1: \documentclass[10pt,a4paper]{article}
2:
3: \usepackage[latin1]{inputenc}
4: \usepackage[T1]{fontenc}
5: \usepackage[english]{babel}
6: \usepackage{amsfonts}
7: \usepackage{euscript,amsthm,pstricks}
8: \usepackage[vcentermath]{youngtab}
9:
10: \setlength{\textheight}{21.5cm}
11:
12: \newtheorem{conj}{Conjecture}[section]
13: \newtheorem{theo}[conj]{Theorem}
14: \newtheorem{hypo}[conj]{Hypothesis}
15: \newtheorem{prop}{Proposition}[section]
16: \newtheorem{coro}[prop]{Corollary}
17: \newtheorem{lemm}{Lemma}[section]
18: \newtheorem{affi}[lemm]{Affirmation}
19: \newtheorem{clai}[lemm]{Claim}
20: \newtheorem{fait}[lemm]{Fact}
21: \newtheorem{defi}{Definition}[section]
22: \newtheorem{nota}[defi]{Notation}
23: \newtheorem{exem}[defi]{Example}
24: \newtheorem{naiv}[defi]{Naive guess}
25: \newenvironment{proo}{{\flushleft \bf Proof :}}{\hfill $\square$ \vspace{2mm}}
26: \newenvironment{rema}{{\flushleft \bf Remark :}}{\hfill}
27: \begin{document}
28:
29: %\input /home/pe/latex/commandes_pe.tex
30: %\newcommand{\color}[1]{}
31: \input commandes_pe.tex
32:
33: \title{Dual varieties of subvarieties of homogeneous spaces}
34: \author{Pierre-Emmanuel Chaput}
35: \maketitle
36:
37: {\def\thefootnote{\relax}
38: \footnote{\hskip-0.6cm
39: {\it AMS mathematical classification \/}: 14N99, 14L35, 14L40. \\
40: {\it Key-words\/}: dual variety, homogeneous space, projective geometry. }}
41:
42: \begin{center}
43: {\bf Abstract}
44: \end{center}
45:
46: If $X \subset \p^n_\C$ is an algebraic complex projective variety, one
47: defines the dual variety $X^* \subset {(\p^n)}^*$ as the set of
48: tangent hyperplanes to $X$.
49: The purpose of this paper is to generalise this notion when $\p^n$ is
50: replaced by a quite general partial flag variety. A similar
51: biduality theorem is proved, and the dual varieties of Schubert varieties
52: are described.
53: \lpara
54:
55: \begin{center}
56: {\bf Introduction}
57: \end{center}
58:
59: Let $X \subset \p V$ be a complex projective algebraic variety, with
60: $V$ a $\C$-vector space. If $h \in \p V^*$ is a hyperplane and $x \in X$
61: is a smooth point, we say that $h$ is tangent to $X$ at $x$ if $h$
62: contains the embedded tangent space of $X$ at $x$. Equivalently, the
63: intersection $X \cap h$ is singular at $x$. The closure of
64: the set of all $h \in \p V^*$
65: which are tangent at some smooth point of $X$
66: is denoted $X^*$ and called the
67: dual variety of $X$; for given $h \in X^*$, the closure
68: of the set of smooth points $x \in X$ such that $h$ is tangent at $x$
69: is called the tangency locus of $h$.
70:
71: This notion of dual varieties is a very classical one, and it is used
72: plentifully in both classical and modern articles. The very powerfull
73: biduality theorem, to the effect that ${(X^*)}^* = X$,
74: and its corollary, which states that the tangency locus
75: at a smooth point $h \in X^*$ is a linear space, are ubiquitous.
76: To state only one example, this result is crucial in Zak's classification
77: of Severi varieties, since it allows proving that the entry locus of
78: a Severi variety is a smooth quadric \cite[proposition IV.2.1]{zak}.
79:
80: This biduality
81: theorem deals with subvarieties of projective space, which have
82: been studied by so many classical algebraic geometers. More recently, work has
83: been done in a new direction which consists in
84: considering subvarieties of other homogeneous
85: spaces. For example, G. Faltings \cite{faltings}
86: and O. Debarre \cite{debarre1,debarre2} have shown
87: some connectivity theorems that hold
88: in an arbitrary homogeneous space, E. Arrondo has proved a classification
89: of some subvarieties of Grassmannians similar to Zak's result
90: \cite{arrondo_severi}, and some
91: topological results on low-codimensional subvarieties of some homogeneous
92: spaces emerge in works of E. Arrondo - J. Caravantes
93: \cite{arrondo_low} and N. Perrin \cite{perrin_low}.
94:
95: \lpara
96:
97: Obviously, to study subvarieties of homogeneous spaces, a similar
98: notion of dual variety and a biduality theorem are lacking. The aim of
99: this article is to fill this gap as much as it is possible.
100:
101: Since homogeneous spaces $G/P$ are by definition projective algebraic
102: varieties, it is certainly possible to embed them in a projective space, and
103: therefore a subvariety $X \subset G/P$ is a fortiori a
104: subvariety of a projective space, so that one can consider
105: the usual dual variety of $X$.
106:
107: However I claim that in many cases this is not the best thing to
108: do. Let us consider an example. Let $V$ be a $\C$-vector space
109: equipped with a non-degenerate quadratic form. If $Q \subset \p V$ is
110: the smooth
111: quadric it defines, then it is well-known that $Q^* \subset \p V^*$ is also a
112: smooth quadric, canonically isomorphic with $Q$.
113: Now, let $r$ be an integer and let us consider the variety $\G_Q(r,V)$
114: parametrising $r$-dimensional isotropic subspaces as a subvariety of
115: a suitable projective space using Plücker embedding. Then clearly we no
116: longer have $\G_Q(r,V)^* \simeq \G_Q(r,V)$. On the contrary,
117: let us consider $\G_Q(r,V)$ as a subvariety of the Grassmannian
118: $\G(r,V)$;
119: proposition \ref{prop_symmetric}
120: shows that for my definition of dual varieties, $\G_Q(r,V) \subset \G(r,V)$
121: has a well-defined
122: dual variety in $\G(r,V^*)$ which is canonically isomorphic with
123: $\G_Q(r,V)$.
124:
125: In fact, homogeneous spaces are often minimally
126: embedded in projective spaces of very big dimension, so that the usual
127: dual variety of a subvariety of a homogeneous space will happen to be
128: very large and often untractable. A notion of dual varieties within
129: homogeneous spaces is probably more suitable if one wants to deal with
130: low-dimensional or low-codimension subvarieties (of course, the price to
131: pay is that the ambient space is a bit more complicated than a projective
132: space).
133:
134: \lpara
135:
136: My definition of dual varieties uses a class of birational
137: transformations called stratified Mukai flops by Namikawa
138: \cite{namikawa}. These are birational maps
139: $\mu:T^*G/P \dasharrow T^*G/Q$ defined in terms of nilpotent orbits
140: for some semi-simple group $G$ and some parabolic subgroups
141: $P,Q$. For
142: given $G,P,Q$, if there exists such a map, then we say that
143: $G/P$ and $G/Q$ allow
144: duality. For $X \subset G/P$, we consider its conormal bundle
145: $N^*X \subset T^*G/P$ and define the dual variety
146: $X^Q = \pi_Q \circ \mu (N^*X) \subset G/Q$ ($\pi_Q : T^* G/Q
147: \rightarrow G/Q$ denotes the projection) if $N^*X$ meets the locus
148: where $\mu$ is defined (in which case we say that $X$ is suitable).
149: For example, if $G = SL(V)$ and
150: $G/P = \p V$, the only possibility for $Q$
151: leads to $G/Q = \p V^*$; any
152: proper subvariety $X \subset \p V$ will be suitable and $X^Q = X^*$.
153: Another example is the fact that a Grassmannian $\G(r,V)$ and its dual
154: Grassmannian $\G(r,V^*)$ allow duality, as one could naturally expect.
155:
156: One advantage of this definition is that it uses the so-called
157: Springer resolutions of the corresponding nilpotent orbit, which are
158: symplectic resolutions, and this
159: article uses heavily informations which come from the study of such
160: resolutions \cite{namikawa,flop_scorza}.
161: Another advantage is that it exhibits
162: the symplectic nature of dual varieties. In fact,
163: $T^*G/P$ and $T^*G/Q$
164: are symplectic varieties and $N^*X$, as a subvariety of
165: $T^* G/P$, is a Lagrangian subvariety. These properties suffice to
166: show very easily the biduality theorem \ref{theo_bidualite} in our
167: setting.
168:
169: \lpara
170:
171: However, this definition also has its drawbacks. The most important is
172: probably that it is not so much intuitive, so that given $x \in X$ and
173: $h \in X^Q$, it is not obvious at all what the sentence ``$h$ is
174: tangent to $X$ at $x$'' should mean. However,
175: in the case of a Grassmannian,
176: using the natural rational
177: map $Hom(\C^r,V) \dasharrow \G(r,V)$, I show that the
178: dual variety of $X \subset \G(r,V)$ can be computed in terms of the usual
179: dual variety of an adequate subvariety of $\p Hom(\C^r,V)$ (see subsection
180: \ref{subsection_grassmannienne}).
181: Therefore, this is a way of understanding more easily
182: dual varieties in the case of Grassmannians.
183: In general however, there are two fundamental
184: differences between our setting and usual duality.
185:
186: First of all, given $G,P$, there may be many different $Q$'s, or none,
187: such that $G/P$ and $G/Q$ allow duality. Therefore, given suitable
188: $X \subset G/P$, we will get a dual variety $X^Q$ for each such $Q$. If one
189: restricts to maximal parabolic subgroups, thanks to
190: \cite{namikawa}, this difficulty disappears
191: because for given $G/P$ there will be at most one parabolic subgroup
192: $Q$ such that
193: $G/P$ and $G/Q$ allow duality. Moreover, section
194: \ref{section_reduction} shows that one can understand all dual
195: varieties if they are understood when $P$ and $Q$ are maximal parabolic
196: subgroups. These cases are therefore called fundamental cases. They
197: include the duality between the Grassmannian $\G(r,V)$ and its dual
198: Grassmannian $\G(r,V^*)$, but also a duality
199: between the two spinor varieties of a quadratic space of dimension
200: $4p+2$, and between the exceptional homogeneous spaces
201: $E_6/P_1 \leftrightarrow E_6/P_6$ and $E_6/P_3 \leftrightarrow E_6/P_5$.
202:
203: The second difference is that not all proper subvarieties
204: $X \subset G/P$ will have a dual variety. Note that $X = \p V$ has no
205: dual variety in $\p V^*$, because for any $x \in X$,
206: no non-zero cotangent form can vanish
207: on $T_xX$. From this point of view, the situation is
208: quite similar in our setting~: too big subvarieties $X$ of $G/P$ don't have
209: dual varieties because for any $x \in X$
210: there is no generic cotangent form in $T^*_x G/P$ which
211: vanishes on $T_xX$.
212:
213: \lpara
214:
215: In the classical setting, a hyperplane $h$ is tangent to $X$ at $x$
216: \iff the intersection $h \cap X$ is singular.
217: As I already alluded to, I have not been able to give a similar
218: geometric notion of ``tangent element''. The only sensible definition
219: seemed to state that $h \in X^Q$ is tangent to $X$ at $x \in X$ if $h$
220: belongs to the image of $N^*_xX$ under $\pi_Q \circ \mu$.
221: Since there is an incidence
222: variety in $G/P \times G/Q$ (the closed $G$-orbit), any $h \in G/Q$
223: still defines, exactly as in the classical situation, a subvariety $I_h
224: \subset G/P$. Lemma \ref{lemm_q} implies that if $h$ is tangent to $x$
225: at $X$, then the intersection $I_h \cap X$ is not transverse, but the
226: reciprocal of this fact is false.
227:
228: Section \ref{section_tangency} deals with this matter. Corollary
229: \ref{coro_incident} states that if $h$ is tangent to $X$ at $x$, then
230: $x \in I_h$. For $x \in X$ with $X$
231: suitable, the tangent cone $\overline{T_xX} \subset G/P$ of $X$ at $x$
232: is defined in a roundabout manner as the dual variety
233: of the variety of $h$'s in $X^Q$ which are
234: tangent to $X$ at $x$. Theorem \ref{theo_cone} implies that
235: $\overline{T_xX}$ is a ``cone'' with vertex $x$, where
236: definition \ref{def_cone}
237: generalises the classical notion of
238: cones from subvarieties of projective space to subvarieties of
239: fundamental homogeneous spaces.
240:
241: \lpara
242:
243: Finally, section \ref{section_exemple} studies dual
244: varieties of Schubert varieties. In the classical setting, the dual
245: variety of a linear subspace is again a linear subspace. In our
246: setting, it is a formal consequence of the definitions that the dual
247: variety of a Schubert variety is again a Schubert
248: variety (see proposition
249: \ref{prop_schubert} which relies on the functorial property of dual
250: varieties given in proposition \ref{foncteur}).
251:
252: Let $B \subset G$ be a Borel subgroup. It turns out that the combinatorial
253: involution $X \mapsto X^Q$ between $B$-stable suitable
254: Schubert subvarieties of $G/P$ and $G/Q$
255: is no longer decreasing, as it was the case for $G/P = \p V$.
256: For this reason, the description of this map is quite intricate. In
257: the case of Grassmannians and spinor varieties, we give explicitly
258: in terms of
259: partitions the map $X \mapsto X^Q$,
260: see propositions \ref{grass_adapte} and \ref{prop_spinoriel}.
261: This relies on a general
262: recipy for finding $X^Q$ when $X$ is a Schubert variety
263: which is given in subsection
264: \ref{subsection_combinatoire}. For the exceptional cases, this recipy
265: theoretically defines the involution (there is only a finite number of
266: calculations to do to compute the dual variety of a Schubert
267: subvariety), but I will not give a more
268: explicit description of it. As a first step, I describe a criterion
269: for a Schubert subvariety to be suitable. Remarkably enough, this
270: criterion can be stated in a uniform way for all the fundamental
271: cases, using the combinatorics of some quivers studied in
272: \cite{carquois}~: see theorem \ref{theo_adapte}.
273:
274: \lpara
275: {\bf Further questions :}
276: Of course this study only gives basic properties of our generalised
277: dual varieties~: if one compares with usual dual varieties, what
278: essentially has been proved is the biduality theorem and the
279: computation of the dual variety of a quadric and a linear subspace. The
280: power of the classical notion of dual varieties gives hope to me that much more
281: can be said on this topic, including~:
282: \begin{itemize}
283: \item
284: Is it true that for a smooth subvariety $X \subset G/P$
285: one has $\dim\ X^Q \geq \dim\ X$ ? This question has been raised by
286: Laurent Manivel.
287: \item
288: Many Fano 3-folds are defined as subvarieties of some homogeneous
289: spaces. What are the dual varieties of these Fano 3-folds ?
290: \item
291: What is the dual variety of a divisor in $G/P$ ? If this is a divisor,
292: what is the degree of this divisor ? The
293: answer to this question for $G/P = \G(2,V)$ or $G/P = E_6/P_1$ and a
294: divisor of degree 1 has been given in \cite{hermitian} (the dual
295: variety is again a divisor of degree 1).
296: \item
297: Classification problems~: for example find all smooth varieties
298: with dual variety a divisor of low degree.
299: \end{itemize}
300:
301: \lpara
302: \noindent
303: {\bf Acknowledgements : }I am very grateful to M. Brion for suggesting
304: that maybe nilpotent orbits could help defining an
305: interesting equivariant rational map
306: $T^*G/P \dasharrow G/Q$, as it was finally exactly the case. Thanks
307: are also due to B. Fu who pointed to me the reference \cite{namikawa}.
308:
309: \tableofcontents
310:
311: \sectionplus{Definition of the dual variety}
312:
313: \label{section_definition}
314:
315: \subsectionplus{Notations and definition}
316:
317: In this subsection, I introduce the (abstract) definition of dual
318: varieties, which allows easy proofs of general results; in subsection
319: \ref{subsection_grassmannienne}, an equivalent but more ``down-to-earth''
320: definition will be given in the case of Grassmannians.
321:
322: \lpara
323:
324: Before giving this definition, which is not so intuitive, I give some
325: ``naive guesses'' and explain why the corresponding notion of dual
326: varieties would not be interesting. In this way, I hope to convince
327: the reader that it is not possible to avoid some technicalities.
328: Let us try our unsuccessful experiments in the case of Grassmannians.
329:
330: So assume $G/P = \G(r,V)$ and $G/Q = \G(r,V^*)$ and assume
331: $2r<\dim V$. Any element
332: $h \in \G(r,V^*)$, representing a codimension $r$
333: subspace of $V$ denoted $L_h$, defines (at
334: least) two subvarieties in $\G(r,V)$. The first (resp. the second)
335: is the subvariety of $x \in \G(r,V)$ such that
336: $L_x \subset L_h$ (resp. $\dim(L_x \cap L_h)>0$). It will be denoted
337: $I_h$ (resp. $h^\bot$).
338: Assume $X \subset \G(r,V)$ is a subvariety and $x \in X$.
339: In the following, we give some naive definitions of the fact that $h$
340: is tangent to $X$ at $x$ in terms of the intersection of $X$, $h^\bot$
341: and $I_h$.
342:
343: \begin{naiv}
344: ``$h$ is tangent to $X$ at $x$ if $x \in I_h$
345: and the intersection $h^\bot \cap X$ is
346: singular at $x$.''
347: \end{naiv}
348: \noindent
349: This is really stupid, because if $x \in I_h$, then
350: $h^\bot$ is singular at $x$, and so is the intersection
351: $h^\bot \cap X$. So any $h$
352: such that $L_x \subset L_h$ will satisfy this condition, regardless to
353: the tangent space $T_xX$.
354:
355: \begin{naiv}
356: ``$h$ is tangent to $X$ at $x$ if $x \in h^\bot$
357: and the intersection $h^\bot \cap X$ is
358: singular at $x$.''
359: \end{naiv}
360: \noindent
361: For the same reason as above, it suffices that $L_h$ contains $L_x$ in
362: order that this condition holds. So if we define $X^*$ as the set of
363: $h$'s satisfying the above condition, we will not have a biduality
364: theorem. In fact, if for example $X=\{x\}$ is a point, then $X^*$ will
365: contain $\{h:L_h \supset L_x\}$ and ${(X^*)}^*$ will certainly not be
366: reduced to $\{x\}$.
367:
368: \lpara
369:
370: Therefore, it seems necessary to use the smooth subvariety $I_h$. In
371: this case, assuming that $I_h \cap X$ is singular is not quite
372: acurate, because $I_h$ has codimension larger than 1, so this
373: condition should be replaced by the fact that the intersection is not
374: transverse~:
375: \begin{naiv}
376: ``$h$ is tangent to $X$ at $x$ if $x \in I_h$ and the
377: intersection $I_h \cap X$ is not transverse at $x$.''
378: \end{naiv}
379: \noindent
380: Again, if we take $X = \{x\}$, then
381: $X^* = \{h : L_h \supset L_x \}$, and
382: ${(X^*)}^* = \{y : \dim(L_x \cap L_y) > 0 \}$. So we don't have a
383: biduality theorem.
384:
385: \lpara
386:
387: Of course we could multiply such definitions; let us just consider one
388: more possibility~:
389: \begin{naiv}
390: ``$X^*$ is the intersection of the usual dual variety of $X$ (in the
391: Plücker embedding) with $\G(r,V^*)$.''
392: \end{naiv}
393: \noindent
394: Already in case $r=2$ and $\dim V$ even, it is easy to see that
395: biduality will not hold. Let again $X = \{x\}$. The usual dual variety
396: of $X$ in the Plücker embedding is a hyperplane; therefore $X^*$ will
397: be a hyperplane section of $\G(2,V^*)$. As it is well-known,
398: the dual variety of
399: $\G(2,V^*) \subset \p \wedge^2 V^*$ is a hypersurface in
400: $\p \wedge^2 V$. Therefore it follows that the usual dual variety of
401: $X^* \subset \p \wedge^2 V^*$
402: will have codimension at most 2 in $\p \wedge^2 V$. Thus
403: its intersection ${(X^*)}^*$ with $\G(2,V)$ cannot be a point.
404:
405: \para
406:
407: I hope that the previous unsuccessfull experiments will convince the
408: reader to accept a more conceptual definition of generalised dual
409: varieties.
410: Let $G$ be a semi-simple simply-connected complex algebraic group with
411: Lie algebra $\g$, and let $\g^*$ be the dual vector space of $\g$.
412: We fix $T\subset B \subset G$ a
413: maximal torus and a Borel subgroup of $G$.
414: If $P \subset G$ is a
415: parabolic subgroup, let $G/P$ denote the corresponding flag variety.
416: If $X$ is a variety, let $T^*X$ denote its cotangent bundle;
417: we denote $t_P:T^*G/P \rightarrow \g^*$ the natual map.
418:
419: \begin{defi}
420: Let $P,Q$ be parabolic subgroups of $G$.
421: We say that $G/P$ and $G/Q$ allow duality if there is a nilpotent
422: orbit ${\cal O} \subset \g^*$ such that
423: $t_P:T^*G/P \rightarrow \g^*$ and $t_Q:T^*G/Q \rightarrow \g^*$ are birational
424: isomorphisms between the cotangent bundles and $\cal O$.
425: \label{defi_dualite}
426: \end{defi}
427:
428: Assume that $G/P$ and $G/Q$ allow duality. The birational map
429: $t_Q^{-1} \circ t_P : T^*G/P \dasharrow T^* G/Q$ will be denoted
430: $\mu$. Let $\co$ be such that $t_P(T^*G/P) = t_Q(T^*G/Q) = \overline \co$.
431: Let $X \subset G/P$ be any subvariety. Let $X^{sm}$ denote its smooth
432: locus and let
433: $N^*X \subset T^* G/P$ denote the conormal bundle to $X^{sm}$~: we have
434: $(x,f) \in N^*X$ \iff $x \in X^{sm}$,
435: $f \in T^*_xG/P$, and $f_{|T_xX} = 0$.
436:
437: \begin{defi}\
438: \label{defi_adapte}
439: \begin{itemize}
440: \item
441: A form $f \in T^*G/P$ (resp. $f\in T^*G/Q$) is called generic if it
442: belongs to $t_P^{-1}(\co)$ (resp. $t_Q^{-1}(\co)$).
443: \item
444: A subvariety $X \subset G/P$
445: is suitable if it is irreducible and there are generic forms
446: in $N^*X$.
447: \item
448: A point $x$ of a suitable variety $X$ is itself suitable if there
449: are generic forms in $N^*_xX$.
450: Let $X^s$ denotable the locus of suitable
451: points of $X$.
452: \end{itemize}
453: \end{defi}
454: \rek
455: One could also consider reducible suitable subvarieties : they
456: would be subvarieties such that every irreducible component is
457: suitable; we could then define the dual variety of a reducible suitable
458: variety as the union of dual varieties of its irreducible
459: components.
460:
461: \begin{nota}
462: Let $\pi_P$ denote the projection $T^* G/P \rightarrow G/P$.
463: \end{nota}
464:
465: \begin{defi}
466: If $X \subset G/P$ is suitable then we define $X^Q \subset G/Q$ as
467: the image of $N^*X$ by the rational map
468: $\pi_Q \circ \mu$.
469: \end{defi}
470:
471: In the rest of the article, $P$ and $Q$ will denote parabolic
472: subgroups of a reductive simply-connected group $G$ allowing duality.
473: Moreover, we denote
474: $p := \pi_P \circ \mu\ : T^* G/Q \dasharrow G/P$,\
475: $q := \pi_Q \circ \mu^{-1}\ : T^* G/P \dasharrow G/Q$
476: the relevant rational maps. Finally, let
477: $\co \subset \g^*$ be the $G$-orbit which is dense in
478: $t_P(T^* G/P) = t_Q(T^* G/Q)$.
479:
480: \begin{defi}\
481: \label{i_h}
482: \begin{itemize}
483: \item
484: Let $x \in X \subset G/P$. We say that $h \in G/Q$ is tangent to $X$
485: at $x$ if $h \in q(N^*_x X)$.
486: \item
487: If $h \in G/Q$, let $I_h$ denote the Schubert variety of elements in
488: $G/P$ which are incident to $h$, in the sense that $x$ is incident to
489: $h$ if the intersection of the stabilisors of $x$ and $h$ (in $G$)
490: contain a Borel subgroup.
491: \item
492: As a corollary of Borel's conjugacy theorem, $I_h$ is homogeneous
493: under the stabilisor of $h$.
494: \end{itemize}
495: \end{defi}
496:
497: \noindent
498: The notion of tangency will be studied in more details in subsection
499: \ref{section_tangency}. Here we only remark the following~:
500:
501: \begin{fait}
502: \label{fait_tangent}
503: If $h$ is tangent to $X$ at $x$, then the intersection
504: $I_h \cap X$ is not transverse at $x$.
505: \end{fait}
506: \noindent
507: The proof of this fact is postponed to section
508: \ref{section_tangency}~: see lemma \ref{lemm_q}.
509: Note that the converse does not hold in general, contrary
510: to the case when $G/P = \p V$.
511:
512:
513:
514:
515: \subsectionplus{Fundamental cases}
516:
517: \label{subsection_fondamental}
518:
519: \begin{defi}
520: Let $P,Q \subset G$ allow duality. We say that $P,Q,G$ is a
521: fundamental case if one of the following hold :
522: \begin{itemize}
523: \item
524: $G=SL_n$, $P$ and $Q$ are the stabilisors of supplementary subspaces
525: of $\C^n$.
526: \item
527: $G=Spin_{4p+2}$, $P$ and $Q$ are the stabilisors of supplementary
528: (and so of different families) isotropic subspaces of $\C^{4p+2}$.
529: \item
530: $G$ is of type $E_6$, and, with Bourbaki's notations
531: \cite[p.261]{bourbaki} $(P,Q)$ correspond either to the roots
532: $(\alpha_1,\alpha_6)$ or $(\alpha_3,\alpha_5)$.
533: \end{itemize}
534: If this holds, $G/P$ and $G/Q$ are called fundamental homogeneous spaces.
535: \label{fondamental}
536: \end{defi}
537:
538: By \cite[theorem 6.1]{namikawa}, these examples are all the examples of
539: maximal parabolic subgroups
540: allowing duality. Recall that the corresponding rational map
541: $\mu : T^* G/P \dasharrow T^* G/Q$ is called a stratified Mukai flop.
542:
543: Moreover, in all the other cases, the rational map
544: $\mu : T^* G/P \dasharrow T^* G/Q$ (and, as we will see in subsection
545: \ref{subsection_sequence}, the dual varieties) may be described using
546: only fundamental stratified Mukai flops~: let us recall
547: this construction \cite[theorem 6.1]{namikawa}.
548: Assume $P,Q \subset G$ are parabolic subgroups
549: included in a common parabolic subgroup
550: $R$. Then we have fibrations
551: $$
552: \begin{array}{ccccc}
553: G/P & & & & G/Q\\
554: & f_P\searrow & & \swarrow f_Q\\
555: && G/R
556: \end{array}
557: $$
558: with fibers $R/P$ and $R/Q$. Let $U(R)$ denote the unipotent radical of
559: $R$ and $Z(R)$ its connected center; let $L=R/Z(R)U(R)$; $R/U(R)$
560: is isomorphic with
561: a levi factor of $R$ and $L$ is semi-simple.
562: Moreover, $R/P$ and $R/Q$ are $L$-homogeneous varieties~:
563: denote $\pi : R \rightarrow L$ the projection, and denote
564: $P_L:=\pi(P)$ (resp. $Q_L:=\pi(Q)$) we have
565: $R/P \simeq L/P_L$ and $R/Q \simeq L/Q_L$. Assume now that $P_L,Q_L$
566: allow duality. Therefore there is a rational map
567: $\mu_L : T^* L/P_L \dasharrow T^* L/Q_L$
568: which can be used to define the stratified Mukai flop.
569:
570: In fact, let $z \in G/R$, and denote $\cf_z:=f_P^{-1}(z)$
571: (resp. $\cg_z:=f_Q^{-1}(z)$), and let $i_z:\cf_z \rightarrow G/P$
572: (resp. $j_z:\cg_z \rightarrow G/Q$) be the natural inclusions. We have
573: $\cf_z \simeq L/P_L$ and $\cg_z \simeq L/Q_L$.
574: Let $L_z = R_z/Z(R_z)U(R_z)$ denote the group
575: isomorphic with $L$ acting on $\cf_z$ and $\cg_z$.
576: Because $\mu_L$ is canonical, it defines an algebraic family
577: of rational maps
578: $\mu_z : T^* \cf_z \dasharrow T^*\cg_z$ parametrised by $G/R$.
579: Now, if $\alpha$ is an element of $T^* G/P$,
580: say $\alpha \in T^*_x G/P$ with $x \in G/P$, then we can restrict this
581: linear form
582: to $T_x \cf_{f_P(x)}$; this gives an element in the bundle
583: $T^*\cf_{f_P(x)}$ which we denote $f_x$.
584: Finally, recall that $\pi_P:T^* G/P \rightarrow G/P$ and
585: $\pi_Q:T^* G/Q \rightarrow G/Q$ denote the bundle projections.
586: With these notations we have the following proposition~:
587:
588: \begin{prop}
589: \label{prop_fibration}
590: If $f \in T_x^* G/P$ belongs to the open $G$-orbit, then
591: $f_x \in T^*\cf_{f_P(x)}$ belongs to the open $L_{f_P(x)}$-orbit, and
592: $\pi_Q(\mu(f)) = j_{f_P(x)} \circ \pi_{Q_L} \circ \mu_x(f_x)$.
593: \end{prop}
594: \noindent
595: Then, using \cite[theorem 4.1]{flop_scorza},
596: one can deduce a description of the flop
597: $T^* G/P \dasharrow T^* G/Q$.
598: \begin{proo}
599: Since both maps of the proposition are equivariant,
600: we can assume that $x$ corresponds to the base point in $G/P$.
601: If the restriction of $f$ to $T_x \cf_{f_P(x)}$ would belong to a
602: closed $L$-stable strict subvariety of $T^* \cf_{f_P(x)}$, then forms
603: in the
604: $G$-orbit of $f$ would restrict to non generic forms; therefore this
605: $G$-orbit could not be dense in $T^* G/P$.
606:
607: Let $\u({\mathfrak r})$ and $\z({\mathfrak r})$ denote the
608: nilpotent part and the centraliser of the Lie algebra
609: ${\mathfrak r}$ of $R$.
610: Let $\plie$ be the Lie algebra of $P$. Under $t_P$,
611: $f$ is mapped to an element in $\g^*$ which is orthogonal to $\plie$
612: and therefore to
613: $\u({\mathfrak r}) \oplus \z({\mathfrak r})$. It thus defines an
614: element $\overline f$ in ${\mathfrak l}^*$, if ${\mathfrak l}$ denotes
615: the Lie algebra of $L$. If $y \in L/Q_L$ denotes the element
616: $\mu_L(\overline f)$, then, by definition of the Mukai flop,
617: $\overline f$ is orthogonal to
618: $\overline \q_y$
619: ($\overline \q_y$ denotes the Lie algebra of the stabilisor of $y$ in
620: $L$). Thus it follows that $f$ vanishes on $\q_{j(y)}$, the Lie
621: algebra of the stabiliser of $j(y)$ in $G/Q$. Therefore, $y$ equals
622: $\pi_Q \circ \mu(f)$.
623: \end{proo}
624:
625:
626: Now, \cite[theorem 6.1]{namikawa} states that for any pair $(P,Q)$
627: of parabolic
628: subgroups allowing duality, we can find a chain
629: $(P_0=P,P_1,\ldots,P_n=Q)$ of parabolic subgroups such
630: that all the pairs $(P_i,P_{i+1})$ are as above and the corresponding
631: pair $P_L,Q_L \subset L$ is a fundamental case. Therefore, the description of
632: stratified Mukai flops in the fundamental cases is enough to understand
633: all stratified Mukai flops. As we will see in section \ref{section_reduction},
634: the same is true as far as dual varieties are concerned.
635:
636:
637: \subsectionplus{Recallections about fundamental homogeneous spaces}
638:
639: \label{subsection_notation}
640:
641: We now introduce some notations and recall some results
642: for fundamental homogeneous spaces
643: which will be used throughout the article.
644: In particular, we give in each case
645: a simple way of understanding the rational map
646: $q : T^* G/P \dasharrow G/Q$.
647:
648: Let $r$ and $n$ be integers with $2r<n$.
649: The Grassmannian parametrising $r$-linear subspace of a fixed
650: $n$-dimensional vector space $V$ will be denoted
651: $\G(r,V)$. The dual Grassmannian, parametrising
652: codimension $r$ subspaces of $V$, will be denoted $\G(r,V^*)$.
653: Let $x \in \G(r,V)$; it represents a linear subspace of $V$ which will
654: be denoted $L_x$. Moreover, we have a natural identification
655: $T^*_x \G(r,V) \simeq Hom(V/L_x,L_x)$. If
656: $\varphi \in Hom(V/L_x,L_x)$ is generic (that is, of rank $r$), then
657: its kernel is a codimension $r$ subspace of $V$ containing $L_x$. In
658: fact, we have $q(\varphi) = \ker \varphi$.
659:
660: \lpara
661:
662: Let $p$ be an integer. Let $V$ be a vector space of dimension $4p+2$,
663: equipped with a quadratic form. In case we need a basis for $V$, we
664: will take a hyperbolic one, of the form
665: $(e_1^+,\ldots,e_{2p+1}^+,e_1^-,\ldots,e_{2p+1}^-)$, such that the
666: quadratic form is given by
667: $Q(\sum x_i^+e_i^+ + \sum x_i^-e_i^-) = \sum x_i^+x_i^-$.
668: Recall that the variety parametrising isotropic subspaces of $V$ of
669: dimension $2p+1$ has two connected components, which will be denoted
670: $G/P = \G_Q^+(2p+1,4p+2)$ and $G/Q = \G_Q^-(2p+1,4p+2)$. As in the
671: case of Grassmannians, for $x \in \G_Q^+(2p+1,4p+2)$ and
672: $h \in \G_Q^-(2p+1,4p+2)$, we denote $L_x,L_h$ the corresponding
673: isotropic subspaces. The relation $x \in I_h$ amounts to
674: $\dim(L_x \cap L_h) = 2p$. Given $x \in \G_Q^+(2p+1,4p+2)$ and
675: $L \subset L_x$ of dimension $2p$,
676: there is exactly one $h \in \G_Q^-(2p+1,4p+2)$ such that
677: $L_x \cap L_h = L$~: this yields a natural isomorphism between
678: $I_x$ and $\p L_x^*$.
679:
680: The map $q$ may be defined as follows. Let $x \in \G_Q^+(2p+1,4p+2)$;
681: the cotangent space $T_x^* \G_Q^+(2p+1,4p+2)$ identifies with
682: $\wedge^2 L_x$. If $\omega \in \wedge^2 L_x$ is a skew form of rank
683: $2p$, let $L_\omega$ be its image. It is a hyperplane in $L_x$;
684: therefore it defines a unique element $h \in \G_Q^-(2p+1,4p+2)$
685: such that $L_x \cap L_h = L_\omega$. We have
686: $q(\omega) = h$.
687:
688: \lpara
689:
690: As far as the exceptional group $E_6$ is concerned, we denote $V_i$
691: the $i$-th fundamental representation of $E_6$, so that
692: $E_6/P_i \subset \p V_i$. We have $V_6 = V_1^*$ and $V_5 = V_3^*$. In
693: terms of this embedding, an element $h \in \p V_1^*$ belongs to
694: $E_6/P_6$ \iff it contains the linear span of two tangent spaces
695: $T_xE_6/P_1,T_yE_6/P_1$, for some distinct $x,y \in E_6/P_1$.
696:
697: We refer to \cite{flop_scorza} for the
698: proofs of the following results. Let $x \in E_6/P_1$. The cotangent
699: space $T_x^* E_6/P_1$ identifies with $\oc \oplus \oc$, if $\oc$
700: denotes the algebra of complexified octonions, an 8-dimensional
701: non-associative and non-commutative algebra over $\C$. This algebra is a
702: normed algebra~: there is a quadratic form $N:\oc \rightarrow \C$ such
703: that $N(z_1z_2) = N(z_1)N(z_2)$ for all $z_1,z_2 \in \oc$. The variety
704: $I_x$ is an 8-dimensional smooth quadric. It is convenient to denote
705: $Z = \C \oplus \oc \oplus \C$ a 10-dimensional space, equipped with
706: the quadratic form $Q(t,z,u) = tu - N(z)$. Then $I_x$ is the smooth
707: quadric defined by $Q$ and $q$ is defined by
708: $q((z_1,z_2)) = [N(z_1):z_1\overline z_2:N(z_2)] \in \p Z$
709: \cite[theorem 3.3 and corollary 3.2]{flop_scorza}.
710:
711: \lpara
712:
713: To visualise the homogneous space $E_6/P_3$ (resp. $E_6/P_5$),
714: we use the fact
715: that its points parametrise
716: projective lines included in $E_6/P_1$ (resp. $E_6/P_6$)
717: \cite[theorem 4.3 p.82]{landsberg}.
718: To avoid confusions between points in $E_6/P_1$ and
719: $E_6/P_3$, we will denote the latters with greek letters.
720: Since the marked Dynkin diagrams of $E_6/P_3$ and $E_6/P_5$ are
721: respectively
722: $\dynkinep{9}{4}{2}$ and $\dynkinep{9}{4}{6}$, we see that for
723: $\kappa \in E_6/P_5$, $I_\kappa \simeq \G(2,5)$. Let us describe this
724: isomorphism $I_\kappa \simeq \G(2,5)$ more explicitly, since this will be
725: needed to describe the rational map $q$.
726: If $\alpha \in E_6/P_3$, we will denote $l_\alpha \subset E_6/P_1$ the
727: corresponding
728: line and $L_\alpha$ the linear subspace it represents.
729: By \cite[proposition 3.6]{flop_scorza}, the span of
730: the affine tangent spaces
731: $\widehat{T_x E_6/P_1}$ in $V_1$ for $x$ in
732: $l_\alpha$ is a 22-dimensional linear subspace in $V_1$ denoted
733: $S_\alpha$.
734: Therefore, any 25-dimensional space which contains $S_\alpha$ defines a
735: pencil of hyperplanes which belong to $E_6/P_6 \subset \p V_1^*$, that
736: is, a point in $E_6/P_5$. Denoting $Q_\alpha = V_1/S_\alpha$,
737: this shows that
738: $I_\alpha = \G(3,Q_\alpha) \simeq \G(2,5)$. Dually, for
739: $\beta \in E_6/P_5$, $I_\beta \simeq \G(2,W_\beta)$,
740: where $W_\beta \subset V_1$ is a
741: 5-dimensional linear subspace such that
742: $\p W_\beta \subset E_6/P_1$.
743:
744: A Levi factor of $P$
745: contains $L' \simeq SL_2 \times SL_5$, and $L_\alpha$ (resp. $Q_\alpha$)
746: is the natural representation of $SL_2$ (resp. $SL_5$).
747: These representations are usefull describing $T^* E_6/P_3$~:
748: let $[e] \in E_6/P_3$ denote the base point; according to
749: \cite[propositions 3.6 and 3.7]{flop_scorza},
750: $T^*_{[e]} E_6/P_3$ is no longer an irreducible
751: $L'$-module, but there are exact sequences of $L'$-representations
752: \begin{equation}
753: \label{suite_exacte}
754: 0 \rightarrow L_\alpha^* \otimes \wedge^2 Q_\alpha \rightarrow
755: T_{[e]} E_6/P_3 \rightarrow Q_\alpha^* \rightarrow 0
756: \end{equation}
757: $$
758: \label{suite_exacte_cotangent}
759: 0 \rightarrow Q_\alpha \rightarrow
760: T^*_{[e]} E_6/P_3 \stackrel{\pi}{\rightarrow}
761: L_\alpha \otimes \wedge^2 Q_\alpha^*
762: \rightarrow 0.
763: $$
764:
765: We now describe the rational map $q$.
766: Choose a base
767: $e_1^*,e_2^*$ (resp. $f_1,\ldots,f_5$) of $L_\alpha^*$ (resp.
768: $Q_\alpha$). The
769: rational map $q:T_{[e]}^* E_6/P_3 \dasharrow I_{[e]}$ factors through
770: $L_\alpha \otimes \wedge^2 Q_\alpha^*$, and the induced rational map
771: $\overline q:L_\alpha \otimes \wedge^2 Q_\alpha^*
772: \dasharrow I_{[e]} = \G(2,Q_\alpha^*)$ is described as
773: follows : let $\varphi \in L_\alpha \otimes \wedge^2 Q_\alpha^* \simeq
774: Hom(L_\alpha \otimes \wedge^2 Q_\alpha^*)$ be generic. Its image in
775: $\G(2,W_5^*)$ under $\overline q$
776: represents the linear subspace generated by
777: \begin{itemize}
778: \item
779: the
780: orthogonal for the alternate form $\varphi(e_2^*)$ of the kernel of
781: $\varphi(e_1^*)$,\hspace{.4cm} and
782: \item the orthogonal for $\varphi(e_1^*)$ of the kernel
783: of $\varphi(e_2^*)$.
784: \end{itemize}
785: This is well-defined \iff $\varphi(e_1^*)$ and
786: $\varphi(e_2^*)$ have maximal rank 4 and the corresponding orthogonals
787: are different lines in $V_2^*$. This is proved in
788: \cite[theorem 4.3]{flop_scorza}.
789:
790:
791:
792:
793: \subsectionplus{Dual schemes}
794:
795: For some purposes (for example \cite{hermitian}),
796: it may be usefull to extend the above definition of
797: dual varieties to more general schemes. The goal of this subsection is
798: to explain how this is possible.
799:
800: Let us first define the cotangent scheme of a subscheme. So let $S$ be
801: an arbitrary scheme and $f:X \rightarrow Y$ a morphism above $S$.
802: The cotangent scheme $T^*X$ of $X$ is
803: ${\bf Spec}\ S\point {\cal H}om(\Omega_{X/S},\co_X)$; it is a scheme
804: over $X$, equipped with a natural section, the zero section.
805: Now $f$ induces a natural morphism of sheaves
806: $f^*\Omega_{Y/S} \rightarrow \Omega_{X/S}$, and so a morphism
807: $f^* T^*Y \rightarrow T^*X$. We finally define the cotangent
808: scheme $N^*_{X,Y}$ as the fiber above the zero section of this map.
809:
810: Let $G$ be a semi-simple Chevalley group scheme over $\Z$, $P$ and $Q$
811: parabolic subgroups.
812:
813: \begin{defi}
814: $P$ and $Q$ allow duality if the complex groups $P(\C),Q(\C)$ do.
815: \end{defi}
816:
817: If $P$ and $Q$ allow duality, although the moment map
818: $T^*G/P \rightarrow \g^*$
819: may fail to be birational in positive caracteristic, there is still a
820: well-defined birational map $T^*G/P \dasharrow T^*G/Q$, defined over
821: $\Z$~:
822:
823: \begin{prop}
824: There is a $G$-equivariant birational map
825: $\mu : T^*G/P \dasharrow T^*G/Q$ defined over $\Z$.
826: \end{prop}
827: \pr
828: By \cite[theorem 6.1]{namikawa} and proposition \ref{prop_fibration},
829: any pair of parabolic subgroups allowing duality is
830: related by a chain of pairs $(P,Q)$ of parabolic subgroups for which the
831: birational map $T^* G/P \dasharrow T^* G/Q$ is locally isomorphic with
832: a family of birational maps given by a
833: fundamental stratified Mukai flop. It is therefore
834: enough to check the proposition
835: for fundamental cases. In these cases it is a consequence of the explicit
836: description of this flop recalled in \ref{subsection_notation}.
837: \qed
838:
839: If $S$ is a scheme and $G,P,Q$ are as above, let $G_S,P_S,Q_S$ the
840: groups obtained by base change $S \rightarrow Spec\ \Z$.
841:
842: \begin{defi}
843: Let $S$ be a reduced irreducible scheme, and
844: let $f: X \rightarrow G_S/P_S$ be an irreducible
845: closed $S$-subscheme. We say that
846: $X$ is suitable if $\mu$ is defined at the generic point
847: of $N^*_{X,G_S/P_S}$.
848: In this case, the dual scheme of $X$ is the
849: scheme-theoretic image of $N^*_{X,G_S/P_S}$ under $\pi_Q \circ \mu$.
850: \end{defi}
851:
852:
853:
854: \subsectionplus{Functorial property of dual varieties}
855:
856: \label{subsection_fonctoriel}
857:
858: We come back to our setting of complex geometry. In the usual setting,
859: if $X_1,X_2 \subset \p V$ are subvarieties, with $X_1 \subset X_2$,
860: there is in general no relation of inclusion between the dual
861: varieties of $X_1$ and $X_2$. Thus dual varieties have bad functorial
862: properties. The only thing one can say is the following obvious result.
863:
864: \begin{prop}
865: Let $P,Q \subset G$ allow duality.
866: Let $X \subset G/P$ be suitable, $g \in G$, and $Y=g(X)$. Then $Y$ is
867: suitable and $g(X^Q)=Y^Q$.
868: \label{foncteur}
869: \end{prop}
870: \pr
871: Let $x \in X^s$ and $f \in N_x^*X \subset \g^*$ an element in the open
872: $G$-orbit. Then
873: $\tr g^{-1}.f \in N_{g(x)}^*Y$ is also in the open $G$-orbit. Therefore,
874: $Y$ is suitable. Moreover, since $q$ is equivariant,
875: $q(\tr g^{-1}.f)=g.q(f)$.
876: Therefore, $g(X^Q) \subset Y^Q$. By symmetry, we have
877: also $g^{-1}(Y^Q) \subset X^Q$, so $g(X^Q) = Y^Q$.
878: \qed
879:
880:
881:
882: \subsectionplus{Dual varieties in type A}
883:
884: \label{subsection_grassmannienne}
885:
886:
887: In this section, I give a description of the dual variety of a subvariety
888: $X \subset \G(r,V)$
889: using an analog of the quotient map $V \dasharrow \p V$ for
890: Grassmannians.
891:
892: \lpara
893:
894: If $A$ and $B$ are vector spaces, $Inj(A,B)$ will
895: denote the sets of linear (resp. linear and injective) maps from $A$ to $B$.
896: Let $\varpi : Hom(\C^r,V) \dasharrow \p Hom(\C^r,V)$ denote the
897: natural rational map, and let
898: $\pi : \p Hom(\C^r,V) \dasharrow \G(r,V)$ map $\varphi$ of rank $r$ on
899: its image.
900: Dually, consider $\varpi' : Hom(V,\C^r) \dasharrow \p Hom(V,\C^r)$
901: and $\pi' : \p Hom(V,\C^r) \dasharrow \G(r,V^*)$ mapping $\varphi'$
902: of rank $r$ on its kernel.
903: If $X \subset \G(r,V)$ is a subvariety, let $\overline X^o$ denote
904: the set $\pi^{-1}(X)$ and $\overline X$ its closure in $\p Hom(\C^r,V)$.
905:
906: \begin{prop}
907: Let $X \subset \G(r,V)$ be a suitable variety. Then
908: $X^Q = \pi' [{(\overline X)}^*]$, where ${(\overline X)}^*$ is the usual dual
909: variety of the subvariety $\overline X \subset \p Hom(\C^r,V)$
910: of a projective space.
911: \label{type_A}
912: \end{prop}
913: \pr
914: We fix a smooth point $x \in X$ and $\overline f \in \p Hom(\C^r,V)$
915: such that $\pi(\overline f)=x$,
916: and start with two easy lemmas.
917: \begin{lemm}
918: $\overline X$ is smooth at $\overline f$.
919: \end{lemm}
920: \begin{proo}
921: In a neighbourhood of $\overline f$ we have $\overline X = \overline X^o$.
922: Moreover,
923: the map $\pi:\overline X^o \rightarrow X$ is locally a trivial fibration
924: with fiber at $x$ the smooth variety $Inj(\C^r,L_x)$.
925: \end{proo}
926:
927: \noindent
928: We denote $f \in Hom(\C^r,V)$ such that
929: $\varpi(f) = \overline f$.
930: \begin{lemm}
931: The affine tangent space $\widehat{T_{\overline f}\overline X}$
932: is the linear space of
933: maps $g:\C^r \rightarrow V$ such that the composition
934: $L_x \stackrel{f^{-1}}{\rightarrow} \C^r
935: \stackrel{g}{\rightarrow}V \rightarrow V/L_x$
936: belongs to $T_xX$.
937: \label{lemme_TxX}
938: \end{lemm}
939: \noindent
940: Recall that for $Z \subset \p W$ a projective variety and $z \in Z$,
941: the affine tangent space $\widehat{T_zZ} \subset W$ is the tangent
942: space of the affine cone over $Z$ at a lift of $z$ in $W$.
943: \begin{proo}
944: Let $\widehat X \subset Hom(\C^r,V)$ be the affine cone over
945: $\overline X$.
946: Since $\widehat X$ is smooth at $f$, any tangent
947: vector is the direction of a curve included in $\widehat X$.
948: Let $\gamma : (C,0) \rightarrow (\widehat X,f)$ be a curve in
949: $\widehat X$ and let $g=\gamma'(0) \in Hom(\C^r,V)$.
950: Under the well-known identification of
951: $T_x \G(r,V)$ with $Hom(L_x,V/L_x)$, the composition of the lemma
952: equals
953: $(\pi \circ \varpi \circ \gamma)'(0)$. Therefore it belongs to
954: $T_xX$. By dimension count, the lemma follows.
955: \end{proo}
956:
957: \noindent
958: \underline{\bf Proof of proposition \ref{type_A} :}
959: The linear subspace $(T_f\widehat X)^\bot \subset Hom(V,\C^r)$ is
960: the set of $g$'s such that for all $h \in T_f\widehat X$, the
961: composition
962: $\C^r \stackrel{h}{\rightarrow} V \stackrel{g}{\rightarrow} \C^r$ is
963: traceless. Since $T_f \widehat X$ contains $Hom(\C^r,L_x)$, this means
964: that $g$ is induced by a morphism $\overline g: V/L_x \rightarrow \C^r$ such
965: that $f \circ \overline g$ is orthogonal to
966: $T_xX \subset Hom(L_x,V/L_x)$, by lemma \ref{lemme_TxX}. Therefore,
967: for $h \in \G(r,V^*)$, we have
968: $h \in \pi' \circ \varpi((T_f \widehat X)^\bot)$
969: \iff
970: $h \in q(N_x^*X)$.
971: \qed
972:
973:
974:
975:
976:
977:
978: %******************************************************************************
979:
980: %******************************************************************************
981:
982: %******************************************************************************
983:
984:
985:
986:
987:
988:
989: \sectionplus{Reduction to fundamental examples}
990:
991: \label{section_reduction}
992:
993: From section \ref{section_definition}, we see that there are a lot of pairs of
994: parabolic subgroups which allow duality. In this section, I will show
995: that to understand all the dual varieties, it is enough to
996: understand dual varieties for fundamental cases.
997:
998: For example, the
999: varieties corresponding to the marked diagrams
1000: $$
1001: \dynkineDeuxMarque{11.5}{4}{0}{2}
1002: \hspace{2cm}
1003: \dynkineDeuxMarque{11.5}{4}{6}{8}
1004: $$
1005: both have dimension 26. Using tables in \cite[p.202]{mcgovern}, we see
1006: that there is a unique nilpotent
1007: orbit of dimension 52 in ${\mathfrak e}_6$ and that the disconnected
1008: centralizer of an element of this orbit is trivial. Therefore, the two
1009: corresponding parabolic subgroups $P,Q \subset E_6$
1010: allow duality. It may seem at first that the corresponding
1011: duality
1012: $X \subset G/P \mapsto X^Q \subset G/Q$ has to do with the
1013: exceptional geometry of $E_6$. However, we will see that it is not
1014: the case; indeed, $X^Q$ can be described using dual varieties in four
1015: classical homogeneous spaces. Indeed, \cite[theorem 6.1]{namikawa}
1016: is verified
1017: in this case thanks to the sequence of parabolic subgroups
1018: $$
1019: \dynkineDeuxMarque{8.5}{4}{0}{2} \rightarrow
1020: \dynkineUneMarqueP{8.5}{4}{0} \rightarrow
1021: \dynkineUneMarqueP{8.5}{4}{8} \rightarrow
1022: \dynkineDeuxMarque{8.5}{4}{6}{8} \hspace{.5cm},
1023: $$
1024: and we will see in
1025: this section how to compute accordingly dual varieties.
1026: We will show that the computation of the dual variety
1027: $X^Q$ for $X \subset G/P$ can be
1028: done in three steps, the first and the last in a family of spinor varieties
1029: $\G_Q^+(5,10)$, and the second in a family $\p^5$'s.
1030:
1031:
1032: \subsectionplus{Biduality theorem}
1033:
1034:
1035: Let $G$ be as above, $P,Q,R \subset G$ be subgroups
1036: such that $P$ and $Q$ allow
1037: duality, and $Q$ and $R$ allow duality.
1038: Then, by definition $P$ and $R$ also allow duality.
1039:
1040: \begin{theo}[Biduality theorem]
1041: Let $X \subset G/P$ be an suitable variety. Then $X^Q$ is suitable
1042: and $\mu(N^*X) = N^*X^Q$. In particular,
1043: ${(X^Q)}^R = X^R$.
1044: \label{theo_bidualite}
1045: \end{theo}
1046: \noindent
1047: If $G=SL_n$, $P=R$ is the stabilisor of a line and $Q$ is the
1048: stabilisor of a hyperplane, we recover the
1049: usual biduality theorem.
1050: \begin{proo}
1051: We follow the argument of \cite[pp.27 to 30]{gkz}.\\
1052: Let $N=\mu(N^*X) \subset T^* G/Q$.
1053: Recall that $T^* G/Q$ is a symplectic variety. Moreover, it is proved
1054: in \cite{gkz} that $N^*X \subset T^* G/P$ is a lagrangien
1055: subvariety of $T^* G/P$. Let $\co \subset \g^*$ denote the
1056: nilpotent orbit which
1057: closure is the image of $T^* G/P$. Since the birational morphisms
1058: $T^* G/P \stackrel{\sim}{\dasharrow} {\cal O}
1059: \stackrel{\sim}{\dasharrow} T^* G/Q$ are symplectic, it follows that
1060: $N$ is also lagrangien.
1061:
1062: Moreover, it has the property that if
1063: $(x,f) \in N$ and $\lambda \in \C$, then $(x,\lambda f) \in N$.
1064: This follows from the fact that the image of $N^*X$ in
1065: $\overline {\cal O}$ is stable under multiplication by scalars.
1066: From \cite[proposition 3.1]{gkz}, we know that
1067: $N=N^*Z$ for $Z=\pi_Q(N)=X^Q$. Therefore, $\mu(N^*X)=N^*X^Q$
1068: and $X^Q$ is suitable.
1069:
1070: Since ${(X^Q)}^R$ (resp. $X^R$) is the image of $N^*X_Q$
1071: (resp. $\mu(N^*X)$) under the
1072: rational map $T^* G/Q \dasharrow G/R$, these varieties are equal.
1073: \end{proo}
1074:
1075: The following corollary shows that the name of biduality theorem for
1076: the above result is justified~:
1077:
1078: \begin{coro}
1079: \label{coro_bidualite}
1080: Let $P,Q \subset G$ allow duality. If $X \subset G/P$ is suitable,
1081: then $X^Q$ is suitable and ${(X^Q)}^P = X$. Moreover,
1082: if $x \in X$ and $h \in X^Q$, then $h$ is
1083: tangent to $X$ at $x$ \iff $x$ is tangent to $X^Q$ at $h$.
1084: \end{coro}
1085: \begin{proo}
1086: To prove that ${(X^Q)}^P = X$,
1087: it is enough to take $R=P$ in theorem \ref{theo_bidualite}, after
1088: observing that for suitable $X \subset G/P$, $X^P = X$.
1089: The second result, that $h$ is tangent to $X$
1090: at $x$ \iff $x$ is tangent to $X^Q$ at $h$ follows from the fact the
1091: first (resp. the second) affirmation means that $(x,h)$ lies in the
1092: image by $(p,\pi_Q)$ of an element in $\mu(N^*X)$ (resp. $N^*X^Q$).
1093: \end{proo}
1094:
1095:
1096:
1097: \subsectionplus{Families of dual varieties}
1098:
1099: \label{subsection_sequence}
1100:
1101:
1102: We consider the following situation :
1103: let $R \subset G$ be a parabolic subgroup.
1104: Let $P,Q \subset R \subset G$ be parabolic subgroups and recall
1105: notations of subsection \ref{subsection_fondamental}.
1106: If $X \subset G/P$ and $z \in G/R$, denote $X_z:=X \cap \cf_z$.
1107: Assume $P_L,Q_L \subset L$ allow duality.
1108: For $z \in G/R$ and suitable $Y \subset \cf_z \simeq L/P_L$, let
1109: $Y^{Q_L} \subset \cg_z \simeq L/Q_L$ denote its generalised dual
1110: variety.
1111:
1112: \begin{theo}
1113: With the previous notations, assume that $P,Q \subset G$ allow
1114: duality, and also $P_L,Q_L \subset L$.
1115: If $X \subset G/P$ is suitable, then
1116: for generic $x \in X$, $X_{f_P(x)} \subset \cf_{f_P(x)}$ is
1117: suitable. Moreover, $X^Q$ is the closure of the union of
1118: the $X_{f_P(x)}^{Q_L}$, for such $x$ in $X$.
1119: \label{sous-groupe}
1120: \end{theo}
1121: \pr
1122: Let $f \in N^*_xX$ an element which $G$-orbit in $T^* G/P$ is dense
1123: and set $z = f_P(x)$.
1124: We have seen in the proof of proposition \ref{prop_fibration}
1125: that the restriction $f_x$ of $f$ to
1126: $T_x \cf_z$ is a generic element in $T^* \cf_z \simeq T^* L/P_L$.
1127: Moreover, this restriction belongs to $N^*_x X_z$, so that $X_z$ is
1128: suitable.
1129:
1130: Let $q_z : T^* \cf_z \dasharrow \cg_z$ be the composition of
1131: $\mu_z : T^* \cf_z \dasharrow T^* \cg_z$ and the projection
1132: $T^* \cg_z \rightarrow \cg_z$. Proposition \ref{prop_fibration}
1133: states that $q(f) = j_z \circ q_z(f_x) \in G/Q$. Therefore it follows
1134: that $q(N^*X_{|X_z}) = j_z(X_z^{Q_L})$.
1135: The description of $X^Q$ in the theorem follows.
1136: \qed
1137:
1138: \lpara
1139:
1140: As a consequence of theorems \ref{theo_bidualite} and \ref{sous-groupe},
1141: if $P=P_1 \times P_2$ and $Q=Q_1 \times Q_2$ are parabolic subgroups
1142: of $G=G_1 \times G_2$,
1143: and if $X=X_1 \times X_2$, then we have
1144: $X^Q = X_1^{Q_1} \times X_2^{Q_2}$.
1145:
1146:
1147:
1148:
1149:
1150: %******************************************************************************
1151:
1152: %******************************************************************************
1153:
1154: %******************************************************************************
1155:
1156:
1157:
1158: \sectionplus{Tangency for fundamental examples}
1159:
1160: \label{section_tangency}
1161:
1162: In this section, if $x \in X \subset G/P$,
1163: I introduce a definition of the embedded tangent cone at $x$,
1164: $\overline{T_xX}$, which is a subvariety of $G/P$ and a cone
1165: at $x$ (in a suitable
1166: sense). I also introduce the cotangent
1167: variety at $x$, $\overline{N_xX}$, which is
1168: a subvariety of $G/Q$. Moreover a notion of ``linear varieties'' is defined
1169: and linear varieties are classified.
1170:
1171: From now on, $P,Q \subset G$ are fundamental subgroups of $G$ allowing
1172: duality.
1173:
1174: \subsectionplus{A tangent element is incident}
1175:
1176: In this subsection, we prove that if $x \in X \subset G/P$ and
1177: $h \in G/Q$ is tangent to $X$ at $x$
1178: (see definition \ref{i_h}), then $h$ is incident to $x$
1179: (in the sense that the stabilisors of $x$ and $h$ contain a common
1180: Borel subgroup). This only holds in fundamental cases.
1181:
1182: \begin{nota}
1183: Let $x \in \g$ nilpotent. Then there exists $y,h \in \g$ such that $(x,y,h)$
1184: is a $\mathfrak {sl}_2$-triple. For $i \in \Z$, let $\g_i$ denote
1185: $\{X \in \g : [h,X]=iX\}$. The parabolic subalgebra
1186: $\plie_x := \oplus _{i \geq 0} \g_i$ does not depend on $y$ and $h$
1187: \cite[theorem 3.8]{mcgovern}, and is
1188: called the canonical parabolic subalgebra of $x$.
1189: \end{nota}
1190:
1191: In the following lemma, I say that $\plie \subset \g$ is a maximal
1192: parabolic subalgebra of $\g$ of fundamental type if the pair
1193: $(\g,\plie)$ is the pair of Lie algebras of groups $(G,P)$ as in
1194: definition \ref{fondamental}.
1195:
1196: \begin{lemm}
1197: Let $x \in \g$ and $\plie$ be a polarisation of $x$. Assume that
1198: $\plie$ is a maximal parabolic subalgebra of fundamental type.
1199: Then $\plie_x \subset \plie$.
1200: \end{lemm}
1201: \begin{proo}
1202: Let $\plie$ be a maximal parabolic subalgebra which is a polarisation of $x$.
1203: Let $(x,y,h)$ be a $\mathfrak {sl}_2$-triplet, $\h$ a Cartan
1204: subalgebra containing $h$ and $\Delta=\{\alpha_1,\ldots,\alpha_r\}$
1205: a basis of the root system such
1206: that $\forall \alpha \in \Delta,\alpha(h) \geq 0$.
1207:
1208: We denote $\plie_1$ the following maximal parabolic subalgebra~:
1209: $$
1210: \plie_1 := \h \oplus \bigoplus_{
1211: \begin{array}{c}
1212: \alpha = \sum_j k_j \alpha_j\\
1213: k_i \geq 0
1214: \end{array}
1215: } \g_{\alpha}\ \ ,
1216: $$
1217: where $i$ is chosen such that $\plie$ is conjugated to $\plie_1$
1218: (such $i$ exists because $\plie$ is a maximal parabolic subalgebra).
1219:
1220: Let us prove that $x \in \u(\plie_1)$.
1221: According to the
1222: decomposition $\g = \h \oplus \bigoplus_{\alpha} \g_{\alpha}$, we can
1223: write $x = h_x + \sum_\alpha x_\alpha$, with $h_x \in \h$ and
1224: $x_\alpha \in \g_\alpha$. Now, since $[h,x]=2x$, we deduce that
1225: $h_x = 0$ and that for any root $\alpha$, either $x_\alpha = 0$ or
1226: $\alpha(h) = 2$.
1227:
1228: \begin{clai}
1229: If $\alpha = \sum k_j \alpha_j$ is a root, then
1230: $\alpha(h)=2 \Longrightarrow k_i > 0$.
1231: \end{clai}
1232: \begin{proo}
1233: This is proved by ad hoc arguments in all cases. Assume first that
1234: $\g = \liesl_n$ and that $\plie$ is the stabilisor of an $r$-dimensional
1235: subspace. Thus $i=r$.
1236: Recall that the weighted diagram of $x$ is by definition the list
1237: of the values $\alpha_j(h)$. The weighted diagrams of nilpotent
1238: elements in $\liesl_n$ are well-known; in our case, since
1239: $x$ is a generic element of $\u(\plie)$ with $\plie$ conjugated to
1240: $\plie_1$, we have
1241: $\alpha_i(h)=\alpha_{n-i}(h)=1$ and the other values $\alpha_j(h)$ equal 0.
1242: The equality $\alpha(h)=2$ with $\alpha = \sum k_j \alpha_j$ amounts
1243: to $k_i + k_{n-i} = 2$, which implies $k_i = k_{n-i} = 1$.
1244:
1245: Assume now that $\g = \spin_{4p+2}$. In this case, there is only one
1246: possibility for the $G$-orbit in $\spin_{4p+2}$
1247: of $x$, and $\alpha_j(h)=1$ \iff $\alpha_j$
1248: is a spin root (ie $j \in \{2p,2p+1\}$); otherwise $\alpha_j(h)=0$. Therefore
1249: $\alpha(h) = 2$ implies that $\alpha$ is not less than the root
1250: $\alpha_{2p-1} + \alpha_{2p} + \alpha_{2p+1}$, which implies the claim.
1251:
1252: If $\g$ is of type $\e_6$ and $\plie$ corresponds to the first root, then
1253: the weighted diagram of $x$ is $\poidsesix 100010$
1254: (see \cite[table p.202]{mcgovern}). Since for all roots
1255: $\sum k_j \alpha_j$ we have $-1\leq k_1,k_6\leq 1$, we again have
1256: $\sum k_j \alpha_j(h) = 2 \Rightarrow k_1=k_6=1$. In case $\plie$
1257: corresponds to the second root, the weighted diagram is
1258: $\poidsesix 010100$. The equality $\alpha(h)=2$ for
1259: $\alpha = \sum k_j \alpha_j$ implies that $k_3 + k_5 = 2$. If
1260: $k_3 = 2$, then necessarily $k_5 \geq 1$ (see the list of roots in
1261: \cite{bourbaki}), so we get a contradiction. Similarly $k_5 \leq 1$.
1262: So $k_3 = k_5 = 1$, and again the claim is proved.
1263: \end{proo}
1264:
1265: This claim therefore proves that if $x_\alpha \not = 0$, with
1266: $\alpha = \sum k_j \alpha_j$, then $k_i > 0$.
1267: This proves that $x$ belongs
1268: to
1269: $$
1270: \bigoplus_{
1271: \begin{array}{c}
1272: \alpha = \sum_j k_j \alpha_j\\
1273: k_i \geq 1
1274: \end{array}
1275: } \g_{\alpha}\ ,
1276: $$
1277: which is readily seen to be $\plie_1^\bot = \u(\plie_1)$.
1278: Thus $x \in \u(\plie_1)$
1279: and $\plie_1$ is a polarisation of $x$.
1280: Now, since the map $T^* G/P \rightarrow \g$ is birational on its
1281: image, there is a unique polarisation of $x$ in the conjugacy class of
1282: $\plie$. Therefore $\plie = \plie_1$.
1283:
1284:
1285: \lpara
1286:
1287: Let us now show that $\plie_x \subset \plie_1$. Since $\plie_x \supset \h$,
1288: it is the sum of $\h$ and some root spaces. Now, assume
1289: $\g_\alpha \subset \plie_x$, with $\alpha = \sum k_j \alpha_j$.
1290: This means that $\sum k_j \alpha_j(h) \geq 0$.
1291: I claim that $k_i \geq 0$. In fact, if $k_i < 0$, then $\alpha$ is a negative
1292: root, so $k_j \leq 0$ for all $j$. We therefore have
1293: $\sum k_j \alpha_j(h) \leq k_i \alpha_i(h)$.
1294: In the proof of the above claim, we have seen that we
1295: allways have $\alpha_i(h)=1$. So we get a contradiction.
1296:
1297: Therefore, we have $k_i \geq 0$, and so $\g_\alpha \subset \plie_1$.
1298: Since $\plie_1 = \plie$, we have proved that $\plie_x \subset \plie$, as
1299: claimed.
1300: \end{proo}
1301:
1302:
1303: \begin{coro}
1304: \label{coro_incident}
1305: If $\plie$ and $\q$ are polarisations of the same nilpotent element
1306: $x$, and are maximal parabolic subalgebras of
1307: fundamental type, then they contain a
1308: common Borel subalgebra.
1309: \end{coro}
1310: \pr
1311: They both contain the canonical parabolic subalgebra $\plie_x$.
1312: \qed
1313:
1314: \lpara
1315:
1316: We now show, with an example, that the above corollary is wrong if one
1317: considers non maximal parabolic subalgebras.
1318:
1319: \begin{exem}
1320: Let $\g = {\mathfrak sl}_n$. Let $x \in \g$ be an element of rank 2,
1321: such that $x^3 = 0$ but $x^2 \not = 0$. Let $\plie$ (resp. $\q$) be the
1322: parabolic subalgebra preserving the image of $x^2$ and the image of
1323: $x$ (resp. the kernel of $x$ and the kernel of $x^2$). Then we have
1324: $x \in \u(\plie)$ and $x \in \u(\q)$. However, since
1325: $\im\ x \not \subset \ker x$, $\plie$ and $\q$ are
1326: not incident.
1327: \end{exem}
1328: \begin{proo}
1329: If $y \in \plie$ (resp. $y \in \q$), then the commutator
1330: $[x,y]$ is strictly upper triangular for the filtration
1331: $\im\ x^2 \subset \im\ x \subset \C^n$ (resp.
1332: $\ker x \subset \ker x^2 \subset \C^n$). Therefore, $[x,y]$ is
1333: traceless and so $x \in \plie^\bot$ (resp. $x \in \q^\bot$).
1334: \end{proo}
1335:
1336: \lpara
1337:
1338: The Schubert varieties $I_h$ (recall definition \ref{i_h})
1339: give a geometric understanding of the
1340: rational map $q : T^* G/P \dasharrow G/Q$ :
1341: \begin{lemm}
1342: \label{lemm_q}
1343: Assume $P$ and $Q$ are maximal parabolic subgroups.
1344: Let $x \in G/P$ and $h \in G/Q$, and
1345: let $f$ be a generic element in
1346: $T_x^* G/P$. Then $q(f) = h$ \iff $x \in I_h$ and the cotangent
1347: form $f$ vanishes on $T_x I_h$.
1348: \end{lemm}
1349: \noindent As a consequence of the lemma, there is a unique $h$ such that
1350: $x \in I_h$ and $f$ vanishes on $T_x I_h$. By definition, if $h$ is
1351: tangent to $X$ at $x$, then there exists $f \in N^*_xX$ such that $q$
1352: is defined at $f$ and $q(f)=h$. Thus the lemma implies that the
1353: intersection $I_h \cap X$ is not transverse at $x$, as was stated in
1354: fact \ref{fait_tangent}.
1355: \begin{proo}
1356: Let $x \in G/P$; $t_P$ restricts to an isomorphism between
1357: $T^*_x G/P$ and $(\g/\plie_x)^* \subset \g^*$ if $\plie_x$ denotes the Lie
1358: algebra of the stabilisor of $x$. Conversely, given $\eta \in {\cal O}$,
1359: $\pi_P(t_P^{-1}(\eta))$ is the unique $x \in G/P$ such that the corresponding
1360: parabolic subalgebra $\plie_x$ is orthogonal to $\eta$.
1361:
1362: Let $x \in G/P$, $f \in T_x^* G/P$ generic and
1363: $\eta = t_P(f) \in \plie_x^\bot$, and let $h=q(f)$.
1364: The previous argument shows that $h$ is the unique element in $G/Q$
1365: such that $\eta$ vanishes on $\q_h$. Moreover, we know by
1366: corollary \ref{coro_incident} that $x \in I_h$.
1367: Note that $T_xG/P = \g / \plie$ and
1368: $T_x I_h \simeq \q_h/(\plie_x \cap \q_h)$.
1369: Since $\eta$ vanishes on $\plie_x$, it will
1370: vanish on $\q_h$ \iff it vanishes on $\q_h/(\plie_x \cap \q_h)$, namely,
1371: \iff $f$ vanishes on $T_xI_h$.
1372: \end{proo}
1373:
1374: \begin{exem}
1375: Let $h \in G/Q$ and let $X = I_h \subset G/P$. Then $X$ is suitable
1376: and $X^Q = \{h\}$. Moreover $p(T^*_hG/Q) = I_h$.
1377: \end{exem}
1378: \begin{proo}
1379: First, let $x \in X$, let $f \in T^*_xX$ be generic and let $h=q(f)$.
1380: By corollary \ref{coro_incident}, $x$ and $h$ are incident, and by
1381: lemma \ref{lemm_q}, $f$ vanishes on $T_x I_h$. Thus, $I_h$ is
1382: suitable. Since $G/Q$ is homogeneous, $I_h$ is suitable for all
1383: $h \in G/Q$.
1384:
1385: Let $x \in X$ and $f \in N^*_xX$ generic. Then by the above $q(f)=:h'$
1386: is well-defined, and by lemma \ref{lemm_q} again, $h'$ is the unique
1387: element in $G/Q$ such that $x \in I_{h'}$ and such that $f$ vanishes on
1388: $T_x I_{h'}$. Since $h$ satisfies these conditions, $h'=h$. Therefore,
1389: $X^Q = \{h\}$.
1390:
1391: For the last point, we note that $p(T^*_h G/Q) = \{h\}^P = I_h$, by
1392: biduality theorem \ref{theo_bidualite},
1393: since we have proved that $I_h^Q = \{h\}$.
1394: \end{proo}
1395:
1396:
1397: \subsectionplus{Dual varieties and cones}
1398:
1399: If $X \subset \p V$ is included in a hyperplane represented by
1400: $h \in \p V^*$, then the dual variety of $X$, which is a subvariety of
1401: $\p V^*$, is a cone over $h$. The aim of this subsection is to prove
1402: an analogous result for our generalised dual varieties. Our first goal
1403: is to define cones.
1404:
1405: \begin{defi}Let $x_1,x_2 \in G/P$
1406: \begin{itemize}
1407: \item
1408: $x_1,x_2$ are linked if there exists $h \in G/Q$ such that
1409: $x_1,x_2 \in I_h$.
1410: \item
1411: If $E \subset G/P$, let
1412: $\displaystyle I_E := \bigcap_{x \in E} I_x \subset G/Q$.
1413: \item
1414: If $x_1,x_2$ are linked, denote
1415: $\displaystyle L(x_1,x_2) = \bigcap_{h \in I_{\{x_1,x_2\}}} I_h$.
1416: \end{itemize}
1417: \end{defi}
1418: \noindent
1419: In $\p V$, all points are linked, and $L(x_1,x_2)$ is the line through
1420: $x_1$ and $x_2$. The difference between $\p V$ and our general
1421: situation is that in general $G$ does not
1422: act transitively on pairs of distinct points $x,y \in G/P$, so that $L(x,y)$
1423: may depend, up to isomorphism, on $x$ and $y$.
1424: However, cones are defined in perfect analogy~:
1425:
1426: \begin{defi}
1427: \label{def_cone}
1428: Let $X \subset G/P$ and $x \in X$. Then $X$ is a cone over $x$ if
1429: for all $y \in X$, $x$ and $y$ are linked and $L(x,y) \subset X$.
1430: \end{defi}
1431: \noindent
1432: An equivalent definition is that for generic $y \in X$ the same
1433: condition holds, as will be clear from the following description of
1434: $L(x,y)$~:
1435:
1436: \begin{prop}
1437: \label{prop_l}
1438: Let $x \not = y \in G/P$. We have~:
1439: \begin{itemize}
1440: \item
1441: If $G/P = \G(r,V)$, then $(x,y)$ are linked \iff
1442: $\codim_V(L_x+L_y) \geq r$, in which case $L(x,y) = \G(r,L_x+L_y)$.
1443: \item
1444: If $G/P = \G_Q^+(2p+1,4p+2)$, then $(x,y)$ are linked \iff
1445: we have $\dim(L_x \cap L_y)=2p-1$, in which case
1446: $L(x,y) = \{ z : L_z \supset L_x \cap L_y \} \simeq \p^1$.
1447: \item
1448: If $G/P = E_6/P_1$, then $(x,y)$ are allways linked. In case a line
1449: passes through $x$ and $y$ in $E_6/P_1$, then $L(x,y)$ is this line;
1450: otherwise, there is a unique smooth 8-dimensional quadric through $x$
1451: and $y$, and $L(x,y)$ is this quadric.
1452: \item
1453: If $G/P = E_6/P_3$, then $(x,y)$ are linked \iff there is a $\G(2,5)$
1454: through them. If $\dim(L_x \cap L_y) = 1$ then
1455: $L(x,y)$ is equal to $\G(2,L_x + L_y) \simeq \p^2$,
1456: otherwise $L(x,y) \simeq \G(2,5)$.
1457: \end{itemize}
1458: \end{prop}
1459: \noindent
1460: In this proposition, for the two exceptional cases, I use the minimal
1461: projective homogeneous embedding $E_6/P_i \subset \p V_i$.
1462: For example, in the case of
1463: $E_6/P_3$, the condition that there is a $\G(2,5)$ through $x$ and $y$
1464: means that there is a linear 10-dimensional subspace $W \subset V_3$
1465: containing $x$ and $y$ and such that $\p W \cap E_6/P_3$ is projectively
1466: isomorphic with a Grassmanian $\G(2,5)$ in its Plücker
1467: embedding. Recall also that $E_6/P_3$ parametrises projective lines in
1468: $V_1$ which are included in $E_6/P_1$. For $x \in E_6/P_3$, the
1469: corresponding 2-dimensional subspace of $V_1$ has been denoted $L_x$.
1470: \begin{proo}
1471: The first case follows directly from the definition. In the second
1472: case, one only has to note that if there exists $h \in G/Q$ such that
1473: $(x,h),(y,h)$ are incident, then
1474: $\dim(L_x\cap L_h) = \dim(L_y \cap L_h) = 2p$, so
1475: $\dim(L_x\cap L_h\cap L_y) = 2p-1$.
1476:
1477: \lpara
1478:
1479: For the exceptional cases one oviously has to use the geometry of the
1480: involved homogeneous spaces. Let us first consider $E_6/P_1$. For all
1481: $x \in E_6/P_1$, $I_x$ is a smooth 8-dimensional quadric. Moreover, for any
1482: $x\not = y \in E_6/P_1$, the intersection of the two quadrics
1483: $I_x$ and $I_y$ is either a point or a $\p^4$. In
1484: fact, this was proved in
1485: \cite[propositions IV.3.2 and IV.3.3]{zak} in the context of Severi
1486: varieties, but also follows easily from the fact that there are three
1487: $E_6$-orbits in $E_6/P_1 \times E_6/P_1$ \cite[proposition 18]{quantique}.
1488: Given $x,y \in E_6/P_1$, we can have $x=y$, $x \not =
1489: y$ and there is a line through $x$ and $y$, or there is no line
1490: through $x$ and $y$. This describes the three orbits in
1491: $E_6/P_1 \times E_6/P_1$.
1492: In the degenerate case when a line passes through $x$ and $y$,
1493: $I_x \cap I_y$ is thus isomorphic with $\p^4$. Dually,
1494: the intersection of all the
1495: $I_h$ for $h$ in this $\p^4$ is a linear space (indeed, $x \in I_h$
1496: \iff $x \in E_6/P_1 \subset \p V_1$ is orthogonal to
1497: $\widehat{T_h E_6/P_6} \subset V_6=V_1^*$) and contains $x$ and $y$;
1498: a direct computation of dimension shows that it is exactly
1499: the line through $x$ and $y$.
1500: In the generic case, $I_x \cap I_y = \{h\}$; therefore
1501: $L(x,y) = I_h$ is the unique 8-dimensional quadric through $x$ and $y$.
1502:
1503: \lpara
1504:
1505: Let $\alpha,\beta \in E_6/P_3$ be linked, and
1506: denote $\kappa \in E_6/P_5$ an element such that
1507: $\alpha,\beta \in I_\kappa$. According to subsection
1508: \ref{subsection_fondamental}, $\alpha$ and
1509: $\beta$ represent 2-dimensional
1510: subspaces of a 5-dimensional subspace of $V_1$ denoted
1511: $W_\kappa$; we have denoted $L_\alpha,L_\beta$ these spaces.
1512:
1513: Assume first that $\dim(L_\alpha \cap L_\beta) = 1$.
1514: It is proved in
1515: \cite[proposition 3.6]{flop_scorza} that the linear
1516: span of all the affine tangent spaces at the points of the projective plane
1517: generated by $L_\alpha$ and $L_\beta$
1518: is 24-dimensional and equal to the span of affine tangent spaces at
1519: points in $l_\alpha \cup l_\beta$.
1520: Thus $(\alpha,\beta)$ defines a projective plane in $E_6/P_6$ and also in
1521: $E_6/P_5$. Moreover
1522: $I_{\alpha,\beta} = I_{\G(2,L_\alpha+L_\beta)} \simeq \p^2$
1523: and $L(\alpha,\beta) = \G(2,L_\alpha+L_\beta) \simeq \p^2$.
1524:
1525: Assume finally that $L_\alpha$ and $L_\beta$ don't meet.
1526: Let $L \subset L_\alpha \oplus L_\beta$ be any 3-dimensional subspace;
1527: the linear span $S_L$ of
1528: the affine tangent spaces at points of $\p L$ is again 24-dimensional,
1529: and any element in $I_{\alpha,\beta}$ must contain it.
1530: Assume that $I_{\alpha,\beta}$
1531: contains two points $\kappa,\lambda \in G/Q$. These points would correspond to
1532: codimension 2 subspaces $L_\kappa,L_\lambda$ of
1533: $V_1$ containing $S_L$; therefore
1534: $L_\kappa$ and $L_\lambda$ would be contained
1535: in a common hyperplane of $V_1$.
1536: Since $\alpha,\beta \in I_{\kappa,\lambda}$, by
1537: the case considered above, this would in turn imply that $L_\alpha$ and
1538: $L_\beta$ meet in dimension 1, which we have excluded.
1539: Therefore we have proved that $I_{\alpha,\beta} = \{ \kappa \}$, so
1540: $L(\alpha,\beta) = I_\kappa$ is isomorphic with $\G(2,5)$.
1541: \end{proo}
1542:
1543:
1544: \begin{theo}
1545: \label{theo_cone}
1546: Let $h \in G/Q$ and let $X \subset G/P$ such that $X \subset I_h$.
1547: Then $X$ is suitable and $X^Q$ is a cone over $h$.
1548: \end{theo}
1549: \begin{rema}
1550: In fact, as the proof will show, in all the cases but in type $A$,
1551: a stronger result holds~: for any $k \in X^Q$, there is a certain
1552: homogeneous subvariety
1553: $q(\C.f + N^*_x I_h) \subset G/Q$, of type given by lemmas
1554: \ref{lemm_cone_gr}, \ref{lemm_cone_spin}, \ref{lemm_cone_e6_1},
1555: and \ref{lemm_cone_e6_2}, containing (eventually stricly) $L(h,k)$,
1556: and included in $X^Q$. Although the idea of proof of this theorem
1557: is uniform, this proof unfortunately ends up with a case by case
1558: analysis.
1559: \end{rema}
1560: \begin{proo}
1561: If $x \in X$, then $N^*_xX$ contains $N^*_x I_h$ on which $q$ is
1562: well-defined generically, so $X$ is suitable.
1563: Assume $X \subset I_h$ and let $k$ be a generic element in $X^Q$.
1564: By definition of $X^Q$ there is an element $x \in X$ and
1565: $f \in N^*_xX$ such that $k = q(f)$. Since $x \in X \subset I_h$, we
1566: have $h \in I_x$. By corollary \ref{coro_incident}, $k \in I_x$;
1567: therefore $h$ and $k$ are linked. Moreover, we have
1568: $f \not \in N_x^* I_h$ (otherwise we would have $q(f)=h$).
1569: Therefore it follows from
1570: the inclusion $q(\C.f + N^*_x I_h) \subset X^Q$ and the following
1571: lemmas \ref{lemm_cone_gr}, \ref{lemm_cone_spin}, \ref{lemm_cone_e6_1}
1572: and \ref{lemm_cone_e6_2} that $L(h,k) \subset X^Q$.
1573: \end{proo}
1574:
1575: \begin{lemm}
1576: \label{lemm_cone_gr}
1577: Let $x \in \G(r,V),\ h\not = k \in \G(r,V^*)$ such that $h,k \in I_x$. Let
1578: $f \in N_x^* I_h$ such that $q$ is defined at $f$. Then
1579: $q(\C.f + N_x^* I_k) = L(h,k)$.
1580: \end{lemm}
1581: \begin{proo}
1582: Let $(e_i)$ be a base of $V$ and $(e_i^*)$ the dual base.
1583: Up to the action of $SL(V)$, we may assume that
1584: $L_x$ is the span of $e_1,\ldots,e_r$,
1585: $L_h$ is the span of $e_{n-r+1}^*,\ldots,e_n^*$, $L_k$ that
1586: of $e_{n-r+1-l}^*,\ldots,e_{n-r}^*,e_{n-r+1}^*,\ldots,e_{n-l}^*$, and
1587: finally that $f \in N_x^* I_h \simeq Hom(L_x^*,L_h)$ is defined by
1588: $f(e_j^*) = e_{n-r+j}^*$. Since
1589: $N_x^* I_k = Hom(L_x^*,L_k)$, a straightforward computation proves
1590: the lemma.
1591: \end{proo}
1592:
1593: \begin{lemm}
1594: \label{lemm_cone_spin}
1595: Let $x \in \G_Q^+(2p+1,4p+2),\ h\not = k \in \G_Q^-(2p+1,4p+2)$ such that
1596: $h,k \in I_x$. Let
1597: $f \in N_x^* I_h$ such that $q$ is defined at $f$.
1598: Then $q(\C.f + N_x^* I_k) \simeq \p^{2p-1}$.
1599: \end{lemm}
1600: \noindent
1601: This lemma implies theorem \ref{theo_cone} in this case since
1602: $q(\C.f + N_x^* I_k)$ is a linear space containing $h$ and $k$,
1603: and will therefore contain the line through $h$ and $k$.
1604: \begin{proo}
1605: We may assume that $x$ represents the isotropic subspace
1606: $L_x$ generated by $e_1^+,\ldots,e_{2p+1}^+$. Since $L_k$ meets $L_x$ along
1607: a hyperplane, we may further assume that this hyperplane is generated by
1608: $e_2^+,\ldots,e_{2p+1}^+$. We therefore have
1609: $N^*_x I_k = \wedge^2 \scal{e_2^+,\ldots,e_{2p+1}^+} \subset
1610: \wedge^2 L_x = T^*_x \G_Q^+(2p+1,4p+2)$. Let
1611: $f \in T^*_x \G_Q^+(2p+1,4p+2)$; since
1612: $f \not \in N^*_x I_k$ (otherwise we would have $h=k$),
1613: the class of $f$ modulo $N^*_x I_h$ is the same as that of some form
1614: $e_1^+ \wedge e$, with $e \in \scal{e_2^+,\ldots,e_{2p+1}^+}$, and we may
1615: assume that $e=e_2^+$.
1616:
1617: Recall that $I_x \simeq \p L_x^*$~:
1618: I claim that $q(\C.f + N_x^* I_k)$ is the orthogonal of $e_2^+$ in
1619: $\p L_x^*$. In fact, let $\wedge^p (\C.f + N_x^* I_k) \subset
1620: \wedge^{2p} L_x \simeq L_x^*$ be the linear span of all the forms
1621: in $\wedge^{2p} L_x$ which can
1622: be written as a wedge product of $p$ forms in
1623: $\C.f \oplus N_x^* I_k$. We have
1624: $\wedge^p (\C.f + N_x^* I_k) \subset {(e_2^+)}^\bot$; therefore
1625: $q(\C.f + N_x^* I_k) \subset \p {(e_2^+)}^\bot$.
1626:
1627: On the other hand, let
1628: $\displaystyle f_0=\Sigma_{i=1}^p e_{2i-1}^+ \wedge e_{2i}^+$; we have
1629: $\displaystyle f_0^{\wedge(p-1)}=\Sigma_{i=1}^p {e_{2i-1}^+}^*
1630: \wedge {e_{2i}^+}^* \wedge e_{2p+1}^*$, from which is follows that the
1631: rational map $\C.f + N_x^* I_k \dasharrow \p {(e_2^+)}^\bot,
1632: g \mapsto [g^{\wedge p}]$
1633: is submersive at $f_0$, which implies the claim and the lemma.
1634: \end{proo}
1635:
1636:
1637: \begin{lemm}
1638: \label{lemm_cone_e6_1}
1639: Let $x \in E_6/P_1,\ h \not = k \in E_6/P_6$ such that
1640: $h,k \in I_x$. Let
1641: $f \in N_x^* I_h$ such that $q$ is defined at $f$.
1642: If there passes a line through $h$ and $k$ in $E_6/P_6$
1643: then $q(\C.f + N_x^* I_k) \simeq \p^4$, otherwise
1644: $q(\C.f + N_x^* I_k) = I_x$
1645: \end{lemm}
1646: \begin{proo}
1647: We adopt the same strategy of proof as for lemma \ref{lemm_cone_spin}.
1648: Let $x \in E_6/P_1$ be fixed.
1649: In subsection \ref{subsection_fondamental}, we saw that
1650: $T^*_xX$ identifies with $\oc \oplus \oc$,
1651: $I_x$ with the projective quadric in $\p (\C \oplus \oc \oplus \C)$
1652: defined by $tu-N(z)=0$.
1653: We can assume that $k \in I_x$ is the class of $(0,0,1)$.
1654: Therefore, $N_x^* I_k = \overline{q^{-1}(k)} = \{(0,z) : z \in \oc\}$.
1655:
1656: Write $f=(z_0,z_1)$. Since
1657: $q((z_0,z_1)) = [N(z_0) : z_0 \overline z_1 : N(z_1)]$,
1658: there will be a line through
1659: $q(f)$ and $k$ in the quadric $I_x$ \iff $N(z_0) = 0$. If this occurs, then
1660: $$q(\C.f + N_x^* I_k) = \{[(0,u,t)]:t\in \C,u\in L(z_0) \},$$
1661: where $L(z_0)$ denotes the set of right multiples of $z_0$~:
1662: $L(z_0)=\{z_0z:z\in \oc\}$.
1663: It is a linear subspace of $\oc$ of dimension 4, so
1664: $q(\C.f + N_x^* I_k)$ is isomorphic with $\p^4$, as desired. If
1665: $N(z_0) \not = 0$, then left multiplication by $z_0$ is invertible, so that
1666: $q : \C.f + N_x^* I_k \dasharrow I_x$ is dominant, and the lemma again
1667: holds.
1668: \end{proo}
1669:
1670:
1671: \begin{lemm}
1672: \label{lemm_cone_e6_2}
1673: Let $\alpha \in E_6/P_2,\ \kappa,\lambda \in E_6/P_5$ such that
1674: $\kappa,\lambda \in I_\alpha$. Let
1675: $f \in N_\alpha^* I_\kappa$ such that $q$ is defined at $f$.
1676: Then $q(\C.f + N_\alpha^* I_\kappa) = I_\alpha$.
1677: \end{lemm}
1678: \begin{proo}
1679: We fix $\alpha \in E_6/P_2$. Let $f_1^*,\ldots,f_5^*$ be a
1680: base of $Q_\alpha^*$ and assume
1681: that $\kappa$ corresponds to the linear subspace
1682: generated by $f_4^*,f_5^*$. Recall that there is a natural
1683: surjective map $\pi:T_\alpha^* E_6/P_2
1684: \rightarrow Hom(L_\alpha^*,\wedge^2 Q_\alpha^*)$. Moreover,
1685: $\pi(N^*_\alpha I_\kappa) = Hom(L_\alpha^*,L)
1686: \subset Hom(L_\alpha^*,\wedge^2 Q_\alpha^*)$, where
1687: $L \subset \wedge^2 Q_\alpha^*$ is
1688: generated by $f_1^* \wedge f_4^*,f_1^* \wedge f_5^*,
1689: f_2^* \wedge f_4^*,f_2^* \wedge f_5^*,f_3^* \wedge f_4^*,f_3^* \wedge f_5^*,
1690: f_4^* \wedge f_5^*$ (for example, this follows from the fact that for any
1691: $\varphi \in Hom(L_\alpha^*,L)$,
1692: $\overline q(\varphi)$, if defined, equals $\kappa$).
1693:
1694: Let $M \subset \wedge^2 Q_\alpha^*$ be generated by
1695: $f_1^* \wedge f_2^*,f_1^* \wedge f_3^*,f_2^* \wedge f_3^*$, so that
1696: $L \oplus M = \wedge^2 Q_\alpha^*$; the class of $\pi(f)$ modulo
1697: $\pi(N_\alpha^* I_\kappa)$ is the class of a unique
1698: $\overline f \in Hom(L_\alpha^*,M)$.
1699: Assume first that the rank of $\overline f$ is 1. We can therefore
1700: assume that $\overline f(e_1^*) = f_1^* \wedge f_2^*$ and
1701: $\overline f(e_2^*) = 0$, where $e_1^*,e_2^*$
1702: is a suitable basis of $L_\alpha^*$.
1703:
1704: In the array below we give, for $\omega \in \wedge^2 Q_\alpha^*$, the value
1705: of the derivative $d\overline q_{\overline f}(\varphi)$, for
1706: $\varphi:L_\alpha^* \rightarrow \wedge^2 Q_\alpha^*$ given by
1707: $\varphi(e_1^*) = \omega$ and $\varphi(e_2^*) = 0$~:
1708: $$
1709: \begin{array}{llll}
1710: f_1^* \wedge f_2^* \mapsto 0 & f_1^* \wedge f_3^* \mapsto 0 &
1711: f_1^* \wedge f_4^* \mapsto 0 & \hspace{-1cm}f_1^* \wedge f_5^* \mapsto 0 \\
1712: f_2^* \wedge f_3^* \mapsto 0 & f_2^* \wedge f_4^* \mapsto f_3^* \wedge f_4^* &
1713: f_2^* \wedge f_5^* \mapsto f_1^* \wedge f_4^* + f_3^* \wedge f_5^*\\
1714: f_3^* \wedge f_4^* \mapsto 0 & f_3^* \wedge f_5^* \mapsto 0 &
1715: f_4^* \wedge f_5^* \mapsto f_1^* \wedge f_5^*.
1716: \end{array}
1717: $$
1718: The following gives similar values for $\varphi$ defined by
1719: $\varphi(e_1^*) = 0$ and $\varphi(e_2^*) = \omega$~:
1720: $$
1721: \begin{array}{llll}
1722: f_1^* \wedge f_2^* \mapsto 0 & f_1^* \wedge f_3^* \mapsto 0 &
1723: f_1^* \wedge f_4^* \mapsto 0 & \hspace{-1cm}f_1^* \wedge f_5^* \mapsto 0 \\
1724: f_2^* \wedge f_3^* \mapsto 0 & f_2^* \wedge f_4^* \mapsto f_2^* \wedge f_3^* &
1725: f_2^* \wedge f_5^* \mapsto f_3^* \wedge f_4^* + f_1^* \wedge f_2^*\\
1726: f_3^* \wedge f_4^* \mapsto 0 & f_3^* \wedge f_5^* \mapsto 0 &
1727: f_4^* \wedge f_5^* \mapsto f_1^* \wedge f_4^*.
1728: \end{array}
1729: $$
1730: It follows from these computations that
1731: $\overline q(\C.\overline f + \pi(N_\alpha^* I_\kappa))$ has
1732: dimension at least 6, so
1733: $\overline q(\C.\overline f + \pi(N_\alpha^* I_\kappa)) = I_\alpha$
1734: in this case.
1735:
1736: In case $\overline f$ has rank 2, the dimension of
1737: $\overline q(\C.\overline f + \pi(N_\alpha^* I_\kappa))$
1738: will not vary if $\overline f$ is replaced by
1739: $g.\overline f$, where $g \in SL(L_\alpha) \times SL(Q_\alpha)$
1740: preserves $\kappa$.
1741: Using a $\C^*$-action we can degenerate
1742: $\overline f \in Hom(L_\alpha^*,M)$
1743: to some element $\overline f_0$ of
1744: rank one, for which we have already seen that
1745: $\dim \overline q(\C.f_0 + \pi(N_\alpha^* I_\kappa)) = 6$.
1746: Since this dimension is
1747: lower semi-continuous, we have
1748: $\dim \overline q(\C.\overline f + \pi(N_\alpha^* I_\kappa)) = 6$ and
1749: the lemma is proved.
1750: \end{proo}
1751:
1752: \subsectionplus{The cotangent space and the tangent cone of a variety}
1753:
1754: \label{subsection_cotangent}
1755:
1756: \begin{defi} Let $x \in X$ be suitable.
1757: \begin{itemize}
1758: \label{cotangent}
1759: \item
1760: The embedded cotangent space of $X$ at $x$ is
1761: $\overline{N_xX}:=q(N^*_xX)\subset G/Q$.
1762: \item
1763: The embedded tangent space of $X$ at $x$ is
1764: $\overline{T_xX}={\overline{N_xX}}^P$.
1765: \item
1766: $X \subset G/P$ is a linear subvariety if $\overline{T_xX}$ does not
1767: depend on $x$ suitable in $X$.
1768: \end{itemize}
1769: \end{defi}
1770:
1771: \rks
1772: \begin{itemize}
1773: \item
1774: The notion of (co)-tangent space (and therefore of linear varieties)
1775: of $X \subset G/P$ could be defined for non maximal parabolic $P$, but
1776: then it would depend on the choice of a parabolic subgroup $Q$.
1777: \item
1778: An equivalent definition of linear subvarieties
1779: is that $\overline{N_xX}$ does not depend on
1780: suitable $x$ in $X$, since $\overline{N_xX} = {\overline{T_xX}}^Q$.
1781: \item
1782: By definition, $X^Q = \overline {\cup_{x \in X^s} \overline{N_xX}}$.
1783: \item
1784: In projective spaces, the tangent cone is the usual embedded tangent
1785: space and linear varieties are linear subspaces. Linear varieties will
1786: be classified in the next subsection.
1787: \end{itemize}
1788:
1789: \begin{exem}
1790: Let $x \in G/P$ and $X=\{x\}$. Then $\overline{T_xX} = \{x\}$.
1791: \end{exem}
1792: \begin{proo}
1793: In fact, $\overline{N_xX} = q(T^*_xG/P) = X^Q$, so this follows from
1794: theorem \ref{theo_bidualite}.
1795: \end{proo}
1796:
1797: \begin{lemm}
1798: For $x \in X^s$,
1799: $\overline{T_xX}$ is a cone over $x$ and therefore $x \in \overline{T_xX}$.
1800: \end{lemm}
1801: \pr
1802: In fact, for $x \in X^s$, we have $\overline{N_xX} \subset I_x$, so
1803: $\overline{T_xX} = {\overline{N_xX}}^P$ is a cone
1804: over $x$ by theorem
1805: \ref{theo_cone}.
1806: \qed
1807:
1808:
1809:
1810: \subsectionplus{Linear subvarieties}
1811:
1812: In this subsection, we classify linear subvarieties.
1813:
1814:
1815: \begin{prop}
1816: \label{prop_lineaire}
1817: The following array gives the list of all linear subvarieties~:
1818: $$
1819: \begin{array}{|c|c|c|}
1820: \hline
1821: \mbox{G/P} & \mbox{Linear varieties} \\
1822: \hline
1823: \G(r,n) & \G(r,p),\ r\leq p \leq n \\
1824: \hline
1825: \G^+_Q(2p+1,4p+2) & \{pt\};I_h,h \in \G_Q^-(2p+1,4p+2) \\
1826: \hline
1827: E_6/P_1 & \{pt\};I_h,h\in E_6/P_6\\
1828: \hline
1829: E_6/P_2 & \{pt\};I_\kappa,\kappa\in E_6/P_5\\
1830: \hline
1831: \end{array}
1832: $$
1833: \end{prop}
1834: \begin{proo}
1835: Let $X \subset G/P$ be linear.
1836: First, we prove that
1837: $\forall x \in X,\overline{T_xX} = X$, and that $X^Q$
1838: is linear.
1839: Let $x \in X^s$. Then
1840: $X^Q = \overline {\cup_{y \in X^s} \overline{N_yX}} = \overline {N_xX}$,
1841: since for all $y \in X^s$, $\overline{N_yX} = \overline{N_xX}$.
1842: Therefore, $X = {\overline{N_xX}}^P = \overline{T_xX}$ by
1843: corollary \ref{coro_bidualite}.
1844: Let $h \in X^Q$ and $x \in X$. Then, by biduality theorem again,
1845: $x \in \overline{N_hX^Q}$ \iff $h \in \overline {N_xX} = X^Q$.
1846: Therefore, $\overline{N_hX^Q} = X$ and $X^Q$ is linear and the claim
1847: is proved.
1848: Since $X = \overline{T_xX}$ for all $x \in X$, $X$ is a cone over all
1849: of its points by theorem \ref{theo_cone}.
1850: \lpara
1851:
1852: We finish the proof case by case.
1853: In the case of Grassmannians, if we denote
1854: $W = \sum_{x \in X} L_x$, since $X$ is a cone over all of its points,
1855: we have $\G(r,W) \subset X$, and so $X = \G(r,W)$.
1856:
1857: In the case of spinor varieties, any $x,y \in X$ must be linked, which
1858: implies that the line through $x$ and $y$ is in $X$, so $X$ is a
1859: linear subspace. As a consequence of the following proposition
1860: \ref{prop_spinoriel}, the only linear subspaces which dual
1861: variety is again a linear
1862: subspace are the point and maximal linear subspaces. Since we have seen
1863: that $X^Q$ must be a linear variety, the proposition follows in this
1864: case.
1865:
1866: Let $X \subset E_6/P_1$ be linear. Let $h \in X^Q$. If there are two
1867: points $x,y \in X$ such that there is no line through $x$ and $y$,
1868: then by lemma \ref{lemm_cone_e6_1} $L(x,y) = I_h$. Since $X \subset I_h$
1869: and $L(x,y) \subset X$, we have $X = I_h$ (and $X^Q$ is a point).
1870:
1871: Otherwise, by theorem \ref{theo_cone}, $X$ is a linear
1872: subspace. If $X^Q$ is not a linear subspace, by the argument above,
1873: $X = {(X^Q)}^P$ is a point. Assume now that both $X$ and $X^Q$ are
1874: linear subspaces, not reduced to a point. By lemma
1875: \ref{lemm_cone_e6_1}, $X^Q$ and $X={(X^Q)}^P$ contain a $\p^4$. But
1876: this implies that $X^Q \subset I_X$ is at most 1-dimensional
1877: (see the proof of theorem \ref{theo_cone}), and we get
1878: a contradiction.
1879:
1880: \lpara
1881:
1882: Let finally $X \subset E_6/P_3$ be linear. Assume $X$ is not reduced to
1883: a point. Let $h \in X^Q$; we have $X \subset I_h$.
1884: On the other hand, since
1885: $X$ is not a point, by lemma \ref{lemm_cone_e6_2}, it must
1886: contain $I_h$. Therefore, $X = I_h$.
1887: \end{proo}
1888:
1889:
1890:
1891:
1892: %******************************************************************************
1893:
1894: %******************************************************************************
1895:
1896: %******************************************************************************
1897:
1898:
1899:
1900:
1901:
1902:
1903:
1904:
1905: \sectionplus{Examples of dual varieties}
1906:
1907: \label{section_exemple}
1908:
1909:
1910: \subsectionplus{Dual varieties of isotropic Grassmannians}
1911:
1912:
1913: Let $V$ be a vector space, $B:V \rightarrow V^*$ a bilinear form.
1914: If $\epsilon = \pm 1$ and
1915: $\tr B = \epsilon B$, we say that $B$ is $\epsilon$-symmetric.
1916: Assume that this is the case. Let $r$ be an integer; we
1917: consider the variety $\G_B(r,V)$ of isotropic subspaces of $V$ of dimension
1918: $r$. The aim of this subsection is to describe the dual of
1919: $\G_B(r,V)$ in $G(r,V^*)$ in case $2r<\dim\ V$ (the other cases would be
1920: similar). Note that we don't assume that $B$ is an isomorphism.
1921:
1922: We have a rational map $\G_B(r,V) \dasharrow \G(r,V^*)$ which maps a linear
1923: subspace to its orthogonal, and which is well-defined at the point $\alpha$
1924: \iff $L_\alpha$ does not meet the kernel of $B$. Assuming there are
1925: such points, we call co-isotropic Grassmannian the image of this
1926: rational map.
1927:
1928: \begin{prop}
1929: \label{prop_symmetric}
1930: Assume $\epsilon=1$. Then $\G_B(r,V)$ is suitable \iff and
1931: $r \leq \rank(B)$. In this case, the dual variety of the isotropic Grassmannian
1932: $G_B(r,V)$ is the co-isotropic Grassmannian.
1933: \end{prop}
1934:
1935: \begin{prop}
1936: \label{prop_skew}
1937: Assume $\epsilon=-1$. Then $\G_B(r,V)$ is suitable \iff $r$ is even
1938: and $r \leq \rank(B)$. In
1939: this case, the dual variety of the isotropic Grassmannian
1940: $G_B(r,V)$ is the co-isotropic Grassmannian.
1941: \end{prop}
1942:
1943: \begin{proo}
1944: We prove propositions \ref{prop_symmetric} and \ref{prop_skew}
1945: simultaneously. Let $x \in \G_B(r,V)$ be generic. Under the natural
1946: isomorphism $T_x \G(r,V) \simeq Hom(L_x,V/L_x)$, we have
1947: the inclusion
1948: $T_x \G_B(r,V) \supset Hom(L_x,L_x^\bot/L_x)$,
1949: where $L_x^\bot$ denotes the orthogonal of $L_x$ with respect to
1950: $B$. It follows that if $\codim\ L_x^\bot < r$, then
1951: $N_x^* \G_B(r,V)$ does not meet the open orbit in $T_x^* \G_B(r,V)$.
1952: If $r > \rank(B)$, this occurs for all $x \in \G_B(r,V)$, hence
1953: $\G_B(r,V)$ is not suitable.
1954:
1955: Assume $r \leq \rank(B)$.
1956: Now, let $x \in \G_B(r,V)$ such that $\codim\ L_x^\bot = r$.
1957: Denote $Q_x = V/L_x$; we have a morphism
1958: $Q_x \rightarrow L_x^*$,
1959: induced by $B$. Clearly, $T_x \G_B(r,V) \subset Hom(T_x,Q_x)$
1960: is the
1961: subspace of linear maps such that the composition
1962: $L_x \rightarrow Q_x \rightarrow L_x^*$
1963: is $(-\epsilon)$-symmetric. Therefore,
1964: the normal space of $\G_B(r,V)$ at $x$ identifies with
1965: $\epsilon$-symmetric maps $L_x^* \rightarrow L_x$.
1966: Since $\G_B(r,V)$ will be suitable \iff there are such maps of rank $r$,
1967: this occurs in all cases if $\epsilon = 1$ and exactly when $r$ is even
1968: when $\epsilon = -1$.
1969:
1970: Now, the computation of the dual variety is straightforward : since
1971: we have already remarked that
1972: $T_x \G_B(r,V) \supset Hom(L_x,L_x^\bot/L_x)$,
1973: the image of a generic conormal form at $x$ under the rational map
1974: $q:T^*_x \G(r,V) \dasharrow G(r,V^*)$ is the element in
1975: $\G(r,V^*)$ corresponding to $L_x^\bot$.
1976: \end{proo}
1977:
1978:
1979: \subsectionplus{Schubert varieties and quivers in the fundamental case}
1980:
1981: \label{subsection_carquois}
1982:
1983: In this subsection, I recall that to a cominuscule homogeneous space one
1984: can naturally associate a quiver, such that Schubert cells are parametrised
1985: by some subquivers. I also recall the Hasse diagram of a representation,
1986: and show how the quiver of a cominuscule homogeneous space can be
1987: identified with the Hasse diagram of a tangent space.
1988: This identification is
1989: due to Nicolas Perrin and
1990: Laurent Manivel.
1991: Then, I show that this identification behaves well as far as Schubert
1992: subvarieties are concerned. Finally, I extend these results to
1993: $E_6/P_3$, which is not a cominuscule homogeneous space.
1994:
1995: \lpara
1996:
1997: The quiver of a cominuscule homogeneous space has been first introduced by
1998: N. Perrin \cite[definition 3.2]{carquois};
1999: here we use the slightly different definition
2000: \cite[definition 2.1]{quantique}. Recall that
2001: $\G(r,V),\G_Q^+(2p+1,4p+2)$ and $E_6/P_1$ are cominuscule spaces (in
2002: fact even minuscule). The quiver is defined using a reduced expression
2003: of $w_{G/P}$, the shortest element in the class of $w_0$ in $W/W_P$
2004: ($w_0$ is the longest element in $W$).
2005: Choose a reduced expression
2006: $w_{G/P} = s_{\beta_1} \cdots s_{\beta_N}$, with $N = \dim G/P$; the
2007: vertices of the quiver
2008: $Q_{G/P}$ are in bijection with $[1,\ldots,N]$, and we refer
2009: to \cite[definition 2.1]{quantique} for the definition of the arrows.
2010: The quivers may be illustrated by relevant
2011: examples as follows~: \vspace{-.8cm}
2012:
2013: $$
2014: \begin{array}{ccccc}
2015: \input{g37} && \input{gq5} && \input{e6} \\
2016: \G(3,7) && \G_Q^+(5,10) && E_6/P_1
2017: \end{array}
2018: $$
2019: In these pictures, all arrows are going down. Moreover, we will use
2020: the definition of height of a vertex of such a quiver. More or less by
2021: definition (see \cite[definition 4.7]{carquois}), it is the height of
2022: the vertex in the above drawing, where by convention the lowest vertex
2023: has height 1 (so the highest vertex has height respectively 6,7,11 for
2024: $\G(3,7),\G_Q^+(5,10),E_6/P_1$).
2025: \lpara
2026:
2027: Later we will have to identify this quiver with a Hasse diagram.
2028: Let $V$ be a representation of a semi-simple group $\Lambda$.
2029: Let us recall that the Hasse diagram
2030: of $V$ is a quiver defined as follows.
2031: The vertices of this quiver are the weights of $V$, and
2032: there is an arrow from $\lambda_1$ to $\lambda_2$ if and only if
2033: $\lambda_2 - \lambda_1$ is a simple root. For example, the Hasse diagram
2034: of the $8$-dimensional representation of $Spin_8$ is given on the left~:
2035:
2036: $$
2037: \begin{array}{ccc}
2038: \input{hasseq8} && \input{racineq10} \\
2039: Spin_8 & \hspace{1cm} & \mbox{Roots for }\Q^8
2040: \end{array}
2041: $$
2042:
2043: \begin{prop}
2044: \label{hasse}
2045: Let $G/P$ be cominuscule and let $x \in G/P$ be the base point.
2046: Let $\Lambda$ be a Levi factor
2047: of the stabilisor of $x$. Then the quiver $Q_{G/P}$
2048: of $G/P$ is isomorphic with the
2049: Hasse diagram $H_{G/P}$ of the $\Lambda$-module
2050: $\widehat{T_xX}/L_x$.
2051: \end{prop}
2052: \noindent
2053: If $G/P \subset V$, recall that $\widehat{T_xG/P} \subset V$
2054: is the affine tangent space at $x$; it contains the
2055: line $L_x \subset V$ represented by $x \in \p V$,
2056: so that it makes sense to consider the quotient
2057: $\widehat{T_xG/P}/L_x$.
2058: We have stated this result without proof in
2059: \cite[proposition 7]{quantique}. In this
2060: article
2061: I need the explicit isomorphism, this is why I sketch the proof,
2062: leaving details to the reader.
2063: \begin{proo}
2064: It is known that to each vertex of the quiver one can associate a root of
2065: $G$. In fact, choose a reduced expression
2066: $w_{G/P} = s_{\beta_1} \ldots s_{\beta_N}$ and set
2067: $$ \alpha_i = s_{\beta_N} \circ \ldots \circ s_{\beta_{i+1}} (\beta_i). $$
2068: Since two different reduced expressions for
2069: $w_{G/P}$ only differ by commutation relations,
2070: it is easy to check that the induced map from the set of vertices of
2071: the quiver
2072: to the set of roots is well-defined (it does not depend on the reduced
2073: expression). In the following, we consider that a reduced expression is
2074: chosen, thus identifying this set of vertices with $[1,N]$.
2075:
2076: For example, if $G/P$ is a smooth 8-dimensional quadric,
2077: then its quiver, and the corresponding roots, are given above
2078: (here we have shitfted the indices,
2079: denoting $(\epsilon_0,\ldots,\epsilon_4)$ a basis of the weight lattice of
2080: $Spin_{10}$). Note that the highest weight of the corresponding
2081: $Spin_{10}$-representation is $\epsilon_0$,
2082: and that we recover the Hasse diagram
2083: of $Spin_8$ by considering the weights $\epsilon_0 - \alpha_i$.
2084:
2085: By \cite[proposition 4.9]{carquois}, we may reduce the proof of our
2086: proposition to the particular case of a quadric
2087: of any dimension, as above, because if
2088: there is an arrow $i \rightarrow j$ in the quiver of $G/P$, then $i$
2089: and $j$ belong to a subquiver of $Q_{G/P}$ isomorphic with the quiver
2090: of a quadric. It is also possible (and probably shorter)
2091: to check directly in each case that if $\omega$ denotes the highest
2092: weight of $\Gamma(G/P,\co(1))$, then the set
2093: $\{\omega - \alpha_i : 1\leq i \leq N\}$ is exactly the set of weights
2094: of the tangent space at the base point of $G/P$, and that the
2095: bijection $i \mapsto \omega - \alpha_i$ is an
2096: isomorphism of quivers $Q_{G/P} \rightarrow H_{G/P}$.
2097: \end{proo}
2098:
2099: \lpara
2100:
2101: Given $[w] \in W/W_P$, we associate the Schubert subvariety
2102: $C_{[w]} \subset G/P$ which is the $B$-orbit closure of
2103: $[w] \in G/P$. Assuming that $w$ is the minimal length representative
2104: of its class, we choose a reduced decomposition of $w$, and this
2105: defines a subquiver $Q_w$ of the quiver $Q_{G/P}$
2106: which is an order ideal
2107: (this means that if $i \rightarrow j$ is an arrow in $Q_{G/P}$ and $i
2108: \in Q_w$, then $j \in Q_w$~:
2109: see \cite[proposition 4.5]{carquois}). We can also consider the
2110: subset $H_w$ of $H_{G/P}$ which elements are the weights of
2111: $w^{-1}.T_{[w]}C_{[w]} \subset T_{[e]} G/P$. The following proposition
2112: will be useful to compute the dual variety of $C_{[w]}$, because it
2113: describes the tangent bundle of $C_{[w]}$~:
2114:
2115: \begin{prop}
2116: \label{prop_tangent_cw}
2117: Under the isomorphism $Q_{G/P} \simeq H_{G/P}$
2118: of proposition \ref{hasse}, we have
2119: $Q_w = H_w$.
2120: \end{prop}
2121: \begin{proo}
2122: Recall that $\omega$ denotes the highest weight of
2123: $\Gamma(G/P,\co(1))$. All the weights of
2124: $\widehat{T_x G/P}/L_x$ are of the
2125: form $\omega + \alpha$, where $\alpha$ are all the roots not in
2126: $\plie = Lie(P)$ (therefore $\alpha$ is a negative root).
2127: All the weights of $T_{[w]} C_{[w]}$ are of the form
2128: $w.\omega + \beta$, with $\beta$ a positive root. Therefore, if
2129: $\omega + \alpha$ is a weight of $w^{-1}.T_{[w]}C_{[w]}$,
2130: $w.\alpha$ must be a positive root. So $\alpha$ must be a negative
2131: root sent by $w$ to a positive root. Denote $l(w)$ the length of $w$;
2132: there are $l(w)$ such roots,
2133: namely $\{-\alpha_i:1\leq i \leq l(w)\}$. Since $l(w)$ is also the
2134: dimension of $C_{[w]}$, it follows that
2135: $H_w$ is exactly the set of weights of the form
2136: $\omega - \alpha_i,1\leq i \leq l(w)$, so the proposition follows.
2137: \end{proo}
2138:
2139: \lpara
2140:
2141: We now consider the case of $E_6/P_3$. Let
2142: $[w] \in W/W_3$; we want to define a quiver $Q_{E_6/P_3}$ and a
2143: subquiver $Q_w$ which pictures the tangent bundle of $C_{[w]}$.
2144: Since $E_6/P_3$ is not cominuscule, the quiver defined as in
2145: \cite[definition 3.2]{carquois} is not well-defined
2146: (it depends on a reduced expression of $w_{E_6/P_3}$), and
2147: as we have already seen, the cotangent bundle $T^* E_6/P_3$ is no
2148: longer irreducible, so its Hasse diagram is not suitable neither.
2149:
2150: But our luck is that for $f \in T^*_\alpha E_6/P_3$, $q(f)$ only
2151: depends on $\pi(f) \in L_\alpha \otimes \wedge^2 Q_\alpha^*$;
2152: therefore, what we care for is not really the conormal bundle of
2153: $C_{[w]}$, but rather its projection to the bundle
2154: $L \otimes \wedge^2 Q^*$. This is why we consider the following~:
2155:
2156: \begin{defi}\
2157: \begin{itemize}
2158: \item
2159: Let $[e] \in E_6/P_3$ be the base point, and let $\Lambda$ denote a Levi
2160: factor of $P_3$.
2161: \item
2162: Let $Q_{E_6/P_3}$ denote the Hasse diagram of the $\Lambda$-module
2163: $L_{[e]}^* \otimes \wedge^2 Q_{[e]} \subset T_{[e]} E_6/P_3$.
2164: \item
2165: For $[w] \in W/W_3$ with $w$ the minimal length representative,
2166: let $Q_w \subset Q_{E_6/P_3}$ denote the set of
2167: weights of
2168: $w^{-1}.T_{[w]} C_{[w]}\ \cap\
2169: (L_{[e]} \otimes \wedge^2 Q_{[e]}^*)$.
2170: \end{itemize}
2171: \end{defi}
2172:
2173: \begin{prop}
2174: \label{prop_order_ideal}
2175: For $[w] \in W/W_3$, $Q_w \subset Q_{E_6/P_3}$ is an order ideal.
2176: \end{prop}
2177: \begin{proo}
2178: For $a \in \{-2,-1,0,1,2\}$, let $\g_k \subset \e_6$ denote
2179: $
2180: \displaystyle
2181: \bigoplus_{
2182: \begin{array}{c}
2183: \alpha = \sum_j k_j \alpha_j\\
2184: k_3 = a
2185: \end{array}
2186: } \g_{\alpha}
2187: $
2188: (by this I mean that the Cartan subalgebra is included in $\g_0$).
2189: We have $\g = \g_{-2} \oplus \g_{-1} \oplus \g_0 \oplus \g_1
2190: \oplus \g_2$ and $Lie(P_3) = \g_0 \oplus \g_1
2191: \oplus \g_2$. The tangent space $T_{[e]} E_6/P_3$ decomposes as
2192: $\g_{-2} \oplus \g_{-1}$; let
2193: ${\cal P}$ denote the weights of
2194: $\widehat{T_{[e]} E_6/P_3}/L_{[e]}$ which are of the
2195: form $\omega + \alpha$, with $\alpha$ a root of $\g_{-1}$ (${\cal P}$
2196: is also the set of weights of $L_{[e]} \otimes \wedge^2 Q_{[e]}^*$).
2197: I claim that $w$ induces an increasing bijection between
2198: ${\cal P}$ and its image.
2199: The proposition follows from this claim because
2200: $Q_{[w]}$ is the set of weights of
2201: $w^{-1}.T_{[w]} C_{[w]}$ which are in ${\cal P}$; arguing as in the
2202: proof of proposition \ref{prop_tangent_cw}, this is the set of roots
2203: of $\g_{-1}$ which are mapped to a positive root by $w$, and this is
2204: obviously an order ideal since $w$ is increasing.
2205:
2206: To prove the claim, we note that
2207: $L_{[e]} \otimes \wedge^2 Q_{[e]}^*$ is a minuscule
2208: $\Lambda$-representation, since $\Lambda$ contains
2209: $SL_2 \times SL_5$. Therefore $W_P$ permutes transitively the roots in
2210: $\g_{-1}$. Let $\alpha_0$ be the highest root of $\g_{-1}$, let
2211: $\alpha_1,\alpha_2$ be roots of $\g_{-1}$ and assume
2212: $\alpha_1 \leq \alpha_2$. We can find $w_1,w_2 \in W_P$ such that
2213: $\alpha_i = w_i . \alpha_0$, assume moreover that $w_1,w_2$ are
2214: minimal such elements. Since $\alpha_1 \leq \alpha_2$, we
2215: have $w_1 \geq w_2$ for the Bruhat order, so that we may assume that a
2216: $w_2$ is a product of reflexions appearing in a reduced expression of
2217: $w_1$. Since $w$ is a minimal length representative in $W/W_3$, the
2218: product $w.w_1$ is still a reduced expression, so
2219: $w.w_1 \geq w.w_2$, and so $w.\alpha_1 \leq w.\alpha_2$, as claimed.
2220: \end{proo}
2221:
2222: \begin{rema}
2223: The same proof works for any $G/P$, as soon as $\g_{-1}$ is a
2224: minuscule $\Lambda$-representation, with the notations of the proof.
2225: \end{rema}
2226:
2227:
2228:
2229:
2230: %******************************************************************************
2231:
2232:
2233:
2234:
2235:
2236: \subsectionplus{Schubert varieties and dual varieties}
2237:
2238: \label{subsection_combinatoire}
2239:
2240: The usual dual variety of a linear subspace is again a linear
2241: subspace. The
2242: goal of this section is to generalise this result for Schubert varieties.
2243:
2244: \begin{prop}
2245: \label{prop_schubert}
2246: Let $X \subset G/P$ be a suitable
2247: Schubert variety. Then $X^Q \subset G/Q$
2248: is a Schubert variety.
2249: \end{prop}
2250: \pr
2251: In fact, $X^Q$ is a $B$-stable (proposition \ref{foncteur})
2252: irreducible closed subvariety of $G/Q$.
2253: \qed
2254:
2255: \lpara
2256:
2257: Recall that $B$-stable
2258: Schubert varieties in $G/P$ are parametrised by the quotient set
2259: $W/W_P$. For $[w] \in W/W_P$ (resp. $[x] \in W/W_Q$),
2260: we denote as in the previous subsection
2261: $C_{[w]} = \overline {B.[w]} \subset G/P$ the corresponding Schubert
2262: subvariety (resp. $D_{[x]} = \overline {B.[x]} \subset G/Q$).
2263: In the rest of this article, I give a description of the $[w]$'s
2264: such that the $C_{[w]}$ is suitable, and of the
2265: element in $W/W_Q$ corresponding to the dual Schubert variety, in the
2266: fundamental cases. According to
2267: section \ref{section_reduction}, this is enough to describe all dual varieties
2268: of Schubert varieties. The strategy for this description is first to use
2269: a $T$-fixed point argument, to reduce the task to a purely
2270: combinatorial one. In the types $A$ and $D$, I then give an
2271: explicit solution of this combinatorial problem. For the exceptional
2272: cases, my description of the dual Schubert varieties is not really
2273: explicit, but to compute them there is in principal only a finite
2274: number of computations to make.
2275:
2276: \lpara
2277:
2278: So we fix the minimal $G$-representation $V$ such that $G/P \subset \p V$. We
2279: denote $V = \oplus V_\lambda$ (resp. $V^* = \oplus V^*_\mu$)
2280: the weight decomposition of $V$ (resp. $V^*$).
2281: Let $C_{[w]} = \overline {B.[w]}$ be a Schubert
2282: variety. Recall that $\overline{N_{[w]} C_{[w]}}$
2283: denotes the variety of $y$'s in
2284: $G/Q$ which are tangent to $C_{[w]}$ at $[w]$, see definition
2285: \ref{cotangent}.
2286:
2287: \begin{lemm}
2288: \label{t_fixe}
2289: Let $[x] \in W/W_Q$ such that
2290: $C_{[w]}^Q = \overline {B.[x]}$. We have
2291: $[x] \in \overline{N_{[w]} C_{[w]}}$.
2292: \end{lemm}
2293: \begin{proo}
2294: First, notice that $C_{[w]}^Q = \overline
2295: {B.\overline{N_{[w]}C_{[w]}}}$. In fact,
2296: $B.\overline{N_{[w]}C_{[w]}}$ contains the set of $y$'s in $G/Q$
2297: which are tangent at a point in $B.[w]$, therefore at a generic point
2298: of $C_{[w]}$.
2299:
2300: Let $\mu_0$ be the highest weight of $V^*$ and denote $\mu = x.\mu_0$.
2301: Let $y \in V^*$ such that $[y] \in C_{[w]}^Q \subset \p V^*$.
2302: Use the weight decomposition
2303: of $V^*$ to write $y = \sum_{\mu'} y_{\mu'}$.
2304: Since $[y] \in C_{[w]}^Q$ which is the closure of the $B$-orbit of the weight
2305: line of weight $\mu$,
2306: $y_{\mu'} = 0$ if $\mu' \not \geq \mu$. Assume that
2307: $\forall [y] \in \overline{N_{[w]} C_{[w]}},y_{\mu}=0$.
2308: It would then follow from
2309: the first point that $\forall y \in C_{[w]}^Q, y_{\mu}=0$, contradicting
2310: $[x] \in C_{[w]}^Q$.
2311:
2312: Therefore, there exists $[y]$ in $\overline{N_{[w]}C_{[w]}}$ such that
2313: $y_\mu \not = 0$. Since $\overline{N_{[w]}C_{[w]}}$
2314: is $T$-stable and $V_\mu$ is
2315: one-dimensional,
2316: we have $[x] \in \overline{N_{[w]} C_{[w]}}$.
2317: \end{proo}
2318:
2319: I now explain how to compute
2320: the element $[x]$ of the previous lemma.
2321: We want to take into account the case of $E_6/P_3$, which cotangent
2322: bundle is not irreducible. Recall that in this case there is a natural
2323: bundle morphism $\pi : T^* E_6/P_3 \rightarrow L \otimes \wedge^2 Q^*$.
2324: To have uniform notations, in the other cases we denote
2325: $\pi : T^* G/P \rightarrow T^* G/P$ the identity and $\overline q = q$.
2326:
2327: Decompose
2328: $\pi(T_{[e]}^* G/P)$ as a sum of weight spaces for the action of a Levi
2329: subgroup of $P$ : $\pi(T_{[e]}^* G/P) = \oplus T^*_\tau$, and write similarly
2330: $\scal{I_{[e]}} = \oplus_\nu N_\nu$, where if $[w'] \in W/W_P$,
2331: $\scal{I_{[w']}} \subset V^*$ denotes the linear span of the Schubert variety
2332: $I_{[w']} \subset G/Q \subset \p V^*$ (see notation \ref{i_h}).
2333: It can be easily checked directly on the
2334: examples that all the weight spaces $T^*_\tau$ and $N_\nu$ have dimension 1.
2335: The rational map
2336: $\overline q : \pi(T_{[e]}^* G/P) \dasharrow G/Q \subset \p V^*$
2337: is given by a list
2338: of polynomial functions $\pi(T_{[e]}^* G/P) \rightarrow N_\nu$
2339: of the same degree $d$ and with values in the complex line $N_\nu$.
2340: The polarisations of these polynomials yield $d$-linear maps
2341: $T_{\tau_1}^* \times \cdots \times T_{\tau_d}^* \rightarrow N_\nu$, which
2342: will be denoted $P_{\tau_1,\ldots,\tau_d;\nu}$; remark that the space of
2343: such $d$-linear maps has dimension 1. Given $w \in W$, we denote
2344: ${\cal P}_w$ the set of weights $\nu$ such that there exist weights
2345: $\tau_1, \ldots , \tau_d$ of
2346: $\pi(w^{-1}.N^*_{[w]} C_{[w]}) \subset \pi(T_{[e]}^* G/P)$
2347: such that $P_{\tau_1,\ldots,\tau_d;\nu}$ does not vanish.
2348: \lpara
2349:
2350: \begin{prop}
2351: \label{prop_combinatoire}
2352: Let $[w] \in W/W_P$, with $w$ its minimal length representative.
2353: The variety $C_{[w]}$ is suitable \iff ${\cal P}_w$ is not empty. In this case,
2354: if we denote $[x] \in W/W_Q$ such that $C_{[w]}^Q = C_{[x]}$,
2355: then $x.\mu_0$ equals $w.\mu_1$, where $\mu_1$ is the
2356: lowest weight in ${\cal P}_w$.
2357: \end{prop}
2358: \begin{proo}
2359: In fact, by lemma \ref{t_fixe} and its proof, $x.\mu_0$ is the lowest weight
2360: $\mu$, if any,
2361: such that the $\mu$-component of the restriction of the rational
2362: map
2363: $\pi(T^*_{[w]} G/P) \dasharrow G/Q$ to $N^* C_{[w]}$
2364: does not vanish identically.
2365:
2366: The weights of $\scal{I_{[w']}}$ for
2367: $[w'] \in W/W_P$ are some weights of $V^*$,
2368: a set on which $W$ acts; therefore it makes sense to talk of
2369: $w''.\mu'$, for $w'' \in W$ and $\mu'$ a weight of $\scal{I_{[w']}}$.
2370: I claim that $w$ induces an increasing bijection between the weights of
2371: $\scal{I_{[e]}}$ and those of $\scal{I_{[w]}}$.
2372: The argument is similar to that of proposition \ref{prop_order_ideal}.
2373: In fact, a weight of $\scal{I_{[e]}}$ can be
2374: written as $v.\mu_0$, with $v \in W_P$ and $\mu_0$ the highest weight
2375: of $V^*$. Given two such weights $v_1.\mu_0 \geq v_2.\mu_0$, we can assume
2376: that $v_2$ is the minimal length representative of its class in
2377: $W_P / W_{P \cap Q}$.
2378: Thus $v_1$ can be written as a product of some reflections which occur
2379: in a reduced expression of $v_2$. Since $v_2 \in W_P$ and $w$ is a minimal
2380: length representative, $l(wv_2) = l(w) + l(v_2)$. Therefore
2381: $wv_1 \leq wv_2$ for the Bruhat order, and thus
2382: $w.(v_1.\mu_0) \geq w.(v_2.\mu_0)$. This proves the claim.
2383:
2384: Since the
2385: rational map $\overline q : \pi(T^* G/P) \dasharrow G/Q$
2386: is $G$-equivariant, the
2387: weight $\mu$
2388: of the proof of lemma \ref{t_fixe}
2389: is also $w.\mu_1$, where $\mu_1$ is the lowest weight such
2390: that the $\mu_1$-component of the restriction of the rational
2391: map $T^*_{[e]} G/P \dasharrow G/Q$ to
2392: $w^{-1}.N^*_{[w]} C_{[w]}$ does not vanish identically.
2393: Obviously $\mu_1$ is the lowest weight $\nu$ such that some
2394: $P_{\tau_1,\ldots,\tau_d;\nu}$
2395: with $\tau_1,\ldots,\tau_d$ weights of $w^{-1}.N^*_{[w]} C_{[w]}$,
2396: does not vanish, so the proposition is proved.
2397: \end{proo}
2398:
2399: We illustrate our method with the easy example $G/P = \p V$.
2400: Let $(e_1,\ldots,e_n)$ be a basis of $V$, let $k$ be an integer and let
2401: $L_k = Vect(e_1,\ldots,e_k),\ M_k = Vect(e_{k+1},\ldots,e_n)$.
2402: We consider the Schubert variety $X = \p L_k \subset \p V$ and compute
2403: its dual variety. The corresponding element of the Weyl group is the
2404: transposition $w=(1k)$. We have
2405: $T_{[w]} X \simeq Hom(e_k,L_k/e_k)$, so
2406: $w^{-1}. T_{[w]} X \simeq Hom(e_1,L_k/e_1)$ and so
2407: $w^{-1}. N^*_{[w]} X \simeq Hom(M_k,e_1)$. Since $q$ is defined taking the
2408: kernel, the lowest weight in $q(w^{-1}. N^*_{[w]})$ is
2409: $\mu_1 = -\epsilon_{k+1}$. We have $w.\mu_1 = \mu_1$, so that the dual
2410: variety of $X$ is the $B$-orbit closure of $e_{k+1}^*$, as
2411: expected. Note that in this example it would have been easier to compute
2412: directly the lowest weight in
2413: $q(N^*_{[w]}C_{[w]})$, instead of applying first $w^{-1}$ and then $w$. In fact,
2414: this is what we will do to compute dual varieties of Schubert varieties
2415: in the cases $G/P = \G(r,V)$ and $G/P = \G_Q^+(2p+1,4p+2)$.
2416: \lpara
2417:
2418: Recall from subsection \ref{subsection_carquois}
2419: the definition of height of a vertex of the quiver
2420: $Q_{G/P} = H_{G/P}$. We denote $h_0$ the maximal $h$ such that there exist
2421: $\tau_1,\ldots,\tau_d \in H_{G/P}$, $\nu$ a weight of $I_{[e]}$ such
2422: that $h(\tau_i) \geq h$ and $P_{\tau_1,\ldots,\tau_d;\nu} \not = 0$.
2423: We have the following values for $h_0$ (I have also indicated the
2424: height $h_{\max}$ of the heighest element of $Q_{G/P}$)~:
2425: $$
2426: \begin{array}{ccccc}
2427: G/P & \G(r,n) & \G_Q^+(2p+1,4p+2) & E_6/P_1 & E_6/P_3\\
2428: h_{\max} & n-1 & 4p-1 & 11 & 8\\
2429: h_0 & \max(r,n-r) & 2p+1 & 8 & 5
2430: \end{array}
2431: $$
2432:
2433: \begin{theo}
2434: \label{theo_adapte}
2435: The Schubert subvariety $C_{[w]}$ is suitable \iff all the vertices of
2436: $Q_w$ have height at most $h_0-1$.
2437: \end{theo}
2438: \begin{proo}
2439: Unfortunately, I don't know how to prove in a uniform way this
2440: theorem. It will follow from propositions \ref{grass_adapte},
2441: \ref{prop_spinoriel}, \ref{prop_e6_1_adapte} and
2442: \ref{prop_e6_2_adapte}. The proof of these propositions also imply the
2443: above given values of $h_0$.
2444: \end{proo}
2445:
2446:
2447:
2448:
2449: %******************************************************************************
2450:
2451:
2452:
2453:
2454: \subsectionplus{Case of Grassmannians}
2455:
2456: \label{subsection_schubert_grassmannienne}
2457:
2458: Recall that $V$ is an $n$-dimensional vector space.
2459: We will
2460: parametrise Schubert varieties in $\G(r,V)$ by increasing lists of $r$
2461: integers,
2462: instead of partitions, because duality will appear easier to formulate
2463: in this way.
2464: The list $(l_i)$ will correspond to the Schubert variety
2465: $C_l \subset \G(r,V)$ (resp. in $D_l \subset \G(r,V^*)$) which is the
2466: $B$-orbit closure of the linear space spanned by the
2467: $l_i$'s $T$-eigenvectors in $V$ (resp. in $V^*$).
2468: For $x \in \{1,\ldots,n\}$, we will write $x \in l$ to mean that there
2469: exists $i$ such that $x = l_i$.
2470: The $T$-fixed points in $V$ (resp. $V^*$) will be denoted $e_i$
2471: (resp. $e_i^*$).
2472: The $T$-fixed point whose $B$-orbit is dense in $C_l$
2473: (resp. $D_l$) will be denoted $x_l$
2474: (resp. $y_l$).
2475: The Bruhat order on Schubert cells is given by
2476: $l \leq m$ \iff $\forall i,l_i \leq m_i$. If $x_1,\ldots,x_r$ are
2477: distinct integers
2478: not necessarily increasing, we denote the list obtained reordering the
2479: $x_i$ as $[x_1,\ldots,x_r]$.
2480:
2481: Let $T_l$ denote $Vect(e_i : i \in l)$ and let
2482: $Q_l$ denote $Vect(e_i : i \not \in l)$.
2483: The tangent
2484: space at $x_l$ identifies with
2485: $Hom(Q_l,T_l)$.
2486: A weight in this space is given by a couple
2487: $(x,y):x \in l,y \not \in l$.
2488: Recall that the rational map
2489: $$T_{x_l}^* \G(m,n) \simeq
2490: Hom(Q_l,T_l) \dasharrow \G(r,V^*)$$
2491: is given by $\varphi \mapsto \ker \varphi$.
2492: Thus the degree of $q$ is $r$ and
2493: with the notations before proposition \ref{prop_combinatoire},
2494: we have~:
2495:
2496: \begin{fait}
2497: \label{fait_P_grassmannienne}
2498: The multilinear form
2499: $P_{(x_1,y_1),\ldots,(x_r,y_r);l'}$ does not vanish \iff the $x_i$'s
2500: and the $y_i$'s are all distinct, and $l'$ is the set of the $y_i$'s.
2501: \end{fait}
2502:
2503: \noindent
2504: Given a list $l$, we consider the list $l^*$ defined inductively by
2505: $$
2506: l^*_i = \min \{ y : y>y_{i-1}, y>x_i, \forall j, y\not = x_j \}.
2507: $$
2508:
2509: \begin{lemm}
2510: We have $\forall i,l_i^* \leq n$ \iff $\forall i \in \{1,\ldots,r\},
2511: l_i<n+2i-2r$.
2512: \end{lemm}
2513: \noindent
2514: In terms of partitions, this means that the $i$-th part
2515: must be at least $r+1-i$.
2516: \begin{proo}
2517: Let $i$ be an integer. The integers $l_j$ for $j>i$ and $l_j^*$ for
2518: $j \geq i$ are strictly greater than $l_i$
2519: and distinct, so the lemma follows.
2520: \end{proo}
2521:
2522: \noindent
2523: As the following proposition
2524: shows, $l \mapsto l^*$ is the combinatorial model for the duality of
2525: Schubert varieties in Grassmannians~:
2526:
2527: \begin{prop}
2528: \label{grass_adapte}
2529: $C_l$ is suitable \iff $\forall i \in \{1,\ldots,r\},l_i < n+2i-2r$.
2530: If $C_l$ is suitable then $C_l^Q = D_{l^*}$.
2531: \end{prop}
2532: \begin{proo}
2533: With the previous notations, the weights of the conormal
2534: space $N^*_{x_l} C_l$ are the couples
2535: $(x,y)$ with $x \in l,y \not \in l$, and $y>l_x$.
2536:
2537: By proposition \ref{prop_combinatoire} and the comment after it,
2538: $C_l$ is suitable \iff there are lists $[y_1,\ldots,y_r]$
2539: and $x_1,\ldots,x_r$ with
2540: $(x_i,y_i)$ a weight of $N^*_{x_l} C_l$ and
2541: $P_{(x_1,y_1),\ldots,(x_r,y_r);l'} \not = 0$.
2542: In this case, if we denote $l'$ the list
2543: such that $C_l^Q = D_{l'}$, then $l'$ is the minimal possible such list.
2544: Moreover, in order that
2545: $P_{(x_1,y_1),\ldots,(x_r,y_r);l'} \not = 0$,
2546: all $x_i$ must be distinct and we must have
2547: $\{x_i\} = l$, so we
2548: may assume by symmetry that $x_i = l_i$.
2549: It is easy to check that the set of such $l'$ is not
2550: empty \iff $\forall i \in \{1,\ldots,r\},l_i<n+2i-2r$. In fact, if
2551: $l_i \geq n+2i-2r$, then the values $y_j$ and $x_j$ for
2552: $i \leq j \leq r$ must be distinct
2553: and between $n+2i-2r$ and $n$, a contradiction. Conversely, if
2554: $\forall i,l_i < n+2i-2r$, one may choose $y_i = l_i^*$.
2555:
2556: We now show that $l^*$ is indeed the minimal list. Let
2557: $[y_1,\ldots,y_r]$ be any list with
2558: $\forall i, y_i>l_i$ and $\forall i,j,y_i \not = l_j$.
2559: Let $(z_1,\ldots,z_r)$ be the corresponding ordered list
2560: (ie $\{y_1,\ldots,y_r\} = \{z_1,\ldots,z_r\}$ and
2561: $z_1 < z_2 < \cdots < z_r$). Then we have $z_1 > l_1$
2562: and $\forall j, z_1 \not = l_j$,
2563: so $z_1 \geq l_1^*$. Say $z_i = y_{\sigma(i)}$. If $z_1 < l_2$ then
2564: $\sigma(1)=1$. Thus in any case $z_2 > l_2$, so $z_2 \geq l^*_2$. By induction
2565: it follows that $\forall i,l^*_i \leq z_i$, so $l^*$ is the minimal possible
2566: list, and proposition \ref{prop_combinatoire} finishes the proof.
2567: \end{proo}
2568:
2569:
2570: We illustrate this proposition with two examples. The array below
2571: computes two dual varieties in $\G(3,8)$. It pictures the fact that
2572: for $l=(2,4,5)$ we have $l^*=\lambda=(3,6,7)$, and that for $m=(2,4,6)$ we have
2573: $m^*=\mu=(3,5,7)$~:
2574: $$
2575: \begin{array}{ccc}
2576: \young(\hfil l\lambda\hfil\hfil\hfil\hfil\hfil,\hfil\hfil\hfil l\hfil\lambda\hfil\hfil,\hfil\hfil\hfil\hfill l\hfil\lambda\hfil)
2577: &
2578: \hspace{1cm}
2579: &
2580: \young(\hfil m\mu\hfil\hfil\hfil\hfil\hfil,\hfil\hfil\hfil m\mu\hfil\hfil\hfil,\hfil\hfil\hfil\hfil\hfil m\mu\hfil)
2581: \end{array}
2582: $$
2583:
2584: \noindent
2585: Note that we have $C_l \subset C_m$ but we don't have
2586: $D_{l^*} \supset D_{m^*}$~: contrary to the case $G/P = \p V$, duality
2587: of Schubert cells is no longer decreasing.
2588:
2589:
2590:
2591: %******************************************************************************
2592:
2593:
2594:
2595: \subsectionplus{Case of spinor varieties}
2596:
2597: \label{subsection_spinoriel}
2598:
2599: Schubert cells in $\G_Q^+(2p+1,4p+2)$
2600: (resp. $\G_Q^-(2p+1,4p+2)$) are parametrised by lists of
2601: $+$ and $-$ signs, with an odd number of $+$ (resp. $-$) signs.
2602: The generic $T$-fixed point corresponding to the list
2603: $(\eta_i)$ is the subspace generated by $e_i^{\eta_i}$, and will be
2604: denoted $x_\eta$.
2605: Schubert cells are also parametrised by strict partitions of size $2p$
2606: (or subsets of $\{1,\ldots,2p\}$), the correspondance
2607: being that we set $x \in \lambda$ ($1 \leq x \leq 2p$)
2608: if $\eta_{2p+1-x} = -\ $.
2609:
2610: \begin{defi}\
2611: \begin{itemize}
2612: \item
2613: If $(\eta_i),1\leq i \leq 2p+1,$ is a sequence of signs and $j$
2614: is an integer, we denote
2615: $\varphi(\eta,j)$ the sequence $\eta'$ of signs such that
2616: $\eta'_i = \eta_i$ for exactly all $i$'s but $j$.
2617: \item
2618: A sequence $(\eta_i),1\leq i \leq 2p+1,$ of signs is admissible if
2619: $$
2620: \forall i\in \{1,\ldots,p\},
2621: \#\{j : 1 \leq j \leq 2i, \eta_j = + \} \geq i.
2622: $$
2623: Assume that $\eta$ is admissible~:
2624: \item
2625: If there exists $i\leq p+1$ such that
2626: $\# \{j\ :\ 1\leq j\leq 2i-1,\eta_j=+\} = i-1$, then let $i_0$ be the
2627: minimal such $i$, and set $\eta^* = \varphi(\eta,2i_0-1)$.
2628: \item
2629: Otherwise there exists $i$ such that
2630: $$
2631: \forall k\geq i,\ \#\{j:j\leq k,\eta_j=+\} > \# \{j:j\leq
2632: k,\eta_j=-\}.
2633: $$
2634: Let $i_0$ be the minimal such $i$ and set $\eta^* =
2635: \varphi(\eta,i_0)$.
2636: \end{itemize}
2637: \end{defi}
2638: \noindent
2639: If there does not exist $i\leq p+1$ such that
2640: $\# \{j\ :\ 1\leq j\leq 2i-1,\eta_j=+\} = i-1$,
2641: then $\# \{ j\ :\ 1 \leq j \leq 2p+1 \} \geq p+1$, so $i=2p+1$
2642: satisfies the condition of the last point of this definition.
2643:
2644: \lpara
2645:
2646: Let $\eta$ be fixed.
2647: Since the positive roots are $\epsilon_i \pm \epsilon_j$ with $i<j$,
2648: the restriction of the Bruhat order on the set of $\varphi(\eta,j)$
2649: is given by~:
2650:
2651: \begin{fait}
2652: \label{fait_ordre_spin}
2653: We have $\varphi(\eta,i) \leq \varphi(\eta,j)$ for the Bruhat order
2654: \iff
2655: $$
2656: \mbox{or }\left \{
2657: \begin{array}{l}
2658: \eta_i = \eta_j = + \mbox{ and } i\leq j\\
2659: \eta_i = + \mbox{ and } \eta_j = -\\
2660: \eta_i = \eta_j = - \mbox{ and } i \geq j.
2661: \end{array}
2662: \right .
2663: $$
2664: \end{fait}
2665:
2666:
2667: \begin{prop}
2668: \label{prop_spinoriel}
2669: $C_\eta$ is suitable \iff $\eta$ is admissible, and in this case
2670: $C_\eta^Q = D_{\eta^*}$.
2671: \end{prop}
2672: \begin{proo}
2673: Recall that $x_\eta \in G/P$ denotes the linear space spanned by
2674: $e_i^{\eta_i}$. It is well-known that $T_{x_\eta} G/P$ identifies
2675: with $\wedge^2 {Vect(e_i^{\epsilon_i})}^*$. Moreover,
2676: $N^*_{x_\eta} C_\eta \subset \wedge^2 {Vect(e_i^{\epsilon_i})}$
2677: is generated by $e_i^{\eta_i} \wedge e_j^{\eta_j}$ for $(i,j)$ such
2678: that $\eta_i = +$ and $i<j$. In fact, with the notations of
2679: \cite[PLANCHE IV]{bourbaki}, the weight of $x_\eta$ is
2680: $\rho_\eta = \frac 12 \sum \eta_i \epsilon_i$, and the weights of
2681: $T_{x_\eta} C_\eta$ are the weights of the form
2682: $\frac 12 \sum \eta'_i \epsilon_i$ which can be expressed as
2683: $\rho_\eta + \alpha$, where $\alpha$ is a positive root. Therefore the
2684: claim follows from the fact that the positive roots are
2685: $\epsilon_i \pm \epsilon_j$ with $i<j$.
2686:
2687: The weights of $T^*_{x_\eta} G/P \simeq \wedge^2 {Vect(e_i^{\epsilon_i})}$
2688: are parametrised by couples $(x,y)$ of integers, with $x<y$.
2689: Now let $x_k,y_k,1 \leq k \leq p,$ be integers with $x_k < y_k$. With
2690: the notations of subsection \ref{subsection_combinatoire}, $\mu$ is given by a
2691: set of polynomials of degree $p$ and the $p$-multilinear map
2692: $P_{(x_k,y_k);\eta'}$ does not vanish \iff the $x_k$'s and the $y_k$'s
2693: are all distinct and $\eta'_i = \eta_i$ for exactly all $i$'s which
2694: belong to the set $U := \{x_k\} \cup \{ y_k \}$.
2695:
2696: Given the previous description of
2697: $N^*_{x_\eta} C_\eta$, the Schubert
2698: variety $C_\eta$ will be suitable \iff we can find
2699: $(x_k,y_k)$ such that
2700: \begin{equation}
2701: \label{condition_spin}
2702: x_k < y_k, \mbox{ the }x_k,y_k \mbox{ are all distinct, and }
2703: \eta_{x_k} = +.
2704: \end{equation}
2705: Therefore, for all
2706: $i$'s with $1\leq i \leq p$, we have the inequality
2707: $$
2708: 2i-1 \leq \#(U \cap [1,2i]) \leq 2\# \{j\ :\ 1\leq j\leq 2i,\eta_j = +
2709: \}.
2710: $$
2711: This implies that $\eta$ should be admissible.
2712:
2713: Conversely, assuming that $\eta$ is admissible, let us consider the
2714: following algorithm which produces a list of distinct elements
2715: $(x_k,y_k)$ with $\eta_{x_k} = +$ and $x_k < y_k$.
2716: If $\forall i>1, \eta_i = +$, set $x_k = 2p$ and $y_k = 2p+1$.
2717: Otherwise, let $i_0$ be the minimal $i>1$ such that $\eta_i = -$;
2718: set $x_1 = i_0 - 1$ (the fact that $\eta$ is admissible garanties that
2719: even in the case $i_0=2$, we have $\eta_{i_0-1} = +$) and $y_1 =
2720: i_0$. Remove $\eta_{x_1}$ and $\eta_{y_1}$ from the list $\eta$ : this
2721: new list is again admissible, as one checks readily. Therefore, it is
2722: possible to define $(x_k,y_k)$ for $k \geq 2$ inductively.
2723:
2724: \lpara
2725:
2726: We therefore have proved that $C_\eta$ is suitable \iff $\eta$ is
2727: admissible. Let us now compute the dual variety. Assume first that
2728: there exists $i \in \{1,p+1\}$ such that
2729: \begin{equation}
2730: \label{condition}
2731: \# \{j : 1 \leq j \leq 2i-1, \eta_j = + \} = i-1.
2732: \end{equation}
2733: Let $i_0$ be the minimal such $i$. Admissibility of $\eta$ implies
2734: that $\# \{j\ :\ 1 \leq j \leq 2i_0-2, \eta_j = + \} = i_0 - 1$ and so
2735: $\eta_{2i_0-1} = -$. Therefore, if $(x_k,y_k)$ is any sequence
2736: satisfying (\ref{condition_spin})
2737: and $U = \{x_k\} \cup \{y_k\}$, there exists $j \leq 2i_0-1$
2738: such that $\eta_j = -$ and $j \not \in U$. Thus if there exists
2739: $(x_k,y_k)$ such that $P_{(x_k,y_k);\eta'} \not = 0$ for some $\eta'$,
2740: this implies $\eta' \geq \varphi(\eta,2i_0-1)$
2741: (recall fact \ref{fait_ordre_spin}).
2742: Conversely, the previous algorithm produces a sequence $(x_k,y_k)$ for
2743: which it is easy to see that
2744: $P_{(x_k,y_k);\varphi(\eta,2i_0-1)} \not = 0$. Thus
2745: $\eta^* = \varphi(\eta,2i_0-1)$ is the lowest
2746: list one can obtain in this way,
2747: so that $C_\eta^Q = D_{\eta^*}$ as claimed in this case.
2748:
2749: Assume finaly that (\ref{condition}) holds for no $i \in \{1,p+1\}$.
2750: Therefore, as we have seen, there exists
2751: $i$ (for example $i=2p+1$) such that
2752: $$
2753: \forall k \geq i, \# \{j\ :\ j \leq k,\eta_j = +\}
2754: >
2755: \# \{j\ :\ j\leq k,\eta_j = -\}.
2756: $$
2757: Let $i_0$ be the minimal such $i$. Obviously, if $i$ is any integer,
2758: and $(x_k,y_k)$ satisfies (\ref{condition_spin})
2759: and $x_k,y_k \not = i$, then
2760: $$\forall k \geq i, \# \{j\ :\ j \leq k,\eta_j = +,j\not = i\}
2761: \geq
2762: \# \{j\ :\ j\leq k,\eta_j = -,j \not = i\},$$ so that
2763: $i \geq i_0$. So $P_{(x_k,y_k);\eta'} \not = 0$ implies
2764: $\eta' \geq \varphi(\eta,i_0)$. Again, the explicit algorithm provides
2765: a sequence $(x_y,y_k)$ such that
2766: $P_{(x_k,y_k);\varphi(\eta,i_0)} \not = 0$, and therefore
2767: $D_{\varphi(\eta,i_0)}$ is the dual variety of $C_\eta$.
2768: \end{proo}
2769:
2770:
2771:
2772: \subsectionplus{Case of $E_{6,I}$}
2773:
2774: \label{subsection_e6_1}
2775:
2776:
2777: We now consider the exceptional cases when $G$ is of type
2778: $E_6$. Recall that there are two possibilities for $(P,Q)$~: either
2779: they correspond to the roots $(\alpha_1,\alpha_6)$ or
2780: $(\alpha_3,\alpha_5)$. In each case I explain in which case
2781: $P_{\tau_1,\ldots,\tau_d;\nu}$ does not vanish. Using subsection
2782: \ref{subsection_combinatoire}, this describes in principle all dual
2783: varieties to Schubert varieties, although I will not give a simple
2784: combinatorial recipy for this correspondance (note however that to
2785: give such a description there is ``only'' a finite number of computations to
2786: do). My description of which $P_{\tau_1,\ldots,\tau_d;\nu}$ don't
2787: vanish will however yield a simple caracterisation of the suitable Schubert
2788: varieties.
2789:
2790: As we have seen in subsection \ref{subsection_fondamental}, a Levi
2791: factor $L$ of $P_1$ is isomorphic with $\C^* \times Spin_{10}$, and
2792: $T_{[e]} E_6/P_1$ identifies with a
2793: $16$-dimensional spinor representation
2794: of $L$. Moreover, the closed $L$-orbit in $\p T^*_{[e]} G/P$
2795: identifies with a $L$-homogeneous spinor variety~: it is a connected
2796: component of the variety
2797: parametrising isotropic linear spaces of dimension 5 in a certain
2798: quadratic vector space of dimension 10 that we will denote $M$.
2799: It is proved in \cite[corollary 3.2]{flop_scorza}
2800: that $I_{[e]} \subset \p M$ is the
2801: corresponding 8-dimensional quadric acted upon by $L$
2802: and that the rational map
2803: $T^*_{[e]} E_6/P_1 \dasharrow I_{[e]}$
2804: is induced by the unique $L$-equivariant
2805: quadratic map $T^*_{[e]} G/P \rightarrow M$. The polarisation
2806: ${\cal P} : T^*_{[e]} E_6/P_1 \times T^*_{[e]} E_6/P_1 \rightarrow M$ of this
2807: equivariant map has the following geometric interpretation~: for
2808: $x,y \in T^*_{[e]} E_6/P_1$ representing points of the spinor variety
2809: corresponding to the isotropic linear spaces $L_x,L_y$, the class
2810: of ${\cal P}(x,y)$ in $\p M$ is the intersection of $L_x$ and $L_y$ if
2811: this intersection has dimension 1, and ${\cal P}(x,y)=0$ otherwise.
2812:
2813: Denote as in subsection \ref{subsection_spinoriel}
2814: $(e_1^+,\ldots,e_5^+,e_1^-,\ldots,e_5^-)$ a base of $M$ such that the
2815: quadratic form $Q$ satisfies
2816: $Q(\sum x_i^+ e_i^+ + \sum x_i^- e_i^-) = \sum x_i^+x_i^-$. An $L$-weight
2817: of $M$ can therefore be denoted
2818: $\nu \in \{1^+,\ldots,5^+,1^-,\ldots,5^-\}$, and a weight
2819: $\eta = (\eta_1,\ldots,\eta_5)$ in
2820: $T^*_{[e]} G/P$ corresponds to a list of plus or minus signs, with an
2821: odd number of plus signs.
2822: The condition for $P_{\eta,\eta';\nu}$ not to vanish is thus that
2823: $\eta$ and $\eta'$ have exactly one sign in common which is $\nu$.
2824:
2825: From this description we can describe the suitable Schubert varieties.
2826: In the array below, I recall the quiver of $E_6/P_1$ and
2827: I define an
2828: element $[w_{max}] \in W/W_P$ by its subquiver~:
2829:
2830: $$
2831: \begin{array}{ccc}
2832: \input{e6} && \input{w_max_1}\\
2833: \mbox{Quiver of }E_6/P_1 & \hspace{2cm} & \mbox{Subquiver of }[w_{max}]
2834: \end{array}
2835: $$
2836:
2837: \noindent
2838: We have the following proposition~:
2839:
2840: \begin{prop}
2841: \label{prop_e6_1_adapte}
2842: Let $[w] \in W/W_P$. Then the Schubert variety
2843: $C_{[w]}$ is suitable \iff $[w] \leq [w_{max}]$.
2844: \end{prop}
2845: \begin{proo}
2846: Below we give the Hasse diagram $H$ (resp. $H^*$) of the $L$-module
2847: $T_{[e]} E_6/P_1$ (resp. $T^*_{[e]} E_6/P_1$), which, by proposition
2848: \ref{hasse}, is isomorphic with the quiver of $E_6/P_1$~:
2849:
2850: $$
2851: \begin{array}{cc}
2852: \input{h} & \input {hd}\\
2853: \\
2854: H & H^*
2855: \end{array}
2856: $$
2857:
2858: Let $\iota : H \rightarrow H^*$ be
2859: induced by the map $\eta \mapsto -\eta$ (in terms of quivers, this
2860: corresponds to the obvious symmetry).
2861: Let $[w] \in G/P$ and $Q_{[w]} \subset H$ the corresponding subquiver, marking
2862: the weights of $w^{-1}.T_{[w]} C_{[w]}$.
2863: Thanks to proposition \ref{prop_combinatoire},
2864: the proposition amounts to the fact that
2865: we can find two weights $\eta,\eta' \in H^* - \iota(Q_{[w]})$
2866: and which have only one sign in common \iff
2867: $Q_{[w]} \subset Q_{[w_{max}]}$. This may be seen as follows~:
2868:
2869: \begin{itemize}
2870: \item
2871: If $Q_{[w]} \subset Q_{[w_{max}]}$, we can set
2872: $\eta = (++--+)$ and $\eta' = (+-++-)$ to check that the corresponding
2873: Schubert variety is suitable (below the subset $\iota(Q_{[w_{max}]})$
2874: is drawn).
2875: \item
2876: If $Q_{[w]}$ contains the vertex corresponding to the
2877: weight $(-+--+)\ $, $\iota(Q_{[w]})$ contains the subset drawn below.
2878: Thus $\eta$ and $\eta'$ are weights which begin with $++$, so
2879: they have two common signs. The corresponding Schubert variety
2880: is not suitable.
2881: \item
2882: The last case is that the subquiver contains the vertex corresponding
2883: to the weight $(--++-)$.
2884: Thus $\eta$ and $\eta'$ have at least 3 plus signs among the
2885: 4 first signs, and
2886: therefore have at least 2 common signs.
2887: \end{itemize}
2888:
2889: $$
2890: \begin{array}{ccccc}
2891: \input{cas0} && \input{cas1} && \input{cas2} \\
2892: \iota(Q_{[w_{max}]}) & \hspace{.5cm} &
2893: (-+--+) \in Q_{[w]} & \hspace{.5cm} & (--++-) \in Q_{[w]}
2894: \end{array}
2895: $$
2896:
2897: \end{proo}
2898:
2899: \begin{exem}
2900: \label{exem_p5e6}
2901: Let $X \subset E_6/P_1$ be a linear subspace of maximal dimension 5.
2902: Then $X^Q \subset E_6/P_6$ is also a linear subspace of dimension 5.
2903: \end{exem}
2904: \noindent
2905: Therefore, this provides, in our setting, a new example of a variety
2906: which is isomorphic to its dual. Similar examples in the usual setting
2907: $X \subset \p^n$
2908: are projective subspaces, quadrics,
2909: $\G(2,2p+1)$, and the spinor variety
2910: $\G_Q^+(5,10)$.
2911: \begin{proo}
2912: Let $X \subset E_6/P_1$ be a linear subspace of dimension 5.
2913: The variety parametrising linear subspaces of maximal dimension 5 is given by
2914: Tits shadows, according to \cite[theorem 4.3]{landsberg}.
2915: In particular, it is a
2916: homogeneous variety, and so we can assume that $X$ is the Schubert
2917: variety corresponding to the Weyl group element
2918: $w=s_6s_5s_4s_3s_1$. The corresponding quiver $Q_w$ and $\iota(Q_w)$
2919: follow; we have also drawn the quiver $Q_{w^*}$ of the dual variety.
2920: $$
2921: \begin{array}{ccccc}
2922: \input{p5} && \input{ip5} && \input{dp5}\\
2923: Q_w & \hspace{1cm} & \iota(Q_w) & \hspace{1cm} & Q_{w^*}
2924: \end{array}
2925: $$
2926:
2927: According to proposition \ref{prop_combinatoire}, we must look for two
2928: weights of $H^*$ not in $\iota(Q_w)$ which have only one common
2929: sign. If this common sign is a minus sign, then among these two
2930: weights there are 6 minus signs. Therefore one of the weight has 4
2931: minus signs which is impossible given $\iota(Q_w)$ and the Hasse
2932: diagram $H^*$. Since $(--+++)$ and $(++--+)$ are weights not in
2933: $\iota(Q_w)$, the lowest weight is $5^+$. Note that this weight is
2934: obtained applying $w_6 = s_3s_4s_5s_6$ to the highest weight.
2935: Therefore, proposition
2936: \ref{prop_combinatoire} shows that the dual variety to $X$ corresponds
2937: to the class of $w.w_6 = s_6s_5s_4s_3s_1s_3s_4s_5s_6$ modulo $W_6$.
2938: Modulo $W_6$, we have
2939: $$
2940: \begin{array}{cccccc}
2941: & s_6s_5s_4s_3s_1s_3s_4s_5s_6 & = & s_6s_5s_4s_1s_3s_1s_4s_5s_6 & = &
2942: s_6s_5s_4s_1s_3s_4s_5s_6s_1\\
2943: \equiv & s_6s_5s_1s_4s_3s_4s_5s_6 & = & s_6s_5s_1s_3s_4s_3s_5s_6 &
2944: \equiv & s_6s_5s_1s_3s_4s_5s_6\\
2945: = & s_6s_1s_3s_5s_4s_5s_6 & \equiv & s_6s_1s_3s_4s_5s_6 &
2946: \equiv & s_1s_3s_4s_5s_6.
2947: \end{array}
2948: $$
2949:
2950: This proves the claim.
2951: \end{proo}
2952:
2953: \subsectionplus{Case of $E_{6,II}$}
2954:
2955: \label{subsection_e6_2}
2956:
2957: Let $\alpha \in E_6/P_3$ be the base point.
2958: Recall that there is a surjection
2959: $\pi : T^*_\alpha E_6/P_3 \rightarrow
2960: L_\alpha \otimes \wedge^2 Q_\alpha^*$ and that the
2961: rational map $q:T^*_\alpha E_6/P_3 \dasharrow I_\alpha$ is induced by a
2962: rational map $\overline q : L_\alpha \otimes \wedge^2 Q^*_\alpha
2963: \dasharrow I_\alpha = \G(2,Q_\alpha^*)$.
2964:
2965: The description of $\overline q$ given in subsection
2966: \ref{subsection_fondamental} implies that
2967: $\overline q$ has degree 6. Consider as in subsection
2968: \ref{subsection_combinatoire} its polarisation, with
2969: coordinates denoted $P$. In order to give a non-vanishing
2970: criterium for $P$, let us introduce some notation.
2971: Let $e_1,e_2$ be a basis of $L_\alpha$ and
2972: $f_1^*,\ldots,f_5^*$ a basis of $Q_\alpha^*$.
2973: The weight of the
2974: vector $e_1 \otimes (f_i^* \wedge f_j^*)$ with $i<j$ will be denoted $ij$,
2975: and the weight of the vector $e_2 \otimes (f_i^* \wedge f_j^*)$ will be
2976: denoted $\underline{ij}$.
2977: Finally, the weight of
2978: $f_i^* \wedge f_j^*$
2979: will be denoted $ij^*$.
2980:
2981: Let $\tau_1,\ldots,\tau_6$ be weights of
2982: $L_\alpha \otimes \wedge^2 Q_\alpha^*$ and $\nu$ a weight of
2983: $\wedge^2 Q_\alpha^*$,
2984: if $P_{\tau_1,\ldots,\tau_6,\nu} \not = 0$, then
2985: $\# \{k:\tau_k \in \{ij\}\ \} =
2986: \# \{k:\tau_k \in \{\underline{ij}\}\ \} = 3$.
2987: So we assume that this is the case and that
2988: $\tau_1,\tau_2,\tau_3$ (resp. $\tau_4,\tau_5,\tau_6$) are of the form
2989: $ij$ (resp. $\underline{ij}$).
2990:
2991: With this setting,
2992: $P_{i_1j_1,i_2j_2,i_3j_3,
2993: \underline{k_1l_1},\underline{k_2l_2},\underline{k_3l_3}
2994: ;mn^*}$
2995: will not vanish if and only if, up to permuting the three first weights and the
2996: three last, we have that $i_1,j_1,i_2,j_2$ (resp. $k_1,l_1,k_2,l_2$)
2997: are all dinstinct; say they take all values in $\{1,\ldots,5\}$ except
2998: $u$ (resp. $v$). Moreover we must have $u \in \{k_3,l_3\}$ (resp.
2999: $v \in \{i_3,j_3\}$), say $\{k_3,l_3\} = \{u,u'\}$ (resp.
3000: $\{i_3,j_3\} = \{v,v'\}$). Finally, we must have $u' \not = v'$ and
3001: $\{u',v'\} = \{m,n\}$.
3002:
3003: In principle, this combinatorial rule describes dual varieties of
3004: Schubert cells in this case. However, as in the case of $E_{6,I}$, we
3005: can be more precise as far as suitability is concerned. The Hasse
3006: diagram $H$ of
3007: $L_\alpha^* \otimes \wedge^2 Q_\alpha$ is given below, as well as a subquiver
3008: denoted $Q_{max}$. I have also indicated the Hasse diagram $H^*$ of
3009: $L_\alpha \otimes \wedge^2 Q_\alpha^*$~:
3010:
3011: $$
3012: \begin{array}{ccccc}
3013: \input{hasse_e62} && \input{qmax} && \input{e6_2}\\
3014: H && Q_{max} && H^*
3015: \end{array}
3016: $$
3017:
3018: In these pictures and the following, the weights $ij$ are drawn in red
3019: and the weights $\underline{ij}$ are drawn in blue.
3020: Given a Schubert cell $C_{[w]}$, with $w$ the minimal length
3021: representative of $[w] \in W/W_P$, recall that we associated
3022: (not injectively) to this Schubert cell
3023: a subquiver $Q_{[w]}$ of $H$ in subsection \ref{subsection_carquois}.
3024:
3025: \begin{prop}
3026: \label{prop_e6_2_adapte}
3027: The Schubert cell $C_{[w]}$ is suitable \iff we
3028: have $Q_{[w]} \subset Q_{max}$.
3029: \end{prop}
3030: \begin{proo}
3031: As in the preceeding subsection, we define
3032: $\iota:H \rightarrow H^*$ given by
3033: $\eta \mapsto -\eta$.
3034: The weights of $H^*$ have been given above.
3035:
3036: Since $q$ is induced by $\overline q$,
3037: a Schubert variety $C_{[w]}$ will be suitable \iff
3038: the rational map
3039: $\overline q:L_\alpha \otimes \wedge^2 Q_\alpha^* \dasharrow I_\alpha$
3040: is defined generically on $\pi(w^{-1}.N^*_{[w]} C_{[w]})$; equivalently,
3041: $\overline q$ should be defined on the orthogonal of
3042: $w^{-1}.T_{[w]} C_{[w]} \cap L_\alpha^* \otimes \wedge^2 Q_\alpha$ in
3043: $L_\alpha \otimes \wedge^2 Q_\alpha^*$.
3044: Equivalently again, we should be able to
3045: find 6 weights $\tau_k$ not in $\iota(Q_{[w]})$ and some integers $i,j$
3046: such that $P_{\tau_k;ij^*}$ does not vanish.
3047:
3048: In case $Q_{[w]}$ is
3049: included in $Q_{max}$, we can consider the weights
3050: $34, 25, 34$, $\underline{15}$, $\underline{24}$, $\underline{15}$
3051: (the corresponding subset $\iota(Q_{max})$ is drawn below), which satisfy
3052: the relation
3053: $P_{34,25,34,\underline{15},\underline{24},\underline{15};45^*} \not
3054: = 0$ and do not belong to $\iota(Q)$. Otherwise, there are four cases~:
3055: \begin{itemize}
3056: \item
3057: If $Q_{[w]}$ contains the weight $15$, by proposition
3058: \ref{prop_order_ideal}, $\iota(Q_{[w]})$ contains the
3059: corresponding subset in the array below. The remaining weights are of
3060: the form $ij$ or $\underline{ij}$ with $1<i<j$, so the Schubert variety cannot
3061: be suitable.
3062: \item
3063: If $Q_{[w]}$ contains the weight $\underline{14}$,
3064: the remaining weights are of the form
3065: $ij$ or $\underline{kl}$ with $2<i<j$ (see the array below), so again
3066: the Schubert variety cannot be suitable.
3067: \item
3068: If $Q_{[w]}$ contains the weight $24$, let
3069: $i_1j_1,i_2j_2,i_3j_3,\underline{i_4j_4},\underline{i_5j_5},
3070: \underline{i_6j_6}$\vspace{1mm} be a list of weights not in $\iota(Q_{[w]})$.
3071: Note that we have
3072: $i_kj_k \in \{12,13,14,15,34\}$ for all $k$. Assume that there exists
3073: $kl^*$ such that
3074: $P_{i_1j_1,i_2j_2,i_3j_3,\underline{i_4j_4},\underline{i_5j_5},
3075: \underline{i_6j_6};kl^*} \not = 0$. This implies that the
3076: only integer $x$ (resp. $y$)
3077: which does not belong to the set $\{i_1,j_1,i_2,j_2\}$
3078: (resp. $\{i_4,j_4,i_5,j_5\}$) is
3079: either 2 or 5. This integer must
3080: therefore belong to the set $\{i_3,j_3\}$ (resp.
3081: $\{i_3,j_3\}$), so we must have $\{i_3,j_3\} = \{1,x\}$
3082: (resp. $\{i_6,j_6\} = \{1,y\}$). This implies that $k=l=1$, a
3083: contradiction.
3084: \item
3085: Assume finaly that $Q_{[w]}$ contains the weight $\underline{23}$.
3086: In this case all the
3087: weights which are not in $\iota(Q_{[w]})$
3088: and of the form $ij$ satisfy $j=5$. Again,
3089: the Schubert variety is not suitable.
3090: \end{itemize}
3091:
3092: $$
3093: \begin{array}{ccccc}
3094: \input{Cas0} & \hspace{.9cm} & \input{Cas1} & \hspace{.9cm} & \input{Cas2}\\
3095: \mbox{Case } \iota(Q_{max}) && \mbox{Case } {15} \in Q_{[w]} &&
3096: \mbox{Case } \underline{14} \in Q_{[w]}
3097: \end{array}
3098: $$
3099: $$
3100: \begin{array}{ccc}
3101: \input{Cas3} && \input{Cas4}\\
3102: \mbox{Case } {24} \in Q_{[w]} & \hspace{1cm} &
3103: \mbox{Case } \underline{23} \in Q_{[w]}
3104: \end{array}
3105: $$
3106:
3107:
3108: \end{proo}
3109:
3110:
3111:
3112:
3113: %******************************************************************************
3114:
3115: %******************************************************************************
3116:
3117: %******************************************************************************
3118:
3119:
3120:
3121: \begin{thebibliography}{99}
3122: \bibitem[Ar 99]{arrondo_severi} Arrondo, E.,
3123: {\it Projections of Grassmannians of lines and
3124: characterization of Veronese varieties.}
3125: J. Algebraic Geom. {\bf 8} (1999), no. 1, 85--101.
3126:
3127: \bibitem[AC 05]{arrondo_low} Arrondo, E., Caravantes, J.
3128: {\it On the Picard group of low-codimension subvarieties},
3129: math.AG/0511267.
3130:
3131: \bibitem[Bou 68]{bourbaki} Bourbaki, N., Groupes et alg\`ebres de Lie.
3132: Chapitre 4,5,6.
3133: Actualit\'es Scientifiques et Industrielles, No. {\bf 1337} Hermann, Paris 1968
3134:
3135: \bibitem[Ch 06]{flop_scorza} Chaput, P.E.,
3136: {\it On Mukai flops for Scorza varieties}, math.AG/0601734
3137:
3138: \bibitem[Ch 07]{hermitian} Chaput, P.E.,
3139: {\it Geometry over composition algebras~: hermitian geometry}, in
3140: preparation.
3141:
3142: \bibitem[CMP 06]{quantique} Chaput, P.E., Manivel, L., Perrin, N.
3143: {\it Quantum cohomology of minuscule homogeneous spaces},
3144: math.AG/0607492, to appear in Transformation Groups.
3145:
3146: \bibitem[De 96a]{debarre1} Debarre, O.,
3147: {\it Théorèmes de connexité pour les produits d'espaces
3148: projectifs et les grassmanniennes},
3149: Amer. J. Math. {\bf 118} (1996), no. 6, 1347--1367
3150:
3151: \bibitem[De 96b]{debarre2} Debarre, O.,
3152: {\it Sur un théorème de connexité de Mumford pour les espaces homogènes},
3153: Manuscripta Math. {\bf 89} (1996), no. 4, 407--425.
3154:
3155: \bibitem[Fa 81]{faltings} Faltings, G.
3156: {\it Formale Geometrie und homogene Räume},
3157: Invent. Math. {\bf 64} (1981), no. 1, 123--165.
3158:
3159: \bibitem[GKZ 94]{gkz} Gelfand, I. M.; Kapranov, M. M.; Zelevinsky, A. V.
3160: Discriminants, resultants, and multidimensional determinants.
3161: Mathematics: Theory and Applications. Birkhäuser Boston, Inc., Boston, MA, 1994
3162:
3163: \bibitem[He 78]{hesselink} Hesselink, W. H.
3164: {\it Polarizations in the classical groups.}
3165: Math. Z. {\bf 160} (1978), no. 3, 217--234.
3166:
3167: \bibitem[LM 03]{landsberg} Landsberg, J., Manivel, L.,
3168: {\it On the projective geometry of rational homogeneous varieties},
3169: Comment. Math. Helv. {\bf 78} (2003), no. 1, 65--100.
3170:
3171: \bibitem[McG 02]{mcgovern} McGovern, W. M.
3172: {\it The adjoint representation and the adjoint action.}
3173: Algebraic quotients. Torus actions and cohomology.
3174: The adjoint representation and the adjoint action, 159--238,
3175: Encyclopaedia Math. Sci., {\bf 131}, Springer, Berlin, 2002.
3176:
3177: \bibitem[Na 06]{namikawa} Namikawa, Y.,
3178: {\it Birational geometry of symplectic resolutions of nilpotent orbits},
3179: to appear in Advanced studies in pure mathematics {\bf 45},
3180: Moduli Spaces and Arithmetic Geometry (Kyoto 2004), p.75-116.
3181:
3182: \bibitem[Pe 05]{perrin_courbe} Perrin N., {\it Rational curves on
3183: minuscule Schubert varieties},
3184: J. Algebra {\bf 294} (2005), no. 2, 431--462.
3185:
3186: \bibitem[Pe 06]{carquois} Perrin N., {\it Small resolutions
3187: of minuscule Schubert varieties}, preprint
3188: arXiv:math.AG/0601117.
3189:
3190: \bibitem[Pe 07]{perrin_low} Perrin N., {\it Small codimension smooth
3191: subvarieties in even-dimensional homogeneous spaces with Picard
3192: group $\Z$}, arXiv:math.AG/0702578.
3193:
3194: \bibitem[Ri 74]{richardson} Richardson, R. W.,
3195: {\it Conjugacy classes in parabolic subgroups of semisimple algebraic groups.}
3196: Bull. London Math. Soc. {\bf 6} (1974), 21--24.
3197:
3198: \bibitem[Za 93]{zak} Zak, F.L.,
3199: Tangents and Secants of Algebraic Varieties.
3200: American Mathematical Society, Providence, RI
3201: (1993).
3202:
3203: \end{thebibliography}
3204:
3205: \vskip .2cm
3206:
3207: \noindent
3208: Author's address :
3209:
3210: \noindent
3211: Laboratoire de Math\'ematiques Jean Leray UMR 6629
3212:
3213: \noindent
3214: 2 rue de la Houssini\`ere - BP 92208 - 44322 Nantes Cedex 3
3215:
3216: \vskip .1cm
3217: \noindent
3218: Pierre-Emmanuel.Chaput@math.univ-nantes.fr
3219:
3220:
3221:
3222: \end{document}
3223:
3224:
3225:
3226: \lpara
3227:
3228: Let us now see that fundamental cases are natural
3229: and in fact linked with usual duality (of $G$-representations)~:
3230:
3231: \begin{prop}
3232: Let $P,Q \subset G$ be non-conjugated maximal parabolic subgroups.
3233: Then there is a unique
3234: invariant hypersurface in $G/P \times G/Q$ \iff $P,Q,G$ are one of the
3235: fundamental examples.
3236:
3237: In this case, if $L$ (resp. $M$) is the
3238: ample line bundle gnerating $Pic(G/P)$ (resp. $Pic(G/Q)$), then the
3239: $G$-representations $\Gamma(G/P,L)$ and $\Gamma(G/Q,M)$ are dual
3240: $G$-representations.
3241: \end{prop}
3242: \pr
3243: It is easy to check that if $P,Q\subset G$ is a fundamental example,
3244: then $\Gamma(G/P,L)$ and $\Gamma(G/Q,M)$ are dual
3245: $G$-representations. One also checks that the variety of orthogonal
3246: pairs in $G/P \times G/Q$ is the unique $G$-invariant hypersurface. In
3247: fact, the orbit of any non-orthogonal pair is dense.
3248: \lpara
3249:
3250: Conversely, assume $H \subset G/P \times G/Q$ is a $G$-invariant
3251: hypersurface. Then we have a $G$-equivariant map which sends a point
3252: $x \in G/P$ on the divisor $\{y \in G/Q,(x,y) \in H\}$. The class of
3253: this divisor in $Pic(G/Q)$ does not depend on $y$; denote it $D$. We therefore
3254: get a map from $G/P$ to $\p \Gamma(G/Q,{\cal O}(D))$. Let $d$ denote
3255: the integer such that $\co(D)$ is $d$ times the ample generator of $Pic(G/Q)$.
3256:
3257: Now, let $\omega_P$ (resp. $\omega_Q$) be the fundamental weight
3258: corresponding to $P$ (resp. $Q$). Call $V_P$ (resp. $V_Q$)
3259: the representation with
3260: highest weight $\omega_P$ (resp. $\omega_Q$)
3261: and $W$ the representation with highest
3262: weight $d.\omega_Q$. Recall the map
3263: $\varphi : G/P \rightarrow \p \Gamma(G/Q,{\cal O}(D)) = \p W^*$. This map is
3264: given by a $G$-invariant subspace
3265: $S\subset \Gamma(X,\varphi^* {\cal O}_{\p W^*}(1))$. There is an
3266: integer $e$ such that
3267: $\varphi^* {\cal O}_{\p W^*}(1) \simeq \co_X(e)$. Thus
3268: $\Gamma(X,\varphi^* {\cal O}_{\p W^*}(1)) = S^{(e)}V^*$
3269: is an irreducible $G$-module and has $\Gamma(Y,\co_{\p W^*}(1))$ and
3270: $W=\Gamma(\p W^*,\co_{\p W^*}(1))$
3271: as subrepresentations, if $Y$ denotes $\varphi(X)$. Therefore,
3272: $S = \Gamma(Y,\co_{\p W^*}(1)) = W$.
3273: So $S^{(d)} V_Q = W = S^{(e)} V_P^*$. It follows that
3274: $d=e$ and $V_P = V_Q^*$. From this, using the classification of the
3275: fundamental representations of simple groups, the proposition follows
3276: easily.
3277: \qed
3278:
3279: Preuve du théo sur les variétés duales quand on a une fibration :
3280:
3281: Let $z$ be any point in $\psi_Q^{-1}[\psi_P(x)]$.
3282: Let $\mathfrak{h}_x$ and $\mathfrak{q}_z \subset \g$ be the Lie
3283: subalgebras of $H_x$ and $Q_z$. Choose Levi subalgebras $\mathfrak{l,m}$
3284: of $\mathfrak{h}_x$ and $\mathfrak{q}_z$ such that
3285: $\mathfrak m \subset \mathfrak l$. Then
3286: $\u(\mathfrak q_z)=\u(\mathfrak h_x)
3287: \oplus \u(\mathfrak q_z \cap \mathfrak l)$.
3288: Therefore, a linear form $f \in \g^*$ vanishes on $\mathfrak q_z$ \iff
3289: it vanishes on $\mathfrak m$, $\u(\mathfrak{h}_x)$ and
3290: $\u(\mathfrak{q}_z \cap \mathfrak{l})$.
3291:
3292: Let $l \in \g^*$ be the image of a cotangent vector in $N^*_xX$. This
3293: implies that $l$ vanishes on $\mathfrak p_x$, and therefore on
3294: $\mathfrak m$ and $\u(\mathfrak{h}_x)$.
3295: Let $Q_z$ be the parabolic group
3296: contained in $H_x$ and which image in $L_x$ is the parabolic subgroup
3297: stabilizing the image of $l_{|T_xF_x}$ in $L_x/Q_y$.
3298: Let $\mathfrak q_z$ denote its
3299: Lie subalgebra. Since the image of $Q_z$ in $L_x$ corresponds to $l_{|T_xF_x}$,
3300: $l$ vanishes on $\u(\mathfrak{q}_z \cap \mathfrak{l})$.
3301: Therefore, it vanishes on $\mathfrak q_z$.
3302:
3303:
3304:
3305:
3306:
3307:
3308:
3309:
3310:
3311:
3312: preuve Nicolas :
3313:
3314: We will show
3315: that this is the case in general.
3316:
3317: Let $\Lambda$ denote the set of $H$-weights of $T_xX$ and let $\omega$
3318: denote the weight of $V$.
3319: We have an injection (?)
3320: $\varphi : [1,N] \rightarrow \Lambda,i\mapsto \omega - \alpha_i$,
3321: where the $G$-weight
3322: $\omega - \alpha_i$ is considered by restriction as an $H$-weight.
3323: Let $ \beta$ denote the only simple root such that
3324: $\langle \beta^v,\omega \rangle = 1$. Because $\alpha_i$ is a root
3325: and, by \cite[lemma 2.13]{perrin_courbe}, $\alpha_i \geq \beta$,
3326: $\omega - \alpha_i$ is a weight of $T_xX$ (?)
3327: Since $\dim T_xX = \dim X = N$ and all $H$-weights in $T_xX$
3328: have multiplicity 1, $\varphi$ is thus a bijection.
3329:
3330: According to lemmas \ref{fleche1} and \ref{fleche2}, $\varphi$
3331: induces an isomorphism of quivers between the quiver $Q_X$
3332: and the Hasse diagram of $T_xX$.
3333: \end{proo}
3334:
3335: \begin{lemm}
3336: \label{fleche1}
3337: If there is an arrow from $i$ to $j$ in the quiver $Q_X$, then
3338: $\alpha_j - \alpha_i$ is a simple root.
3339: \end{lemm}
3340: \begin{proo}
3341: Let $i,j$ be such that there is an arrow $i \rightarrow j$. Assume first
3342: that $j$ has a predecessor $p(j)$. We consider the quiver $Q'_X$ defined
3343: as in \cite{carquois}. Then
3344: according to proposition 4.9 in \cite{carquois}, the interval
3345: $[p(j),j]$ for the order induced by $Q_X$ is of the form
3346:
3347:
3348: \begin{pspicture*}(-.5,-.5)(4.50,10.50)
3349:
3350: \psline[linecolor=black](0.12,9.88)(0.88,9.12)
3351: \psline[linecolor=black](1.12,8.88)(1.88,8.12)
3352: \psline[linecolor=black](2.12,7.88)(2.88,7.12)
3353: \psline[linecolor=black](3.0,6.88)(3.0,6.12)
3354: \psline[linecolor=black](3.12,6.88)(3.88,6.12)
3355: \psline[linecolor=black](3.0,5.88)(3.0,5.12)
3356: \psline[linecolor=black](3.88,5.88)(3.12,5.12)
3357: \psline[linecolor=black](2.88,4.88)(2.12,4.12)
3358: \psline[linecolor=black](1.88,3.88)(1.12,3.12)
3359: \psline[linecolor=black](0.88,2.88)(0.12,2.12)
3360: \pscircle*(0.0,10.0){.17}
3361: \pscircle*(0.0,2.0){.17}
3362: \pscircle*(1.0,9.0){.17}
3363: \pscircle*(1.0,3.0){.17}
3364: \pscircle*(2.0,8.0){.17}
3365: \pscircle*(2.0,4.0){.17}
3366: \pscircle*(3.0,7.0){.17}
3367: \pscircle*(3.0,6.0){.17}
3368: \pscircle*(3.0,5.0){.17}
3369: \pscircle*(4.0,6.0){.17}
3370: \end{pspicture*}
3371:
3372: (with an arbitrary even number of vertices denoted $L$).
3373:
3374: We may moreover
3375: assume that $p(j)=i+L-1$. In fact, this is stated without proof as
3376: remark 4.10 in \cite{carquois}, and we sketch an argument for
3377: the convenience of the reader.
3378:
3379: Consider first the case $L=4$. We have four vertices, three of which are
3380: $p(j),i,j$. Let $k$ denote the other vertex. For $p(j) \leq t \leq j$
3381: and $t \not \in \{i,k\}$,
3382: the reflexions $s_{\beta_t}$ commute with $\beta_j$ because there are exactly
3383: two arrows coming from $p(j)$ in $Q'_X$ and these arrow point to $i,k$.
3384: Therefore, we may already assume that $j=i+1$ and $k=p(j)+1$.
3385:
3386: Let ${\cal D}$ denote the dynkin diagram of $G$; $\beta_i$ and
3387: $\beta_k$ are in two distinct connected components of ${\cal D}-\{\beta_j\}$.
3388: Since two reflexions corresponding to simple roots in different connected
3389: components of ${\cal D}-\{\beta_j\}$ commute, we may assume that there is an
3390: integer $u$ such that $\forall l \in ]k,u]$, $\beta_l$ is in the same
3391: connected component as $\beta_i$, and $\forall l \in ]u,i[$,
3392: $\beta_l$ is not. It follows that $\beta_l,l \in ]k,u]$ commute with
3393: $\beta_k$, and $\beta_l, l \in ]u,i[$ commute with $\beta_i$. So we are done
3394: in this case.
3395:
3396: In the general case, since again there are at most two arrows from a vertex
3397: in $Q'_X$, all the reflexions $s_{\beta_t}$ with $t \in ]p(j),j[$
3398: commute with the $L/2-1$ first reflexions in the interval $]p(j),j[$ for
3399: the quiver order. Therefore, the only difficulty is to commute
3400: with the middle reflexions, and the argument is analogous to the first case.
3401: Therefore, we deduce that
3402: $\alpha_j - \alpha_i = \alpha_{p(j)+1} - \alpha_{p(j)}$.
3403:
3404: \smallskip
3405:
3406: Assume now that $j$ has no predecessor. Then $i$ has also no predecessor.
3407: In fact, if $k$ is the predecessor of $i$, then by definition of $Q'_X$
3408: there is an arrow from $k$ to $j$, and from $i$ to $j$ in $Q'_X$. Therefore,
3409: $j$ is a hole in $Q'_X$ (according to definition 4.7 \cite{carquois}).
3410: This is a contradiction, because $Q'_X$ has no holes (this follows from
3411: \cite[proposition 4.11]{carquois}).
3412: Moreover, we can again assume that $j=i+1$.
3413: Therefore, direct computation shows that
3414: $\alpha_j-\alpha_i = \beta_j$, so it is a simple
3415: root.
3416: It follows by induction that $\alpha_j - \alpha_i$ is allways a
3417: simple root, so that the lemma is proved.
3418: \end{proo}
3419:
3420: \begin{lemm}
3421: \label{fleche2}
3422: If $\alpha_j-\alpha_i$ is a simple root, then there is an arrow from
3423: $i$ to $j$ in the quiver $Q_X$.
3424: \end{lemm}
3425: \begin{proo}
3426: Assume that $\gamma_i := \alpha_j - \alpha_i$ is a simple root.
3427: If $j$ has no predecessor, then
3428: there is a unique path from the first vertex in the quiver to $j$, so
3429: that all the roots $\alpha_t$ are either smaller or bigger than $\alpha_j$,
3430: according to $t < j$ or $t > j$ for the order in $Q'_X$.
3431: Therefore, there is a path from $i$ to
3432: $j$, actually an arrow.
3433:
3434: We now assume that $j$ has a predecessor. Since $Q'_X$
3435: is connected, there is an arrow from a vertex, say $k$, to $j$. If $k=i$,
3436: we are done, so we assume $k \not = i$.
3437: By lemma
3438: \ref{fleche1}, $\gamma_k := \alpha_j - \alpha_k$ is a simple root. Moreover,
3439: the equality $\gamma_i = \gamma_k$ would imply $\alpha_i = \alpha_k$, which
3440: is not the case. Set $\alpha = \alpha_j - \beta_i - \beta_k$. Since
3441: $s_{\beta_i}$ and $s_{\beta_k}$ commute (?), $\alpha$ is a root. Therefore,
3442: there exists an integer $l$ such that $\alpha = \alpha_l$. Now, by induction,
3443: we can assume that there is an arrow from $l$ to $i$ and from $l$ to $k$ in
3444: $Q'_X$. Since $j$ has a predecessor,
3445: \cite[proposition 4.9]{carquois} implies that there is an
3446: arrow from $i$ to $j$.
3447: \end{proo}
3448: