math0411104/main.tex
1: \documentclass[11pt]{article}
2: 
3: \textwidth=6.5in
4: \textheight=8.9in
5: 
6: \oddsidemargin=0cm
7: \topmargin=-1cm
8: 
9: \usepackage[all]{xy}
10: 
11: \usepackage{mathrsfs} %
12: \usepackage{bbm}
13: \usepackage{enumerate} %
14: 
15: \input{commands.tex}
16: 
17: \begin{document}
18: \setcounter{table}{1}
19: 
20: \begin{center}
21: {
22: \Large\bf
23: 
24: Jordan algebras, exceptional groups,\\[2mm]
25: and higher composition laws
26: }
27: 
28: \vspace{\baselineskip}
29: 
30: {\large Sergei Krutelevich}
31: 
32: \vspace{\baselineskip}
33: 
34: \end{center}
35: 
36: \begin{abstract}
37: 
38: We consider an integral version of the Freudenthal construction relating
39: Jordan algebras and exceptional algebraic groups. We show how this
40: construction is related to higher composition laws of M.~Bhargava in
41: number theory \cite{bh1}.
42: 
43: We propose an algorithmic approach to studying orbit spaces of
44: groups underlying higher composition laws.
45: Using this method we discover two new
46: examples of spaces sharing similar properties, and indicate
47: several more examples of spaces where such composition laws may be
48: introduced.
49: \end{abstract}
50: 
51: \tableofcontents
52: 
53: 
54: \section{Introduction}
55: \subsection{Integral representations of exceptional groups}
56: \label{i:fc}
57: 
58: It is a well-known fact in number theory that there is a one-to-one
59: correspondence between
60: the set of $\SL_2(\zet)$-equivalence classes of integral binary quadratic
61: forms and the set of (narrow) ideal classes in quadratic rings. About two
62: hundred years ago Gauss discovered the law of composition of binary
63: quadratic forms, which turns
64: the set of equivalence classes of primitive forms of a given discriminant
65: $D$ into a group isomorphic to the ideal class group of the quadratic ring
66: of discriminant $D$. This correspondence is a very important tool
67: for doing computations in the ideal class group.
68: 
69: A few years ago M.~Bhargava discovered several more examples of the same
70: kind, which he referred to as higher composition laws. More precisely, he
71: showed that there are other examples of linear groups $G_\zet$
72: and their integral representations $V_\zet$ such that $G_\zet$-orbits in $V_\zet$
73: are in one-to-one correspondence with ideal classes in the rings of
74: integers in number fields, see \cite{bh-phd, bhau, bh1}.
75: M.~Bhargava also noted a surprising connection between spaces underlying higher
76: composition and exceptional Lie groups.
77: 
78: In the present paper we investigate this connection with exceptional
79: groups. We show that there is a natural construction which assigns
80: a cubic Jordan algebra to every space underlying higher composition laws
81: associated with quadratic rings. We study the appropriate orbit spaces using an
82: algorithmic procedure reminiscent of the Gaussian elimination algorithm
83: for the usual integral matrices. This approach allows us to provide two
84: new examples of spaces sharing similar properties, and indicate
85: several more examples of spaces where such composition laws may be
86: introduced.
87: 
88: \medskip
89: Our interest in the study of integral representations of exceptional groups was
90: originally motivated by a question on a standard form of a charge vector in a certain
91: physical model, which appeared in a paper on BPS black holes in string theory by
92: H.~Maldacena, G.~Moore, A.~Strominger \cite{mms}.
93: This question is equivalent to the question on a normal form of a vector
94: in the $27$-dimensional integral representation of the split group of type
95: $E_6$.
96: Such a representation may be constructed via the exceptional Jordan
97: algebra of $3\times 3$ Hermitian matrices over the algebra of octonions.
98: The answer, generalizing the classical results on the Smith normal
99: form for regular integral matrices, was given in our earlier paper
100: \cite{k}, see also Theorem~$\ref{np-orbits}$ below.
101: 
102: It was suggested by G.~Moore that another interesting space to look
103: at in this context is the $56$-dimensional integral representation of the
104: split group of type $E_7$. This representation may be constructed using the
105: $27$-dimensional exceptional Jordan algebra. This procedure is known
106: as the Freudenthal construction \cite{b, fr14},
107: and it can be applied to other cubic Jordan algebras.
108: Given a cubic Jordan algebra~$\J$, the Freudenthal  construction
109: produces a semisimple algebraic group and an irreducible module over it.
110: We will denote the group obtained by $\gr$ and the module by $\mj$
111: (see Subsection~\ref{ss:fc} for details).
112: Moreover, this module is equipped with a quartic form and a symplectic
113: form, which are invariant under the action of the group.
114: When $\J$ is the $27$-dimensional exceptional Jordan
115: algebra, the group $\gr$ is $E_7$ and $\mj$ is its $56$-dimensional
116: representation.
117: 
118: A natural
119: way to construct a Jordan algebra with a cubic form is to consider the
120: space of $3\times 3$-Hermitian matrices over a composition algebra
121: over a field $F$, the cubic form being the determinant of such matrices
122: (see Example~$\ref{ex:jcubic}$ for details). When $F$ is algebraically
123: closed, there are four composition algebras, and they produce four cubic
124: Jordan algebras of dimension $6, 9, 15, 27$.
125: 
126: The application of the Freudenthal construction yields a module of
127: dimension $14, 20, 32, 56$ for a certain simple algebraic group
128: of type $C_3, A_5, D_6, E_7$.
129: This module is in fact a prehomogeneous vector space for this
130: algebraic group, and it is natural to study orbits under the action of
131: the group.
132: In the case $F=\com$, the classification of orbits of the one-dimensional
133: subspaces\footnote{We avoid
134: using the term ``projective" here, since it is used in completely
135: different sense in the rest of the paper.}
136: arising this way was obtained by J.-L.~Clerc~\cite{clerc}.
137: 
138: Lie algebras of the algebraic groups of type $C_3, A_5, D_6, E_7$ appear
139: in the third row of the Freudenthal-Tits magic square. The groups,
140: produced by the Freudenthal construction, also
141: appear in the last row of the
142: ``magic triangle" of subgroups associated to the exceptional
143: series of P.~Deligne, see \cite{dg}. In addition, the representation $\m$
144: of the group $\gr$ is the {\em preferred} representation of this group
145: in the sense of \cite{dg}.
146: 
147: In the present paper we consider an integral version of the Freudenthal
148: construction.
149: In this case the quartic form (the {\em norm}) takes on integer values.
150: We study integral forms of the split groups of type $A_5, D_6, E_7$
151: determined by the Freudenthal construction, and
152: we obtain the following result
153: on the structure of integral orbits in the $\zet$-modules of
154: dimension $20, 32, 56$.
155: \medskip
156: 
157: \hspace{-\parindent}{\bf Theorem~$\ref{thm:main}$}
158: {\em
159: Let $(G_\zet, \mz)$ be one the following pairs
160: $$
161: \Bigl(\SL_6(\zet), \wedge^3(\zet^6)\Bigr),\quad
162: \Bigl(D_6(\zet), \mbox{\rm half-spin}_\zet\Bigr),\quad
163: \Bigl(E_7(\zet), V(\omega_7)_\zet\Bigr),
164: $$
165: Then
166: \begin{itemize}
167: 
168: \item
169: The $G_\zet$-invariant quartic form (the norm) on the module $\mz$
170: has values congruent to $0$ or $1\, (\!\mod 4)$.
171: 
172: \item
173: Let $n$ be an integer $\equiv 0$ or $1\, (\!\mod 4)$.
174: The group $G_\zet$ acts
175: transitively on the set of {\em projective} elements of norm $n$.
176: 
177: \item
178: If $n$ is a fundamental discriminant\footnote{
179: An integer $n$ is called a fundamental discriminant if $n$ is squarefree
180: and $\equiv 1 (\mod 4)$
181: or $n=4k$, where $k$ is a squarefree integer that is $\equiv 2$ or $3 (\mod 4)$.
182: The result stated in the theorem applies in the case $n=1$.
183: },
184: then every element of norm $n$ is projective, and hence in this case
185: $G_\zet$ acts transitively on the set of elements of norm $n$.
186: \end{itemize}
187: }
188: 
189: 
190: The assertion of this theorem for $\SL_6(\zet)$-orbits
191: on $\wedge^3(\zet^6)$ (when $n\ne 0$) was proved by M.~Bhargava in
192: \cite[Theorem 7]{bh1}, using the correspondence with (narrow) ideal
193: classes in quadratic orders.
194: 
195: Our approach, based on the Freudenthal construction, allows us to treat
196: all values of the norm (including $n=0$) uniformly.
197: Our statement of Theorem~$\ref{thm:main}$
198: was motivated by \cite{bh1}, but the proof
199: presented here is completely independent.
200: It is algorithmic in nature,
201: and may be thought of as a more sophisticated version of
202: the Gaussian algorithm bringing an integer matrix to the Smith
203: normal form by elementary row and column transformations.
204: The extended version of Theorem~$\ref{thm:main}$ for the degenerate orbits
205: (corresponding to the case $n=0$)
206: is given in Theorem~$\ref{thm2}$.
207: 
208: 
209: The concept of a projective element was introduced in \cite{bh1}.
210: The idea is that these elements are mapped to invertible ideal classes under
211: Bhargava's correspondence. In the case of $\wedge^3(\zet^6)$ they were
212: defined via their $\SL_6(\zet)$ orbit representatives
213: (cf. Definition~$\ref{def:proja}$(a,b)).
214: It follows from our considerations that projective elements have a very convenient
215: description in terms of the partial derivatives of the quartic invariant
216: of the module (see Corollary~$\ref{cor:nabla}$). This assertion is proved for the spaces
217: $\mz$ as
218: in Theorem~$\ref{thm:main}$, but it remains valid in other spaces associated to ideal
219: classes in quadratic orders. (see Subsection~\ref{i:hcl}).
220: 
221: %
222: We also indicate a link between the Freudenthal construction and the
223: original Gauss's composition of quadratic forms in the Appendix at the end
224: of the paper.
225: 
226: \medskip
227: As a by-product of our considerations we obtain the classification of
228: orbits in the four ``Freudenthal" modules in the case of a field. Similar
229: results were obtained in \cite{clerc} and \cite[Section 5.3]{lm}
230: for orbits of one-dimensional subspaces in the case of complex numbers.
231: Our
232: techniques however are purely algebraic, which allows us to extend those
233: results to orbits of {\em elements} over arbitrary field of char$\ne 2,3$.
234: 
235: \medskip
236: \hspace{-\parindent}{\bf Theorem~$\ref{thm1}$}
237: {\em
238: \begin{enumerate}
239: \item[\rm (i)]
240: {
241: Let $F$ be a field of char$\ne 2,3$, let $\calg$ be the split composition algebra
242: $\bin, \quat, \oct$ of dimension $2,4,8$ over $F$, and let $\J=\h\calg$.
243: Let $(G, \m)$ be the pair (group, module) produced from $\J$
244: by the Freudenthal construction.
245: 
246: Then
247: 
248: \begin{itemize}
249: \item
250: There exists a $G$-invariant quartic form (the norm) on the module $\m$.
251: 
252: \item
253: The group $G$ acts
254: transitively on the sets of elements of rank $1,2$, and $3$ in the module
255: $\m$.
256: 
257: \item
258: In the case of rank\/ $4$ the group $G$ acts transitively on the set of
259: elements of a given norm $k$, for any $k\in F$, $k\ne 0$.
260: \end{itemize}
261: All these orbits are distinct, and the union of these orbits and $\{0\}$
262: is the whole module~$\m$.
263: }
264: 
265: \item[\rm (ii)]
266: If in addition every element of $F$ is a square,
267: then the same results apply to the pair $(G, \m)$ obtained
268: from the Jordan algebra $\J=\h F$.
269: \end{enumerate}
270: This construction yields the classification of orbits
271: of the irreducible representations
272: of simple algebraic groups listed in the following table
273: $$
274: \begin{array}{|c|c|c|}
275: \hline
276: \qquad \J \qquad       &  \mbox{Type of }G & \mbox{Highest weight of }\ \m\\
277: \hline
278: \h{\F}    &  C_3 & \omega_3\\
279: \h{\bin}  &  A_5 & \omega_3\\
280: \h{\quat} &  D_6 & \omega_5\ {\rm or}\ \omega_6\\
281: \h{\oct}  &  E_7 & \omega_7\\
282: \hline
283: \end{array}
284: $$
285: }
286: 
287: 
288: 
289: 
290: 
291: \subsection{The Freudenthal construction and higher composition laws}
292: \label{i:hcl}
293: 
294: M.~Bhargava showed that Gauss's composition law
295: is one in a series of at least 14 examples of the same kind (higher
296: composition laws) \cite{bh-phd, bhau, bh1}.
297: There is a certain integral linear group $G_\zet$ and
298: a module $V_\zet$ over it in each of his examples
299: such that $G_\zet$-orbits in $V_\zet$ can be described
300: in terms of ideal classes of orders in a number field.
301: Spaces underlying higher composition laws are closely related to
302: prehomogeneous vector spaces classified by M.~Sato and T.~Kimura \cite{SK}. In
303: particular, each of them is equipped with a polynomial, which is invariant
304: under the action of the appropriate group.
305: 
306: An examination of the table of higher composition laws \cite[Table 1]{bhau}
307: shows that each
308: space associated to a quadratic ring (except Gauss's composition) has a
309: polynomial invariant of degree four. A more detailed analysis suggests
310: that for each pair $(G_\zet, \ V_\zet)$ associated to a quadratic ring,
311: there exists a cubic Jordan algebra $\J_\zet$, such that
312: $(G_\zet, \ V_\zet)$ is essentially the pair produced by the Freudenthal
313: construction. These observations are summarized in Table 1 below.
314: 
315: %
316: \bigskip
317: \caption{The Freudenthal construction and higher composition laws}
318: \nopagebreak
319: $$
320: \begin{array}{|c|c|c|c|c|c|c|}
321: \hline
322: \# &
323: \J &
324: %
325: \phantom{\Bigl(}
326: {\dim}\ \J&
327: %
328: {\rm Group}\ \gr &
329: {\rm Rep.\ \m(\J)} &
330: {\dim}\ \m(\J)  &
331: (\lie, \gamma) \\
332: \hline
333: \hline
334: 
335: 
336: 1 & F\phantom{\Bigl(}   & 1  & \SL_2 & {\rm Sym}^3 V_2     & 4  & G_2, \alpha_2 \\
337: 2 & F\oplus F   & 2  & (\SL_2)^2 & V_2\tensor {\rm Sym}^2 V_2     & 6  & B_3, \alpha_2 \\
338: 3 & \h{0}=F\oplus F\oplus F\phantom{\Bigl(}
339: & 3 & (\SL_2)^3& V_2\otimes V_2\otimes V_2 & 8  & D_4, \alpha_2 \\
340: %
341: 4 & F\oplus Q_4& 5 & \SL_2\times \SL_{4} & V_2\tensor \wedge^2 V_4     & 12  & D_5, \alpha_2\\
342: 5 & \h\bin  & 9  & \SL_6 & V(\omega _3)     & 20  & E_6, \alpha_2 \\
343: \hline
344: \hline
345: 6 & \h\quat & 15 & D_6 & \mbox{half-spin} & 32  & E_7, \alpha_1  \\
346: 7 & \h\oct  & 27 & E_7 & {\rm minuscule}     & 56  & E_8, \alpha_8  \\
347: \hline
348: \hline
349: 8 & F\oplus Q_n, n\ge 3 & 1+n & {\rm SL_2\times SO}_{n+2} & V_2\tensor V(\omega_1)
350: & 2n+4  &{\mathfrak s\mathfrak o}_{n+6}, \alpha_2  \\
351: %
352: 9 & \h{F}   & 6  & C_3 & V(\omega _3)     & 14  & F_4, \alpha_1 \\
353: \hline
354: \end{array}
355: $$
356: %
357: 
358: \renewcommand{\abstractname}{Summary of Table 1}
359: \nopagebreak
360: \begin{abstract}
361: {\bf Headers:} the table lists a semisimple
362: Jordan algebra $\J$ with a cubic form, its
363: dimension, the group $\gr$ (up to a finite subgroup or finite covering)
364: and its module $\mj$ produced by the Freudenthal
365: construction, and the dimension of $\mj$.
366: The last column lists a certain simple Lie
367: algebra $\lie$ and its simple root $\gamma$ associated to $\gr, \mj$ (see
368: subsection \ref{lieconn}).
369: 
370: {\bf Notation:} in the second column, $\h{\calg}$ stands for the Jordan
371: algebra of $3\times 3$-Hermitian matrices over the composition algebra
372: $C$; and $Q_n$ stands for the simple Jordan algebra of a quadratic form
373: (details are found in Subsection~$\ref{ss:ex}$). $V_n$ denotes the standard $n$-dimensional
374: module for $\SL_n$, and $V(\omega_i)$ denotes the simple module with
375: highest weight $\omega_i$ over the appropriate group.
376: Note that rows $2,3,4$ are special cases of row $8$ with $n=1,2,4$.
377: 
378: \medskip
379: 
380: There is a natural integral structure in each Jordan algebra $\J$ above. It
381: induces an integral structure in $\mj$ and $\gr$.
382: The first five rows of the table contain the integral group
383: and a module, which appear in the list of
384: higher composition laws
385: (see \cite[Table 1]{bhau} or \cite{bh1}).
386: 
387: Rows $6$ and $7$ do not appear in the list of higher composition laws \cite{bhau}.
388: However, Theorem~$\ref{thm:main}$  implies that the structure of orbits
389: of the projective elements is
390: the same as in row $\#5$. These are the first examples
391: of groups, more complex than a direct product of $\SL$'s,
392: acting in the spaces underlying higher composition laws.
393: 
394: Finally, the Freudenthal construction produces two more examples (rows $8$
395: and $9$). We conjecture that for $\#8$ the orbit structure can be
396: described in terms of known examples of small dimension. And the row $9$
397: could possibly produce a new example of a space with a composition law.
398: \end{abstract}
399: 
400: It would be interesting to describe orbits of the group $D_6(\zet)$ in the
401: $32$-dimensional module and of the group $E_7(\zet)$ in the
402: $56$-dimensional module in terms of the ideal classes of the appropriate
403: rings in a way similar to how this is done in \cite{bh1} for $\SL_6(\zet)$
404: orbits in $\wedge^3 \zet^6$.
405: 
406: Finally, we note that our algorithmic approach allows one to develop,
407: at least in some cases, a reduction theory similar to that for binary
408: quadratic forms.
409: 
410: \subsection{Connection with Lie algebras}
411: \label{lieconn}
412: 
413: Another interesting feature of the spaces underlying higher composition
414: laws is their remarkable connection with exceptional groups.
415: M. Bhargava showed that each of his pairs (group, space) may be described
416: in terms of the Levi decomposition of parabolic subgroups
417: of exceptional groups.
418: 
419: This connection can be stated more precisely for the Freudenthal
420: construction, and hence for orbit spaces associated to quadratic rings.
421: Here we present a more algebraic (as opposed to Lie group) realization of
422: this construction.
423: Namely, we consider a simple Lie algebra $\lie$ with a $5$-grading associated
424: to the minimal nilpotent orbit in $\lie$. From this setup it
425: is possible to extract an algebraic group $G$ and a $G$-module $M$ with
426: $G$-invariant quartic and skew-symmetric bilinear form, such that $(G,M)$
427: will produce all pairs arising from the Freudenthal construction.
428: 
429: The detailed description of this procedure is given below. Most of the
430: material of this subsection has appeared earlier elsewhere. We used
431: \cite[Section~2]{gw} and \cite{clerc} as the references.
432: For simplicity we assume in this subsection that the ground field $F$ is
433: the field of complex numbers $\com$.
434: 
435: \medskip
436: Let $\lie$ be a complex simple Lie algebra, and we assume that $\lie\ne
437: A_n, C_n$.
438: Let $\Phi$ be its root system with respect to a Cartan subalgebra
439: $\cart$; let $\Phi^+$ be a set of positive roots,
440: and let $\Delta$ be its collection of simple roots.
441: Let $\beta$ be the
442: highest root of $\Phi ^+$. We normalize the inner product
443: $\langle\cdot,\cdot\rangle$ on the real span of roots by requiring
444: $\langle\beta,\beta\rangle=2$. Then it follows that for any
445: $\alpha\in\Phi^+$, $\langle\alpha,\beta\rangle=0,1,2$; and
446: $\langle\alpha,\beta\rangle=2$ iff $\alpha=\beta$.
447: 
448: 
449: We consider the grading
450: \begin{equation}\label{grading}
451: \lie=\lie_{-2}\oplus \lie_{-1}\oplus\lie_{0}\oplus\lie_{1}\oplus\lie_{2},
452: \end{equation}
453: 
454: given by
455: $$
456: \lie_0=\cart\oplus\bigoplus_{\stackrel{\alpha\in\Phi}
457: {\langle\alpha,\beta\rangle=0}}
458: \lie_\alpha,\qquad
459: \lie_k=\bigoplus_{\stackrel{\alpha\in\Phi}
460: {\langle\alpha,\beta\rangle=k}}
461: \lie_\alpha\quad
462: k=\pm 1, \pm 2.
463: $$
464: 
465: In particular, we have $\lie_2=\lie_\beta,\
466: \lie_{-2}=\lie_{\beta}$.
467: 
468: We will also need to consider the extended Dynkin diagram of $\lie$
469: (with the additional vertex corresponding to the root $-\beta$).
470: We let $\gamma$ denote the unique simple
471: root whose vertex is connected to the
472: extended vertex $-\beta$. The standard references for root systems
473: and Dynkin diagrams is \cite{b1}.
474: The appropriate diagrams are also provided in~\cite{gw}.
475: 
476: The grading (\ref{grading}) has the following property:
477: the root space  $\lie_\alpha$ is contained in $\lie_k$
478: if and only if
479: the decomposition of $\alpha$ in the basis of simple roots $\Delta$
480: has coefficient $k$ at the root $\gamma$
481: $(k=0,\pm 1, \pm 2)$.
482: 
483: There exists a subalgebra $\ml\subset\lie_0$
484: such that $\lie_0=[\lie_{-2}, \lie_2]\oplus \ml$.
485: $\ml$ is a semisimple Lie
486: algebra, and its Dynkin diagram can be obtained from the Dynkin diagram of
487: $\lie$ by removing the vertex corresponding to the simple root~$\gamma$.
488: 
489: We let $M_\com$ be a complex connected Lie group, whose Lie algebra is
490: $\ml$. The group $M_\com$ and the Lie algebra $\ml$ act on the space
491: $\lie_1$. The resulting pairs $(M_\com, \lie_1)$ are tabulated in
492: \cite[Table~2.6]{gw}. An examination of that table shows that when
493: $\lie\ne A_n, C_n$, pairs $(M_\com, \lie_1)$ are the same (up to a finite
494: covering) as the pairs $(\gr, \m)$ arising from the Freudenthal construction
495: and listed in Table~1 above. In particular, the last column of that table
496: lists the simple Lie algebra $\lie$ and the simple root $\gamma$ such that
497: the ``Freudenthal" pair $(\gr,\m)$ is obtained from the construction
498: described above.
499: 
500: \medskip
501: Next, we are going show how one can describe the ``Freudenthal" quartic and
502: symplectic forms on $\lie_1\ (=\m)$ invariant with respect to $M_\com$.
503: 
504: The Lie bracket on $\lie$ induces a map $\wedge^2 \lie_1\to \lie_\beta$
505: which is $\lie_0$-equivariant. Since $\lie_\beta$ is one-dimensional, this
506: map defines a symplectic bilinear form $\{\cdot, \cdot\}$ on $\lie_1$.
507: In addition, we have $[\ml, \lie_2]=0$, and this implies that
508: $\{\cdot, \cdot\}$ is invariant with respect to $\ml$ (and hence $M_\com$).
509: 
510: Finally, for $X\in \lie_1$ we consider the map
511: $$
512: ({\adlie\,}X)^4: \ \lie_{-2}\ \longrightarrow\ \lie_2.
513: $$
514: 
515: For fixed root vectors $x_{\pm \beta}\in \lie_{\pm 2}$ we have
516: \begin{equation}\label{qmap}
517: ({\adlie\,}X)^4: \ x_{-\beta}\ \mto \ P(X)\,x_\beta \qquad
518: \mbox{for some }P(X)\in\com.
519: \end{equation}
520: 
521: The function $P:\lie_1\to\com$ defined by (\ref{qmap}) is in fact non-zero
522: homogeneous
523: polynomial function of degree four, invariant with respect to $\ml$ and
524: $M_\com$ \cite[Proposition~3.1]{clerc}. Since $\ml$ acts irreducibly on
525: $\lie_1$, we can normalize the polynomial $P$ so that it coincides with
526: the Freudenthal quartic form.
527: 
528: \medskip
529: With a little more work one can even extract the cubic Jordan algebra $\J$
530: from $\ml$. Namely, in each such $\ml$ there will be a simple root
531: $\alpha_0$, which will induce a $3$-grading on $\ml$ similar to the one we had
532: for $\lie$:
533: \begin{equation}\label{grading2}
534: \ml=\ml_{-1}\oplus\ml_{0}\oplus\ml_{1}.
535: \end{equation}
536: 
537: Then the space $\ml_1$ will have a structure of a Jordan algebra.
538: 
539: One way
540: to prove it is by considering a compact real form of (\ref{grading2}) and
541: noticing that it produces a Hermitian symmetric space, which is of tube
542: type (see \cite[Section 4]{clerc} for details).
543: 
544: The structure of a Jordan algebra on $\ml_1$
545: can be described in purely algebraic terms,
546: see \cite{tits}, or \cite[Lemma~$4$]{kac}, which is a more accessible
547: reference.
548: 
549: 
550: \medskip
551: To summarize, given a cubic Jordan algebra $\J$, the Freudenthal
552: construction produces a group $\gr$ and its module $\m$ with
553: $\gr$-invariant quartic and symplectic forms. Also there exists a
554: simple Lie algebra $\lie\ (\ne A_n, C_n)$ with grading (\ref{grading}),
555: which yields the same data (and $\m=\lie_1)$. Moreover the subalgebra $\ml
556: \subset\lie_0$ possesses grading (\ref{grading2}) such that $\ml_1$
557: has the structure of a Jordan algebra isomorphic to $\J$.
558: 
559: 
560: \subsection{Organization of the paper}
561: We provide basic information about split
562: composition algebras and their integral structures
563: in the beginning of Section~$2$. Then we give basic information about Jordan
564: algebras.
565: We present the Springer construction of cubic Jordan algebras and
566: describe the three main examples of Jordan algebras possessing
567: an admissible cubic form (Subsection~$\ref{ss:ex}$). In this paper
568: we are mainly interested in cubic Jordan algebras $\h\calg$ of
569: $3\times 3$ Hermitian matrices over composition algebras
570: (Example~$\ref{ex:jcubic}$).
571: In Subsection~$\ref{ss:33herm}$ we describe the
572: groups $\str\J$ and $\npg\J$ associated with these algebras, and describe
573: orbits in $\h\calg$ under the action of the norm-preserving group
574: $\npg{\h\calg}$ (Proposition~$\ref{prop:np-o}$).
575: We conclude Section~$2$ with the description of the integral structures in
576: Jordan algebras of Subsection~$\ref{ss:ex}$, and the description of
577: integral orbits in $\h\calg$ (Theorem~$\ref{np-orbits}$).
578: \medskip
579: 
580: We begin Section~3 with the description of the Freudenthal construction.
581: Given a cubic Jordan algebra $\J$, this construction produces a group
582: $\gr$ and its representation $\mj$.
583: We identify groups $\gr$  and its representations $\mj$
584: in Proposition~$\ref{prop:gener}$ and Remark~$\ref{rem:semis}$, see also
585: Example~$\ref{ex:sl6}$.
586: 
587: We present the concept of rank of elements in the module $\mj$ in
588: Subsection~$\ref{ss:rank}$, and we use it in the classification of
589: $\gr$-orbits in $\m$ in Subsection~$\ref{ss:canf}$
590: (Theorem~$\ref{thm1}$). The key step in our approach is the computation of
591: Lemma~$\ref{lcomp}$, which allows to do a complete reduction of elements of
592: $\m$ using purely algebraic considerations.
593: 
594: Most of the assertions in Subsection~$\ref{ss:canf}$ are proved under the
595: assumption that the ground field $F$ is an arbitrary field of
596: characteristic $\ne 2,3$. The restriction char\,$F\ne 2$, is essential
597: in the paper since it is used in many intermediate computations (see also
598: Remark~$\ref{badred}$). As for characteristic $3$, it appears that this
599: restriction may be dropped without impairing the statements. However this
600: assumption was made in \cite{b}, which is our main reference concerning
601: the Freudenthal construction, and therefore we had to incorporate it in
602: our considerations.
603: 
604: \medskip
605: Section~$3$ may also be viewed as a testing ground (or a simpler version) of
606: the techniques that we apply to studying the integral case in Section~$4$.
607: We introduce the integral version of the Freudenthal construction  in
608: Subsection~$\ref{ss:intstr}$. We develop a reduction procedure for
609: elements of $\mz$, which is somewhat similar to the reduction of
610: $n\times n$ matrices with integer entries under the elementary row and
611: column transformations. This is done in Lemma~$\ref{bl-fz}$, which is one
612: of the main technical results in Section~4. This lemma does not provide a
613: complete reduction of the elements in $\mz$, but it is sufficient, for
614: example, to describe generators of the group $\grz$ (Proposition~$\ref{genz})$
615: in a way similar the case of a field.
616: 
617: Our elementary approach does not seem to be sufficient for the complete
618: description of orbits in $\mz$, and we restrict our attention to the
619: orbits of projective elements. This concept was introduced by M.Bhargava
620: \cite{bh-phd, bh1}
621: in the context of orbit spaces associated with ideal classes in quadratic
622: rings. In Subsection~$\ref{ss:proj}$ we show how this concept may be
623: transferred to elements of the modules $\mz$. We also provide a simple
624: test for the projectivity of elements in terms of the quartic invariant
625: of the module $\mz$ (Corollary~$\ref{cor:nabla}$,
626: Remark~$\ref{rem:nabla}$).
627: 
628: In Subsection~$\ref{ss:fred}$ we explain how one can do the complete
629: reduction for the projective elements (Lemma~$\ref{red2}$) and summarize
630: all the previous results in the proof of the main result of the paper
631: on the structure of orbits of the projective elements
632: (Theorem~$\ref{thm:main}$).
633: 
634: In the last subsection we analyze degenerate orbits of elements of $\mz$.
635: We are able to obtain a complete classification of orbits of elements of
636: rank $1$ and $2$. These results complement results of the previous
637: subsection; they are stated in Theorem~$\ref{thm2}$.
638: 
639: %
640: \medskip
641: We conclude the paper with an Appendix, where we describe a link between
642: the Freudenthal construction and Gauss's law of composition of quadratic
643: forms via the Cube Law of M.~Bhargava.
644: 
645: 
646: \subsection{Notation and conventions}
647: 
648: We work over a ground field $F$ of characteristic $\ne 2,3$.
649: The vector space of $n\times n$ matrices over $F$ is denoted by $M_n(F)$.
650: The identity matrix in $M_n(F)$ is denoted by $I_n$. For an arbitrary
651: vector space $V$ over $F$, the symbol $\id_V$ denotes the identity linear
652: transformation in $\End_F(V)$.
653: 
654: We use the word {\em space} to denote a set with an additional structure.
655: Depending on a context such a space may be a vector space, a
656: $\zet$-module, or a set of orbits under the action of a group
657: (an {\em orbit space}).
658: 
659: Many spaces that we consider have a certain integral structure introduced
660: in the paper. For a space~$V$,
661: the corresponding integral structure will be denoted by $V_\zet$.
662: Very often such spaces will arise as spaces of representation
663: of certain algebraic groups. In such cases $G(\zet)$ will denote the
664: set of $\zet$-rational points of $G$ with respect to the integral
665: structure in $V_\zet$. This is an {\em integral form\/} of the group $G$.
666: 
667: We will occasionally make references to simple split finite-dimensional
668: Lie algebras (algebraic groups) and their root systems.
669: The labeling of their simple roots, fundamental weights, etc. corresponds
670: to that of \cite{b1}.
671: 
672: 
673: \bigskip
674: {\bf Acknowledgements.}
675: The author is grateful to E.~Zelmanov and M.~Racine for their help and
676: encouragement. Many thanks to
677: B. Allison,
678: M.~Bhargava,
679: O.~Loos,
680: E.~Neher
681: for helpful discussions and comments on the content of the paper.
682: The author is also grateful to M.~Bhargava for providing a copy of his
683: manuscript~\cite{bh1}.
684: 
685: 
686: 
687: 
688: \section{Cubic Jordan algebras and associated structures}
689: \label{s:jordan}
690: 
691: The Freudenthal construction that we will describe in Section~\ref{sec:FC}
692: was originally introduced in the case of the $27$-dimensional exceptional
693: Jordan algebra. We are going to apply this construction to other examples
694: of the so called cubic Jordan algebras, and we introduce the appropriate
695: concepts in this section.
696: 
697: The basic information concerning Jordan algebras is given in
698: Subsection~$\ref{ss:jbasic}$. However, considering the Freudenthal
699: construction, we will not make much use of the Jordan product $\jprod$.
700: Instead, we will be looking at the cubic form $N$, the trace form and
701: the trace bilinear form, the ``sharp" operation $\#$ and its linearization
702: $\times$.
703: 
704: All of these operations (as well as the Jordan product $\jprod$) in a cubic
705: Jordan algebra may be defined starting with an admissible cubic form.
706: We will give the appropriate construction in Subsection~$\ref{ss:sc}$,
707: and provide explicit examples of admissible cubic forms (and Jordan
708: algebras) in Subsection~$\ref{ss:ex}$.
709: 
710: 
711: \subsection{Composition algebras and their integral structures}
712: \label{ss:compint}
713: 
714: A (not
715: necessarily associative) algebra $\calg$ with a unit element
716: is called a {\em composition} algebra
717: if it has a non-degenerate quadratic form $\ncomp$ ({\em the norm})
718: which permits composition, i.e.,
719: $$%
720: \ncomp(xy)=\ncomp(x)\ncomp(y),\qquad
721: \mbox{for all }x,y\in\calg.
722: $$%
723: 
724: A famous theorem of Hurwitz states that
725: such an algebra is always finite-dimensional,  %
726: and moreover,
727: its dimension can only be equal to $1, 2, 4$, or $8$
728: (see, e.g.,  \cite[Section~II.2.6]{atoj} or \cite[Theorem~1.6.2]{SpV}).
729: In this paper we
730: restrict our attention to {\em split} composition algebras only (a
731: composition algebra is split, if it contains non-zero elements of zero
732: norm; in that case the quadratic form $\ncomp$ has maximal Witt index).
733: For a field $F$ there is a unique up to isomorphism split composition
734: algebra of a given dimension $(2,4,$ or $8)$ \cite[Theorem 1.8.1]{SpV}.
735: We will call it the algebra of split {\em binarions, quaternions, octonions},
736: and denote it by
737: $\bin, \quat, \oct$, respectively.
738: We will use symbol $\calg$ to denote an arbitrary fixed composition
739: algebra.
740: 
741: Next we will present the construction of the three split
742: composition algebras and show how to define an integral structure in each
743: of them.
744: 
745: Our starting point for constructing composition algebras will be the
746: algebra of $2\times 2$ matrices over the field $F$. This algebra has
747: dimension $4$ and in fact it can be taken as a model for the algebra of
748: split quaternions $\quat$. The quadratic form $\ncomp$ in this algebra
749: is the usual determinant of $2\times 2$ matrices $\det$. The composition
750: property for $\ncomp$ is just the multiplicativity of the determinant.
751: 
752: One can define the concepts of {\em trace} and {\em conjugation} of
753: elements in a composition algebra. In the case of quaternions they are
754: the trace and the symplectic involution of a $2\times 2$ matrix:
755: \begin{equation}
756: \tcomp (\bfa)=a+d,\qquad
757: \conj{\bfa}=\maketwo{d}{-b}{-c}a,\qquad
758: {\rm for}\
759: {\bfa }=\maketwo{a}bcd\in \quat.
760: \end{equation}
761: 
762: One can define the algebra of binarions $\bin$ to be the subalgebra of
763: $\quat$, which consists of the diagonal elements
764: \begin{equation}\label{binar}
765: \bin=\left\{\left.
766: \maketwo{a}00d\right|
767: a,d\in F\right\}
768: \end{equation}
769: 
770: The quadratic form, trace, and the conjugation operation in $\bin$ are
771: induced by those in $\quat$.
772: 
773: We may consider the ground field $F$ to be embedded in the binarions
774: \begin{equation}
775: F\cong \left\{\left.\maketwo{a}00a\right|a\in F\right\}
776: \subset\bin,
777: \end{equation}
778: and consider the induced structure of a one-dimensional composition algebra.
779: %
780: 
781: \medskip
782: Finally, we define the algebra of octonions using the Cayley-Dickson
783: duplication process \cite{atoj, SpV}. $\oct$ is defined to be the
784: direct sum $\quat\oplus\quat$, and we write an arbitrary octonion with the
785: aid of a formal variable $v$ (the ``imaginary unit") as
786: \begin{equation}
787: {\bfa}+\bfb v,\qquad \bfa , \bfb\in \quat.
788: \end{equation}
789: 
790: One then defines
791: $$
792: ({\bf a+b}v)\cdot ({\bf c+d}v)\defeq
793: {\bf (ac-\conj{d}b)+(da+b\conj{c})}v,
794: \qquad {\bf a,b,c,d}\in \quat.
795: $$
796: \begin{equation}
797: \ncomp({\bf a+b}v) \defeq \det({\bf a})+\det({\bf b}),\quad
798: \tcomp({\bf a+b}\it v) \defeq \tcomp({\bf a}), \quad
799: \conj{{\bf a+b}v} \  \defeq \conj{{\bf a}}-{\bf b}v.
800: \end{equation}
801: 
802: These operations turn $\oct$ into a composition algebra.
803: 
804: We note that the multiplication of binarions is both associative and
805: commutative; the multiplication of quaternions is associative, but not
806: commutative; and, finally, we lose both the associative and commutative
807: property of the multiplication of the octonions.
808: 
809: We identify elements of the ground field $F$ with the appropriate
810: multiples of the identity element in the composition algebra. These are
811: the only elements which are fixed under the conjugation.
812: 
813: 
814: \bigskip
815: 
816: To define an integral structure in these algebras, we again start with
817: the quaternions. We define integral quaternions to be the $2\times 2$
818: matrices with integral entries
819: \begin{equation}
820: \quatz=
821: \left\{\left.
822: \maketwo{a}bcd\right|
823: a,b,c,d\in \zet\right\}.
824: \end{equation}
825: 
826: This integral structure is extended to the binarions and octonions in a
827: natural way
828: \begin{equation}
829: \binz=
830: \left\{\left.
831: \maketwo{a}00d\right|
832: a,d\in \zet\right\},\qquad
833: \octz=
834: \Bigl\{{\bfa}+\bfb v\Bigr|\ \bfa , \bfb\in \quatz
835: \Bigr\}.
836: \end{equation}
837: 
838: It is easy to see
839: that each of these $\zet$-modules has the structure of a composition
840: algebra (over~$\zet$), and the trace and the norm defined on them have
841: integer values.
842: 
843: 
844: \subsection{Basics of Jordan algebras}
845: \label{ss:jbasic}
846: 
847: We are going to present basic definitions and constructions concerning
848: Jordan algebras in this subsection. The classical reference for the
849: subject is N.~Jacobson's book \cite{ja}. A modern exposition of
850: the theory of Jordan algebras may be found in the recent monograph by
851: K.~McCrimmon \cite{atoj}.
852: 
853: \begin{defn}
854: {\rm
855: A (linear) {\em Jordan algebra} $\J$ over a field $F$ of char$\ne 2$ is a vector
856: space over $F$ with a bilinear product $\jprod$
857: (the {\em Jordan product}) satisfying the following two axioms:
858: \begin{equation} \label{j-ax}
859: x\jprod y =y\jprod x;\qquad
860: x^2\jprod (x\jprod y) = x\jprod (x^2 \jprod y)\qquad
861: {\rm \ for \ }x,y\in \J
862: \end{equation}
863: ($x^2$ is defined as $x\jprod x$).
864: }
865: \end{defn}
866: The Jordan product is commutative by definition, but it does not have to satisfy
867: the associative law.
868: 
869: 
870: A prototypical example of a Jordan algebra is the algebra $A^+$
871: obtained from any associative algebra $A$ (e.g., a matrix algebra)
872: by defining the Jordan product to be
873: \begin{equation}  \label{j-assoc}
874: x\jprod y =\frac{1}{2}(xy+yx).
875: \end{equation}
876: Here and below the product $xy$ represents the multiplication in the
877: original (associative) algebra~$A$.
878: 
879: \medskip
880: We will be mostly interested in the so-called cubic Jordan
881: algebras, i.e., Jordan algebras,
882: in which every element satisfies a cubic polynomial equation.
883: This concept can be introduced via the concept of {\em generic norm} and
884: {\em generic minimal polynomial} \cite[Section VI.3]{ja}.
885: 
886: However in this paper we will use a different approach. Namely, we will
887: consider certain examples of Jordan algebras,
888: which are constructed from a vector
889: space with a cubic form. The details of the process are given in the
890: following subsection.
891: 
892: \subsection{The Springer construction of cubic Jordan algebras}
893: \label{ss:sc}
894: 
895: In this subsection we will present a construction of a class of Jordan
896: algebras obtained from a cubic form on a vector space, known as the
897: Springer construction. Our exposition will follow
898: \cite[Section~I.3.8]{atoj}. The original references are
899: \cite{spr}, \cite[Section 4]{mc-sp}. We are still working under the
900: assumption that char$F\ne 2,3$, and we will use it occasionally to
901: simplify the constructions.
902: 
903: \bigskip
904: A cubic form $\mnorm$ on a vector space $V$ over $F$ (char\,$F\ne 2,3$) is a map
905: $\mnorm:V\to F$ such that
906: \begin{itemize}
907: \item $\mnorm(\alpha x)=\alpha^3 \mnorm(x)$ for $\alpha\in F, x\in V$;
908: \item $\mnorm(x,y,z)$ is a trilinear function $V\times V \times V\to F$;
909: \end{itemize}
910: where $\mnorm(x,y,z)$ is the full linearization of $\mnorm$ given by
911: $$
912: \mnorm(x,y,z)=\frac{1}{6}
913: \Bigl(\mnorm(x+y+z)-\mnorm(x+y)-\mnorm(x+z)-\mnorm(y+z)+\mnorm(x)+\mnorm(y)+\mnorm(z)\Bigr).
914: $$
915: In particular, we have
916: $$
917: \mnorm (x,x,x) = \mnorm (x).
918: $$
919: 
920: We say that $c\in V$ is a {\em basepoint} for $\mnorm$, if $\mnorm(c)=1$. One then
921: can define the following maps
922: \begin{itemize}
923: \item
924: a linear map (trace) $V\to F$: $\tr(x)=3\,\mnorm(c,c,x)$;
925: \item
926: a quadratic map $V\to F$: $S(x)=3\mnorm(x,x,c)$;
927: \item
928: a bilinear map $V\times V\to F$: $S(x,y)=6\mnorm(x,y,c)$;
929: \item
930: a trace bilinear  form $V\times V\to F$: $(x,y)=\tr(x)\tr(y)-S(x,y)$.
931: \end{itemize}
932: 
933: In particular, we have
934: $$
935: \mnorm(c)=1;\qquad
936: S(c)=3;\qquad
937: \tr(c)=3.
938: $$
939: 
940: \begin{defn}
941: {\em
942: A cubic form with a basepoint $(\mnorm,c)$ on a finite-dimensional vector space
943: $V$ over a field $F$ of char$\ne 2,3$ is said to be {\em admissible}
944: or a {\em Jordan cubic}, if
945: \begin{enumerate}
946: \item[(1)]
947: $\mnorm$ is nondegenerate at the basepoint $c$ in the sense that
948: the trace bilinear form $(x,y)$ is nondegenerate;
949: \item[(2)]
950: The quadratic {\em adjoint} (or {\em sharp}) map $V\to V$,
951: defined uniquely by $(x^\#, y)=3\mnorm(x,x,y)$, satisfies the adjoint identity:
952: \begin{equation}\label{adid}
953: (x^\#)^\#=\mnorm(x)x.
954: \end{equation}
955: %
956: %
957: \end{enumerate}
958: We will also use the term {\em Jordan cubic} when referring to the
959: associated Jordan algebra (see Proposition~\ref{prop:jcubic}).
960: }
961: \end{defn}
962: 
963: The following relation holds for all $x$ in $V$:
964: \begin{equation}\label{trace-spur}
965: \tr(x^\#)=S(x).
966: \end{equation}
967: 
968: \bigskip
969: We define the linearization of the sharp map
970: \begin{equation}
971: x\cross y = (x+y)^\#-x^\#-y^\#.
972: \end{equation}
973: Note that
974: $$
975: x\cross x = 2 x^\#.
976: $$
977: 
978: 
979: 
980: \begin{prop}\label{prop:jcubic}
981: {\em \cite[Section I.3.8]{atoj}}
982: Every vector space with an admissible cubic form gives rise to a Jordan
983: algebra with unit $\one=c$ and the Jordan product given by
984: \begin{equation}\label{j-prod}
985: x\jprod y =
986: \frac{1}{2}\Bigl( x\cross y+\tr(x)y+\tr(y)x-S(x,y)\one \Bigr).
987: \end{equation}
988: Every element of this Jordan algebra satisfies the cubic polynomial:
989: \begin{equation} \label{3poly}
990: x^3-\tr(x)x^2+S(x)x-\mnorm(x)\one=0.
991: \end{equation}
992: We also have
993: $$
994: x^\#=x^2-\tr(x)x+S(x)\one.
995: $$
996: \end{prop}
997: 
998: Taking the trace of the last expression and then using~$(\ref{trace-spur})$,
999: one gets
1000: $$
1001: \tr(x^2)=\tr(x)^2-2S(x),
1002: $$
1003: 
1004: which linearizes to
1005: \begin{equation} \label{traces}
1006: \tr(x\jprod y)=(x,y).
1007: \end{equation}
1008: 
1009: \medskip
1010: The following simple example illustrates the concepts introduced above.
1011: 
1012: \begin{example}\label{ex:mat3}
1013: {\rm
1014: Let $V$ be the vector space $M_{3}(F)$ of $3\times 3$ matrices over $F$.
1015: We let the cubic form $\mnorm$ be the determinant, and we let $c$ be the
1016: identity matrix.
1017: 
1018: Then the linear map $\tr$ is the regular trace of matrices. The sharp map
1019: $x^\#$ produces the classical adjoint (transposed cofactor) matrix of $x$.
1020: Equation (\ref{j-prod}) yields the Jordan product, which coincides with the
1021: product in the Jordan algebra $M_{3}(F)^+$ given by (\ref{j-assoc}):
1022: $$
1023: x\jprod y=\frac{1}{2}(xy+yx).
1024: $$
1025: The trace bilinear form $(x,y)$ is equal to $\tr(x\jprod y)$.
1026: It coincides in this
1027: example with the standard trace form in the matrix algebra.
1028: Finally, we notice that the equation (\ref{3poly}) becomes just the~
1029: Cayley-Hamilton equation for $3\times 3$ matrices.
1030: }
1031: \ee
1032: \end{example}
1033: 
1034: 
1035: \subsection{Three main examples}
1036: \label{ss:ex}
1037: 
1038: We will provide several more examples of cubic Jordan algebras in this
1039: subsection.
1040: 
1041: 
1042: The following example is a simplified version of the {\em reduced cubic
1043: factor example} \cite[Section~I.3.9]{atoj}.
1044: \begin{example}\label{ex:jcubic}
1045: {\rm
1046: Let $\calg$ be a composition algebra over a field $F$
1047: with a quadratic form $\ncomp$ and an involution $\bar{\ }$.
1048: 
1049: We let $V$ be the space $\h{\calg}$ of $3\times 3$ Hermitian matrices over $\calg$.
1050: An arbitrary element $A$ in $\h{\calg}$ has the form
1051: $$
1052: A=\makeherm{a}bc{\bf x}{\bf y}{\bf z},\qquad
1053: \begin{array}{l}
1054: \mbox{where }\ a,b,c \in \F, \\
1055: \mbox{and }{\bf x,y,z}\in\calg.
1056: \end{array}
1057: $$
1058: The basepoint $c$ is defined to be the identity (diagonal) matrix in
1059: $\h\calg$, and the cubic form
1060: $$
1061: N:\h{\calg}\to F
1062: $$
1063: is given by the expression
1064: reminiscent of the regular determinant
1065: \begin{equation} \label{cubic1}
1066: \mnorm(A) = \ abc-a\, {\bf x\conj{x}}-b\, {\bf y\conj{y}}-c\, {\bf z\conj{z}}+
1067: {\bf (xy)z+\bar{\bf z}{(\bar{\bf y}\bar{\bf x})}}.
1068: \end{equation}
1069: 
1070: The trace form $\tr$, as defined in the previous section from $\mnorm$,
1071: coincides with the regular trace $\tr(A)=a+b+c$ and
1072: $\btr(A,B)=\tr\Bigl(\frac{1}{2}(AB+BA)\Bigr)$.
1073: The ``sharp" operation produces the regular adjoint matrix:
1074: $$
1075: A^\#=
1076: \left(
1077: \begin{array}{lll}
1078: 
1079:    bc-\ncomp({\bf x}) \quad                       & \conj{\bf y}\ \conj{\bf x}-c\,{\bf z} \quad   & {\bf z\, x}-b\,\conj{{\bf y}} \\
1080:    {\bf x}\ {\bf y}-c\,\conj{\bf z}     & ac-\ncomp({\bf y})                          & {\bf \conj{z}\ \conj{y}} -a\,{\bf x}\\
1081:    {\bf \conj{x}\, \conj{z}}-b\,{\bf y} & {\bf y \, z} -a\,{\bf x}                &  ab-\ncomp({\bf z})
1082: \end{array}
1083: \right).
1084: $$
1085: 
1086: The cubic form $\mnorm$ defined by $(\ref{cubic1})$ is admissible.
1087: }\ee
1088: \end{example}
1089: 
1090: 
1091: We note that Example \ref{ex:mat3} can be viewed as
1092: a special case of Example \ref{ex:jcubic}
1093: as explained in the following
1094: 
1095: \begin{rem}\label{rem:hbm3}
1096: {\rm
1097: If the composition algebra $\calg$ is the algebra of split binarions
1098: $\bin$, then an arbitrary element of $\h\bin$ has the form
1099: (see~$(\ref{binar})$)
1100: $$
1101: \makethree
1102: {a}                 {\bfa_{12}}         {\bfa_{13}}
1103: {\conj{\bfa_{12}}}  {b}                 {\bfa_{23}}
1104: {\conj{\bfa_{13}}}  {\conj{\bfa_{23}}}  {c}
1105: ,\qquad
1106: \mbox{where }
1107: \bfa_{ij}=\maketwo{a_{ij}}{\scriptstyle 0}{\scriptstyle 0}{a_{ji}}\in
1108: \bin,\quad
1109: a,b,c, \in F.
1110: $$
1111: 
1112: We then define a linear map $\h\bin\to M_3(F)$
1113: \begin{equation}\label{bin3iso}
1114: \makethree
1115: {a}                 {\bfa_{12}}         {\bfa_{13}}
1116: {\conj{\bfa_{12}}}  {b}                 {\bfa_{23}}
1117: {\conj{\bfa_{13}}}  {\conj{\bfa_{23}}}  {c}
1118: %
1119: \longmapsto
1120: \makethree
1121: {a}       {a_{12}}  {a_{13}}
1122: {a_{21}}  {b}       {a_{23}}
1123: {a_{31}}  {a_{32}}  {c}.
1124: \end{equation}
1125: 
1126: It is easy to see that this isomorphism of vector spaces
1127: is a norm isometry, where the cubic form in $M_3(F)$ is
1128: the regular determinant.
1129: Hence $(\ref{bin3iso})$ defines an isomorphism of
1130: Jordan algebras.
1131: \ee
1132: }
1133: \end{rem}
1134: 
1135: \begin{example}\label{ex:quad}
1136: {\rm
1137: \cite[Section~$4$]{mc-sp}
1138: Let $Q_n$ be a vector space of dimension $n$ over a field $F$,
1139: and let $B_0$ be a
1140: non-degenerate quadratic form on $Q_n$ with a basepoint $c_0$
1141: (i.e., $B_0(c_0)=1$). We form a vector space $V$ by
1142: taking the direct sum of a copy of the
1143: ground field $F$ and $Q_n$:
1144: $$
1145: V=F \oplus Q_n.
1146: $$
1147: 
1148: We then define a cubic form $\mnorm$ on $V$ by
1149: $$
1150: \mnorm(\alpha, x_0)=\alpha\, B_0(x_0), \qquad
1151: \mbox{for }\alpha \in F, x_0\in Q_n.
1152: $$
1153: We let $c=(1,c_0)$. A simple verification
1154: shows that $c$ is a basepoint for $\mnorm$
1155: and that the cubic form $\mnorm$ is admissible.
1156: 
1157: In particular, the following formulas hold for $x,y \in V$:
1158: \begin{eqnarray*}
1159: x^\# &=& \Bigl(B_0(x_0),\ \alpha x_0^*\Bigr),\\
1160: (x,y)&=& \alpha\beta + B_0(x_0^*, y_0),
1161: \end{eqnarray*}
1162: where $x=(\alpha, x_0), y=(\beta, y_0)$, $x_0^*=B_0(x_0, c_0)c_0-x_0$, and
1163: $B_0(\cdot,\cdot)$ is a linearization of the quadratic form $B_0(\cdot)$:
1164: $$
1165: B_0(u,v)=B_0(u+v)-B_0(u)-B_0(v).
1166: $$
1167: }\ee
1168: \end{example}
1169: 
1170: \begin{example}\label{ex:f1}
1171: {\rm
1172: We let $V$ be the one-dimensional vector space $V=F$. We define a cubic
1173: form $\mnorm$ on $V$ by
1174: $$
1175: N(\alpha)=\alpha^3, \qquad
1176: \alpha\in V
1177: $$
1178: with the obvious choice of the base point $c=1\in F$. We have
1179: $\tr(\alpha)=3\alpha$, $(\alpha, \beta)=3\alpha\beta$,
1180: $\alpha^\#=\alpha^2$ for $\alpha, \beta\in F$, and evidently
1181: $N$ is an admissible cubic form.
1182: }\ee
1183: \end{example}
1184: 
1185: 
1186: 
1187: \subsection{Two groups associated to cubic Jordan algebras}
1188: \label{ss:33herm}
1189: 
1190: Next we will introduce two groups associated to a Jordan algebra $\J$. The
1191: definitions, due to N.~Jacobson, are valid for all finite-dimensional
1192: Jordan (or power-associative) algebras, but we will use them only for
1193: Jordan algebras with a cubic form.
1194: 
1195: \begin{defn}\label{def-groups}
1196: 
1197: The {\em norm-preserving group}
1198: $$
1199: \npg\J=\Bigl\{g\in GL(\J)\ |\  \mnorm(g A)=\mnorm(A)\mbox{ for all }A\in \J\Bigr\}
1200: $$
1201: is the group of all invertible $F$-linear transformations of the
1202: vector space $\J$, which preserve the norm~$\mnorm$.
1203: 
1204: Similarly we define the {\em group of norm similarities} or
1205: the {\em structure group}
1206: $$
1207: \str\J=\Bigl\{g\in GL(\J)\ |\  \mnorm(g A)=\chi(g)\, \mnorm(A)\mbox{ for all }A\in \J\Bigr\},
1208: $$
1209: where $\chi(g)$ is a scalar in $F$, which depends on the group element $g$
1210: only.
1211: \end{defn}
1212: 
1213: We have the obvious inclusion $\npg\J \subset \str\J$.
1214: 
1215: \bigskip
1216: 
1217: 
1218: \begin{rem}\label{groups}
1219: {\rm
1220: Here we will provide a description of these groups for cubic Jordan
1221: algebras of the form $\h\calg$
1222: (\cite[Ch.~$14$]{sp-alg}, see also \cite{J3}, \cite[VI.7--9]{ja}).
1223: %
1224: %
1225: 
1226: \begin{itemize}
1227: \item[(i)]
1228: {\bf Case} $\J=\h{F}$.
1229: 
1230: $\str\J$ is the group of transformations of the form
1231: $$
1232: X\mto \gamma AXA^t,
1233: $$
1234: 
1235: where $X\in\h{F}$, $A\in GL_3(F)$, $\gamma\in F^\mult$, $A^t$ is the transpose of $A$.
1236: %
1237: 
1238: $\npg\J$ consists of transformations for which
1239: $\gamma^3(\det A)^2=1$.
1240: 
1241: 
1242: \item[(ii)]
1243: {\bf Case} $\J=\h{\bin}$.
1244: 
1245: In this case $\J$ is isomorphic to the Jordan algebra of all $3\times 3$
1246: matrices over $F$ (Remark~$\ref{rem:hbm3}$).
1247: 
1248: The group $\str\J$ is generated by the transformations of the form
1249: $$
1250: \eta(A,B): X\mto AXB^{-1}\quad
1251: \mbox{ and }\quad
1252: t: X\mto X^t,
1253: $$
1254: 
1255: where $X\in M_3(F)$, $A,B\in GL_3(F)$, and $X^t$ denotes the transpose of $X$.
1256: 
1257: $\eta(A,B)$ acts trivially iff $A=B=\alpha\id, \ \alpha\in F,\alpha\ne 0$.
1258: 
1259: $\npg\J$ consists of transformations $\eta(A,B)$ for which $\det A=\det B$.
1260: 
1261: %
1262: %
1263: 
1264: We will let $\str \J\cc$ denote the subgroup of $\str\J$ that
1265: consists of the transformation $\eta(A,B)$. It is a subgroup of index two
1266: in $\str\J$, and we have $\str\J\cong\str\J\cc\rtimes \{1, t\}$.
1267: Similarly we define $\npg\J\cc=\npg\J \cap \str\J\cc$.
1268: 
1269: \item[(iii)]
1270: {\bf Case} $\J=\h{\quat}$.
1271: 
1272: We have the isomorphisms of Jordan algebras
1273: $\h{\quat}\cong {\cal H}(M_{6}(F), {\rm symp})$ (the $6\times 6$ symplectic
1274: symmetric matrices) (see, e.g., \cite[p. 65]{J3}).
1275: 
1276: The group $\str\J$ is the group of transformations of the form
1277: $$
1278: X\mto \gamma AXA^{\rm symp}
1279: $$
1280: 
1281: where $X\in{\cal H}(M_{6}(F), {\rm symp})$,
1282: $A\in GL_6(F)$, $\gamma\in F$, $\gamma\ne 0$, $A^{\rm symp}$ is the transpose of $A$
1283: with respect to the symplectic involution.
1284: 
1285: $\npg\J$ consists of transformations for which $\gamma^3 (\det A)^2=1$.
1286: 
1287: \item[(iv)]
1288: {\bf Case} $\J=\h{\oct}$.
1289: 
1290: In this case we have that $\npg{\J}$ is a simply connected
1291: simple algebraic group of type $E_6$, whose center is
1292: isomorphic to the group $\bfmu _3(F)$
1293: of third roots of unity in $F$.
1294: 
1295: The group $\str\J$ in this case is equal to the product
1296: $\npg\J \cdot (F^\mult\id_\J)$ with $\npg\J \cap F^\mult\id_\J=\bfmu_3(F)\id$.
1297: \end{itemize}
1298: 
1299: Later we will view groups $\str\J$ and $\npg\J$ as the group of
1300: $F$-rational points of the algebraic groups $\astr \J$ and $\anpg\J$,
1301: respectively. These group are connected in all cases, except $\J=\h\bin$.
1302: For $\J=\h\bin$, the group $\astr{\J}$ (respectively, $\anpg{\J}$)
1303: has two connected components; and the set of $F$-rational points of the
1304: component of the identity coincides with $\str{\J}\cc$ (respectively,
1305: $\npg{\J}\cc$).
1306: \ee
1307: }
1308: \end{rem}
1309: 
1310: The following definition generalizing the concept of rank for the usual
1311: $3\times 3$ matrices goes back to N.~Jacobson.
1312: 
1313: \begin{defn}\label{def:jrank}
1314: Let $\J$ be a cubic Jordan algebra and let $\mnorm$ be the cubic form on $\J$.
1315: The {\em rank} of an arbitrary element $A\in\J$ is an integer between zero and
1316: three, which is defined by the following relations:
1317: \begin{itemize}
1318: \item
1319: \rank$A =3$ iff \ $N(A)\ne 0$;
1320: \item
1321: \rank$A \le 2$ iff \ $N(A)=0$;
1322: \item
1323: \rank$A \le 1$ iff \ $A^\#=0$;
1324: \item
1325: \rank$A=0$ iff \ $A=0$.
1326: \end{itemize}
1327: \end{defn}
1328: 
1329: It is known that the rank is invariant under the action of the
1330: groups $\npg\J$ and $\str\J$, see, e.g., \cite[Section~$2$]{J3}.
1331: 
1332: \medskip
1333: We will conclude this subsection by describing orbits in $\h\calg$
1334: under the action of the \np\ group $\npg{\h\calg}$.
1335: The classification is based on the concept of rank
1336: and is analogous to the classification of orbits in $M_3(F)$
1337: under the action of the elementary row and column transformations.
1338: 
1339: \begin{prop}\label{prop:np-o}
1340: 
1341: \
1342: \begin{enumerate}[\bf (i)]
1343: \item
1344: Let $\calg$ be the split composition algebra over a field $F$,
1345: $\calg=\bin, \quat, \oct$, char $F\ne 2,3$.
1346: Let $\h{\calg}$ be the cubic Jordan algebra
1347: of $3\times 3$ Hermitian matrices over $\calg$.
1348: 
1349: Then the group $\npg{\h{\calg}}$ acts transitively on the
1350: sets of elements of rank $1$ and $2$. In the case of rank\/ $3$,
1351: the group $\npg{\h{\calg}}$ acts transitively on the elements
1352: of a given norm $k$, $k\in F$, $k\ne 0$.
1353: 
1354: All these orbits are distinct and the union of these orbits and $\{0\}$ is
1355: $\h{\calg}$.
1356: \item
1357: If in addition every element of $F$ is a square,
1358: the same result holds for $\h{F}$.
1359: \end{enumerate}
1360: \end{prop}
1361: 
1362: It follows that the orbit representatives for the action of
1363: $\npg{\h{\calg}}$ may be chosen to be the following
1364: diagonal matrices:
1365: $$
1366: 0, \quad
1367: \makethree 1.. .0. ..0,\quad
1368: \makethree 1.. .1. ..0,\quad
1369: \makethree 1.. .1. ..k,\quad
1370: k\ne 0.
1371: $$
1372: 
1373: {\sc Proof.}
1374: 
1375: {\bf (i)}
1376: The statement of the theorem is obvious for matrices in
1377: $\h\bin\cong M_3(F)$ (Remark~$\ref{rem:hbm3}$).
1378: For other split composition algebras the statement was
1379: essentially known to N.~Jacobson, see e.g. \cite[Section~$2$]{J3}.
1380: Alternatively, one may view this proposition as a corollary (of the proof)
1381: of the main theorem of \cite{k}, see also Theorem~$\ref{np-orbits}$ below.
1382: 
1383: {\bf (ii)}
1384: The proof of (i) does not work in the case of Hermitian matrices over the
1385: ground field $F$. The result here depends on the arithmetic properties
1386: of the ground field. When $F$ satisfies the assumptions of (ii), the
1387: assertion is well known, see, e.g. \cite[Theorems~$6.5, 6.6$]{jba}.
1388: \ep
1389: 
1390: \subsection{Operations and identities in cubic Jordan algebras}
1391: \label{ss:id}
1392: 
1393: Given an arbitrary Jordan algebra $\J$, one can define the
1394: {\em Jordan triple product}
1395: \begin{equation}\label{jtp}
1396: \{x,y,z\}  =  (x\jprod y)\jprod z+ x\jprod (y\jprod z)-(x\jprod z)\jprod
1397: y,\qquad
1398: x,y,z\in\J.
1399: \end{equation}
1400: 
1401: When $\J\subseteq A^+$ for some associative algebra $A$, the Jordan triple
1402: product has a simple expression in terms of the associative operation in
1403: $A$:
1404: \begin{equation}\label{a:triple}
1405: \{x,y,z\}=\frac{1}{2}(xyz+zyx),\qquad
1406: x,y,z\in \J\subseteq A^+.
1407: \end{equation}
1408: This operation is very important for generalizations of Jordan algebras.
1409: We mentioned it here, since we need to introduce another operation
1410: $V_{x,y}:\J\to\J$ defined by
1411: \begin{eqnarray}\label{a:v}
1412: V_{x,y} (z) & = & \{x, y, z\},\qquad
1413: x,y,z\in \J.
1414: \end{eqnarray}
1415: We will use the operation $V_{x,y}$ in
1416: Subsections~$\ref{ss:rank}$ and~$\ref{ss:if}$ in the definition
1417: of rank of elements of the module~$\mj$.
1418: 
1419: \medskip
1420: 
1421: Next we list several identities which relate the triple product, the cross
1422: product, and the trace bilinear form in an arbitrary {\em cubic} Jordan
1423: algebra~$\J$.
1424: 
1425: It follows from $(\ref{3poly})$ that
1426: $$
1427: X\jprod X^\#=\mnorm(X)\,\one
1428: $$
1429: 
1430: for any element $X$ in $\J$.
1431: 
1432: 
1433: %
1434: 
1435: One can linearize this identity
1436: (see, e.g., \cite[Section~${2}$]{J3}, \cite[Section~5.2]{SpV})
1437: and get
1438: 
1439: \begin{eqnarray} \label{a:id1}
1440: \{X,Y,X\}+2Y\cross (X^\#)&=&(X,Y)X\\ \label{a:id2}
1441: (X,Z)Y+(Y,Z)X&=&2\{X,Z,Y\}+(X\cross Y)\cross Z\\ \label{a:id3}
1442: N(Y)X+(X,Y^\#)Y&=&(X\cross Y)\cross Y^\#\\ \label{a:id4}
1443: (X,Z^\#)X+\{Z,X^\#,Z\}&=&(X\cross Z)^\#.
1444: \end{eqnarray}
1445: 
1446: 
1447: One can define the * operation
1448: for an arbitrary element $s$ of the \np\ group
1449: $\npg\J$ by the relation:
1450: \begin{equation} \label{a:adj}
1451: \Bigl(\ s   (X),\ Y\ \Bigr) \ =
1452: \ \Bigl(\ X,\ s^* (Y)\ \Bigr) \qquad \mbox{for any }X,Y\in \J.
1453: \end{equation}
1454: 
1455: It satisfies the following identities:
1456: \begin{equation} \label{a:crossid}
1457: {s^{*-1}}(X\cross Y)= s (X) \cross s (Y) \qquad \mbox{for any }s\in\npg\J,\ X,Y\in\J.
1458: \end{equation}
1459: \begin{equation} \label{a:tripleid}
1460: s \Bigl(\{X,Y,Z\}\Bigr)= \Bigl\{\ s (X),\ {s^{*-1}}(Y),\ s (Z)\ \Bigr\} \qquad
1461: \mbox{for any }s\in\npg\J,\ X,Y,Z\in\J.
1462: \end{equation}
1463: 
1464: 
1465: \subsection{The integral case and orbits under the \np\ group}
1466: \label{ss:33int}
1467: 
1468: The integral structure $C_\zet$
1469: of the composition algebras $C$ in Subsection~$\ref{ss:compint}$ induces an
1470: integral structure in the spaces $\h{C_\zet}$.
1471: 
1472: We note that $\h{C_\zet}$ is not closed under the Jordan product
1473: because of the factor $\frac{1}{2}$, but
1474: it is easy to see from the expressions given in Example~$\ref{ex:jcubic}$,
1475: that the cubic form $\mnorm$, the trace $\tr$ and the trace bilinear form
1476: $\btr(\cdot, \cdot)$ take values in (and onto) $\zet$. We also have that
1477: $\h{C_\zet}$ is closed under the sharp operation $\#$, and
1478: $X\cross Y\in\h{C_\zet}$ for any $X,Y\in \h{C_\zet}$.
1479: 
1480: 
1481: We consider the \np\ groups defined in Subsection~$\ref{ss:33herm}$,
1482: and we look at
1483: their subgroups of elements which preserve the integral submodule
1484: $\h{C_\zet}$. This subgroup is an integral form of the appropriate group.
1485: We consider the action of each of these groups on the space of
1486: integral Hermitian matrices.
1487: 
1488: The structure of the orbits in $\h{C_\zet}$ under the action
1489: of the \np\ groups is described in
1490: Theorem~\ref{np-orbits} below. Before stating the theorem
1491: we will recall the following well-known
1492: 
1493: \begin{defn}
1494: {\rm
1495: We say that an $n\times n$ matrix $A$ is in the {\em Smith normal form},
1496: if $A$ is a diagonal matrix
1497: $$
1498: A=\diag{d_1, d_2}{\dots}{d_n},
1499: $$
1500: where
1501: $$
1502: \mbox{all }d_i\mbox{'s are integers, }\quad
1503: d_i \,|\, d_{i+1},\quad d_i\ge 0\quad
1504: \mbox{ for }1\le i \le n-1
1505: $$
1506: and all zeros on the diagonal are located in the lower right corner.
1507: }
1508: \end{defn}
1509: 
1510: \begin{theorem} \label{np-orbits}
1511: Let $\h{\calg_\zet}$ be the $\zet$-module of $3\times 3$ Hermitian matrices
1512: over the split composition {\em ring} $\calg_\zet$, $\calg_\zet=\bin_\zet,
1513: \quat_\zet, \oct_\zet$.
1514: 
1515: Then every element of $\h{\calg_\zet}$ is equivalent to an element in the
1516: {\em Smith normal form} under the action of the group $\npg{\h{\calg_\zet}}$.
1517: Distinct elements in the Smith normal form lie in distinct
1518: $\npg{\h{\calg_\zet}}$-orbits.
1519: \end{theorem}
1520: 
1521: {\sc Proof.}
1522: The assertion of this theorem when $C_\zet=\oct_\zet$
1523: was proved in \cite{k} (we used the term {\em canonical diagonal form}
1524: to represent Smith normal form).
1525: 
1526: The reasoning of that paper also applies in the case of integer
1527: quaternions and was stated as Corollary there \cite[p.~294]{k}.
1528: This assertion may also be stated in terms of integer skew-symmetric
1529: matrices (cf. Remark~\ref{groups}(iii)), see \cite[Theorem~IV.1]{newman}.
1530: 
1531: In case of integer binarions, one could also repeat the argument of
1532: \cite{k} to arrive to the same conclusion. An alternative way
1533: to prove the theorem is to notice that the isomorphism of
1534: Remark~$\ref{rem:hbm3}$ yields a $\zet$-linear norm isometry of
1535: $\zet$-modules
1536: $\h{\binz}$ and $M_{3}(\zet)$
1537: A consequence of this fact is that the action of the \np\
1538: group in $\h\binz$ can be expressed in terms of the elementary row and columns
1539: transformations of the regular $3\times 3$ matrices over $\zet$.
1540: Hence in the case of
1541: integral binarions the assertion of the theorem is equivalent to the
1542: assertion on the Smith normal form for the usual $3\times 3$ matrices over
1543: $\zet$.
1544: \ep
1545: 
1546: \begin{rem}
1547: {\rm
1548: An important feature of the theorem is that we have here a chain of
1549: embedded spaces
1550: \begin{equation}
1551: \h{\bin} \subset \h{\quat} \subset \h{\oct},
1552: \end{equation}
1553: and the action of the corresponding groups there.
1554: Theorem~\ref{np-orbits} gives us
1555: a uniform description of orbits for all three spaces
1556: in terms of diagonal matrices contained in each of these spaces.
1557: \ee
1558: }
1559: \end{rem}
1560: 
1561: 
1562: We also have the following trivial corollary to Theorem~\ref{np-orbits}
1563: (which also applies in the case of general $n\times n$ matrices over
1564: $\zet$ and their orbits under the elementary row and column transformations).
1565: 
1566: \begin{cor}
1567: Let $\h{\calg_\zet}$ and $G_\zet$ be as in Theorem~$\ref{np-orbits}$,
1568: and let $n$ be an integer, $n\ne 0$.
1569: The group $G_\zet$ acts transitively on set of matrices of determinant $n$
1570: if and only if $n$ is a squarefree integer.
1571: \end{cor}
1572: 
1573: 
1574: 
1575: \section{The Freudenthal construction}
1576: \label{sec:FC}
1577: 
1578: \subsection{Preliminary facts on the module $\mj$ and the group $\gr$}
1579: \label{ss:fc}
1580: 
1581: In this section we consider a certain class of modules and
1582: linear groups acting on them.
1583: Historically, the first example of this kind was introduced by
1584: H.~Freudenthal
1585: in \cite{fr14} in the process of constructing the $56$-dimensional
1586: representation of the group $E_7$ from the $27$-dimensional exceptional
1587: Jordan algebra.
1588: These modules were studied axiomatically (under the name of the Freudenthal
1589: triple systems) in \cite{b}, \cite{fau}, \cite{fer}.
1590: 
1591: We will consider examples of Freudenthal triple systems
1592: of the form $\mj$, where $\J$ is a cubic Jordan algebra
1593: of Section~$\ref{s:jordan}$.
1594: We will say that the module $\mj$ and its automorphism group $\gr$ (see below)
1595: are obtained from $\J$ using the~{\bf Freudenthal construction}.
1596: 
1597: 
1598: \medskip
1599: We will use \cite{b} as the main reference for this subsection.
1600: We consider a vector space $\m=\mj$ constructed from the space $\J$ in the
1601: following way
1602: \begin{equation}\label{eq:mj}
1603: \mj=F\oplus F \oplus \J \oplus \J,
1604: \qquad \mbox{ where $\J$ is a cubic Jordan algebra over $F$.}
1605: \end{equation}
1606: 
1607: We have $\dim \m=2\, \dim\J+2$, and an arbitrary element $x$ of the space
1608: $\m$ has the form
1609: \begin{equation} \label{gen}
1610: x=\mel\alpha \beta A B,\qquad
1611: {\rm where\ } \alpha, \beta \in F,\quad A,B \in \J.
1612: \end{equation}
1613: 
1614: We have the following skew-symmetric bilinear and quartic forms
1615: on $\m$ defined by
1616: $$
1617: \{ x,y\}= \alpha\delta - \beta\gamma+ (A,D)-(B,C)
1618: $$
1619: $$
1620: q(x)= 8 (A^\#, B^\#) -8\alpha \mnorm(A) - 8\beta \mnorm(B)
1621: -2\Bigl((A,B)-\alpha\beta\Bigr)^2.
1622: $$
1623: Here we have
1624: $x=\mel\alpha \beta A B$, $y=\mel\gamma \delta C D$.
1625: $(\cdot, \cdot)$ is the trace bilinear form,
1626: $\mnorm$ is the norm, and $\#$ is the
1627: sharp map in the Jordan algebra $\J$
1628: (see Subsections~$\ref{ss:sc}$,~$\ref{ss:ex}$ for detail).
1629: 
1630: Later on, when we turn to the integral case, it will be more convenient
1631: for us to consider the modified form $q'$:
1632: %
1633: \begin{equation}\label{qmod}
1634: q'(x)= -4 (A^\#, B^\#) +4\alpha \mnorm(A) +4\beta \mnorm(B)
1635: +\Bigl((A,B)-\alpha\beta\Bigr)^2,
1636: \qquad q=-2q'.
1637: \end{equation}
1638: 
1639: We will often refer to the form $q$ as the {\em norm form} or just the
1640: {\em norm} in the module~$\m$.
1641: 
1642: We can linearize the form $q$ and get a symmetric four-linear form
1643: $q(x,y,z,w)$ such that $q(x,x,x,x) = q(x)$. It follows that both the bilinear
1644: and four-linear form are non-degenerate. And hence we can define a
1645: trilinear operator $T:\m \times \m \times \m \to \m$ by the following
1646: rule: for given $x,y,z \in \m$, $T(x,y,z)$ is the unique element in $\m$
1647: such that
1648: \begin{equation} \label{deft}
1649: \{ T(x,y,z), w\} = q(x,y,z,w)\qquad \mbox{for any $w\in \m$}.
1650: \end{equation}
1651: 
1652: \begin{defn}
1653: The group $\gr$ is defined to be the group of all invertible $F$-linear
1654: transformations of $\m$ that preserve these forms, i.e.,
1655: \begin{equation} \label{invar}
1656: \Bigl\{ \sigma (x),\sigma (y)\Bigr\} = \{ x,y\},\qquad q\Bigl(\sigma (x)\Bigr) =
1657: q(x)\quad
1658: \end{equation}
1659: for any $\sigma \in \gr$.
1660: \end{defn}
1661: 
1662: It follows immediately from $(\ref{deft})$ and $(\ref{invar})$ that for
1663: any $\sigma \in \gr$
1664: \begin{equation} \label{autot}
1665: T\Bigl(\sigma (x), \sigma (y), \sigma (z)\Bigr)=
1666: \sigma \Bigl( T(x,y,z) \Bigr).
1667: \end{equation}
1668: 
1669: 
1670: We have the following four types of transformations in the group $\gr$:
1671: \begin{itemize}
1672: 
1673: \item[]
1674: For any $C\in \J$
1675: 
1676: \begin{equation} \label{phi}
1677: \phi(C)\quad : \quad \mel \alpha \beta A B\  \mto
1678: \mel {\alpha+(B,C)+(A, C^\#)+\beta \mnorm(C)} {\quad    \beta}
1679:  {\quad A+\beta C} {\quad  B+A\cross C +\beta C^\#}.
1680: \end{equation}
1681: 
1682: \item[]
1683: For any $D\in \J$
1684: 
1685: \begin{equation} \label{psi}
1686: \psi(D) \quad : \quad
1687: \mel \alpha \beta A B\  \mto
1688: \mel \alpha {\quad \beta + (A,D) + (B, D^\#) + \alpha \mnorm(D)}
1689: {A+ B\cross D+\alpha D^\#} {B+\alpha D}.
1690: \end{equation}
1691: 
1692: \item[]
1693: In addition, for every norm similarity $s\in \str\J$
1694: (cf. Definition~$\ref{def-groups}$)
1695: we have the transformation $\tf s \in \gr$ defined by:
1696: \begin{eqnarray} \label{t}
1697: \tf s& : &
1698: \mel \alpha\beta A B\  \mto
1699: \mel {\lambda ^{-1}\alpha}{\quad \lambda\beta}
1700:      {\quad s(A)}{\quad s^{*^{-1}}(B)}.
1701: \end{eqnarray}
1702: Here $\lambda \in \F$ is such that $\mnorm(s(x))=\lambda \mnorm(x)$,
1703: and $s^*$ is the linear transformation adjoint to $s$ with respect
1704: to the trace bilinear form $(\cdot, \cdot)$.
1705: We will mostly
1706: use such transformations when $s\in \npg\J$ and so $\lambda =1$.
1707: 
1708: \item[]
1709: Finally, we have the transformation $\tau$, which acts by:
1710: \begin{equation} \label{tau}
1711: \tau :  \quad
1712: \mel\alpha\beta A B \ \mto \ \mel {-\beta} {\alpha} {-B} A.
1713: \end{equation}
1714: \end{itemize}
1715: 
1716: The transformations $\phi(\cdot)$ and $\psi(\cdot)$ are conjugate to each
1717: other by $\tau$:
1718: $$
1719: \tau \phi(C) \tau^{-1}=\psi(-C).
1720: $$
1721: In addition we have the following relations in $\gr$:
1722: \begin{equation}\label{comm-rel}
1723: \begin{array}{rcl}
1724: \phi(-\one)\psi(\one)\phi(-\one) &=& \tau\\
1725: \tau^2 &=& -\id_\m\\
1726: \tau T(s) &=& T(s^{*-1})\tau\\
1727: T(s)\phi(C) &=& \phi(\lambda^{-1}\, s(C))\, T(s)\\
1728: T(s)\psi(C) &=& \psi(\lambda\, s^{*-1}(C))\, T(s).
1729: %
1730: %
1731: \end{array}
1732: \end{equation}
1733: Here $C\in\J$ and $s\in \str\J$ satisfies
1734: $\mnorm(s(D))=\lambda \mnorm (D)$.
1735: 
1736: 
1737: Note that when matrices $C$ and $D$ above have rank 1, the transformations
1738: $\phi$ and $\psi$ have the following simpler form:
1739: \begin{eqnarray}\label{t1s}
1740: \phi(C) & : &
1741: \mel\alpha\beta A B\  \mto\
1742: \mel{\alpha+(B,C)}{\quad    \beta}{\quad A+\beta C}{\quad  B+A\cross C}\\
1743:      \label{t2s}
1744: \psi(D) & : &
1745: \mel \alpha \beta A B\  \mto \
1746: \mel \alpha{\quad \beta + (A,D)}{\quad A+ B\cross D}{\quad    B+\alpha D}.
1747: \end{eqnarray}
1748: 
1749: \medskip
1750: The following proposition gives a more precise description of the group
1751: $\gr$ and its representation $\mj$. We will let $\astr\J$ (respectively,
1752: $\ag\m$) denote the algebraic group whose group of $F'$-rational points is
1753: $\astr{\J\tensor_F F'}$ (respectively, $\ag{\m\tensor_F F'}$) for all extension
1754: fields $F'$ of $F$. The symbol $\aJ$ will denote the vector (algebraic)
1755: group defined by $\J$, i.e., $\aJ(F')=\J\tensor_F F'$ taken with the
1756: additive group structure.
1757: 
1758: 
1759: 
1760: 
1761: \begin{prop}\label{prop:gener}
1762: \
1763: Let $\J$ be a Jordan algebra $\h{\calg}$ with $\calg=F,\bin, \quat, \oct$
1764: over a field $F$ (char$F=p\ne 2,3)$
1765: and let $\m=\mj$.
1766: 
1767: \begin{itemize}
1768: 
1769: \item[{\rm (i)}]
1770: 
1771: Then the group $\gr$
1772: is generated by elements $\phi (C)$, $\psi (D)$, $\tf s$.
1773: 
1774: \item[{\rm (ii)}]
1775: The  group $\gr$ is the set of $F$-points
1776: of an absolutely almost simple linear algebraic group $\ag\m$,
1777: which is defined over $F$ and $F$-split.
1778: 
1779: The group $\gr$ is connected (except the case $\J=\h\bin$). It has
1780: a two-element center, and its quotient modulo the
1781: center is a simple group of adjoint type.
1782: 
1783: In the case $\J=\h\bin$ the same result is true for the connected component
1784: $\ag\m\cc$, which is a subgroup of index $2$ in $\gr$.
1785: 
1786: \item[{\rm (iii)}]
1787: The following table lists the types of the group $\ag\m$ as well
1788: as the highest weight of its irreducible representation on the space $\m$.
1789: 
1790: $$
1791: \begin{array}{|c|c|c|}
1792: \hline
1793: \qquad \J \qquad       &  \mbox{Type of }\ag\m & \mbox{H.w. of }\ \m\\
1794: \hline
1795: \h{\F}    &  C_3 & \omega_3\\
1796: \h{\bin}  &  A_5 & \omega_3\\
1797: \h{\quat} &  D_6 & \omega_5\ {\rm or}\ \omega_6\\
1798: \h{\oct}  &  E_7 & \omega_7\\
1799: \hline
1800: \end{array}
1801: $$
1802: 
1803: \end{itemize}
1804: 
1805: \end{prop}
1806: 
1807: 
1808: {\sc Proof.}
1809: 
1810: 
1811: (i) was proved in \cite[Theorem~3]{b}. The theorem is stated there when $\J$
1812: is $27$-dimensional exceptional Jordan algebra, but it remains true for
1813: other algebras in the list. We will give a somewhat different proof
1814: of this statement when we describe the generators of the group $\gr$ in
1815: the integral case (Proposition~$\ref{genz}$).
1816: 
1817: \medskip
1818: (ii)
1819: It is not hard to see that algebraic equations defining the group $\ag\m$
1820: have integer coefficients.
1821: Then reducing modulo $p$, we can assume that it is defined over the prime
1822: field $F_p$, and hence over~$F$.
1823: 
1824: The $F$-split torus in $\ag\m$ is the image of the diagonal split torus of
1825: $\astr\J$ under the mapping
1826: $T:\astr\J\to\ag\m$, see (\ref{t}). Such a torus has the
1827: ``right rank", and it was described explicitly in Remark~\ref{groups}
1828: (the case of $\J=\h\oct$ was considered in \cite[Theorem 3.5]{gar}).
1829: 
1830: Next we notice that ${\phi(\aJ)}, {\psi(\aJ)}, T{(\astr\J)}$
1831: are closed subgroups
1832: of $\ag{\m}$, since they are homomorphic images of the algebraic
1833: groups $\aJ, \aJ, \astr\J$, respectively.
1834: These groups are connected when
1835: $\J=\F, \quat, \oct$ (see Remark~\ref{groups}).
1836: Since by (i) the group $\ag\m$ is generated by these subgroups,
1837: it is connected in these cases.
1838: 
1839: If $\J=\bin$, the group ${\astr\J}$ has two connected components, and in fact
1840: is isomorphic to the semi-direct product ${\astr\J}\cc\rtimes C_2$, where
1841: $C_2$ is the group of order $2$ generated by the element $t$
1842: corresponding to the transpose operation (Remark~\ref{groups}(ii)).
1843: This fact and commutation relations~$(\ref{comm-rel})$ imply that
1844: $\ag\m$ may have one or two components. The {\em algebraic} subgroup $\alg{H}$,
1845: generated by ${\phi(\aJ)}, {\psi(\aJ)}, T({\astr\J}\cc)$,
1846: is connected and closed. In addition, it is known
1847: (see, e.g., \cite[Section~$7.5$]{hum})
1848: %
1849: that $\alg H$ is
1850: generated by ${\phi(\aJ)}, {\psi(\aJ)}, T({\str\aJ}\cc)$ as an
1851: abstract group.
1852: The analysis of the action of the group $\gr$
1853: in the $20$-dimensional module $\mj$ (cf.
1854: Example~\ref{ex:sl6}) shows that the element $T(t)$ does not lie
1855: in $\alg H$.
1856: Hence we have a decomposition into two cosets
1857: $\ag\m=\alg H \cup T(t)\alg H$, which implies that $\alg H$ is the connected
1858: component of $\ag \m$.
1859: 
1860: The group $\gr$ contains the central element $\tau^2$, which acts as
1861: the scalar $-1$
1862: on the module $\m$. The statement about the center and the simplicity
1863: of the quotient modulo the center was proved in \cite[Theorem~6]{b}
1864: (the theorem was stated there for
1865: $\J=\h\oct$, but it remains true for other cases as well).
1866: %
1867: 
1868: \medskip
1869: (iii) If char$F=0$,
1870: the group $\ag\m$ may be identified by its Lie algebra, which is
1871: the Tits-Kantor-Koecher construction of $\J$ \cite{koe}.
1872: 
1873: An analysis of a slightly different version of the Freudenthal construction
1874: may be found in \cite[$2.22-2.26$]{sp-alg}
1875: with the resulting groups being identified in Section
1876: $14.31$ of the same book.
1877: 
1878: The corresponding irreducible representation can be identified by its
1879: highest weight vector and its dimension. The highest weight vector in such
1880: a representation may be chosen to be $(1,0,0,0)$ in $\mj$.
1881: \ep
1882: 
1883: \begin{rem} \label{rem:semis}
1884: {\rm
1885: Another example of an admissible cubic Jordan algebra
1886: $$
1887: \J=F\oplus Q_n, n\ge 1
1888: $$
1889: was given in Example~$\ref{ex:quad}$.
1890: 
1891: In this case the Freudenthal construction
1892: (assuming the quadratic form has maximal Witt index)
1893: produces the semi-simple $F$-split
1894: algebraic groups of type $\SL_2\times \SO_{n+2}$ acting on the
1895: tensor product $V_2 \tensor V(\omega_1)$ of the
1896: irreducible $\SL_2$-module $V_2$ of dimension two and the irreducible
1897: $\SO(n+2)$-module of highest weight $\omega_1$ (and dimension $n+2$).
1898: 
1899: For small $n$ the action of the group $\gr$ in $\m$ is essentially isomorphic
1900: to the action of the group $G$ in $\m$ listed in the table below.
1901: 
1902: $$
1903: \begin{array}{|l|l|l|l|}
1904: \hline
1905: n & G & \m & {\rm Comment}\\
1906: \hline
1907: 1 & \SL_2\times \SL_2 & V_2\tensor {\rm Sym}^2 V_2 &
1908: using\ Spin_3\cong \SL_2\\
1909: \hline
1910: 2 & (\SL_2)^3 & V_2\tensor V_2\tensor V_2 &
1911: using\ Spin_4\cong \SL_2\times \SL_2\\
1912: \hline
1913: 4 & \SL_2\times \SL_4 & V_2\tensor  \wedge^2(F^4) &
1914: using\ Spin_6\cong \SL_4\\
1915: \hline
1916: \end{array}
1917: $$
1918: 
1919: These cases are listed separately in Table $1$.
1920: }\ee
1921: \end{rem}
1922: 
1923: \begin{rem}\label{badred}
1924: {\rm
1925: We note that when char\,$F=2$, the quartic form $q'$ will reduce to
1926: $$
1927: q'(x)=(\alpha\beta-(A,B))^2,\qquad
1928: x\in \m.
1929: $$
1930: Hence the group $\gr$ will become the group of transformations preserving
1931: the symplectic form $\{\cdot, \cdot\}$ and the quadratic form
1932: $q'_2=(\alpha\beta-(A,B))$. Next we notice that the linearization of
1933: $q'_2$ will produce exactly the symplectic form $\{\cdot, \cdot\}$.
1934: Hence the group $\gr$ in the case char\,$F=2$ coincides with the orthogonal
1935: group on the vector space $\m$ with respect to the quadratic form $q'_2$.
1936: }\ee
1937: \end{rem}
1938: 
1939: We will conclude this subsection with an explicit description of the
1940: module $\m=\mj$ and the group $\gr$ for the Jordan algebra
1941: $\J=M_3(F)\cong \h\bin$.
1942: We will let $\gr\cc$ denote the subgroup generated by
1943: $\phi(\J), \psi(\J), T(\str\J\cc)$ for $\J=M_3(F)$
1944: (cf. Remark~$\ref{groups}$(ii)). This is a subgroup of index two in $\gr$,
1945: and it coincides with the group of $F$-rational points of the connected
1946: component of the algebraic group $\ag\m$ (cf. Proposition~$\ref{prop:gener}$).
1947: 
1948: \begin{example} \label{ex:sl6}
1949: {\rm
1950: We let $\J=M_3(F)$, and
1951: we are going to describe how the action of the group $\gr$
1952: in the $20$-dimensional module $\m=\mj$ is related to the action
1953: of the group $\SL_6(F)$ in the space $\wedge^3 F^6$.
1954: 
1955: Note that the group $\SL_6(F)$ has a center isomorphic to $\bfmu_6(F)$,
1956: and the quotient $\SL_6(F)/\bfmu_3(F)I_6$ acts faithfully in
1957: $\wedge^3 F^6$. We are going to explicitly describe an isomorphism of
1958: vector spaces $\theta: \mj\to \wedge^3 F^6$ and an isomorphism of groups
1959: $\theta':\gr\cc\to \SL_6(F)/(\bfmu_3(F)I_6)$ satisfying:
1960: \begin{equation}\label{ex:sl3iso}
1961: \theta(g\cdot v)=\theta' (g)\cdot \theta(v)\quad
1962: \mbox{ for }g\in \gr\cc, v\in\mj.
1963: \end{equation}
1964: With a slight abuse of notation we will still use $6\times 6$ matrices to
1965: represent elements of the quotient $\SL_6(F)/\bfmu_3(F)I_6$.
1966: 
1967: We begin by introducing the following notation. Let
1968: $$
1969: \{e_1, e_2, e_3, f_1, f_2, f_3\}
1970: $$
1971: be a standard basis of $F^6=F^3\oplus F^3$. Next we introduce the
1972: following ``dual" set of linearly independent vectors in $\wedge^2 F^6$:
1973: $$
1974: e_1^*=e_2\wedge e_3,\quad
1975: e_2^*=e_3\wedge e_1,\quad
1976: e_3^*=e_1\wedge e_2,
1977: $$
1978: $$
1979: f_1^*=f_2\wedge f_3,\quad
1980: f_2^*=f_3\wedge f_1,\quad
1981: f_3^*=f_1\wedge f_2.
1982: $$
1983: 
1984: Then we define a correspondence $\theta$ between bases of $\m$ and
1985: $\wedge^3F^6$ by
1986: \begin{eqnarray}\label{ex:sl1}
1987: \theta: \hspace{2cm}
1988: \mel 1000 & \mapsto & e_1\wedge e_2 \wedge e_3,\nonumber \\
1989: \mel 0100 & \mapsto & f_1\wedge f_2 \wedge f_3,\\
1990: \mel 00{E_{ij}}0 & \mapsto & e_i\wedge f_j^*,\nonumber\\
1991: \mel 000{E_{ij}} & \mapsto & f_i\wedge e_j^*,
1992: \hspace{4cm} 1\le i,j\le 3.
1993: \nonumber
1994: \end{eqnarray}
1995: 
1996: Next we define the homomorphism $\theta'$. First we will do it assuming
1997: that every element of $F$ is a cube of another element of $F$.
1998: The group $\gr\cc$ is generated by transformations $\phi(\J), \psi(\J),
1999: T(\str\J\cc)$.
2000: First we define the map $\theta'$ for $\phi(\J), \psi(\J)$ by
2001: \begin{equation}\label{ex:sl2}
2002: \theta':\quad
2003: \phi(A)\mto
2004: \maketwoblock   {I_3}{0}A{I_3},\qquad
2005: \psi(B)\mto
2006: \maketwoblock   {I_3}B{0}{I_3},
2007: \end{equation}
2008: where $A,B\in\J=M_3(F)$, and each block in
2009: $\maketwoblock\cdot\cdot\cdot\cdot$ represents
2010: a matrix in $M_3(F)$.
2011: 
2012: Next we define a homomorphism
2013: $$
2014: \theta'': GL_3(F)\times GL_3(F)\to \SL_6(F)/\bfmu_3(F)I_6
2015: $$
2016: by
2017: \begin{equation}\label{ex:sl3}
2018: \theta'':
2019: (A,B)\ \mapsto\
2020: \zeta_A\zeta_B
2021: \maketwoblock{A/\det (A)}0
2022: 0{B/\det (B)},\qquad
2023: A,B\in \GL_3(F), \zeta_A, \zeta_B\in F^\mult,
2024: \end{equation}
2025: where $\zeta_A, \zeta_B$ are chosen so that
2026: $$
2027: \zeta_A^3=\det(A), \quad
2028: \zeta_B^3=\det(B).
2029: $$
2030: 
2031: Since the map $\theta''$ has values in $\SL_6(F)/\bfmu_3(F)I_6$,
2032: the relation $(\ref{ex:sl3})$ does not depend on the choice of the third
2033: roots $\zeta_A, \zeta_B$. We also note that $\ker\theta''=F(I_3, I_3)$.
2034: Recall that $\str\J\cc \cong GL_3(F)\times GL_3(F)/F^\mult(I_3, I_3)$
2035: (cf. Remark~$\ref{groups}$(ii)),
2036: and hence the map $\theta''$ produces a well defined homomorphism
2037: \begin{equation}\label{ex:sl31}
2038: \theta': T(\str\J\cc) \to \SL_6(F)/\bfmu_3(F)I_6.
2039: \end{equation}
2040: 
2041: A direct computation shows that the generators of $\gr\cc$ and
2042: $\SL_6(F)/\bfmu_3(F)I_6$
2043: associated via $(\ref{ex:sl2}),(\ref{ex:sl3}), (\ref{ex:sl31})$
2044: define the
2045: same linear transformations in the (isomorphic) vector spaces
2046: $\m$ and $\wedge^3 \F^6$.
2047: Hence the maps $\theta, \theta'$ define an ``isomorphism" of the pairs
2048: $(\gr, \m)$ and $(\SL_6(F)/\bfmu_3(F)I_6,\ \wedge^3 F^6)$ in the sense
2049: of~$(\ref{ex:sl3iso})$.
2050: 
2051: In the case of an arbitrary field $F$, the quantities $\zeta_A,
2052: \zeta_B$ are elements of a suitable field extension of $F$.
2053: However, since $\zeta_A^3, \zeta_B^3$ are elements of $F$ and
2054: the matrix in~$(\ref{ex:sl3})$ acts in $\wedge^3 F^6$ (not just in $F^6$),
2055: the expression~$(\ref{ex:sl3})$ will produce a well defined $F$-linear
2056: transformation of $\wedge^3 F^6$.
2057: \ee
2058: }
2059: \end{example}
2060: 
2061: \subsection{The rank of elements of $\m$}
2062: \label{ss:rank}
2063: 
2064: \begin{defn}\label{def:rank}
2065: Let $\m=\m(\J)$, where $\J$ is a cubic Jordan algebra. The {\em
2066: rank} of an element $x\in \m$ is an integer between $0$ and $4$,
2067: which is uniquely defined by the following relations:
2068: \begin{itemize}
2069: \item
2070: rank $x = 4$  iff \ $q(x)\ne 0$;
2071: \item
2072: rank $x \le 3$  iff \ $q(x)=0$;
2073: \item
2074: rank $x \le 2$  iff \ $T(x,x,x)=0$;
2075: \item
2076: rank $x \le 1$  iff \  $3\,T(x,x,y)+\{x,y\}\,x=0$
2077: for all $y\in \m$;
2078: \item
2079: rank $x = 0$ iff \ $x=0$.
2080: \end{itemize}
2081: \end{defn}
2082: 
2083: The expressions defining rank $2,3,$ and $4$ are quite natural; they
2084: appeared elsewhere before, see, e.g., \cite{clerc}. %
2085: However our coordinate-free
2086: relation defining $\rank x\le 1$ appears to be new in this
2087: context.
2088: 
2089: The following lemma is an immediate corollary of the definitions of $q(x)$
2090: and $T(x,x,x)$.
2091: 
2092: \begin{lemma}\label{lem:rank}
2093: The rank of elements is preserved under the action of the group $\gr$.
2094: \end{lemma}
2095: 
2096: The expression defining the rank of $x$ are given by homogeneous
2097: algebraic equations in terms of the coordinates of $x$. We list
2098: the appropriate polynomials in the following
2099: 
2100: \begin{rem}\label{rem:rp}
2101: {\rm
2102: These polynomials will be described in terms of their value at an
2103: arbitrary element $x\in \m=\mj$ of the form
2104: \begin{equation}\label{elem1}
2105: x=\mel\alpha \beta{\ A}B.
2106: \end{equation}
2107: 
2108: \begin{itemize}
2109: \item
2110: The {\em quartic rank polynomial} is the quartic form $q$:
2111: \begin{equation} \label{p4}
2112: q(x)= 8 (A^\#, B^\#) -8\alpha \mnorm(A) - 8\beta \mnorm(B)
2113: -2[(A,B)-\alpha\beta]^2.
2114: \end{equation}
2115: 
2116: \item
2117: It was computed in \cite{b} that
2118: \begin{eqnarray}  \label{p3} \label{txxx}
2119: T(x,x,x)        \nonumber
2120: &=& \biggl( -\alpha ^2 \beta + \alpha\, (A,B)
2121: - 2 \mnorm(B), \qquad
2122: \alpha \beta ^2 -\beta\,(A,B) + 2 \mnorm(A),  \\
2123: && \phantom{-}
2124: 2B \cross A^\#-2\beta\, B^\# -[(A, B) - \alpha\beta]A,\\
2125: &&                 \nonumber -2A\cross B^\#+2\alpha\, A^\#
2126: +[(A,B)-\alpha\beta]B \biggr).
2127: \end{eqnarray}
2128: 
2129: \item
2130: We define {\em quadratic rank polynomials} to be the following expressions:
2131: \begin{eqnarray} \label{p2}
2132: \alpha A-B^\#, \quad \beta B - A^\# ,\qquad Q(x),
2133: \end{eqnarray}
2134: where $Q$ is a quadratic polynomial function with values in End$(\J)$,
2135: i.e., $Q(x)$ is a linear transformation of $\J$ defined by
2136: \begin{equation}\label{q}
2137: Q(x) : \quad C \quad \mto\quad (\alpha\beta - (A,B)) C +2\, V_{A,B}(C),
2138: \qquad \quad C\in \J.
2139: \end{equation}
2140: where $V_{A,B}$ was defined in $(\ref{a:v})$.
2141: \ee
2142: \end{itemize}
2143: }
2144: \end{rem}
2145: 
2146: \begin{lemma}\label{lem:rank1}
2147: Let $x=\mel\alpha\beta A B$ be an element of $\m=\m(\J)$.
2148: The quadratic rank polynomials
2149: \begin{equation} \label{e1r1}
2150: \alpha A-B^\#, \quad \beta B - A^\#,\quad Q(\bx) \qquad
2151: \mbox{ are equal to zero}
2152: \end{equation}
2153: if and only if
2154: \begin{equation} \label{e2r1}
2155: 3\,T(x,x,y)+\{x,y\}\,x \quad
2156: \mbox{ is equal to zero \ \ for all }y\in \m.
2157: \end{equation}
2158: \end{lemma}
2159: {\sc Proof.}
2160: It was computed in \cite[p.~88]{b} that linearization of $T(x,x,x)$ yields
2161: \begin{eqnarray*}  \label{p3lin}
2162: 3\,T(x,x,y)
2163: &=&
2164: \Bigl( -2\alpha \beta \gamma -\alpha^2 \delta + \gamma\, (A,B)
2165: +\alpha\, (C,B) +\alpha\, (A,D) - 6 \mnorm(B,B,D), \qquad\\
2166: && \quad\
2167: \gamma \beta ^2 +2\alpha\beta\delta-\delta\,(A,B)-\beta\, (C,B)
2168: -\beta\, (A,D)+ 6 \mnorm(A,A,C),  \\[2mm]
2169: &&\ \
2170: -[(C, B)+(A,D)-\alpha\delta - \beta\gamma]A
2171: -(A,B)C +\alpha\beta C-\\
2172: &&\ \ %
2173: - 2\delta B^\# -2\beta B\cross D
2174: +2 D \cross A^\#+2B\cross(A \cross C),\\[2mm]
2175: &&\quad\
2176: [(C,B)+(A,D)-\alpha\delta - \beta\gamma]B+
2177: (A,B)D -\alpha\beta D+\\
2178: &&\quad\
2179: +2\gamma A^\#+2\alpha A\cross C
2180: -2C\cross B^\#-2A\cross(B\cross D)
2181:  \Bigr)
2182: \end{eqnarray*}
2183: for $x=\mel\alpha \beta A B,\ y=\mel \gamma\delta C D$.
2184: 
2185: We can rewrite it using the identity ~$(\ref{a:id3})$
2186: and the definition of the form $\{\cdot, \cdot\}$ as
2187: \begin{equation} \label{r1eq}
2188: 3\,T(x,x,y)+\{x,y\}\,x =
2189: \end{equation}
2190: \vspace{-8mm}
2191: \begin{eqnarray*} \label{r1int-main}
2192: &=&
2193: \Bigl(
2194: -[3\alpha\beta-(A,B)]\gamma+2(\alpha A -B^\#, D),\\
2195: && \quad\
2196: [3\alpha\beta-(A,B)]\delta-2(\beta B -A^\#, C),\\
2197: && \quad\
2198: [3\alpha\beta-(A,B)]C- 2 (\beta B-A^\#)\cross D
2199: +2(\alpha A-B^\#)\delta-2Q(x)(C),\\
2200: && \ \,
2201: -[3\alpha\beta-(A,B)]D+ 2 (\alpha A-B^\#)\cross C
2202: -2(\beta B-A^\#)\gamma+2Q(x')(D)
2203: \Bigr).
2204: \end{eqnarray*}
2205: %
2206: 
2207: Here $x'=\mel\beta \alpha B A$.
2208: 
2209: \medskip
2210: It follow from the last expression (using that char\,$F\ne 2$) that
2211: $$
2212: 3\,T(x,x,y)+\{x,y\}\,x=0 \quad
2213: \mbox{for any }y\in \m
2214: $$
2215: 
2216: if and only if
2217: $$
2218: \alpha A-B^\#=0, \quad \beta B - A^\# =0,\quad
2219: Q(x)=0,\ Q(x')=0, \quad
2220: 3\alpha\beta-(A,B)=0.
2221: $$
2222: 
2223: This proves the ``$\Leftarrow$" implication of the lemma.
2224: 
2225: \medskip
2226: 
2227: Now suppose (\ref{e1r1}) holds. To prove ``$\Rightarrow$",
2228: we need to show that $Q(x')=0$ and $3\alpha\beta-(A,B)=0$.
2229: 
2230: We start with the latter one. We have $Q(x)(C)=0$ for any $C\in\J$.
2231: In particular, this is true for $C=\one$.
2232: We then compute using~(\ref{q}) and~$(\ref{traces})$:
2233: $$
2234: 0=\tr(Q(x)(\one))=3\alpha\beta-3(A,B)+2\tr(A\jprod  B)=
2235: 3\alpha\beta-3(A,B)+2(A, B)=3\alpha\beta-(A,B).
2236: $$
2237: 
2238: To prove $Q(x')=0$ we use the definition of the Jordan triple
2239: product~$(\ref{jtp})$ to notice that
2240: $$
2241: Q(x)(C)-Q(x')(C)=2Q(x)(\one)\jprod C,
2242: $$
2243: and the statement again follows from the fact that $Q(x)(C)=0$
2244: for any $C\in \J$.
2245: \ep
2246: 
2247: 
2248: \subsection{The canonical form in the case of a field}
2249: \label{ss:canf}
2250: 
2251: \begin{lemma} \label{bl-f}
2252: Let $\J$ be as in Proposition~$\ref{prop:gener}$ and $\m=\mj$.
2253: Every non-zero element of the module $\m$ can be brought to the form
2254: $$
2255: \mel 1 \beta {\ A} 0
2256: \qquad for\ some\ \beta\in \F,\ A\in\J
2257: $$
2258: by an appropriate element of $\gr$.
2259: \end{lemma}
2260: 
2261: {\sc Proof.} We start with an arbitrary element $x_1$ of the form
2262: $\mel {\alpha_1}{\beta_1} {A_1} {B_1}$.
2263: 
2264: First we show that we can transform it to an element in which the
2265: component $B_1$
2266: is not zero. If $B_1\ne 0$, there is nothing to do. If $A_1 \ne
2267: 0$, the transformation $\tau$ does the trick.
2268: %
2269: Now let us assume $A_1= B_1 = 0$ in $x_1$. After application of $\tau$ if
2270: necessary, we may assume that $\alpha _1\ne 0$. Then applying $\psi (D)$
2271: with any non-zero $D$, we get an element $x_2$ in which the
2272: component at the position $B_1$ is not zero.
2273: 
2274: 
2275: Thus we proceed with an element $x_2$ of the form $(\alpha _2,\beta_2,
2276: A_2, B_2)$ with $B _2\ne 0$. If $\alpha _2\ne 1$ we argue as follows.
2277: Since elements of $\J$ of rank $1$ span the whole $\J$, and the trace form
2278: in $\J$ is non-degenerate, we may find an element $C\in\J$ of rank 1, such
2279: that $(B_2,C)\ne 0$. Scaling $C$, we may assume that $\alpha_2+(B_2,C)=1$.
2280: The condition ${\rm rank}\,C=1$ implies $C^\# =0$ and $N(C)=0$. After
2281: applying the transformation $\phi (C)$ to $x_2$, we get an element $x_3$
2282: in which the first component is equal to $\alpha _2+(B_2,C)=1$.
2283: 
2284: We arrived at the element $x_3$ of the form
2285: $$
2286: \mel 1 {\beta_3}  {A_3} {B_3}.
2287: $$
2288: 
2289: The application of $\psi(D)$ with $D=-B_3$ brings this element to the
2290: desired form. \ep
2291: 
2292: \begin{lemma}
2293: \label{lcomp} An element
2294: $$
2295: \mel \alpha  \beta {\ \diag{a_1}{a_2}{a_3}} 0, \quad\alpha\ne 0
2296: $$
2297: can be mapped by a transformation in $\gr$ to the elements
2298: \begin{itemize}
2299: \item[\rm(i)]
2300: \hspace{2cm} $
2301: \mel \alpha {\ \beta-2\frac{a_2a_3}{\alpha}\,c}
2302: {\quad \diag{a_1+\beta c-\frac{a_2a_3}{\alpha}\, c^2}{\ a_2}{\ a_3}} 0,
2303: $
2304: \item[\rm(ii)]
2305: \hspace{2cm} $
2306: \mel \alpha {\ \beta-2\frac{a_1a_3}{\alpha}\,c}
2307: {\quad \diag{a_1}{\ a_2+\beta c-\frac{a_1a_3}{\alpha}\, c^2}{\ a_3}} 0, $
2308: \item[\rm(iii)]
2309: \hspace{2cm} $
2310: \mel \alpha {\ \beta-2\frac{a_1a_2}{\alpha}\,c}
2311: {\quad \diag{a_1}{\ a_2}{\ a_3+\beta c-\frac{a_1a_2}{\alpha}\, c^2}} 0, $
2312: \end{itemize}
2313: where $c$ is an arbitrary element in $\F$.
2314: 
2315: This lemma is also valid in the case $F=\zet$ assuming
2316: $\alpha|a_2, \alpha|a_3$ in {\rm (i)};
2317: $\alpha|a_1, \alpha|a_3$ in {\rm (ii)};
2318: $\alpha|a_1, \alpha|a_2$ in {\rm (iii)}.
2319: \end{lemma}
2320: 
2321: {\sc Proof.} The proof of this lemma is a direct computation. The action
2322: of
2323: $$
2324: \phi (C),\ C=\diag{c}00
2325: $$
2326: and then
2327: $$
2328: \psi(D),\ D=-\frac{1}{\alpha}\, \diag0{a_3c}{a_2c}
2329: $$
2330: gives the first element. The appropriate modifications of these yield the
2331: other two elements. \ep
2332: 
2333: \begin{rem} \label{rem1}
2334: {\rm
2335: When the component $B$ of an element of the form $(\ref{elem1})$ in $\m$ is
2336: equal to zero, the relations $(\ref{p4})-(\ref{p2})$ in the definition of
2337: the rank have the following simpler form
2338: \begin{itemize}
2339: \item
2340: rank $x \le 1$  iff
2341: \begin{equation} \label{r1s}
2342: \alpha A=0,\quad A^\# =0,\quad \alpha\beta=0;
2343: \end{equation}
2344: \vspace{-7mm}
2345: \item
2346: rank $x \le 2$  iff
2347: \begin{equation} \label{r2s}
2348: \alpha ^2\beta=0, \quad \alpha\beta ^2 +2N(A)=0, \quad \alpha\beta
2349: A=0,\quad \alpha A^\# =0;
2350: \end{equation}
2351: \vspace{-7mm}
2352: \item
2353: rank $x \le 3$  iff
2354: \begin{equation} \label{r3s}
2355: 8\alpha N(A)+2\alpha ^2\beta ^2=0\quad (i.e., q(x)=0);
2356: \end{equation}
2357: \vspace{-7mm}
2358: \item
2359: rank $x = 4$  iff
2360: \begin{equation} \label{r4s}
2361: q(x)\ne 0.
2362: \end{equation}
2363: \end{itemize}
2364: }\ee
2365: \end{rem}
2366: 
2367: \begin{theorem} \label{thm1}
2368: \
2369: 
2370: \begin{enumerate}
2371: \item[\rm (i)]
2372: {
2373: Let $F$ be a field of char\ $\ne 2,3$, let $\calg$ be the split composition algebra
2374: $\bin, \quat, \oct$ over $F$, and let $\J=\h\calg$.
2375: Let $(G, \m)$ be the pair (group, module) produced from $\J$
2376: by the Freudenthal construction.
2377: 
2378: Then
2379: 
2380: \begin{itemize}
2381: \item
2382: There exists a $G$-invariant quartic form (the norm) on the module $\m$.
2383: 
2384: \item
2385: The group $G$ acts
2386: transitively on the sets of elements of rank $1,2$, and $3$ in the module
2387: $\m$.
2388: 
2389: \item
2390: In the case of rank\/ $4$ the group $G$ acts transitively on the set of
2391: elements of a given norm $k$, for any $k\in F$, $k\ne 0$.
2392: \end{itemize}
2393: All these orbits are distinct, and the union of these orbits and $\{0\}$
2394: is the whole module~$\m$.
2395: }
2396: 
2397: \item[\rm (ii)]
2398: If in addition every element of $F$ is a square,
2399: then the same results apply to the pair $(G, \m)$ obtained
2400: from the Jordan algebra $\J=\h F$.
2401: \end{enumerate}
2402: This construction yields the simple algebraic groups and their irreducible
2403: representations described in Proposition~$\ref{prop:gener}$.
2404: \end{theorem}
2405: 
2406: {\sc Proof.}
2407: 
2408: The $G$-invariant quartic form $q$ defined on the space $\m$ is a part of
2409: the Freudenthal construction. One defines the rank of elements of $\m$ using the
2410: quartic form (Definition~$\ref{def:rank}$).
2411: 
2412: We are going to show that the group $\gr$ acts transitively on the set of
2413: elements of a given rank/given norm. Namely, we will show that every
2414: element of $\m$ is $\gr$-equivalent to one of the following elements
2415: in the ``canonical" form:
2416: \begin{eqnarray}
2417: \label{r1}
2418: \rank 1 & : & \mel 10 {\diag 000} 0,\\
2419: \label{r2}
2420: \rank 2 & : & \mel 10 {\diag 100} 0,\\
2421: \label{r3}
2422: \rank 3 & : & \mel 10 {\diag 110} 0,\\
2423: \label{r4}
2424: \rank 4 & : & \mel 10 {\diag 11k} 0, \quad k\in F, k\ne 0.
2425: \end{eqnarray}
2426: These elements have distinct rank/norm,
2427: and therefore by Lemma~$\ref{lem:rank}$ and the definition of $\gr$
2428: they lie in distinct orbits.
2429: 
2430: 
2431: We start with an arbitrary non-zero element $x$ in $\m$. By
2432: Lemma~\ref{bl-f} we can transform $x$ to the element
2433: \begin{equation} \label{semican}
2434: x_1=\mel 1 \beta {\ A} 0.
2435: \end{equation}
2436: We have rank\,$x$=rank\,$x_1$, and we can use Remark~\ref{rem1} when
2437: computing the rank of $x_1$. We also make use of the rank of elements
2438: of $\J$ (Definition~$\ref{def:jrank}$).
2439: \medskip
2440: 
2441: If rank\,$x_1=1$, then it follows from Remark~\ref{rem1} that $\beta=0$
2442: and $A=0$. So the element $x_1$ is already in the form~$(\ref{r1})$.
2443: 
2444: If rank\,$x_1=2$, then it follows from Remark~\ref{rem1} that $\beta=0$.
2445: The condition $\alpha A^\# =0$ of~$(\ref{r2s})$ implies rank\,$A=1$.
2446: By Proposition~$\ref{prop:np-o}$
2447: there exists $s\in \npg\J$ such that
2448: $$
2449: s(A)=\diag{1}00.
2450: $$
2451: The action of the transformation $\tf{s}\in \gr$ on the element $x_1$
2452: brings it to the form~$(\ref{r2})$ as desired.
2453: \medskip
2454: 
2455: Let us now consider the case of rank\,$x_1=3$. We start by showing that we
2456: can assume that the $\beta$-component of $x_1$ is zero. If $\beta\ne 0$,
2457: it follows from $(\ref{r3s})$ that
2458: $$
2459: N(A)=-\frac{1}{4}\beta ^2\ne 0.
2460: $$
2461: Again, by Proposition~$\ref{prop:np-o}$
2462: %
2463: we can find an element $s\in \npg\J$ that brings $A$ to the diagonal
2464: form:
2465: $$
2466: s(A)=\diag{1}1n\quad {\rm with\ } n=N(A).
2467: $$
2468: And hence we have
2469: $$
2470: \tf{s}(x_1)=\mel 1 \beta {\ \diag{1}1n} 0.
2471: $$
2472: Taking $c=\beta/2$ in Lemma~\ref{lcomp}(iii) we see that we can map
2473: %
2474: the above element to the element
2475: $$
2476: x_2=\mel 1 0 {\ A'} 0
2477: $$
2478: with component $\beta$ equal to zero.
2479: 
2480: We still have $\rank x_2=3$, and
2481: it follows from $(\ref{r3s})$ that $N(A')=0$. On the other hand, $A^\#
2482: \ne 0$, since the opposite would imply that rank\,$x_2 \le 2$. Hence
2483: rank\,$A'=2$. Again we can find an $s'\in \npg\J$ that brings $A'$ to the
2484: diagonal form:
2485: $$
2486: s'(A')=\diag{1}10,
2487: $$
2488: and hence
2489: $$
2490: \tf{s'}(A')=\mel 1 0 {\ \diag{1}10} 0.
2491: $$
2492: This is an element of the form~$(\ref{r3})$,
2493: and we are done with the case of rank 3.
2494: 
2495: \medskip
2496: The last case is the case of rank $4$.
2497: 
2498: We are still working with an element $x_1$ of the form $(\ref{semican})$.
2499: Since $\rank x_1=4$, we have $q(x_1)\ne 0$.
2500: 
2501: There are three possible subcases for the element
2502: $$
2503: x_1=\mel 1 \beta {\ A} 0.
2504: $$
2505: 
2506: \medskip
2507: {\em Subcase $1$.} $\beta =0$.
2508: 
2509: In this subcase the expression for the norm becomes
2510: $$
2511: q(x_1)=-8\, N( A).
2512: $$
2513: Since $q(x_1)\ne 0$, we have $N(A)\ne 0$, and we can find an $s\in \npg\J$
2514: that brings $A$ to the form $\diag{1}1k$, $k\ne 0$. It follows that
2515: $\tf{s}(x_1)$ has the form~$(\ref{r4})$ as desired.
2516: 
2517: \medskip
2518: {\em Subcase $2$.} $\beta \ne 0$ and rank\,$A\ge 2$.
2519: 
2520: We start by bringing $A$ to the diagonal form of Proposition~$\ref{prop:np-o}$
2521: by an appropriate
2522: element $s\in \npg\J$. Then $s(A)$ has the form $\diag{1}1n$, $n\in \F$. We
2523: can apply Lemma~\ref{lcomp}(iii) with $c=\beta /2$ to the element
2524: $\tf{s}(x_1)$, and get a new element with $\beta$-component equal to zero.
2525: This brings us to the subcase 1, and we are done.
2526: 
2527: \medskip
2528: {\em Subcase $3$.} $\beta \ne 0$ and rank\,$A\le 1$.
2529: 
2530: We again start by bringing $A$ to the diagonal form by an
2531: appropriate element $s\in \npg\J$. Then $s(A)$ has the form
2532: $\diag{\epsilon}00$, $\epsilon=0$ or $1$ depending on the rank\,$A$. We
2533: can apply Lemma~\ref{lcomp}(ii) and (i) (if necessary) with $c=1/\beta$,
2534: and obtain element
2535: $$
2536: x_2=\mel \alpha \beta {\diag 110} 0.
2537: $$ This brings us to
2538: the subcase 2, and we are done again.
2539: 
2540: We showed that an arbitrary non-zero element can be brought to one of the
2541: elements in the canonical form~$(\ref{r1})-(\ref{r4})$.
2542: Hence the union of the orbits described in the hypothesis
2543: is the whole~$\m$.
2544: \ep
2545: 
2546: \section{The Integral version of the Freudenthal construction}
2547: \subsection{The integral structure in the module $\mj$}
2548: \label{ss:intstr}
2549: 
2550: In this section we consider the integral structure $\mz$ in the module
2551: $\m$ and the integral form $\grz$ of the group $\gr$ that acts on it.
2552: We define
2553: \begin{equation}\label{mz}
2554: \mz\defeq\m(\jz)=\zet\oplus\zet\oplus\jz\oplus\jz,
2555: \end{equation}
2556: where $\jz$ is one of
2557: \begin{equation} \label{zlist}
2558: \jz:\qquad
2559: \h{\octz},\  \h{\quatz},\  \h{\binz},\
2560: \h{\zet},\  \zet\oplus \zet \oplus \zet,
2561: \end{equation}
2562: see Subsection~$\ref{ss:33int}$
2563: 
2564: Most of the definitions and assertions of this section apply also in the
2565: cases
2566: $$
2567: \zet\ \mbox{ and }\ \zet\oplus Q_n(\zet) \quad
2568: \mbox{(cf.
2569: Remark~\ref{rem:semis})}.
2570: $$
2571: However these cases require somewhat different treatment and
2572: we will not consider them in this paper.
2573: 
2574: 
2575: An arbitrary element of the module $\mz$ has the form
2576: $$
2577: \mel\alpha\beta AB,
2578: \qquad
2579: \mbox{where } \alpha, \beta \in \zet \mbox{ and } A,B \in \jz.
2580: $$
2581: 
2582: Since $\jz$ is closed under $\#$, and $\tr$ and $\btr(\cdot, \cdot)$ are
2583: integer-valued,
2584: we have that the quartic form $q$, the modified form $q'$
2585: (see \ref{qmod})), and the skew-symmetric form~$\{\cdot,\cdot\}$ have integer
2586: values on $\mz$. The expression~$(\ref{txxx})$ implies
2587: that $T(x,x,x)\in \mz$ for $x\in\mz$.
2588: 
2589: \begin{defn}
2590: {\rm
2591: We define the {group $\grz$} to be the group of invertible
2592: $\zet$-linear transformations of $\mz$ that preserve the quartic
2593: form $q$ and the bilinear form $\{\cdot, \cdot \}$.
2594: }
2595: \end{defn}
2596: It is immediate that transformations
2597: \begin{equation} \label{elz}
2598: \phi(C), \psi(D),\ \tf{s},\ \tau \quad
2599: {\rm with}\ C,D\in \jz,\ s\in \str{\jz}
2600: \end{equation}
2601: preserve $\mz$, and therefore lie in the group $\grz$.
2602: 
2603: \begin{defn}
2604: {\rm
2605: We will call the transformations of the form $(\ref{elz})$ the
2606: {\em elementary\/} transformations of $\mz$.
2607: }
2608: \end{defn}
2609: 
2610: If $F$ is the field of characteristic zero, then
2611: the group $\grz$ may also be defined as the set of integral points
2612: of the group $\gr$ with respect to the integral structure on the
2613: module $\m$ determined by (\ref{mz}).
2614: 
2615: We will prove in Proposition~$\ref{genz}$ below that $\grz$ is generated by
2616: the elementary transformations~$(\ref{elz})$.
2617: We will make use of this assertion in the next examples, which give an
2618: explicit description of $\mz$ and $\grz$ in two special cases.
2619: 
2620: \begin{example} \label{ex:sl6int}
2621: {\rm
2622: This example gives a  description of the integral version
2623: of Example~$\ref{ex:sl6}$.
2624: Here we have
2625: $\jz=M_3(\zet)\cong\h\binz$ and
2626: $\mz=\m(\jz)=\zet\oplus\zet\oplus\jz\oplus\jz$.
2627: By analogy with Example~$\ref{ex:sl6}$, we define the group $\grz\cc$ to
2628: be the subgroup of $\grz$ generated by
2629: $\phi(\jz)$, $\psi(\jz)$, $T(\str\jz\cc)$.
2630: We are going to establish the isomorphism of the integral pairs
2631: $(\grz\cc, \jz)\cong(\SL_6(\zet), \wedge^3\zet^6)$ in the sense
2632: of~$(\ref{ex:sl3iso})$
2633: 
2634: We keep the notation for bases and define the isomorphism $\theta$ of
2635: $\zet$-modules $\mz$ and $\wedge^3 \zet^6$ as in Example~$\ref{ex:sl6}$.
2636: 
2637: We have $\bfmu_3(\zet)=\{1\}$ and hence
2638: $\SL_6(\zet)=\SL_6(\zet)/\bfmu_3(\zet)I_6$ acts faithfully in
2639: $\wedge^3 \zet^6$.
2640: We are going to define the map
2641: $\theta':\grz\cc\to\SL_6(\zet)$ in a way similar to Example~$\ref{ex:sl6}$.
2642: 
2643: 
2644: First we define $\theta'$ for $\phi(\jz), \psi(\jz)$ by
2645: \begin{equation}\label{ex:sl2int}
2646: \theta':\quad
2647: \phi(A)\mto
2648: \maketwoblock   {I_3}{0}A{I_3},\qquad
2649: \psi(B)\mto
2650: \maketwoblock   {I_3}B{0}{I_3},
2651: \end{equation}
2652: where $A,B\in\jz=M_3(\zet)$.
2653: 
2654: As for the group $\str\jz\cc$, we notice that
2655: it consists of the transformations of the form
2656: $$
2657: X\mapsto AXB^{-1}
2658: $$
2659: with $X\in\jz, A,B\in \GL_3(\zet)$ and $(A,B)$ acting trivially iff
2660: $A=B=\pm I_3$ (cf. Remark~$\ref{groups}$(ii)).
2661: 
2662: Next we will define a homomorphism
2663: $$
2664: \theta'': \GL_3(\zet)\times \GL_3(\zet)\to \SL_6(\zet).
2665: $$
2666: Since in this case the determinants of the matrices involved are $\pm 1$,
2667: we can define $\theta''$ by the following simpler formula:
2668: \begin{equation}\label{ex:sl3int}
2669: \theta'':
2670: (A,B)\ \mapsto\ \maketwoblock{(\det B)A}0
2671: 0{(\det A)B},\qquad
2672: A,B\in \GL_3(\zet).
2673: \end{equation}
2674: 
2675: The map $\theta''$ factors through the quotient
2676: $\GL_3(\zet)\times \GL_3(\zet)/\{\pm(I_3, I_3)\}$ and yields a well-defined
2677: map
2678: \begin{equation}\label{ex:sl31int}
2679: \theta': T(\str\jz\cc)\to \SL_6(\zet).
2680: \end{equation}
2681: 
2682: A direct computation shows that the generators of the groups $\grz\cc$ and
2683: $\SL_6(\zet)$
2684: associated via $(\ref{ex:sl2int}),(\ref{ex:sl3int}), (\ref{ex:sl31int})$
2685: define the
2686: same linear transformations in the (isomorphic) $\zet$-modules
2687: $\mz$ and $\wedge^3 \zet^6$.
2688: Hence the maps $\theta, \theta'$ define an ``isomorphism" of the pairs
2689: $(\grz\cc, \mz)$ and $(\SL_6(\zet),\ \wedge^3 \zet^6)$ as desired.
2690: 
2691: It will be proved in Proposition~$\ref{genz}$ that every element in $\grz$
2692: is a product of the elementary transformations~$(\ref{elz})$. Commutation
2693: relations~$(\ref{comm-rel})$ imply that every element in $\grz$ has
2694: the form $g_0$ or $T(t) g_0$, where $g_0\in \grz\cc$ and $t$ is the
2695: transpose operation, see Remark~$\ref{groups}$(ii).
2696: Hence we have $\grz=\grz\cc \cup T(t)\grz\cc$.
2697: \ee
2698: }
2699: \end{example}
2700: 
2701: \begin{example}\label{ex:FFF}
2702: {\rm
2703: Let $\J=F\oplus F \oplus F$ and hence $\jz=\zet\oplus\zet\oplus\zet$,
2704: $\mz=\m(\jz)=\zet\oplus \zet\oplus \jz\oplus \jz$.
2705: In this example we are going to explicitly describe the relation between
2706: the action of the group $\grz$ in the $8$-dimensional module $\mz$ and the
2707: action of the group $\SL_2(\zet)\times \SL_2(\zet)\times \SL_2(\zet)$ in the space
2708: $\zet^2\tensor \zet^2 \tensor \zet^2$.
2709: 
2710: We will use the more compact notation $\Gamma$ to denote the
2711: group $\SL_2(\zet)\times \SL_2(\zet)\times \SL_2(\zet)$.
2712: Note that the group
2713: $\Gamma$
2714: has a center that consists of the matrices $\{(\pm I_2, \pm I_2, \pm I_2)\}$,
2715: and the quotient
2716: $\Gamma/K_4$ acts faithfully in
2717: $\zet^2\tensor \zet^2 \tensor \zet^2$, where
2718: $$
2719: K_4=\{(I_2, I_2, I_2),(I_2, -I_2, -I_2),(-I_2, I_2, -I_2),(-I_2, -I_2, I_2),\}.
2720: $$
2721: 
2722: 
2723: Let $G_0$ be the subgroup in $\grz$ generated by transformations
2724: $\phi(\jz)$, $\psi(\jz)$, and $T(\pm \id_\jz)$.
2725: We are going to describe explicitly an isomorphism of $\zet$-modules
2726: $\theta:\mz \to \zet^2\tensor\zet^2\tensor\zet^2$ and an isomorphism
2727: of groups
2728: $\theta':G_0 \to\Gamma/K_4$ satisfying
2729: \begin{equation}\label{ex:F0}
2730: \theta(g\cdot v)=\theta' (g)\cdot \theta(v)\quad
2731: \mbox{ for }g\in G_0, v\in\mz.
2732: \end{equation}
2733: With a slight abuse of notation we will use triples of $2\times 2$
2734: matrices to
2735: represent elements of the quotient $\Gamma/K_4$.
2736: 
2737: 
2738: First we will establish an isomorphism of $\zet$-modules
2739: $\mz\cong\zet^2\tensor\zet^2\tensor\zet^2$.
2740: 
2741: 
2742: 
2743: Let $\{E_1, E_2, E_3\}$
2744: be the standard basis of $\jz=\zet\oplus\zet\oplus\zet$,
2745: and let $\{e_1, e_2\}$ be the standard basis of $\zet^2$.
2746: We define $\theta$ on bases and extend it by $\zet$-linearity
2747: \begin{eqnarray} \label{ex:F1}
2748: &&
2749: \mel 1000 \mapsto e_1\tensor e_1\tensor e_1,\quad\ \
2750: \mel 0100 \mapsto e_2\tensor e_2\tensor e_2,\nonumber \\
2751: &&
2752: \mel 00{E_1}0 \mapsto e_1\tensor e_2\tensor e_2,\quad
2753: \mel 00{E_2}0 \mapsto e_2\tensor e_1\tensor e_2,\quad
2754: \mel 00{E_3}0 \mapsto e_2\tensor e_2\tensor e_1,\quad\\
2755: &&
2756: \mel 000{E_1} \mapsto e_2\tensor e_1\tensor e_1,\quad
2757: \mel 000{E_2} \mapsto e_1\tensor e_2\tensor e_1,\quad
2758: \mel 000{E_3} \mapsto e_1\tensor e_1\tensor e_2.\nonumber
2759: \end{eqnarray}
2760: 
2761: Next we define the map $\theta'$ on generators of $G_0$
2762: (and onto generators of $\Gamma/K_4$) by
2763: \begin{eqnarray} \label{ex:F2}
2764: \phi(a_1, a_2, a_3) & \mapsto&
2765: \Biggl(
2766: {%
2767: \maketwo{1}{a_1}{0}{1},
2768: \maketwo{1}{a_2}{0}{1},
2769: \maketwo{1}{a_3}{0}{1}}
2770: \Biggr)\nonumber \\
2771: \psi(a_1, a_2, a_3) & \mapsto &
2772: \Biggl(
2773: \maketwo{1}{0}{a_1}{1},\maketwo{1}{0}{a_2}{1},\maketwo{1}{0}{a_3}{1}
2774: \Biggr)\\
2775: T(\pm \id_\jz) & \mapsto &
2776: \pm
2777: \Bigl(
2778: I_2, I_2, I_2
2779: \Bigr),\nonumber
2780: \end{eqnarray}
2781: where $(a_1, a_2, a_3)\in \jz$.
2782: 
2783: A direct computation shows that the generators of $G_0$ and
2784: $\Gamma/K_4$
2785: associated via $(\ref{ex:F2})$
2786: define the
2787: same linear transformations in the (isomorphic) $\zet$-modules
2788: $\mz$ and $\zet^2\tensor \zet^2\tensor \zet^2$.
2789: Hence the maps $\theta, \theta'$ define an ``isomorphism" of the pairs
2790: $(G_0, \mz)$ and $(\Gamma/K_4,\ \zet^2\tensor \zet^2\tensor \zet^2)$
2791: in the sense of~$(\ref{ex:F0})$.
2792: }\ee
2793: \end{example}
2794: 
2795: 
2796: \subsection{Reduction in the module $\m(\jz)$}
2797: \label{ss:red}
2798: 
2799: \begin{defn}
2800: {\rm
2801: Given a free $\zet$-module $M$,
2802: we say that an integer $d$ {\em divides} an
2803: element $x\in M$ if $x=dx'$ for some $x'\in M$.
2804: 
2805: We define the $\gcd$ of a collection of elements in
2806: $M$ to be the greatest integer that
2807: divides all these elements. By definition the $\gcd$ is  a positive
2808: integer, if at least one of the elements is nonzero.
2809: }
2810: \end{defn}
2811: 
2812: \begin{defn}\label{prim}
2813: {\rm
2814: An element $x$ of a free $\zet$-module $M$ is said to be {\em primitive},
2815: if $\gcd (x)=1$.
2816: }
2817: \end{defn}
2818: Evidently, for any non-zero element $x$ we have
2819: $
2820: x = \gcd(x)\ x',
2821: $
2822: where $x'$ is primitive.
2823: 
2824: \begin{lemma} \label{gcd}
2825: Let $M$ be a free $\zet$-module and
2826: let $g$ be an element in {\rm End}${}_\zet(M)$ such that $g^{-1}$ exists and
2827: also lies in {\rm End}${}_\zet(M)$.
2828: Then for any $x$ in $M$ and any non-zero integer\/ $d$
2829: $$
2830: d\ divides\ x\qquad
2831: \mbox{\rm if and only if}\qquad
2832: d\ divides\ g(x).
2833: $$
2834: \end{lemma}
2835: 
2836: {\sc Proof.}
2837: Obvious.
2838: 
2839: \bigskip
2840: 
2841: %
2842: %
2843: %
2844: %
2845: 
2846: \begin{defn}\label{def:reduced}
2847: {\rm
2848: An element $x$ of the module $\mz=\m (\jz)$ is said to be {\em reduced},
2849: if it is of the form
2850: \begin{equation} \label{reduced}
2851: x=\mel\alpha\beta{\ A}0,\qquad
2852: \alpha,\beta\in \zet,\ A\in\jz,
2853: \end{equation}
2854: with $\alpha>0,\quad \alpha |\beta, \ \alpha |A$. We say that $x$ is a
2855: {\em diagonal reduced} element, if in addition $A$ is a diagonal matrix.
2856: }
2857: \end{defn}
2858: 
2859: We note that for a reduced element $x$ as in (\ref{reduced}) we have
2860: $\gcd(x)=\alpha$. And a reduced $x$ is primitive if and only if
2861: $\alpha=1$.
2862: 
2863: 
2864: 
2865: %
2866: %
2867: 
2868: \begin{lemma}[\sc Reduction Lemma I]  \label{bl-fz}
2869: \
2870: 
2871: \hspace{-\parindent}%
2872: Let $\jz$ be one of $\h{\binz}, \h{\quatz}, \h{\octz}, \zet\oplus\zet\oplus\zet$.
2873: Every non-zero element ${\bx}$ of the module $\mz=\m (\jz)$
2874: is equivalent to a {\em diagonal reduced} element under the action of
2875: a series of elementary transformations
2876: \begin{equation} \label{elem-tr}
2877: \phi (C), \psi(D), \tf{s}, \tau \quad
2878: with \ C, D \in \jz,\ s\in \npg{\jz}.
2879: \end{equation}
2880: \end{lemma}
2881: 
2882: {\sc Proof.}
2883: 
2884: {\bf (a)}
2885: First we are going to prove the assertion when $\jz$
2886: is one of $\h{\binz}, \h{\quatz}, \h{\octz}$.
2887: 
2888: We are going to apply transformations $\phi (C), \psi(D), \tf{s}$
2889: to obtain a reduction
2890: procedure similar to the Gaussian elimination process for
2891: regular integer matrices.
2892: 
2893: The proof consists of two parts.
2894: We describe four reduction steps in the first part, and in the second part
2895: we show how we use these steps to bring an arbitrary non-zero element to a
2896: reduced form.
2897: 
2898: {\bf Part I.}
2899: 
2900: We start the proof of the lemma by explaining how we can use the
2901: elementary transformations~(\ref{elem-tr}) to get the reduction.
2902: We will use the
2903: relations $(\ref{phi})$, $(\ref{psi})$, and the ability to bring any
2904: element in $\jz$ to the (diagonal) Smith normal form by an element
2905: in $\npg \jz$ (see Theorem~\ref{np-orbits}).
2906: Note that transformations $s\in \npg\jz$ preserve the norm, and hence
2907: we have $\lambda=1$ in (\ref{t}).
2908: 
2909: The four reduction steps are summarized below:
2910: \medskip
2911: 
2912: (RED1): \
2913: $\mel\alpha\beta{\ A}B \longrightarrow \mel{g_1}**0$
2914: with $0<g_1\le\ \min \{|\alpha|,\gcd( B)\}$
2915: \medskip
2916: 
2917: (RED2): \
2918: Assume $\alpha>0$, \ $\alpha \not | \gcd(\beta, A)$;
2919: 
2920: \ \phantom{(RED2): }%
2921: $\mel\alpha\beta{\ A}0 \longrightarrow \mel{*}{\beta'}{A'}{{}*}$
2922: with $0<\min\{\beta', \gcd(A')\}<\alpha$
2923: 
2924: \bigskip
2925: (RED3): \
2926: $\mel\alpha\beta{\ A}B \longrightarrow \mel *{g_2}0{{}*}$
2927: with $0<g_2\le \min \{|\beta|,\gcd(A)\}$
2928: 
2929: \medskip
2930: (RED4): \
2931: Assume $\beta>0$, \ $\beta \not | \gcd(\alpha, B)$;
2932: 
2933: \ \phantom{(RED4): }%
2934: $\mel\alpha \beta{\ 0} B \longrightarrow \mel{\alpha'} *{*} {B'}$
2935: with $0<\min\{\alpha', \gcd(B')\}<\beta$
2936: 
2937: 
2938: \medskip
2939: The symbol ``$*$" above represents some element at the appropriate
2940: position. We will not be interested in that entry for the moment.
2941: 
2942: The idea of (RED1) is to use the interaction of the elements $(\alpha, B)$
2943: in the left column to replace them by $(g_1, 0)$. (RED2) is designed to be
2944: applied after (RED1), and its idea is to (eventually) reduce the
2945: pair $(\beta, A)$ in the right column modulo $g_1$.
2946: 
2947: (RED3) and (RED4) are mirror images of (RED1) and (RED2) with the roles of
2948: the left and right columns reversed.
2949: 
2950: We now proceed to describing these steps in detail.
2951: \medskip
2952: 
2953: (RED1): \
2954: $\mel\alpha\beta{\ A} B \longrightarrow \mel{g_1} * * 0$
2955: with $0<g_1\le\ \min \{|\alpha|,\gcd( B)\}$
2956: 
2957: We start with an element $x\in \mz$ of the form
2958: \begin{equation}
2959: x=\mel\alpha\beta{\ A} B,\qquad
2960: \alpha,\beta\in \zet,\ A,B\in\jz.
2961: \end{equation}
2962: 
2963: We can use the transformation $\psi (D)$ with an appropriately chosen $D \in
2964: \jz$ to reduce entries of the element $B$ modulo $\alpha$. This
2965: essentially means that the $9, 15,$ or $27$ integers in the entry $B+\alpha D$ of
2966: $\psi(D)({\bx})$ will lie in the interval $0 \dots |\alpha | -1$.
2967: 
2968: Conversely, we can reduce the entry $\alpha$ modulo $m_2$, where
2969: $m_2 = \gcd (B)$.
2970: Without loss of generality we can assume that $B$ is in the Smith normal
2971: form (to get it, first find an $s\in \npg\jz$ such that $s(B)$ is in the
2972: Smith canonical form; then replace $x$ by $\tf{s^{*^{-1}}}\ ({\bx}))$;
2973: the value of $\alpha$ will not be affected by this). This remark
2974: implies that the component $B$ of the (new) element $x$ will have the form
2975: $$
2976: B=\diag{m_2}**\qquad{\rm with}\ m_2=\gcd(B).
2977: $$
2978: Next, we find
2979: an integer $c_1$ such that $\alpha +m_2 c_1$ is between $1$ and $m_2$,
2980: and let $C$ be the diagonal matrix $\diag{c_1}00$.
2981: The ``$1,1$"-entry of the element $\phi (C) ({\bx})$ is
2982: $$
2983: \alpha+(B,C)+(A, C^\# ) +\beta \mnorm(C)
2984: $$
2985: and since rank $C$=1, it is equal to
2986: $$
2987: \alpha+(B,C)=\alpha+m_2 c_1,
2988: $$
2989: which is in $\{1,\dots, m_2\}$,
2990: and we have reduced $\alpha$ modulo $\gcd(B)$
2991: (note that we chose the set of residues,
2992: which excludes $0$, but contains $m_2$, since we want to keep this
2993: component non-zero).
2994: 
2995: We can now reduce the entries $\alpha$ and $B$ modulo each other, and
2996: continue to do so as long as the entry at the position $B$ is non-zero.
2997: The value of the component $\alpha$ will decrease after each pair of
2998: iterations, while remaining positive. This condition guarantees that the
2999: process will terminate at some step. This procedure
3000: will bring $x$ to the desired form, and since at each step we reduced
3001: modulo $\alpha$ or $\gcd(B)$, the condition on $g_1$ is satisfied. This
3002: completes the description of (RED1).
3003: 
3004: \bigskip
3005: 
3006: (RED2): \
3007: Assume $\alpha>0$, \ $\alpha \not | \gcd(\beta, A)$;
3008: 
3009: \ \phantom{(RED2): }%
3010: $\mel \alpha\beta A0 \longrightarrow \mel *{\beta'}{A'}{{}*}$
3011: with $0<\min\{\beta', \gcd(A')\}<\alpha$.
3012: 
3013: We start with an element $x$ as stated in the assumption.
3014: We can apply a transformation $\tf{s}$ to bring the entry $A$ to the
3015: Smith normal form without changing $\alpha$ and $\beta$, so without
3016: loss of generality we assume that $A$ is already in the Smith normal form
3017: $A=\diag{a_1}{a_2}{a_3}$, $\ a_1|a_2,\ a_2|a_3,\ \alpha>0$.
3018: 
3019: Let us now take matrix $D=\diag{0}1{d_3}$ (we will choose the value for
3020: $d_3$ later). Then, since $B=0$ and $\mnorm(D)=0$, we have
3021: $$
3022: \psi(D)({\bx})=
3023: \mel\alpha {\ \beta +(A,D)}{\quad  A+\alpha D^\# }{\ \alpha D}.
3024: $$
3025: 
3026: We have
3027: $D^\# =\diag{d_3}00$, and hence the ``$1,2$"-entry of the above element
3028: is
3029: $$
3030: \diag{a_1+d_3 \alpha}{\quad a_2}{\quad a_3}.
3031: $$
3032: We can find $d_3\in \zet$ that brings $a'_1=a_1+d_3 \alpha$ into the
3033: integer interval $1,\dots,\alpha$.
3034: If $a'_1<\alpha$, then we are done. And if $a'_1=\alpha$, it means
3035: $\alpha | A$. Then
3036: we can use the element $a'_1$, ideas from (RED3), and the non-divisibility
3037: assumption (which translates now into $\alpha \not| \beta$),
3038: to bring the entry in the positions $\beta$
3039: to the interval $1,\dots,\alpha -1$.
3040: This completes the description of (RED2).
3041: 
3042: The steps (RED3) and (RED4) are mirror images of (RED1) and (RED2).
3043: They are performed in a similar way.
3044: 
3045: \medskip
3046: 
3047: {\bf Part 2.}
3048: 
3049: Now we show how we can use the above reduction steps
3050: to get the $\gcd ({\bx})$
3051: of a non-zero element
3052: $$
3053: {\bx}=\mel\alpha\beta A B,
3054: $$
3055: 
3056: at the position $\alpha$ or $\beta$.
3057: 
3058: In the loop described below
3059: we will apply our reductions to the appropriate matrix, but we will
3060: concentrate our attention on either the pair in the right column or
3061: the pair in the left column of the matrix $\mel\alpha\beta AB$.
3062: 
3063: Without loss of generality we can assume that either of $\alpha$ or $B$ is
3064: non-zero (otherwise we apply $\tau$ and the requirement is satisfied).
3065: \medskip
3066: 
3067: \hspace{10mm}
3068: \begin{minipage}{135mm}
3069: \renewcommand{\baselinestretch}{1}
3070: \renewcommand{\arraystretch}{1}
3071: \normalsize
3072: 
3073: {\sc Loop.}
3074: 
3075: {\em Step 1.}
3076: We have an element of the form
3077: $$
3078: \mel{\alpha'} {\beta'} {A'} {B'}
3079: $$
3080: with non-zero pair $\alpha', B'$ in the left column.
3081: 
3082: By (RED1) we can bring it to the form
3083: $$
3084: \mel {g_1}\beta A 0
3085: $$
3086: with some (new) values of $\beta, A$ and
3087: $0<g_1\le \min\{|\alpha'|, \gcd (B')\}$. It follows from (\ref{end2})
3088: below that $g_1$ is strictly smaller than
3089: the value of $g_2$ from the previous step
3090: ({\em ignore this remark if you just entered the loop}).
3091: 
3092: We need to consider two cases:
3093: \begin{itemize}
3094: 
3095: \item
3096: $g_1 | \beta, g_1 |A$
3097: 
3098: We are done: leave the loop.
3099: 
3100: \item
3101: $g_1$ does not divide either $\beta$ or $A$
3102: 
3103: It follows that $g_1 \not | \gcd (\beta, A)$.
3104: The assumptions of (RED2) are now satisfied, and we apply it to
3105: $\mel {g_1} \beta A 0$. The result is a new element of the form
3106: $\mel *{\beta''}{A''} {{}*}$ with
3107: \begin{equation} \label{end1}
3108: 0<\min\{ \beta'',\gcd (A'')\}<g_1.
3109: \end{equation}
3110: 
3111: 
3112: \end{itemize}
3113: Proceed to {\em Step 2}.
3114: %
3115: \vspace{3mm}
3116: 
3117: %
3118: %
3119: \end{minipage}
3120: 
3121: \hspace{10mm}
3122: \begin{minipage}{135mm}
3123: \renewcommand{\baselinestretch}{1}
3124: \renewcommand{\arraystretch}{1}
3125: \normalsize
3126: 
3127: 
3128: {\em Step 2.}
3129: 
3130: This step is a mirror image of the previous step, with the roles of elements in the
3131: left column and the right column exchanged.
3132: 
3133: We have an element of the form
3134: $$
3135: \mel{\alpha''} {\beta''} {A''} {B''}
3136: $$
3137: with elements in the right column satisfying (\ref{end1}).
3138: 
3139: Using (RED3) we can bring it to the form
3140: $$
3141: \mel\alpha{g_2} 0 B
3142: $$
3143: with some (new) values of $\alpha, B$ and
3144: $0<g_2\le \min\{|\beta''|,\ \gcd(A'')\}$. And (\ref{end1}) implies that~
3145: $g_2$ is strictly smaller than the value of $g_1$ from the previous step.
3146: 
3147: We consider two cases again:
3148: \begin{itemize}
3149: 
3150: \item
3151: $g_2 |\alpha, g_2 | \gcd (B)$
3152: 
3153: We are done: leave the loop.
3154: 
3155: \item
3156: $g_2$ does not divide either $\alpha$ or $B$
3157: 
3158: 
3159: It follows that $g_2 {}\not\!|{\ } \gcd (\alpha, B)$. %
3160: The assumptions of (RED4) are now satisfied, and we apply it to
3161: $\mel \alpha {g_2} 0 B$. The result is a new element of the form
3162: $\mel {\alpha'} **{B'}$ with
3163: \begin{equation} \label{end2}
3164: 0<\min\{ \alpha',\gcd (B')\}<g_2.
3165: \end{equation}
3166: \end{itemize}
3167: Proceed to {\em Step 1}.
3168: \end{minipage}
3169: 
3170: \vspace{4mm}
3171: As we run this loop the values of $g_1$ and $g_2$ will decrease
3172: at each even and odd step, respectively.
3173: On the other hand, they both remain positive
3174: integers, and this guarantees us that we leave the loop at some step.
3175: 
3176: If we left the loop at the first step, then the first entry of the
3177: resulting element is $\gcd({\bx})$. And if we left it at the second
3178: step, then the $\gcd({\bx})$ is at the second position. In the latter
3179: case we apply $\tau^{-1}$ to bring it to the first entry.
3180: 
3181: Let us change the notation again (we have done it many times recently),
3182: and denote the obtained element by
3183: $$
3184: \mel\alpha\beta A0
3185: $$
3186: (it follows from the above that the ``$2,1$"-entry is equal to zero).
3187: 
3188: We can apply a transformation from $\npg\jz$,
3189: if necessary, to convert $A$ to the Smith normal form. Then $A$
3190: turns into a diagonal matrix, and the last requirement of the lemma is
3191: satisfied.
3192: 
3193: This element has the desired form.
3194: \bigskip
3195: 
3196: {\bf (b)}
3197: We now consider the case $\jz=\zet\oplus\zet\oplus\zet$
3198: (and hence $\mz=\m(\zet\oplus\zet\oplus\zet)$).
3199: In this case our ``matrices" in $\jz$ are already in the ``diagonal" form.
3200: But the norm-preserving group is too small in this case,
3201: and it is not true that every element of $\jz$ may be converted to a Smith
3202: normal form. However, one can still use the ideas from the algorithm in part (a)
3203: with minor modifications to prove the assertion of the lemma.
3204: 
3205: We will not pursue the route described in the previous paragraph to prove~(b).
3206: Instead we will use the relation between the pairs
3207: $\Bigl(\grz, \mz\Bigr)$ and
3208: $\Bigl(\SL_2(\zet)\times \SL_2(\zet)\times \SL_2(\zet),\
3209: \zet^2\tensor \zet^2\tensor \zet^2\Bigr)$ described in Example~$\ref{ex:FFF}$.
3210: 
3211: The reduction procedure described in \cite[Appendix]{bh1}
3212: explains how an arbitrary element of $\zet^2\tensor \zet^2\tensor \zet^2$
3213: may be transformed to a certain $5$-parameter form by a transformation in
3214: the group $\SL_2(\zet)\times \SL_2(\zet)\times \SL_2(\zet)$.
3215: Using the isomorphisms of Example~$\ref{ex:FFF}$,
3216: this procedure may be restated in terms of
3217: transformations in $\grz$ acting on the elements on $\m(\jz)$,
3218: which yields the desired result.
3219: \ep
3220: 
3221: An immediate corollary of this reduction lemma is Proposition~$\ref{genz}$
3222: that describes the generators of the group $\grz$.
3223: 
3224: \begin{prop} \label{genz}
3225: Let $\jz$ be one of $\h{\binz}, \h{\quatz}, \h{\octz},
3226: \zet\oplus\zet\oplus\zet$, and let
3227: $\mz=\m(\jz)$.
3228: The group $\grz$ is generated by the elementary transformations
3229: \begin{equation} \label{ele}
3230: \phi(C),\ \psi(D),\ \tf{s},\ \tau
3231: \end{equation}
3232: with $C,D\in \jz,\ s\in \str\jz$.
3233: \end{prop}
3234: 
3235: {\sc Proof.}
3236: 
3237: Let $\rho$ be an arbitrary element in the group $\grz$. We need to prove
3238: that $\rho$ can be represented as a product of elementary transformations
3239: $(\ref{ele})$.
3240: 
3241: Let $f_1$ be
3242: the element $\mel 1000\in \mz$. Let us consider the element
3243: $$
3244: \bx \defeq \rho (f_1).
3245: $$
3246: 
3247: By Lemma~$\ref{bl-fz}$ we can use a product $\eta$ of elementary
3248: transformations to bring $\bx$ to the (diagonal) reduced form
3249: $$
3250: \bx'=\mel\alpha\beta A 0,\qquad \alpha>0.
3251: $$
3252: 
3253: %
3254: %
3255: %
3256: %
3257: %
3258: We have $\bx' = \eta\, \rho (f_1)$, and it follows from %
3259: Lemma~$\ref{lem:rank}$ that
3260: $$
3261: {\rm rank\ }\bx'={\rm rank\ }f_1=1.
3262: $$
3263: 
3264: Then it follows from the relations in Remark~$\ref{rem1}$
3265: that $\beta = 0$, $A=0$, and hence
3266: $$
3267: \bx'=\mel{\alpha}000.
3268: $$
3269: 
3270: We have $\eta, \rho\in \grz$, and it follows from Lemma~$\ref{gcd}$ that
3271: $\gcd(\bx')=\gcd(f_1)$,
3272: and hence we get that
3273: $$
3274: \alpha=1 \qquad {\rm and} \qquad \bx'=f_1,
3275: $$
3276: 
3277: and hence
3278: $$
3279: \eta \rho (f_1)=f_1.
3280: $$
3281: 
3282: Repeating the argument of \cite[Lemma 12,13]{b} and using the fact
3283: that our transformations preserve the integral structure,
3284: we prove that $\eta \rho$ has the form
3285: $$
3286: \eta \rho = \tf{s}\phi(C) \quad
3287: {\rm for\ some \ }s\in \str\jz, C\in \jz,
3288: $$
3289: and since $\eta$ was the product of elementary transformations, $\rho$ has
3290: the desired form.
3291: \ep
3292: 
3293: 
3294: 
3295: 
3296: \subsection{Projective elements in the module $\m(\jz)$}
3297: \label{ss:proj}
3298: 
3299: 
3300: In this subsection we discuss the concept of projective elements,
3301: which plays the central role in the classification of orbits in the
3302: integral case.
3303: 
3304: \begin{defn}\label{def:proja}{\bf (a)}
3305: 
3306: {\rm
3307: Let $\jz=
3308: \zet\oplus\zet\oplus\zet$, and let $\mz=\m(\jz)$.
3309: An element $x\in \mz$
3310: $$
3311: x=
3312: \mel\alpha\beta
3313: {({a_1}, {a_2}, {a_3})}
3314: {({b_1}, {b_2}, {b_3})},\qquad
3315: \alpha, \beta, a_i, b_i\in \zet
3316: $$
3317: is said to be {\em projective}, if each of the three binary quadratic forms
3318: associated to this element
3319: \begin{eqnarray*}
3320: \biggl(\alpha a_1-b_2 b_3\biggr)x^2 +
3321: \biggl(-a_1b_1+a_2 b_2+ a_3 b_3 -\alpha\beta\biggr)xy +
3322: \biggl(\beta b_1-a_2a_3\biggr) y^2,\\[2mm]
3323: \biggl(\alpha a_2-b_3 b_1\biggr)x^2 +
3324: \biggl(\phantom{-}a_1b_1-a_2 b_2+ a_3 b_3 -\alpha\beta\biggr)xy +
3325: \biggl(\beta b_2-a_3a_1\biggr) y^2,\\[2mm]
3326: \biggl(\alpha a_3-b_1 b_2\biggr)x^2 +
3327: \biggl(\phantom{-}a_1b_1+a_2 b_2- a_3 b_3 -\alpha\beta\biggr)xy +
3328: \biggl(\beta b_3-a_1a_2\biggr) y^2.
3329: \end{eqnarray*}
3330: is primitive, i.e., the $\gcd$ of its coefficients is equal to one.
3331: 
3332: }
3333: \end{defn}
3334: 
3335: The concept of projective element was first introduced by M.~Bhargava
3336: \cite{bh-phd, bh1} for elements of
3337: $\zet^2\tensor\zet^2\tensor\zet^2$. The equivalence of the two definitions
3338: follows from the isomorphism
3339: $$
3340: \zet^2\tensor\zet^2\tensor\zet^2 \cong \m(\zet\oplus\zet\oplus\zet)
3341: $$
3342: established in Example~\ref{ex:FFF}.
3343: 
3344: The concept of projective element was extended to
3345: $\wedge^3\zet^6$ (and several more examples) in~\cite{bh1}.
3346: More precisely, M.~Bhargava showed that there is a natural injection
3347: \begin{equation}\label{inc1}
3348: \zet^2\tensor\zet^2\tensor\zet^2 \hookrightarrow \wedge^3\zet^6
3349: \end{equation}
3350: and in addition every element of $\wedge^3\zet^6$ is
3351: $\SL_6(\zet)$-equivalent to an element in the image of
3352: $\zet^2\tensor\zet^2\tensor\zet^2$ under the embedding (\ref{inc1}).
3353: An element of $\wedge^3\zet^6$ was defined to be {\em projective},
3354: if its $\SL_6$-orbit contains an image under (\ref{inc1}) of a
3355: projective element in $\zet^2\tensor\zet^2\tensor\zet^2$.
3356: 
3357: In our case we have a chain of natural inclusions of cubic vector spaces
3358: ($\zet$-modules):
3359: $$
3360: \zet\oplus\zet\oplus\zet \subset
3361: \h\binz \subset \h\quatz \subset \h\octz
3362: $$
3363: (the first inclusion being the diagonal embedding), which induces the
3364: natural chain
3365: \begin{equation} \label{chain3}
3366: \m(\zet\oplus\zet\oplus\zet) \subset
3367: \m(\h\binz) \subset \m(\h\quatz) \subset \m(\h\octz).
3368: \end{equation}
3369: 
3370: The first two modules in the last chain are isomorphic to
3371: $\zet^2\tensor\zet^2\tensor\zet^2$ and $\wedge^3\zet^6$, respectively
3372: (see Examples~$\ref{ex:FFF}, \ref{ex:sl6int}$). We proved in Lemma~\ref{bl-fz}
3373: that an arbitrary element of
3374: each of the modules $\mz$ in (\ref{chain3}) is $\grz$-equivalent to
3375: a {\em diagonal reduced} element,
3376: which can be thought of as an element of the smallest submodule
3377: $\m(\zet\oplus\zet\oplus\zet)$ in the chain~$(\ref{chain3})$.
3378: This lemma allows us to extend
3379: the concept of a projective element to $\m(\h\binz)\cong \wedge^3\zet^6$,
3380: using an argument different from Bhargava's, and further extend this
3381: concept to $\m(\h\quatz)$ and  $\m(\h\octz)$.
3382: 
3383: %
3384: \hspace{-\parindent}{\bf Definition~\ref{def:proja} (b)}
3385: {\rm
3386: Let $\jz$ be one of $\h{\binz}, \h{\quatz}, \h{\octz}$, and let
3387: $\mz=\m(\jz)$. An element $x\in\mz$ is said to be {\em projective}, if its
3388: $\grz$-orbit contains a diagonal reduced element (cf. Lemma~$\ref{bl-fz}$)
3389: \begin{equation}\label{proj-el}
3390: \mel\alpha\beta{\diag{a_1} {a_2} {a_3}}{0},\qquad
3391: \alpha, \beta, a_i\in \zet,\quad
3392: \alpha>0,\quad \alpha |\beta, \ \alpha |a_i,
3393: \end{equation}
3394: which is projective.
3395: }
3396: %
3397: 
3398: We note that for an element of the form (\ref{proj-el}) the projectivity
3399: conditions of Definition~\ref{def:proja}(a) become
3400: \begin{eqnarray} \label{def:projexpr}
3401: & \gcd(\alpha a_1, \alpha\beta, a_2a_3)=1 & \nonumber\\
3402: & \gcd(\alpha a_2, \alpha\beta, a_1a_3)=1 &\\
3403: & \gcd(\alpha a_3, \alpha\beta, a_1a_2)=1 & \nonumber
3404: \end{eqnarray}
3405: Taking into account the divisibility conditions of reduced elements
3406: (Definition~\ref{def:reduced}), it follows that if element (\ref{proj-el})
3407: is projective, then $\alpha=1$, i.e., $x$ is primitive. Nevertheless we
3408: kept $\alpha$ in the expressions (\ref{def:projexpr}) to emphasize the fact
3409: that they are homogeneous expressions of degree $2$.
3410: \medskip
3411: 
3412: The definition of a projective element we just gave is not quite
3413: satisfactory, since
3414: it is not convenient to work with orbit representatives of the group
3415: $\grz$, which may be quite large.
3416: Next we would like to argue that the concept of a projective
3417: element is related to certain equations of degree $3$, which will allow us
3418: to get a simple projectivity test in {\em almost} all cases.
3419: 
3420: \begin{prop} \label{prop:proj}
3421: {\rm
3422: Let $\jz$ be as in Definition~$\ref{def:proja}$(a,b) and $\mz=\m(\jz)$.
3423: Let $x$
3424: be in $\mz$ and let $T(x,x,x)$ be defined as in~$(\ref{deft})$.
3425: \begin{enumerate}[(i)]
3426: \item
3427: If $\gcd\, T(x,x,x)=1$ then $x$ is projective;
3428: \item
3429: If $\gcd\, T(x,x,x)\ge 3$ or $T(x,x,x)=0$ then $x$ is not projective.
3430: \item
3431: When $q'(x)$ is odd, $\gcd T(x,x,x)=1$ iff $x$ is projective.
3432: \end{enumerate}
3433: }
3434: \end{prop}
3435: 
3436: {\sc Proof.}
3437: 
3438: First we notice that the quantity $\gcd T(x,x,x)$ is invariant with
3439: respect to the action of the group $\grz$.
3440: This assertion follows from the
3441: $\gr$-invariance of the quartic form $q(x)$ and the symplectic form
3442: $\{\cdot,\cdot\}$,
3443: the definition of $T(x,x,x)$ $(\ref{deft})$, and Lemma~$\ref{gcd}$
3444: (see Lemma~$\ref{di}$ below for a detailed argument).
3445: Hence, for a diagonal reduced element
3446: \begin{equation}\label{x1a}
3447: x_1=\mel\alpha\beta{\diag{a_1} {a_2} {a_3}}{0},\qquad
3448: \alpha, \beta, a_i\in \zet,\quad
3449: \alpha>0, \alpha|\beta, \alpha| a_i.
3450: \end{equation}
3451: contained in the $\grz$-orbit of $x$ by Lemma~$\ref{bl-fz}$ we have
3452: $$
3453: \gcd T(x,x,x)=\gcd T(x_1,x_1,x_1).
3454: $$
3455: 
3456: It is more convenient to work with $x_1$,
3457: since the components
3458: of $T(x_1,x_1, x_1)$ have the following simple form (see~$(\ref{txxx})$):
3459: $$
3460: T(x_1,x_1,x_1)=
3461: \mel{\alpha^2\beta}{\ \alpha\beta^2+2\mnorm(A)}{\ \alpha\beta A}{\ 2\alpha
3462: A^\#},
3463: $$
3464: 
3465: where $A=\diag{a_1} {a_2} {a_3}$,
3466: $\mnorm (A)=a_1a_2a_3,\ A^\#=\diag{a_2a_3}{a_3a_1}{a_1a_2}$.
3467: 
3468: 
3469: {\bf (i)}
3470: Let $\gcd T(x_1,x_1,x_1)=1$, and suppose $x_1$ is not projective. It means
3471: that the $\gcd$ of at least one of the expressions in~$(\ref{def:projexpr})$
3472: is greater than $1$. Without loss of generality we assume that there
3473: exists a prime $p>1$ such that
3474: $$
3475: p\ \Bigl|\  \alpha a_3,\  \alpha\beta,\ a_1a_2.
3476: $$
3477: This implies that $p$ divides each component of $T(x_1, x_1, x_1)$,
3478: contrary to our assumption.
3479: 
3480: {\bf (ii)}
3481: We are given $x$ such that $\gcd T(x, x, x)\ge 3$, and need to show that
3482: $x$ is not projective. Let $x_1$ be {\em any} diagonal reduced element of
3483: the form~$(\ref{x1a})$ in the $\grz$-orbit of $x$. Hence
3484: $$
3485: \gcd T(x_1, x_1, x_1)=
3486: \gcd
3487: \mel{\alpha^2\beta}{\alpha\beta^2+2\mnorm(A)}{\alpha\beta A}{2\alpha A^\#}
3488: \ge 3.
3489: $$
3490: 
3491: First we notice that if $\alpha >1$, then $x_1$ is not primitive.
3492: Non-primitivity implies non-projectivity, and hence there is nothing to
3493: prove in this case.
3494: Thus from now on we may assume
3495: $\alpha=1$, and hence
3496: $$
3497: T(x_1, x_1, x_1)=
3498: \mel{\beta}{\beta^2+2\mnorm(A)}{\beta A}{2 A^\#}.
3499: $$
3500: 
3501: Suppose that there exists a prime $p\ge 3$ that divides $\gcd T(x_1,
3502: x_1, x_1)$. Then we have $p|\beta$. In addition, the condition $p|A^\#$
3503: implies that at least two of the numbers $a_i$ are divisible by $p$. Then
3504: $p$ divides at least one of the expressions~$(\ref{def:projexpr})$, and
3505: hence $x_1$ is not projective.
3506: 
3507: If the prime $p$ such as in the previous paragraph does not exist, it
3508: follows from the conditions that $\gcd T(x_1, x_1, x_1)$ is a power of $2$
3509: (and is greater than $2$). Then we have $4|\beta$, $2|A^\#$, and repeating
3510: the argument of the previous paragraph, we get that $2$ divides at least
3511: one of the expressions~$(\ref{def:projexpr})$. Hence $x_1$ is not
3512: projective in this case either.
3513: 
3514: We have proved that if $\gcd T(x,x, x)\ge 3$, then every diagonal reduced
3515: element in $\grz$-orbit of $x$ is not projective, and hence $x$ is not
3516: projective.
3517: 
3518: This argument also applies in the case $T(x,x,x)=0$, if one replaces the
3519: condition ``divisible by $p$" with the condition ``equal to $0$".
3520: 
3521: {\bf (iii)}
3522: The implication ``$\Rightarrow$" was proved in (i). To prove the converse
3523: implication, we again look at the diagonal reduced projective
3524: $x_1$ of the form $(\ref{x1a})$ in the $\grz$-orbit of $x$.
3525: 
3526: It follows from (ii) that $\gcd T(x_1,x_1,x_1)$ is equal to $1$ or $2$,
3527: and we will show that the latter option is not possible.
3528: 
3529: Since $x_1$ is projective, it is primitive, and hence $\alpha=1$.
3530: We then have $q'(x_1)=\beta^2+4\mnorm (A)$, and since $q'(x)=q'(x_1)$ is
3531: odd, we have that $\beta$ is odd. This conclusion implies that $2$ cannot
3532: divide $\gcd T(x_1,x_1,x_1)$, and hence $\gcd T(x_1,x_1,x_1)=\gcd T(x,x,x)=1$.
3533: \ep
3534: 
3535: \begin{rem}
3536: {\rm
3537: The above proposition  provides a projectivity test for all cases
3538: except $\gcd\,$ %
3539: $T(x,x,x)=2$. Elements
3540: \begin{equation}
3541: x_1=\mel{1}{2}{\diag{1}12}0 \qquad
3542: x_2=\mel{1}{2}{\diag{1}22}0
3543: \end{equation}
3544: both have $\gcd \Bigl(T(x_i,x_i,x_i)\Bigr)=2$.
3545: However $x_1$ is projective and $x_2$ is not projective in the sense of
3546: Definition~$\ref{def:proja}(a)$.
3547: }
3548: \end{rem}
3549: 
3550: \begin{cor}\label{cor:anyrep}
3551: {\rm
3552: Proposition~$\ref{prop:proj}$ implies that if $\gcd T(x,x,x)$ is
3553: different from $2$ then the (non)proje\-ctivity of $x$ is determined by
3554: {\em any} representative of its $\grz$-orbit
3555: in $\m(\zet\oplus\zet\oplus\zet)$.
3556: }
3557: \end{cor}
3558: 
3559: It appears that the assertion of Corollary~$\ref{cor:anyrep}$
3560: remains true for $x$ with $\gcd\, T(x,x,x)=2$,
3561: however the quantity $T(x,x,x)$ does not seem to be precise enough to
3562: treat this case. Apparently, the difficulty of proving this fact
3563: is related to the ``bad" reduction
3564: of the quartic form in the case when char\,$F=2$ (cf. Remark~$\ref{badred}$).
3565: \medskip
3566: 
3567: Our next remark is concerned with one more characterization of projective
3568: elements.
3569: There is a natural choice of a basis and coordinates in the module
3570: $\mz=\m(\jz)$ with $\jz=\zet\oplus\zet\oplus\zet,\h\binz,\h\quatz,\h\octz$.
3571: In each of these cases the quartic form $q'(x)$ may be viewed as a
3572: homogeneous polynomial in $n$ variables, $n=8, 20, 32, 56$, respectively.
3573: And the components of $T(x,x,x)$ are nothing else but the $n$ partial
3574: derivatives of the polynomial $q'(x)$ (divided by $2$, since all
3575: these partial derivatives have even coefficients).
3576: In other words, each component of the formal gradient vector
3577: $\frac{1}{2}\nabla q'(x)$ is a polynomial with integer coefficients,
3578: and these expressions are the same as the components of $T(x,x,x)$.
3579: These observations and Proposition~$\ref{prop:proj}$ yield the following
3580: 
3581: \begin{cor}\label{cor:nabla}
3582: {\rm
3583: Let $x$ be in $\mz=\m(\jz)$ with $\jz$ as in Definition~$\ref{def:proja}$(a,b).
3584: Let $q'$ be the quartic invariant of the module $\mz$, see~$(\ref{qmod})$.
3585: Then $\gcd T(x,x,x)=\gcd \frac{1}{2}\nabla q'(x)$ and hence
3586: \begin{itemize}
3587: \item
3588: If $\gcd\, \frac{1}{2}\nabla q'(x)=1$ then $x$ is projective;
3589: \item
3590: If $\Bigl(\gcd\, \frac{1}{2}\nabla q'(x)\ge 3\Bigr)$
3591: or $\Bigl(\nabla q'(x)=0\Bigr)$ then $x$ is not projective.
3592: \end{itemize}
3593: }
3594: \end{cor}
3595: 
3596: 
3597: 
3598: \begin{rem}\label{rem:nabla}
3599: {\rm
3600: It follows
3601: from general considerations that if $p(x)$ is a (quartic) form on a
3602: $\zet$-module $V_\zet$ invariant with respect to a group $G_\zet$, then
3603: $\gcd \nabla p(x)$ is invariant with respect to $G_\zet$. This observation
3604: suggests that Corollary~$\ref{cor:nabla}$ may be used to describe
3605: (non)projective elements in spaces ${\rm Sym}^3\zet^2,
3606: \zet^2\tensor{\rm Sym}^2\zet^2, \zet^2\tensor\zet^4$
3607: underlying higher composition laws
3608: related to quadratic rings (see Table~1 and \cite{bh1} for details).
3609: }
3610: \end{rem}
3611: 
3612: 
3613: \subsection{Further reduction and the classification of the projective orbits}
3614: \label{ss:fred}
3615: 
3616: \begin{lemma}[\sc Reduction Lemma II]  \label{red2}
3617: \
3618: 
3619: \hspace{-\parindent}%
3620: Let $\jz$ be one of $\h{\binz}, \h{\quatz}, \h{\octz}$
3621: and $\mz=\m(\jz)$.
3622: Let ${\bx}\in \mz$ be a projective element.
3623: 
3624: 
3625: Then $x$ is equivalent to an element
3626: \begin{equation} \label{proj1}
3627: \mel 1 \varepsilon{\ \diag{1}1k} 0,\qquad
3628: \varepsilon\in\{0,1\},\quad
3629: k\in \zet
3630: \end{equation}
3631: 
3632: under a series of elementary transformations
3633: \begin{equation} \label{elem-tr1}
3634: \phi (C), \psi(D), \tf{s}, \tau \quad
3635: with \ C, D \in \jz,\ s\in \npg{\jz}.
3636: \end{equation}
3637: 
3638: 
3639: The values of $\varepsilon$ and $k$ in~$(\ref{proj1})$ are uniquely
3640: determined by $q'(x)$.
3641: \end{lemma}
3642: 
3643: {\sc Proof.}
3644: 
3645: By the definition of a projective element, $x$ is equivalent to a
3646: diagonal reduced projective element
3647: $$
3648: x_1=\mel{\alpha_1} {\beta_1}{\ A_1} 0, \qquad
3649: {\rm where}\ \alpha_1=\gcd(x_1)=\gcd(x).
3650: $$
3651: under the action of transformations (\ref{elem-tr1}). Since any projective
3652: element is primitive, we have $\alpha_1=1$. In addition, acting by $T(s)$
3653: if necessary we can assume that $A_1$ is in the Smith normal form
3654: (Theorem~$\ref{np-orbits}$).
3655: Summarizing these remarks we conclude that $x_1$ is of the form
3656: \begin{equation}\label{x1}
3657: x_1=\mel 1{\beta_1}{\ \diag{a_1}{a_2}{a_3}}0
3658: \end{equation}
3659: with $\diag{a_1}{a_2}{a_3}$ in the Smith normal form.
3660: 
3661: The definition of the projective element
3662: (Definition~$\ref{def:proja}$(a), see also~$(\ref{def:projexpr})$) implies that
3663: \begin{equation} \label{proj3}
3664: \gcd\{ \beta_1,\ a_1 a_2, \ a_1 a_3, \ a_2 a_3\}=1.
3665: \end{equation}
3666: Since $A_1$ is in the Smith normal form with $a_1 | a_2,\ a_2|a_3$,
3667: the relation (\ref{proj3}) is equivalent to
3668: \begin{equation} \label{proj4}
3669: \gcd\{ \beta_1,\ a_1 a_2\}=1.
3670: \end{equation}
3671: 
3672: 
3673: {\bf Step 1.}
3674: We show that $a_1$ can be taken to be equal to $1$.
3675: 
3676: Assume this is not the case. Then $a_1>1$
3677: and $a_1|a_2, a_1|a_3$ (case $a_1=a_2=a_3=0$ needs slightly
3678: different treatment; we skip details).
3679: Relation (\ref{proj4}) implies $\gcd(\beta_1, a_1)=1$.
3680: 
3681: Then we apply Lemma~\ref{lcomp} (iii) to $x_1$ (with $c=1$)
3682: and transform it to
3683: \begin{equation}\label{proj5}
3684: \mel 1
3685: {\beta_1-2a_1a_2}{\quad \diag{a_1}{\ a_2}{\ a_3+\beta_1-{a_1a_2}}}0,
3686: \end{equation}
3687: It follows from the above that
3688: $\gcd\{{a_1},{a_2},\ a_3+\beta_1-{a_1a_2}\}=
3689: \gcd\{{a_1},{a_2}, \beta_1\}=1$.
3690: The $\gcd$ condition implies that the Smith canonical form
3691: of $$\diag{a_1}{a_2}{\ a_3+\beta_1-{a_1a_2}}$$ looks like
3692: $\diag{1}**$.
3693: Then we can apply an appropriate $T(s)$
3694: to (\ref{proj5}),
3695: and it will yield the desired result, completing the first step.
3696: 
3697: \medskip
3698: Step 1 implies that we can assume that $a_1$ in the element~$(\ref{x1})$
3699: is equal to $1$ and we proceed to
3700: 
3701: {\bf Step 2.}
3702: We show that $a_2$ in~$(\ref{x1})$ can be taken to be equal to $1$.
3703: 
3704: Assume this is not the case. Then $a_2>1$ and we still have $a_2|a_3$
3705: (case $a_2=a_3=0$ treated similarly).
3706: Relation (\ref{proj3}) implies $\gcd(\beta_1, a_2)=1$.
3707: 
3708: We again apply Lemma~\ref{lcomp} (iii) to $x_1$ (with $c=1$)
3709: and transform it to
3710: \begin{equation}\label{proj6}
3711: \mel 1
3712: {\beta_1-2a_2}{\quad \diag{1}{\ a_2}{\ a_3+\beta_1-{a_2}}} 0.
3713: \end{equation}
3714: 
3715: We have $\gcd\{ {a_2}, a_3+\beta_1-{a_2} \}=
3716: \gcd\{ {a_2},\beta_1\}=1$.
3717: Similarly to Step~1, we can apply an appropriate $T(s)$
3718: to (\ref{proj6}), an get an element of the form
3719: $$
3720: \mel 1{\beta_1-2a_2}{\quad \diag{1}{1}{*}}0.
3721: $$
3722: The second step is complete.
3723: 
3724: It follows from the above steps that element (\ref{x1}) may be taken to be
3725: of the form
3726: \begin{equation}\label{proj7}
3727: \mel 1 {\beta_1}{\quad \diag{1}{1}{a_3}}0.
3728: \end{equation}
3729: 
3730: {\bf Step 3.}
3731: We show that $\beta_1$ in the element (\ref{proj7}) may be taken to be $0$ or
3732: $1$.
3733: 
3734: To do it we again apply Lemma~\ref{lcomp} (iii):
3735: $$
3736: \mel 1{\beta_1-2c}{\quad \diag{1}{\ 1}{\ a_3+\beta_1 c-c^2}}0.
3737: $$
3738: 
3739: Obviously, there is a $c$, which makes the second component $0$ or
3740: $1$.
3741: \bigskip
3742: 
3743: The three steps above show that an arbitrary projective element can be
3744: brought to the form~(\ref{proj1}). It remains to prove the uniqueness
3745: assertion.
3746: 
3747: When $x'$ is of the form (\ref{proj1}), we obtain from
3748: $(\ref{qmod})$ that
3749: $$
3750: q'(x') = 4 k+\varepsilon ^2 .
3751: $$
3752: 
3753: This equation determines the parity of $\varepsilon$ uniquely,
3754: and hence $k$ is unique as well.
3755: \ep
3756: 
3757: Now we have all the necessary tools to prove the main result of the paper:
3758: the theorem concerning orbits in the spaces associated to cubic Jordan
3759: algebras of Hermitian matrices over split composition algebras.
3760: 
3761: \begin{theorem}\label{thm:main}
3762: 
3763: Let $(G_\zet, \mz)$ be one the following pairs
3764: $$
3765: \Bigl(\SL_6(\zet), \wedge^3(\zet^6)\Bigr),\quad
3766: \Bigl(D_6(\zet), \mbox{\rm half-spin}_\zet\Bigr),\quad
3767: \Bigl(E_7(\zet), V(\omega_7)_\zet\Bigr).
3768: $$
3769: Then
3770: \begin{itemize}
3771: \item
3772: The $G_\zet$-invariant quartic form (the norm) on the module $\mz$
3773: has values congruent to $0$ or $1\, (\!\mod 4)$.
3774: 
3775: \item
3776: Let $n$ be an integer $\equiv 0$ or $1\, (\!\mod 4)$.
3777: The group $G_\zet$ acts
3778: transitively on the set of {\em projective} elements of norm $n$.
3779: 
3780: \item
3781: If $n$ is a fundamental discriminant\footnote{
3782: An integer $n$ is called a fundamental discriminant if $n$ is squarefree
3783: and $\equiv 1 (\mod 4)$
3784: or $n=4k$, where $k$ is a squarefree integer that is $\equiv 2$ or $3 (\mod 4)$.
3785: The result stated in the theorem also applies in the case $n=1$.
3786: },
3787: then every element of norm $n$ is projective, and hence in this case
3788: $G_\zet$ acts transitively on the set of elements of norm $n$.
3789: \end{itemize}
3790: \end{theorem}
3791: 
3792: {\sc Proof.}
3793: It was shown in Proposition~$\ref{prop:gener}$ that for $\J=\h\bin,
3794: \h\quat, \h\oct$ the Freudenthal construction yields an
3795: absolutely almost simple connected algebraic
3796: group $G=\gr$ of type $A_5, D_6, E_7$, respectively.
3797: Each of these groups acts on the vector space $\m=\mj$, producing
3798: an irreducible representation
3799: whose type is given in the statement of the theorem.
3800: 
3801: The integral structure in the group $G_\zet=\grz$
3802: and the module $\mz$ is induced by the integral structure in $\jz$.
3803: The case $\jz=\h\binz$ requires special treatment.
3804: It was shown in Example~$\ref{ex:sl6int}$ that $\SL_6(\zet)$ is isomorphic
3805: to a subgroup (of index two) in $\g{\m(\h{\bin_\zet})}$. We will do the proof
3806: of the theorem for the whole group $\grz$, and address the issue of
3807: $\SL_6(\zet)$ at the end of the proof.
3808: 
3809: The module $\mz=\m(\jz)$ comes equipped with the quartic form
3810: $$
3811: q'(x)= \Bigl((A,B)-\alpha\beta\Bigr)^2
3812: -4 (A^\#, B^\#) +4\alpha \mnorm(A) +4\beta \mnorm(B),\qquad
3813: x=\mel{\alpha}\beta A B \in \mz,
3814: $$
3815: invariant with respect to $G_\zet$. It was noted in Subsection~$\ref{ss:33int}$
3816: that the operations $ \mnorm$ and $(\cdot,\cdot)$ in $\jz$ have integer
3817: values, which implies that $q'(x)$ is always congruent
3818: to $0$ or $1$ modulo $4$.
3819: 
3820: Next, let $x$ be a projective element in $\mz$. It follows from
3821: Lemma~$\ref{red2}$, that $x$ can be transformed to an element of the
3822: form
3823: \begin{equation}
3824: \mel{1}\varepsilon {\diag{1}1k}0,\qquad
3825: \varepsilon\in\{0,1\},\quad
3826: k\in \zet
3827: \end{equation}
3828: with values of $\varepsilon$ and $k$ uniquely determined by the value of
3829: $q'(x)$. This immediately implies that the group $\grz$ acts transitively
3830: on the set of projective elements of norm $q'(x)$.
3831: 
3832: Finally, let an integer $n$ be a fundamental discriminant, and let
3833: $x\in\mz$ be such that $q'(x)=n$.
3834: By Lemma~$\ref{bl-fz}$, $x$ is equivalent to a diagonal reduced element of the
3835: form
3836: \begin{equation}\label{red-thm}
3837: x_1=\mel \alpha\beta {\diag{a_1}{a_2}{a_3}} 0,\qquad
3838: \alpha>0,\alpha|\beta,\alpha|a_i.
3839: \end{equation}
3840: We need to show that $x$ is projective, and by
3841: Definition~$\ref{def:proja}$(b), it is sufficient to show that $x_1$ is
3842: projective.
3843: 
3844: We have $q'(x)=q'(x_1)$, and hence
3845: \begin{equation}\label{n1}
3846: n=\alpha^2\beta^2+4a_1a_2a_3.
3847: \end{equation}
3848: 
3849: First we show that $\alpha$ must be equal to $1$.
3850: 
3851: Note that $(\ref{red-thm})$ and $(\ref{n1})$ imply $\alpha^3|n$.
3852: 
3853: If $\alpha >2$, this remark implies that $n$ is not square-free. If
3854: $\alpha=2$, we get $16|n$. And in both cases we get that $n$ is not a
3855: fundamental discriminant. Hence $\alpha=1$.
3856: 
3857: It means we can rewrite $x_1$ in the form
3858: $x_1=\mel 1 \beta {\diag{a_1}{a_2}{a_3}} 0$, and
3859: $$
3860: n=\beta^2+4a_1a_2a_3.
3861: $$
3862: 
3863: To complete the proof we will show that if $x_1$ is not projective, then
3864: $n$ is not a fundamental discriminant.
3865: 
3866: So suppose that $x_1$ is not projective. Then the $\gcd$ in one of the
3867: expressions~$(\ref{def:projexpr})$ is greater than $1$. Without loss of
3868: generality (and using $\alpha=1$) we will assume that
3869: $$
3870: \gcd (a_3, \beta, a_1 a_2)>1.
3871: $$
3872: Let $p$ be a prime dividing $\gcd (a_3, \beta, a_1 a_2)$. It follows that
3873: $$
3874: p|\beta \quad \mbox{ and \ $p$ divides at least two of the $a_i$'s}.
3875: $$
3876: This remark implies $p^2|n$. If $p>2$ it already implies that $n$ is not a
3877: fundamental discriminant. And if $p=2$, it implies that
3878: $$
3879: \beta=2\beta_1, \qquad
3880: a_1a_2a_2=4c
3881: $$
3882: for some integers $\beta_1$ and $c$. Hence $n$ may be rewritten in the
3883: form
3884: $$
3885: n=4\Bigl(\beta_1^2+4c\Bigr).
3886: $$
3887: Depending on the parity of $\beta_1$, the quantity $\beta_1^2+4c$ is
3888: congruent to $0$ or $1$ modulo $4$. In both cases it implies that $n$ is
3889: not a fundamental discriminant.
3890: 
3891: We thus proved that if $n$ is a fundamental discriminant, then every
3892: element of norm $n$ is projective, and hence $G_\zet$ acts transitively on
3893: the set of such elements.
3894: \medskip
3895: 
3896: Our last remark is concerned with the case $\jz=\h\binz$. It was noted
3897: in Example~$\ref{ex:sl6int}$ that $\SL_6(\zet)\cong \g{\jz}\cc$, which is
3898: a subgroup of index two in $\grz$. We need to show that orbits under
3899: $\g{\jz}\cc$ are the same as the orbits under the action of the whole
3900: group $G_\zet=\grz$.
3901: 
3902: For this we take an arbitrary element $x\in\mz$ and let $x_1$ be a
3903: diagonal reduced element
3904: $$
3905: x_1=\mel \alpha\beta {\diag{a_1}{a_2}{a_3}} 0,\qquad
3906: \alpha>0,\alpha|\beta,\alpha|a_i
3907: $$
3908: lying in the $\grz$-orbit of $x$, i.e., $x_1=g(x)$ for some $g\in\grz$.
3909: It was noted in Example~$\ref{ex:sl6int}$ that every $g\in\grz$ has the
3910: form
3911: $$
3912: g=g_0\quad
3913: \mbox{or}\quad
3914: g=T(t) g_0
3915: $$
3916: with $g_0\in\grz\cc$. There is nothing to prove in first case, and in the
3917: second case we have
3918: $$
3919: x_1=T(t)(x_1)=T(t)^2\,g_0 (x)=g_0(x)
3920: $$
3921: using that $T(t)$ acts as the transpose operation
3922: on the matrices at the two off-diagonal
3923: entries of $x_1$, and the fact that $T(t)^2=\id_\mz$. Hence $\grz$-orbits
3924: are the same as $\grz\cc$-orbits when $\jz=\h\binz$, and in this case
3925: the theorem remains true for the subgroup of index two isomorphic to
3926: $\SL_6(\zet)$.
3927: \ep
3928: 
3929: 
3930: \subsection{Invariant factors
3931: in the module $\mz$ and the degenerate orbits}
3932: \label{ss:if}
3933: 
3934: In this subsection we introduce the ``invariants" $d_i, i=1,2,3,4$,
3935: reminiscent of invariant factors
3936: for regular matrices over integers, and use them to
3937: describe the orbits of degenerate elements, i.e., those whose norm is
3938: equal to zero.
3939: We define these invariants using the concept of the rank polynomials introduced
3940: earlier in $(\ref{p4})-(\ref{p2})$.
3941: 
3942: \medskip
3943: 
3944: \begin{defn} \label{def:di}
3945: {\rm Let $\jz$ be as in~$(\ref{zlist})$, and let $\mz=\m(\jz)$.
3946: For an element $\bx\in\mz$ of the form
3947: $$
3948: \bx = \mel\alpha\beta A B
3949: $$
3950: the functions $d_i:\ \mz\ \to \ \zet$ are defined by the following
3951: expressions
3952: \begin{itemize}
3953: \item
3954: $d_1 (\bx) =\gcd(\bx)$;
3955: \item
3956: $d_2 (\bx) = \gcd \Bigl(3\,T(x,x,y)+\{x,y\}\,x\Bigr) \quad
3957: \mbox{for all }y\in \mz$;
3958: %
3959: \item
3960: $d_3 (\bx) = \gcd ( T(\bx,\bx,\bx) )$;
3961: \item
3962: $d_4 (\bx) = q'(\bx)$.
3963: \end{itemize}
3964: }
3965: \end{defn}
3966: 
3967: \begin{lemma} \label{di}
3968: Let $\mz$ be as in Definition~$\ref{def:di}$.
3969: \begin{enumerate}[{\rm (i)}]
3970: \item
3971: The functions $d_i$ are invariant under the action of the
3972: group $\grz$.
3973: \item
3974: \begin{equation}\label{d2}
3975: d_2(x)= \gcd\Bigl\{
3976: 3\alpha\beta-(A,B),\quad
3977: 2(\alpha A-B^\# ), \quad 2(\beta B - A^\# ),\quad
3978: 2Q(x)\Bigr\}.
3979: \end{equation}
3980: \end{enumerate}
3981: 
3982: \end{lemma}
3983: 
3984: {\sc Proof.}
3985: 
3986: {\bf (i)}
3987: The statement for $d_1$ follows from Lemma~$\ref{gcd}$.
3988: 
3989: The function $d_4$ is the norm $q'$, and the statement follows from the
3990: definition of the groups $\gr$ and $\grz$.
3991: 
3992: The statement for $d_3$ follows from the following computation
3993: $$
3994: d_3(\eta(x))= d_1\biggl(\  T\Bigl(\eta(x), \eta(x),\eta(x)\Bigr)\
3995: \biggr)=
3996: $$
3997: $$
3998: =d_1\biggl(\ \eta \Bigl( T(x,x,x) \Bigr)\ \biggr)= d_1\biggl( T(x,x,x)
3999: \biggr)= d_3(x) \qquad \mbox{for any }\eta\in \grz.
4000: $$
4001: 
4002: In this computation we used the definition of $d_3$, the relation
4003: $(\ref{autot})$, and the invariance statement for $d_1$.
4004: 
4005: \medskip
4006: The statement for $d_2$ follows from the invariance of the skew-symmetric
4007: form $\{\cdot, \cdot\}$ and the argument analogous to that in the previous
4008: paragraph.
4009: 
4010: 
4011: {\bf (ii)}
4012: The computation ~$(\ref{r1eq})$ of Lemma~$\ref{lem:rank1}$ implies that
4013: $$
4014: \mbox{an integer $d$ divides}\quad
4015: 3\,T(x,x,y)+\{x,y\}\,x \quad
4016: \mbox{for any }y\in \mz
4017: $$
4018: if and only if $d$ divides each of the following expressions
4019: $$
4020: 3\alpha\beta-(A,B),\quad
4021: 2(\alpha A-B^\# ), \quad 2(\beta B - A^\# ),\quad
4022: 2Q(x),\quad 2Q(x'),
4023: $$
4024: where $x=\mel \alpha\beta A B$, $x'=\mel \beta \alpha B A$, and
4025: $Q(x)\in \End_\zet(\jz)$ was defined in~$(\ref{q})$.
4026: 
4027: As in the proof of Lemma~$\ref{lem:rank1}$ we have
4028: $$
4029: Q(x')(C)=Q(x)(C)-2Q(x)(\one)\jprod C\quad
4030: \mbox{for any }C\in \jz.
4031: $$
4032: This implies that $d$ divides $2Q(x')$ whenever $d$ divides $2Q(x)$,
4033: and the statement (ii) of the lemma follows.
4034: \ep
4035: 
4036: 
4037: \begin{rem}\label{rem:rpz}
4038: {\rm
4039: It is possible to define alternative quadratic invariant $\dvap$
4040: via
4041: \begin{equation}\label{alt-d2}
4042: \dvap(x)=\gcd\Bigl(
4043: \alpha A-B^\# , \quad \beta B - A^\# ,\quad Q(\bx)
4044: \Bigr).
4045: \end{equation}
4046: 
4047: However the short proof of $\grz$-invariance of $d_2$ in the previous
4048: lemma does not work for $\dvap$. One can still prove that $\dvap$ is
4049: invariant under each of the transformations $\phi(C), \psi(D), T(s)$, and
4050: then using Proposition~$\ref{genz}$ conclude that $\dvap$ is invariant
4051: under the whole $\grz$. This route is quite technical, it was implemented
4052: in Lemma~$2.3.5$ of \cite{kphd}.
4053: 
4054: Our primary use for the invariants $d_i$ is to distinguish elements of $\mz$
4055: lying in different orbits.
4056: We should note that $\dvap$ is ``finer" than $d_2$ in this sense.
4057: For example, using $\dvap$ one can conclude that elements
4058: $$
4059: \mel 1 2 {\diag{1}00} 0
4060: \quad {\rm and}\quad \mel 1 2 {\diag{2}00} 0
4061: $$
4062: lie in distinct $\grz$-orbits, though all $d_i$'s are equal for these two
4063: elements.
4064: 
4065: In this paper will make use of more ``coarse" relations~$(\ref{d2})$,
4066: which are sufficient for our purposes. \ee
4067: }
4068: \end{rem}
4069: 
4070: \begin{rem} \label{rem2}
4071: {\rm
4072: By analogy with Remark~$\ref{rem1}$ we notice that when $x$ is a reduced
4073: element
4074: $$
4075: x=\mel \alpha\beta A 0
4076: $$
4077: then
4078: $%
4079: d_2(x)=\gcd\Bigl(\alpha\beta,\ 2\alpha A,\ 2A^\# \Bigr)
4080: $. \ee%
4081: }
4082: \end{rem}
4083: 
4084: The following theorem in an extended version of Theorem~$\ref{thm:main}$.
4085: It provides the description of degenerate $\grz$-orbits (corresponding the
4086: zero value of the quartic form) in the module $\mz$.
4087: 
4088: \begin{theorem} \label{thm2}
4089: Let $\jz$ be one of $\h{\binz}, \h{\quatz}, \h{\octz}$
4090: and $\mz=\m(\jz)$.
4091: 
4092: 
4093: \begin{itemize}
4094: 
4095: \item
4096: Every element $\bx$ of rank $1$ in the module $\mz$ can be brought to the
4097: form
4098: \begin{equation} \label{thr1}
4099: \mel \alpha 000 \quad {\rm where}\  \alpha = d_1(\bx)
4100: \end{equation}
4101: by an element in the group $\grz$.
4102: 
4103: The set
4104: \begin{equation} \label{thset1}
4105: \left.\left\{\mel k 0 0 0\ \right|\ k\in\zet, k\ge 1\right\}
4106: \end{equation}
4107: is a complete set of {\em distinct\/} orbit representatives of elements of
4108: rank $1$ in $\mz$.
4109: 
4110: \item
4111: Every element $\bx$ of rank $2$ in the module $\mz$ can be brought to the
4112: form
4113: \begin{equation} \label{thr2}
4114: \mel \alpha 0 {\ \diag{a}00} 0
4115: \quad {\rm where}\  \alpha|a,\ \alpha =
4116: d_1(\bx), \ a= d_2(\bx)/\alpha
4117: \end{equation}
4118: by an element in the group $\grz$.
4119: 
4120: The set
4121: \begin{equation} \label{thset2}
4122: \left\{
4123: \left.\mel k 0 {\diag{m}00} 0 \
4124:  \right.|\ k,m\in\zet, k,m>0, k|m\right\}
4125: \end{equation}
4126: is a complete set of {\em distinct\/} orbit representatives of elements of
4127: rank $2$ in $\mz$.
4128: 
4129: \item
4130: In the case $\rank x=3$ or $4$, the group $\grz$ acts transitively on the
4131: set of {\em projective} elements of norm $n$.
4132: Every such element may brought to the form
4133: \begin{equation}\label{hirank}
4134: \mel 1 \varepsilon{\ \diag{1}1k} 0,\qquad
4135: \varepsilon\in\{0,1\},\quad
4136: k\in \zet,
4137: \end{equation}
4138: where $k=\frac{q'(x)-\varepsilon^2}{4}$ and $\varepsilon\equiv
4139: q'(x)\,(\mod 2)$ are uniquely determined by $x$.
4140: \end{itemize}
4141: \end{theorem}
4142: 
4143: {\sc Proof.}
4144: 
4145: We will often use in the proof the facts that transformations in the group
4146: $\grz$ preserve the rank and the invariants $d_i$ (Lemma~$\ref{lem:rank}$,
4147: $\ref{di}$) of elements of $\mz$.
4148: 
4149: We note that if $\rank x \le 2$, then $T(x,x,x)=0$
4150: (Definition~$\ref{def:rank}$).
4151: Such an $x$ is not projective by Proposition~$\ref{prop:proj}$(ii),
4152: and hence Theorem~$\ref{thm:main}$ gives no information about orbits
4153: of such elements.
4154: 
4155: \medskip
4156: {\em The case of\/} rank $1$.
4157: 
4158: Let $\bx$ be an arbitrary element of rank $1$. First we show that $\bx$
4159: can be brought to the form $(\ref{thr1})$. Then we show that elements of
4160: the form $(\ref{thset1})$ with distinct $k$'s lie in distinct orbits of
4161: the group $\grz$.
4162: 
4163: By Lemma~$\ref{bl-fz}$ there exists $\sigma\in \grz$ such that
4164: $$
4165: \sigma (\bx) = \mel \alpha \beta A 0 ,
4166: \quad \qquad \alpha>0,\
4167: \alpha|\beta,\ \alpha |A.
4168: $$
4169: 
4170: Since rank $\bx =1$, it follows from $(\ref{r1s})$ that $\beta =0$ and
4171: $A=0$. So
4172: $$
4173: \sigma (\bx) = \mel \alpha 000
4174: $$
4175: 
4176: is in the desired form.
4177: 
4178: \medskip
4179: Now suppose we have two elements $\bx_1=\mel {k_1}000$ and
4180: $\bx_2=\mel{k_2}000$
4181: of the form $(\ref{thset1})$ lying in the same $\grz$-orbit. Then
4182: $d_1(\bx _1)=k_1$ and $d_1(\bx _2)=k_2$, and since $d_1$ is constant on
4183: orbits, we have $k_1=k_2$, and hence $x_1=x_2$.
4184: The proof in the case of rank $1$ is complete.
4185: 
4186: 
4187: 
4188: \medskip
4189: {\em The case of\/} rank $2$.
4190: 
4191: Let $\bx$ be an arbitrary element of rank $2$. First we show that $\bx$
4192: can be brought to the form~$(\ref{thr2})$. Then we show that elements of
4193: the form~$(\ref{thset2})$ with distinct $k$'s and $m$'s lie in distinct
4194: orbits of the group $\grz$.
4195: 
4196: By Lemma~$\ref{bl-fz}$ there exists $\sigma\in \grz$ such that
4197: $$
4198: \sigma (\bx) = \mel\alpha\beta A 0, \quad \qquad \alpha>0,\
4199: \alpha|\beta,\ \alpha |A.
4200: $$
4201: 
4202: Since rank $\bx =2$, it follows from $(\ref{r2s})$ that $\beta =0$ and
4203: $A^\# =0$. The last relation implies that rank $A\le 1$. On the other
4204: hand, $A\ne 0$ as $A=0$ would imply that rank $x \le 1$ (see
4205: $(\ref{r1s})$). Hence rank $A=1$.
4206: 
4207: By Theorem~$\ref{np-orbits}$,
4208: there exists $\xi\in \npg\jz$ which brings $A$ to the
4209: Smith normal form. Since $\rank A =1$,
4210: this normal form has only one nonzero (in fact, positive)
4211: element on the diagonal. We thus get
4212: 
4213: $$
4214: \tf{\xi}\, \sigma \ (\bx)= \mel \alpha 0 {\diag{a_1}00} 0.
4215: $$
4216: 
4217: By construction $a_1 = \gcd(A)$, and so $\alpha | a_1$.
4218: 
4219: Hence we have brought $\bx$ to the desired form~$(\ref{thr2})$.
4220: \medskip
4221: 
4222: Now let us take two elements $x_1, x_2$ of the form $(\ref{thset2})$
4223: $$
4224: x_i= \mel {k_i} 0 {\diag{m_i}00} 0\ \qquad k_i,m_i\in\zet, k_i, m_i > 0,
4225: k_i|m_i, i=1,2
4226: $$
4227: 
4228: lying in the same $\grz$-orbit. We want to show that $x_1=x_2$.
4229: 
4230: We have $d_1(x_i)=k_i$, and using the fact that $d_1$ is constant on
4231: orbits, we get
4232: $$
4233: k_1=d_1(x_1)=d_1(x_2)=k_2.
4234: $$
4235: 
4236: Next we notice that Remark~$\ref{rem2}$ implies
4237: $$
4238: d_2(x_i)\ =\ 2\, k_i\, m_i
4239: $$
4240: 
4241: and hence
4242: $$
4243: 2\,k_1\, m_1=d_2(x_1)=d_2(x_2)=2\, k_2\, m_2,
4244: $$
4245: 
4246: which implies that in this case $m_1=m_2$, and hence $x_1=x_2$.
4247: This completes the proof in the case of rank $2$.
4248: 
4249: \medskip
4250: {\em The case of\/} rank $>2$.
4251: A complete reduction procedure is not known in the case of elements of
4252: rank $3$ and $4$, and the structure of $\grz$-orbits may be quite complicated.
4253: For example, when $\jz=\h\binz$, the structure of orbits
4254: in $\mz$ is essentially equivalent to the structure of $\SL_6(\zet)$-orbits
4255: in $\wedge^3(\zet^6)$ (see Example~$\ref{ex:sl6int}$).
4256: It is as complicated as the structure of
4257: (balanced) triples of ideal classes in quadratic rings
4258: \cite[Theorem~$18$]{bh1}.
4259: 
4260: Orbits of the projective elements are the most interesting from the
4261: viewpoint of number theory in this case, and the transitivity result
4262: stated in the theorem follows from Lemma~$\ref{red2}$ (it was also treated
4263: in more detail in Theorem~$\ref{thm:main}$). The case $k=0$ corresponds to
4264: the projective elements of rank~$3$ ($n=q'(x)=0$), and the case $k\ne 0$
4265: corresponds to non-degenerate orbits ($q'(x)\ne 0$).
4266: 
4267: We notice that in the case $\jz=\h\binz$ the result of the theorem remains
4268: valid for the subgroup $\grz\cc$ isomorphic to $\SL_6(\zet)$.
4269: \ep
4270: 
4271: \input{appendix.tex}
4272: 
4273: %
4274: 
4275: 
4276: \begin{thebibliography}{99}
4277: 
4278: \bibitem{b1} %
4279: N. Bourbaki. {\it \'El\'ements de math\'ematique. Groupes et alg\`ebres de Lie.} Chapitre
4280: IV--VI. Hermann, Paris, 1968.
4281: 
4282: \bibitem{bh-phd} %
4283: M. Bhargava. {\em Higher composition laws.} Ph.D. Thesis, Princeton
4284: University, 2001.
4285: 
4286: \bibitem{bhau} %
4287: M. Bhargava. {\em Gauss composition and generalizations.}
4288: Algorithmic Number Theory: 5th International Symposium, Sydney, Australia.
4289: Lecture Notes in Computer Science, 2002.
4290: %
4291: 
4292: \bibitem{bh1} %
4293: M. Bhargava.
4294: {\it Higher composition laws I:
4295: a new view on Gauss composition, and quadratic generalizations.}
4296: Ann. of Math.(2), {\bf 159}(2004), no. 1, pp. 217--250.
4297: 
4298: 
4299: \bibitem{b} %
4300: R. Brown. {\it Groups of type $E_{7}$.} J. Reine Angew. Math. {\bf 236}(1969), pp.
4301: 79--102.
4302: 
4303: \bibitem{clerc} %
4304: J.-L. Clerc.
4305: {\em Special prehomogeneous vector spaces associated
4306: to $F\sb 4, E\sb 6, E\sb 7, E\sb 8$ and
4307: simple Jordan algebras of rank $3$.}
4308: J. Algebra,  {\bf 264}(2003), no. 1, pp. 98--128.
4309: 
4310: \bibitem{dg} %
4311: P. Deligne, B.H. Gross.
4312: {\em On the exceptional series, and its descendants.}
4313: C. R. Math. Acad. Sci. Paris, {\bf 335}(2002), no. 11, pp. 877--881.
4314: 
4315: \bibitem{fau} %
4316: J. Faulkner.
4317: {\em A geometry for $E_7$.}
4318: Trans. Amer. Math. Soc. {\bf 167}(1972), pp. 49--58.
4319: 
4320: \bibitem{fer}  %
4321: J.C. Ferrar.
4322: {\em Strictly regular elements in Freudenthal triple systems.}
4323: Trans. Amer. Math. Soc., {\bf 174}(1972), pp. 313--331 (1973).
4324: 
4325: \bibitem{fr14}  %
4326: H. Freudenthal.
4327: {\em Beziehungen der $E_7$ and $E_8$ zur Oktavenebene.} I--IV.
4328: Indagationes Math.
4329: {\bf 16}(1954) pp. 218--230,
4330: {\bf 16}(1954) pp. 363--368,
4331: {\bf 17}(1955) pp. 151--157,
4332: {\bf 17}(1955) pp. 277--285.
4333: 
4334: \bibitem{gar} %
4335: R.S. Garibaldi.
4336: {\em Structurable algebras and groups of type $E\sb 6$ and $E\sb 7$.}
4337: J. Algebra {\bf 236}(2001), no.~2, pp. 651--691.
4338: 
4339: \bibitem{gw} %
4340: B. Gross, N. Wallach.
4341: {\em On quaternionic discrete series representations, and their
4342: continuations.}
4343: J. Reine Angew. Math. {\bf 481} (1996), pp. 73--123.
4344: 
4345: \bibitem{hum} %
4346: J.E.~Humphreys.
4347: {\em Linear algebraic groups.} GTM 21.
4348: Springer-Verlag, 1975.
4349: 
4350: \bibitem{J3} %
4351: N. Jacobson.
4352: {\em Some Groups of transformations defined by Jordan algebras. III.}
4353: {J.~Reine Angew. Math.} {\bf 207}(1961), pp. 61--85.
4354: 
4355: \bibitem{ja} %
4356: N. Jacobson. {\em Structure and representations of Jordan algebras.}
4357: AMS Colloquium Publications, Vol. XXXIX. Providence, R.I., 1968.
4358: 
4359: \bibitem{jba} %
4360: N. Jacobson.
4361: {\em Basic algebra. I. Second edition.}
4362: W. H. Freeman and Company, New York, 1985.
4363: 
4364: \bibitem{kac} %
4365: V. Kac.
4366: {\em Classification of simple $Z$-graded
4367: Lie superalgebras and simple Jordan superalgebras.}
4368: Comm. Algebra. {\bf 5} (1977), no. 13, pp. 1375--1400.
4369: 
4370: \bibitem{koe} %
4371: M. Koecher. {\em Imbedding of Jordan algebras into Lie algebras. I.}
4372: Amer. J. Math, {\bf 89}(1967), pp. 787--816.
4373: 
4374: \bibitem{k} %
4375: S. Krutelevich. {\em On a canonical form of a $3\times 3$ Hermitian matrix over the
4376: ring of integral split octonions.} J. Algebra, {\bf 253}(2002), no. 2,
4377: pp. 276--295.
4378: 
4379: \bibitem{kphd} %
4380: S. Krutelevich. {\em Orbits of exceptional groups and Jordan systems.}
4381: Ph.D. thesis, Yale University, 2003.
4382: 
4383: \bibitem{lm} %
4384: J. Landsberg, M. Manivel.
4385: {\em The projective geometry of Freudenthal's
4386: magic square.}
4387: J. Algebra, {\bf 239} (2001), no. 2, pp. 477--512.
4388: 
4389: \bibitem{mms} %
4390: J. Maldacena, G. Moore, A. Strominger.
4391: {\it Counting BPS Blackholes in Toroidal
4392: Type II String Theory.}
4393: arXiv:hep-th/9903163.
4394: 
4395: 
4396: 
4397: \bibitem{mc-sp} %
4398: K. McCrimmon.
4399: {\em The Freudenthal-Springer-Tits constructions of exceptional Jordan
4400: algebras.}
4401: Trans. Amer. Math. Soc. {\bf 139} 1969, pp. 495--510.
4402: 
4403: 
4404: \bibitem{atoj} %
4405: K. McCrimmon.
4406: {\em A taste of Jordan algebras.}
4407: Universitext. Springer-Verlag, New York, 2004.
4408: 
4409: \bibitem{newman} %
4410: M. Newman. {\em Integral matrices.}
4411: Pure and Applied Mathematics, v. 45.
4412: Academic Press, 1972.
4413: 
4414: \bibitem{SK} %
4415: M. Sato and T. Kimura.
4416: {\em A classification of irreducible prehomogeneous vector spaces and their
4417: relative invariants.}
4418: Nagoya Math. J. {\bf 65} (1977), pp. 1--155.
4419: 
4420: \bibitem{spr} %
4421: T.A. Springer.
4422: {\em  Characterization of a class of cubic forms.}
4423: Nederl. Akad. Wetensch. Proc. Ser. A 65
4424: = Indag. Math. {\bf 24} 1962, pp. 259--265.
4425: 
4426: \bibitem{sp-alg} %
4427: T.A. Springer.
4428: {\em Jordan algebras and algebraic groups.}
4429: Springer-Verlag, 1973 (reprinted in 1998).
4430: 
4431: \bibitem{SpV} %
4432: T.A. Springer, F.D. Veldkamp.
4433: {\em Octonions, Jordan algebras and exceptional groups.}
4434: Springer-Verlag, Berlin, 2000.
4435: 
4436: \bibitem{tits} %
4437: J. Tits.
4438: {\em Une classe d'alg\`ebres de Lie en relation avec les alg\`ebres de
4439: Jordan.}
4440: Nederl. Akad. Wetensch. Proc. Ser. A 65 = Indag. Math.,
4441: {\bf 24} 1962, pp. 530--535.
4442: 
4443: \end{thebibliography}
4444: 
4445: {\sc
4446: Department of Mathematics and Statistics,
4447: University of Ottawa
4448: 
4449: 585 King Edward Ave.,
4450: Ottawa, ON, K1N 6N5, Canada
4451: }
4452: 
4453: \medskip
4454: E-mail: {\tt sergei.krutelevich@science.uottawa.ca}
4455: \end{document}
4456: