensor S_8^-),\]
3487: so we obtain
3488: \be \e_8 \iso \so(16) \oplus S^+_{16} \label{e8.5} \ee
3489: or in more octonionic language,
3490: \[
3491: \e_8 \iso \so(\O \oplus \O) \oplus (\O \otimes \O)^2
3492: \label{e8.6} \]
3493: where we use $\so(V)$ to mean the Lie algebra of skew-adjoint real-linear
3494: transformations of the real inner product space $V$.
3495:
3496: The really remarkable thing about equation (\ref{e8.5}) is that the Lie
3497: bracket in $\e_8$ is entirely built from natural maps involving
3498: $\so(16)$ and $S^+_{16}$:
3499: \[ \so(16) \tensor \so(16) \to \so(16) , \qquad
3500: \so(16) \tensor S^+_{16} \to S^+_{16} , \qquad
3501: S^+_{16} \tensor S^+_{16} \to \so(16) .\]
3502: The first of these is the Lie bracket in $\so(16)$, the second is the
3503: action of $\so(16)$ on its right-handed spinor representation, and the
3504: third is obtained from the second by duality, using the natural inner
3505: product on $\so(16)$ and $S^+_{16}$ to identify these spaces with their
3506: duals. In fact, this is a very efficient way to {\it define} $\e_8$.
3507: If we take this approach, we must verify the Jacobi identity:
3508: \[ [[a,b],c] = [a,[b,c]] - [b,[a,c]] .\]
3509: When all three of $a,b,c$ lie in $\so(16)$ this is just the Jacobi
3510: identity for $\so(16)$. When two of them lie in $\so(16)$, it boils
3511: down to fact that spinors indeed form a representation of $\so(16)$.
3512: Thanks to duality, the same is true when just one lies in $\so(16)$. It
3513: thus suffices to consider the case when $a,b,c$ all lie in $S_{16}^+$.
3514: This is the only case that uses anything special about the number 16.
3515: Unfortunately, at this point a brute-force calculation seems to be
3516: required. For two approaches that minimize the pain, see the books by
3517: Adams \cite{Adams} and by Green, Schwarz and Witten \cite{GSW}. It
3518: would be nice to find a more conceptual approach.
3519:
3520: Starting from $\e_8$, we can define $\E_8$ to be the simply-connected
3521: Lie group with this Lie algebra. As shown by Adams \cite{Adams}, the
3522: subgroup of $\E_8$ generated by the Lie subalgebra $\so(16) \subset
3523: \e_8$ is $\Spin(16)/\Z_2$. This lets us define the {\bf octooctonionic
3524: projective plane} by
3525: \[ (\O \tensor \O)\P^2 = \E_8\,/\,(\Spin(16)/\Z_2) . \]
3526: By equation (\ref{e8.5}), the tangent space at any point
3527: of this manifold is isomorphic to $S_{16}^+ \iso (\O \tensor \O)^2$.
3528: This partially justifies calling it `octooctonionic projective plane',
3529: though it seems not to satisfy the usual axioms for a projective plane.
3530:
3531: We can put an $\E_8$-invariant Riemannian metric on the octooctonionic
3532: projective plane by the technique of averaging over the group action.
3533: It then turns out \cite{Besse} that
3534: \[ \E_8 \iso \Isom((\O \tensor \O)\P^2) \]
3535: and thus
3536: \[ \e_8 \iso \isom((\O \tensor \O)\P^2) . \label{e8.7} \]
3537:
3538: Summarizing, we have the following octonionic descriptions of
3539: $\E_8$:
3540: \begin{thm} \et \label{e8-description} The compact real form of
3541: $\e_8$ is given by
3542: \ban
3543: \e_8 &\iso& \isom((\O \tensor \O)\P^2) \\
3544: &\iso& \Der(\O) \oplus \Der(\h_3(\O)) \oplus
3545: (\Im(\O) \!\tensor\! \sh_3(\O)) \\
3546: &\iso& \Der(\O) \oplus \Der(\O) \oplus \sa_3(\O \tensor \O) \\
3547: &\iso& \so(\O \oplus \O) \oplus (\O \tensor \O)^2 \\
3548: &\iso& \so(\O) \oplus \so(\O) \oplus (\O \tensor \O)^3
3549: \ean
3550: where in each case the Lie bracket on $\e_8$ is built from
3551: natural bilinear operations on the summands.
3552: \end{thm}
3553:
3554: \section{Conclusions}
3555:
3556: It should be clear by now that besides being a fascinating mathematical
3557: object in their own right, the octonions link together many important
3558: phenomena whose connections would otherwise be completely mysterious.
3559: Indeed, the full story of these connections is deeper and more elaborate
3560: than I have been able to explain here! It also includes:
3561: \begin{itemize}
3562: \item Attempts to set up an octonionic analogue of the theory of
3563: analytic functions (see \cite{GT} and the references therein).
3564: \item The role of Jordan pairs, Jordan triple systems and
3565: Freudenthal triple systems in the construction of exceptional Lie groups
3566: \cite{Brown,Faulkner,FF,GKN,GT,McCrimmon,Meyberg}.
3567: \item Constructions of the $\E_8$ lattice and Leech lattice using
3568: integral octonions \cite{Coxeter,EG}.
3569: \item Tensor-categorical approaches to normed division algebras
3570: and the invariant of framed trivalent graphs coming from the
3571: quantum group associated to $\G_2$ \cite{Boos,Bremner,Kuperberg,Rost}.
3572: \item Octonionic constructions of vertex operator algebras \cite{FFrenkel}.
3573: \item Octonionic constructions of the exceptional simple Lie superalgebras
3574: \cite{Sudbery2}.
3575: \item Octonionic constructions of symmetric spaces \cite{Besse}.
3576: \item Octonions and the geometry of the `squashed 7-spheres', that is,
3577: the homogeneous spaces $\Spin(7)/\G_2$, $\Spin(6)/\SU(3)$, and
3578: $\Spin(5)/\SU(2)$, all of which are diffeomorphic to $S^7$ with its
3579: usual smooth structure \cite{CD}.
3580: \item The theory of `Joyce manifolds', that is, 7-dimensional Riemannian
3581: manifolds with holonomy group $\G_2$ \cite{Joyce}.
3582: \item The octonionic Hopf map and instanton solutions
3583: of the Yang-Mills equations in 8 dimensions \cite{GKS}.
3584: \item Octonionic aspects of 10-dimensional superstring theory and
3585: 10-dimensional super-Yang-Mills theory
3586: \cite{CH,Deligne,Evans,KT,Schray,Sierra}.
3587: \item Octonionic aspects of 11-dimensional supergravity and supermembrane
3588: theories, and the role of Joyce manifolds in compactifying 11-dimensional
3589: supergravity to obtain theories of physics in 4 dimensions \cite{Duff}.
3590: \item Geoffrey Dixon's extension of the Standard Model based on the
3591: algebra $\C \tensor \H \tensor \O$ \cite{Dixon}.
3592: \item Other attempts to use the octonions in physics
3593: \cite{CMT,GT,LPS,Okubo}.
3594: \end{itemize}
3595:
3596: \noindent I urge the reader to explore these with the help of the references.
3597:
3598: \subsection*{Acknowledgements}
3599:
3600: I thank John Barrett, Toby Bartels, Robert Bryant, Geoffrey Dixon, James
3601: Dolan, Tevian Dray, Bertram Kostant, Linus Kramer, Pertti Lounesto,
3602: Corinne Manogue, John McKay, David Rusin, Tony Smith, Anthony Sudbery, and
3603: Matthew Wiener for useful discussions.
3604:
3605: \begin{thebibliography}{10}
3606:
3607: \bibitem{Adams0} John F.\ Adams,
3608: On the non-existence of elements of Hopf invariant
3609: one, {\sl Ann.\ Math.\ }{\bf 72} (1960), 20--104.
3610:
3611: \bibitem{Adams} John F.\ Adams, {\sl Lectures on Exceptional Lie
3612: Groups}, eds.\ Zafer Mahmud and Mamoru Mimira, University of
3613: Chicago Press, Chicago, 1996.
3614:
3615: \bibitem{AH} Michael Atiyah and Friedrich Hirzebruch,
3616: Bott periodicity and the parallelizability of the spheres.
3617: {\sl Proc.\ Cambridge Philos.\ Soc.\ }{\bf 57} (1961), 223--226.
3618:
3619: \bibitem{AM} Helena Albuquerque and Shahn Majid, Quasialgebra structure
3620: of the octonions, preprint available as math.QA/9802116.
3621:
3622: \bibitem{BS} Chris H.\ Barton and Anthony Sudbery, Magic squares of Lie
3623: algebras, preprint available as math.RA/0001083.
3624:
3625: \bibitem{Besse} Arthur L.\ Besse, {\sl Einstein Manifolds}, Springer,
3626: Berlin, 1987, pp.\ 313--316.
3627:
3628: \bibitem{vdB} F.\ van der Blij, History of the octaves, {\sl Simon
3629: Stevin} {\bf 34} (1961), 106---125.
3630:
3631: \bibitem{vdBS} F.\ van der Blij and Tonny A.\ Springer, Octaves and
3632: triality, {\sl Nieuw Arch.\ v.\ Wiskunde} {\bf 8} (1960), 158--169.
3633:
3634: \bibitem{Boos} Dominik Boos, {\sl Ein tensorkategorieller Zugang zum Satz
3635: von Hurwitz}, Diplomarbeit, ETH Zurich, March 1998.
3636:
3637: \bibitem{Borel} Armand Borel, Le plan projectif des octaves et les sph\'eres
3638: commes espaces homog\`enes, {\sl Compt.\ Rend.\ Acad.\ Sci.\ }{\bf 230}
3639: (1950), 1378--1380.
3640:
3641: \bibitem{BM} Raoul Bott and John Milnor, On the parallelizability
3642: of the spheres, {\sl Bull.\ Amer.\ Math.\ Soc.\ } {\bf 64} (1958)
3643: 87--89.
3644:
3645: \bibitem{Bremner} Murray Bremner, Quantum octonions,
3646: {\sl Comm.\ Alg.} {\bf 27} (1999), 2809--2831.
3647:
3648: \bibitem{Brown} Robert B.\ Brown, Groups of type $\E_7$, {\sl
3649: Jour.\ Reine Angew.\ Math.\ }{\bf 236} (1969), 79--102.
3650:
3651: \bibitem{Cartan0} \'Elie Cartan, {\sl Sur la structure des groupes
3652: de tranformations finis et continus}, Th\`ese, Paris, Nony, 1894.
3653:
3654: \bibitem{Cartan} \'Elie Cartan, Les groupes r\'eels simples finis et
3655: continus, {\sl Ann.\ Sci.\ \'Ecole Norm.\ Sup.\ }{\bf 31} (1914),
3656: 255--262.
3657:
3658: \bibitem{Cartan2} \'Elie Cartan, Nombres complexes, in {\sl
3659: Encyclop\'edie des sciences math\'ematiques}, {\bf 1}, ed.\ J.\ Molk,
3660: 1908, 329--468.
3661:
3662: \bibitem{Cartan3} \'Elie Cartan, Le principe de dualit\'e et la
3663: th\'eorie des groupes simple et semi-simples, {\sl Bull.\ Sci.\
3664: Math.\ }{\bf 49} (1925), 361--374.
3665:
3666: \bibitem{Cayley} Arthur Cayley, On Jacobi's elliptic functions,
3667: in reply to the Rev.\ B.\ Bronwin; and on quaternions, {\sl Philos.\
3668: Mag.\ }{\bf 26} (1845), 208--211.
3669:
3670: \bibitem{Cayley2} Arthur Cayley, On Jacobi's elliptic functions, in
3671: reply to the Rev.\ B.\ Bronwin; and on quaternions (appendix only), in
3672: {\sl The Collected Mathematical Papers}, Johnson Reprint Co., New York,
3673: 1963, p.\ 127.
3674:
3675: \bibitem{CMT} Sultan Catto, Carlos J.\ Moreno and Chia-Hsiung Tze,
3676: {\sl Octonionic Structures in Physics}, to appear.
3677:
3678: \bibitem{CS} Claude Chevalley and Richard D.\ Schafer, The exceptional
3679: simple Lie algebras $\F_4$ and $\E_6$, {\sl Proc.\ Nat.\ Acad.\ Sci.\ USA}
3680: {\bf 36} (1950), 137--141.
3681:
3682: \bibitem{CD} Yvonne Choquet-Bruhat and C\'ecile DeWitt-Morette,
3683: {\sl Analysis, Manifolds and Physics}, part II, Elsevier, Amsterdam,
3684: 2000, pp.\ 263--274.
3685:
3686: \bibitem{Clifford} William K.\ Clifford, Applications of Grassmann's
3687: extensive algebra, {\sl Amer.\ Jour.\ Math.\ }{\bf 1} (1878), 350--358.
3688:
3689: \bibitem{Cohen} Frederick R.\ Cohen, On Whitehead squares, Cayley--Dickson
3690: algebras and rational functions, {\sl Bol.\ Soc.\ Mat.\ Mexicana}
3691: {\bf 37} (1992), 55--62.
3692:
3693: \bibitem{CH} E.\ Corrigan and T.\ J.\ Hollowood, The exceptional Jordan
3694: algebra and the superstring, {\sl Comm.\ Math.\ Phys.\ }{\bf 122} (1989),
3695: 393--410.
3696:
3697: \bibitem{Coxeter} Harold Scott MacDonald Coxeter, Integral Cayley
3698: numbers, {\sl Duke Math.\ Jour.\ }{\bf 13} (1946), 561--578.
3699:
3700: \bibitem{Crowe} Michael J.\ Crowe, {\sl A History of Vector Analysis},
3701: University of Notre Dame Press, Notre Dame, 1967.
3702:
3703: \bibitem{Curtis} C.\ W.\ Curtis, The four and eight square problem
3704: and division algebras, in {\sl Studies in Modern Algebra}, ed.\
3705: A.\ Albert, Prentice--Hall, Englewood Cliffs, New Jersey, 1963, pp.\ 100--125.
3706:
3707: \bibitem{Deligne} Pierre Deligne {\it et al}, eds., {\sl Quantum
3708: Fields and Strings: A Course for Mathematicians}, 2 volumes, Amer.\
3709: Math.\ Soc., Providence, Rhode Island, 1999.
3710:
3711: \bibitem{Dickson} Leonard E.\ Dickson, On quaternions and their
3712: generalization and the history of the eight square theorem, {\sl
3713: Ann.\ Math.\ }{\bf 20} (1919), 155--171.
3714:
3715: \bibitem{Dixon} Geoffrey M.\ Dixon, {\sl Division Algebras: Octonions,
3716: Quaternions, Complex Numbers and the Algebraic Design of Physics},
3717: Kluwer, Dordrecht, 1994.
3718:
3719: \bibitem{Duff} M.\ J.\ Duff, ed., {\sl The World in Eleven Dimensions:
3720: Supergravity, Supermembranes and M-Theory}, Institute of Physics
3721: Publishing, Bristol, 1999.
3722:
3723: \bibitem{EG} Noam Elkies and Benedict H.\ Gross, The exceptional
3724: cone and the Leech lattice, {\sl Internat.\ Math.\ Res.\ Notices}
3725: {\bf 14} (1996), 665--698.
3726:
3727: \bibitem{Emch} Gerard G.\ Emch, {\sl Algebraic Methods in Statistical
3728: Mechanics and Quantum Field Theory}, Wiley-Interscience, New York, 1972.
3729:
3730: \bibitem{Evans} J.\ M.\ Evans, Supersymmetric Yang--Mills theories and
3731: division algebras, {\sl Nucl.\ Phys.\ }{\bf B298} (1988), 92--108.
3732:
3733: \bibitem{Faulkner} John R.\ Faulkner, A construction of Lie algebras from
3734: a class of ternary algebras, {\sl Trans.\ Amer.\ Math.\ Soc.\ }{\bf 155}
3735: (1971), 397--408.
3736:
3737: \bibitem{FF} John R.\ Faulkner and Joseph C.\ Ferrar,
3738: Exceptional Lie algebras and related algebraic and geometric structures,
3739: {\sl Bull.\ London Math.\ Soc.\ }{\bf 9} (1977), 1--35.
3740:
3741: \bibitem{FFrenkel} Alex J.\ Feingold, Igor B.\ Frenkel, and John F.\ X.\
3742: Ries, {\sl Spinor Construction of Vertex Operator Algebras, Triality, and
3743: $\E_8^{(1)}$}, Contemp.\ Math.\ 121, Amer.\ Math.\ Soc.,
3744: Providence, Rhode Island, 1991.
3745:
3746: \bibitem{Freudenthal4} Hans Freudenthal, Oktaven, Ausnahmegruppen und
3747: Oktavengeometrie, mimeographed notes, 1951. Also available in
3748: {\sl Geom.\ Dedicata} {\bf 19} (1985), 7--63.
3749:
3750: \bibitem{Freudenthal} Hans Freudenthal, Zur ebenen Oktavengeometrie,
3751: {\sl Indag.\ Math.\ }{\bf 15} (1953), 195--200.
3752:
3753: \bibitem{Freudenthal2} Hans Freudenthal, Beziehungen der $\e_7$ und
3754: $\e_8$ zur Oktavenebene, I, II, {\sl Indag.\ Math.\ }{\bf 16} (1954),
3755: 218--230, 363--368. III, IV, {\sl Indag.\ Math.\ }{\bf 17} (1955),
3756: 151--157, 277--285. V --- IX, {\sl Indag.\ Math.\ }{\bf 21} (1959),
3757: 165--201, 447--474. X, XI, {\sl Indag.\ Math.\ }{\bf 25} (1963) 457--487.
3758:
3759: \bibitem{Freudenthal3} Hans Freudenthal, Lie groups in the foundations of
3760: geometry, {\sl Adv.\ Math.\ }{\bf 1} (1964), 145--190.
3761:
3762: \bibitem{Freudenthal5} Hans Freudenthal, Bericht \"uber die Theorie
3763: der Rosenfeldschen elliptischen Ebenen, in {\sl Raumtheorie}, Wege
3764: Der Forschung, CCLXX, Wissenschaftliche Buchgesellschaft, Darmstadt,
3765: 1978, pp. 283--286.
3766:
3767: \bibitem{Garner} Lynn E.\ Garner, {\sl An Outline of Projective Geometry},
3768: North Holland, New York, 1981.
3769:
3770: \bibitem{Graves} Robert Perceval Graves, {\sl Life of Sir William Rowan
3771: Hamilton}, 3 volumes, Arno Press, New York, 1975.
3772:
3773: \bibitem{GSW} Michael B.\ Green, John H.\ Schwarz and Edward Witten,
3774: {\sl Superstring Theory}, volume 1, Cambridge University Press,
3775: Cambridge, 1987, pp.\ 344--349.
3776:
3777: \bibitem{GKS} B.\ Grossman, T.\ E.\ Kephart, and James D.\ Stasheff,
3778: Solutions to Yang-Mills field equations in eight dimensions and the
3779: last Hopf map, {\sl Comm.\ Math.\ Phys.\ }{\bf 96} (1984), 431--437.
3780:
3781: \bibitem{Gunaydin} Murat G\"unaydin, Generalized conformal and
3782: superconformal group actions and Jordan algebras,
3783: {\sl Mod.\ Phys.\ Lett.\ }{\bf A8} (1993), 1407--1416.
3784:
3785: \bibitem{GKN} Murat G\"unaydin, Kilian Koepsell, and Hermann Nicolai,
3786: Conformal and quasiconformal realizations of exceptional Lie groups,
3787: preprint available as hep-th/0008063.
3788:
3789: \bibitem{GPR} Murat G\"unaydin, C.\ Piron and H.\ Ruegg, Moufang plane
3790: and octonionic quantum mechanics, {\sl Comm.\ Math.\ Phys.\ }{\bf 61}
3791: (1978), 69--85.
3792:
3793: \bibitem{GT} Feza G\"ursey and Chia-Hsiung Tze, {\sl On the Role of Division,
3794: Jordan, and Related Algebras in Particle Physics}, World Scientific,
3795: Singapore, 1996.
3796:
3797: \bibitem{Hamilton} William Rowan Hamilton, Four and eight square
3798: theorems, in Appendix 3 of {\sl The Mathematical Papers of William
3799: Rowan Hamilton}, vol.\ 3, eds.\ H. Halberstam and R. E. Ingram,
3800: Cambridge University Press, Cambridge, 1967, pp.\ 648--656.
3801:
3802: \bibitem{Hankins} Thomas L.\ Hankins, {\sl Sir William Rowan Hamilton},
3803: John Hopkins University Press, Baltimore, 1980.
3804:
3805: \bibitem{Harvey} F.\ Reese Harvey, {\sl Spinors and Calibrations},
3806: Academic Press, San Diego, 1990.
3807:
3808: \bibitem{Hurwitz} Adolf Hurwitz, \"Uber die Composition der quadratischen
3809: Formen von beliebig vielen Variabeln, {\sl Nachr.\ Ges.\ Wiss.\ G\"ottingen}
3810: (1898) 309--316.
3811:
3812: \bibitem{Husemoller} Dale Husemoller, {\sl Fibre Bundles}, Springer, Berlin,
3813: 1994.
3814:
3815: \bibitem{Jordan} Pascual Jordan, \"Uber eine Klasse nichtassociativer
3816: hyperkomplexer Algebren, {\sl Nachr.\ Ges.\ Wiss.\ G\"ottingen} (1932),
3817: 569--575.
3818:
3819: \bibitem{Jordan2} Pascual Jordan, \"Uber eine nicht-desarguessche
3820: ebene projektive Geometrie, {\sl Abh.\ Math.\ Sem.\ Hamburg}
3821: {\bf 16} (1949), 74--76.
3822:
3823: \bibitem{JNW} Pascual Jordan, John von Neumann, Eugene Wigner,
3824: On an algebraic generalization of the quantum mechanical formalism,
3825: {\sl Ann.\ Math.\ }{\bf 35} (1934), 29--64.
3826:
3827: \bibitem{Joyce} Dominic Joyce, {\sl Compact Manifolds with Special
3828: Holonomy}, Oxford U.\ Press, Oxford, 2000.
3829:
3830: \bibitem{KS} I.\ L.\ Kantor and A.\ S.\ Solodovnikov, {\sl Hypercomplex
3831: Numbers --- an Elementary Introduction to Algebras,} Springer, Berlin, 1989.
3832:
3833: \bibitem{Kervaire} Michel Kervaire, Non-parallelizability of the $n$
3834: sphere for $n > 7$, {\sl Proc.\ Nat.\ Acad.\ Sci.\ USA} {\bf 44} (1958),
3835: 280--283.
3836:
3837: \bibitem{Killing} Wilhelm Killing, Die Zusammensetzung der stetigen
3838: endlichen Transformationsgruppen I, {\sl Math.\ Ann.\ } {\bf 31} (1888),
3839: 252--290. II, {\bf 33} (1889) 1--48. III, {\bf 34} (1889), 57--122.
3840: IV {\bf 36} (1890), 161--189.
3841:
3842: \bibitem{Kuperberg} Greg Kuperberg, Spiders for rank 2 Lie algebras,
3843: {\sl Comm.\ Math.\ Phys.\ }{\bf 180} (1996), 109--151.
3844:
3845: \bibitem{KT} T.\ Kugo and P.--K.\ Townsend, Supersymmetry and the
3846: division algebras, {\sl Nucl.\ Phys.\ }{\bf B221} (1983), 357--380.
3847:
3848: \bibitem{LM} J.\ M.\ Landsberg and L.\ Manivel: The projective geometry of
3849: Freudenthal's magic square, preprint available as math.AG/9908039.
3850:
3851: \bibitem{LPS} Jaak Lohmus, Eugene Paal, and Leo Sorgsepp,
3852: {\sl Nonassociative Algebras in Physics}, Hadronic Press, Palm
3853: Harbor, Florida, 1994.
3854:
3855: \bibitem{MD} Corinne A.\ Manogue and Tevian Dray, Octonionic
3856: M\"obius transformations, {\sl Mod.\ Phys.\ Lett.\ }{\bf A14} (1999),
3857: 1243--1256.
3858:
3859: \bibitem{MS} Corinne A.\ Manogue and J\"org Schray, Finite Lorentz
3860: transformations, automorphisms, and division algebras, {\sl Jour.\
3861: Math.\ Phys.\ }{\bf 34} (1993), 3746--3767.
3862:
3863: \bibitem{MS2} Corinne A.\ Manogue and J\"org Schray, Octonionic
3864: representations of Clifford algebras and triality, {\sl Found.\
3865: Phys.\ }{\bf 26} (1996), 17--70.
3866:
3867: \bibitem{McCrimmon} Kevin McCrimmon, Jordan algebras and their applications,
3868: {\sl Bull.\ Amer.\ Math.\ Soc.\ } {\bf 84} (1978), 612--627.
3869:
3870: \bibitem{Meyberg} K.\ Meyberg, Eine Theorie der Freudenthalschen
3871: Tripelsysteme, I, II, {\sl Indag.\ Math.\ }{\bf 30} (1968),
3872: 162--190.
3873:
3874: \bibitem{Moreno} R.\ Guillermo Moreno, The zero divisors of the
3875: Cayley--Dickson algebras over the real numbers, preprint available
3876: at q-alg/9710013.
3877:
3878: \bibitem{Moufang} Ruth Moufang, Alternativk\"orper und der Satz vom
3879: vollst\"andigen Vierseit, {\sl Abh.\ Math.\ Sem.\ Hamburg} {\bf 9}
3880: (1933), 207--222.
3881:
3882: \bibitem{OV} A.\ L. Onishchik and E.\ B.\ Vinberg, eds., {\sl Lie Groups
3883: and Lie Algebras III}, Springer, Berlin, 1991, pp.\ 167--178.
3884:
3885: \bibitem{Okubo} Susumu Okubo, {\sl Introduction to Octonion and Other
3886: Non-Associative Algebras in Physics}, Cambridge University Press,
3887: Cambridge, 1995.
3888:
3889: \bibitem{PR} Roger Penrose and Wolfgang Rindler, {\sl Spinors and
3890: Space-Time}, 2 volumes, Cambridge U.\ Press, Cambridge, 1985-86.
3891:
3892: \bibitem{Porteous} Ian R.\ Porteous, {\sl Topological Geometry},
3893: Cambridge U.\ Press, 1981.
3894:
3895: \bibitem{Rosenfeld1} Boris A.\ Rosenfeld, Geometrical interpretation of
3896: the compact simple Lie groups of the class $\E$ (Russian), {\sl
3897: Dokl.\ Akad.\ Nauk.\ SSSR} (1956) {\bf 106}, 600-603.
3898:
3899: \bibitem{Rosenfeld} Boris A.\ Rosenfeld, {\sl Geometry of Lie Groups},
3900: Kluwer, Dordrecht, 1997.
3901:
3902: \bibitem{Rost} Markus Rost, On the dimension of a composition algebra,
3903: {\sl Doc.\ Math.\ } {\bf 1} (1996), 209--214.
3904:
3905: \bibitem{Salzmann} Helmut Salzmann {\it et al}, {\sl Compact Projective
3906: Planes: With an Introduction to Octonion Geometry}, de Gruyter, Berlin,
3907: 1995.
3908:
3909: \bibitem{Schafer1} Richard D.\ Schafer, On algebras formed by the
3910: Cayley--Dickson process, {\sl Amer.\ Jour.\ of Math.\ }{\bf 76} (1954)
3911: 435--446.
3912:
3913: \bibitem{Schafer} Richard D.\ Schafer, {\sl Introduction to Non-Associative
3914: Algebras}, Dover, New York, 1995.
3915:
3916: \bibitem{Schray} J\"org Schray, {\sl Octonions and Supersymmetry},
3917: Ph.D.\ thesis, Department of Physics, Oregon State University, 1994.
3918:
3919: \bibitem{Sierra} G.\ Sierra, An application of the theories of Jordan
3920: algebras and Freudenthal triple systems to particles and strings,
3921: {\sl Class.\ Quant.\ Grav.\ }{\bf 4} (1987), 227--236.
3922:
3923: \bibitem{Springer} Tonny A.\ Springer, The projective octave plane, I--II,
3924: {\sl Indag.\ Math.\ }{\bf 22} (1960), 74--101.
3925:
3926: \bibitem{Springer2} Tonny A.\ Springer, Characterization of a class of
3927: cubic forms, {\sl Indag.\ Math.\ }{\bf 24} (1962), 259--265.
3928:
3929: \bibitem{Springer3} Tonny A.\ Springer, On the geometric algebra of
3930: the octave planes, {\sl Indag.\ Math.\ }{\bf 24} (1962), 451--468.
3931:
3932: \bibitem{SV} Tonny A.\ Springer and Ferdinand D.\ Veldkamp,
3933: {\sl Octonions, Jordan Algebras and Exceptional Groups}, Springer,
3934: Berlin, 2000.
3935:
3936: \bibitem{Stevenson} Frederick W. Stevenson, {\sl Projective Planes}, W.\ H.\
3937: Freeman and Company, San Francisco, 1972.
3938:
3939: \bibitem{Sudbery2} Anthony Sudbery, Octonionic description of exceptional
3940: Lie superalgebras, {\sl Jour.\ Math.\ Phys.\ }{\bf 24} (1983), 1986--1988.
3941:
3942: \bibitem{Sudbery} Anthony Sudbery, Division algebras, (pseudo)orthogonal
3943: groups and spinors, {\sl Jour.\ Phys.\ }{\bf A17} (1984), 939--955.
3944:
3945: \bibitem{Tits} Jacques Tits, Le plan projectif des octaves et les groupes
3946: de Lie exceptionnels, {\sl Bull.\ Acad.\ Roy.\ Belg.\ Sci.\ }{\bf 39} (1953),
3947: 309--329.
3948:
3949: \bibitem{Tits2} Jacques Tits, Le plan projectif des octaves et les groupes
3950: exceptionnels $\E_6$ et $\E_7$, {\sl Bull.\ Acad.\ Roy.\ Belg.\
3951: Sci.\ }{\bf 40} (1954), 29--40.
3952:
3953: \bibitem{Tits3} Jacques Tits, Alg\`ebres alternatives, alg\`ebres de
3954: Jordan et alg\`ebres de Lie exceptionnelles, {\sl Indag.\ Math.\ }{\bf 28}
3955: (1966), 223--237.
3956:
3957: \bibitem{Tits4} Jacques Tits, {\sl Buildings of Spherical Type
3958: and Finite BN-Pairs, } Lecture Notes in Mathematics, Vol.\ 386,
3959: Springer, Berlin, 1974.
3960:
3961: \bibitem{Varadarajan} V.\ S.\ Varadarajan, {\sl Geometry of Quantum
3962: Theory}, Springer-Verlag, Berlin, 1985.
3963:
3964: \bibitem{Vinberg} E.\ B.\ Vinberg, A construction of exceptional
3965: simple Lie groups (Russian), {\sl Tr.\ Semin.\ Vektorn.\ Tensorn.\
3966: Anal.\ }{\bf 13} (1966), 7--9.
3967:
3968: \bibitem{Zorn} Max Zorn, Theorie der alternativen Ringe,
3969: {\sl Abh.\ Math.\ Sem.\ Univ.\ Hamburg} {\bf 8} (1930), 123--147.
3970:
3971: \bibitem{Zorn2} Max Zorn, Alternativk\"orper und quadratische Systeme,
3972: {\sl Abh.\ Math.\ Sem.\ Univ.\ Hamburg} {\bf 9} (1933), 395--402.
3973:
3974: \end{thebibliography}
3975: \end{document}
3976:
3977:
3978:
3979:
3980: