∎
equivariantly Hirzebruch homogènes irreducible
11institutetext: T. Mason 22institutetext: Department of Mathematical Sciences,
Georgia Southern University,
Statesboro, GA 30460
22email: [email protected]
33institutetext: F. Ziegler (corresponding author) 44institutetext: Department of Mathematical Sciences,
Georgia Southern University,
Statesboro, GA 30460
44email: f[email protected]
Explicit Pseudo-Kähler Metrics on Flag Manifolds
Abstract
The coadjoint orbits of compact Lie groups each carry a canonical (positive definite) Kähler structure, famously used to realize the group’s irreducible representations in holomorphic sections of appropriate line bundles (Borel-Weil theorem). Less studied are the (indefinite) invariant pseudo-Kähler structures they also admit, which can be used to realize the same representations in higher cohomology of the sections (Bott’s theorem). Using “eigenflag” embeddings, we give a very explicit description of these metrics in the case of the unitary group. As a byproduct we show that has exactly invariant complex structures, a count which seems to have hitherto escaped attention.
Keywords:
Coadjoint orbit Unitary group Pseudo-Kähler manifold Homogeneous complex manifold Flag manifoldMSC:
14M15 17B08 32M10 32Q15 53C50Contents
toc
1 Introduction
One of A. Borel’s early claims to fame was his discovery of a complex structure on the quotient of a compact Lie group by its maximal torus (B, 53, §29):
On the off chance, let us point out a property common to the and certain torsion free homogeneous spaces of (cf. §31, No. 1), but without knowing whether it is related to the topological question which occupies us;
admits a complex manifold structure invariant under the homeomorphisms from .
His original argument was that coincides with the quotient of the complexified group by what we now call a Borel subgroup,
(1) |
and the pointer to §31 referred to
(2) |
which he described as the manifold “of flags” whose generating element consists of nested subspaces of . Soon after, H. C. Wang showed that the property of admitting (finitely many) invariant complex structures is characteristic of homogeneous spaces of the form
(3) |
with compact and the centralizer of a torus W (54); Borel observed that these spaces admit invariant Kähler metrics B (54), and with Hirzebruch, gave the theory its classic exposition (B, 58, §§12–13).
Notably, this contains a section 13.7 Number of invariant complex structures which actually only gives the count in the two extreme cases where or , i.e. (in (2)) or .
The purpose of this paper is to fill this gap and give as explicit a description as possible of all invariant complex structures in the case of the unitary group. (Extension to other types is a seemingly intricate problem.) An ulterior motive is to get a concrete handle on Bott-Borel-Weil modules for various purposes in representation theory; for example, our results should afford a constructive proof of Bott’s theorem in the spirit of T (14).
Paper organization and terminology
The multi-faceted nature of flag manifolds has led different authors to different choices of a working definition, with different connotations. For instance (1) singles out a complex structure, and all of (1–3) a base point, foreign to the nested subspaces definition. In this paper we follow B (87) and work with
a coadjoint orbit of the compact Lie group () | (4) |
(§2). This is base-point free but singles out a symplectic form (of Kirillov-Kostant-Souriau).111Thus the historical order is reversed, in which one finds first complex structures, then metrics (Fubini-Study S (05), Cartan-Ehresmann C (29); E (34)), and only finally symplectic forms. In that setting turns out to have a preferred complex structure and attendant Kähler metric (§3), and classifying other invariant complex structures is tantamount to classifying compatible pseudo-Kähler metrics (§4). Embeddings into products of Grassmannians then allow us to give for these the very explicit formulas we are after (§5).
We note that Theorems 3.1 to 4.2 are well-known and indeed easily generalized to any compact connected in (4): see (B, 82, Exerc. 4.8 and 6.13). For notational simplicity and uniformity, we resisted the temptation to present them in more generality than Theorems 4.3 to 5.2, which we emphatically do not know how to extend beyond type A.
As recently pointed out by G. Nawratil N (17), the idea of nested subspaces and the name flag itself originate with R. de Saussure (in Esperanto! Fig. 1). They were used only sporadically, mainly by associates of projective geometer H. de Vries W (11), (W, 36, p. 15), (F, 49, p. 22), (F, 69, p. 415), until A. Borel revived them in his Thesis.
1.2cm
2 The coadjoint orbits
2.1 The unitary group and its complexification
Throughout this paper will denote the group of unitary matrices in , and will always mean the adjoint (a.k.a. complex conjugate transpose) of any row, column or matrix . The Lie algebra splits as the sum of the skew-adjoint and the self-adjoint:
(5) |
where .
2.2 The trace form and duality
We write for the symmetric, complex bilinear form defined by
(6) |
This form is -invariant, i.e. it satisfies and infinitesimally
(7) |
where and . These formulas hold for all and , and we have
Proposition 1
The restriction is real-valued, real bilinear, -invariant and positive definite. This allows us to identify with (and hence with ) so that duality and the coadjoint action read, for ,
(8) |
(The restriction has the same properties, except it is negative definite.)∎
2.3 The orbits
A coadjoint orbit is an orbit of the action (8) of on , or in other words, a conjugacy class of self-adjoint matrices. Since such matrices have real eigenvalues and an orthonormal basis of eigenvectors, we have
Proposition 2
Each orbit meets, exactly once, the dominant Weyl chamber
(9) |
consisting of nonincreasing real diagonal matrices.∎
Here denotes the scalar matrix of a certain size , i.e. we are lumping equal eigenvalues together: while the map is nominally , it is constant on the members of a partition
(10) |
of into consecutive segments whose cardinalities are the ; hence it induces a map which we write again .
Example
For as in (36) below, .
2.4 The stabilizer and its center
Proposition 3
We note that , where is the maximal torus of all diagonal matrices in , and equality holds when all (nondegenerate eigenvalues). Again the trace form (6) allows us to identify with ; under this identification, the projections
(12) |
consist in taking the diagonal part, resp. the block average
(13) |
2.5 The tangent space and its complexification
Under the identifications of Proposition 1, the last formula in (8) says that the tangent space to at is the image of the map
(14) |
As this image sits in the real part of , we can complexify it “in place” as
(15) |
And as (14) is skew-adjoint (see (7)), its image is the orthogonal of its kernel relative to , i.e. we have
Proposition 4
(16) |
Remark 1
When is (or , or ), coadjoint orbits are just 2-spheres. Then (16) is the statement that the tangent space at a point is the orthogonal to the axis of rotations around that point (Fig. 2). Counterclockwise rotation by provides one of the complex structures we are about to describe.
2cm
3 The canonical complex structure
Let be the coadjoint orbit with dominant element , as in (9). The restriction of (14) to its tangent space (16) has kernel , hence is a (still skew-adjoint) linear bijection we shall denote
(17) |
Recall that carries the Kirillov-Kostant-Souriau (KKS) -form , defined by .
Theorem 3.1
The KKS -form of is given by
(18) |
Moreover the formulas
(19) |
where make part of a -invariant Kähler structure :
-
(a)
is an (integrable) complex structure,
-
(b)
is a positive definite metric,
-
(c)
we have and .
Proof
Fix and put . Then (17, 14, 8) give
(20) |
whence the definition of the KKS 2-form and (8) give us (18):
(21) |
Next we note that and are the (commuting) positive definite and unitary part of the polar decomposition of . So they depend smoothly on (S, 70, 6.70) and , being again skew-adjoint, is an almost complex structure: . Now (c) is clear by plugging into (18, 19), and so is (b) since is negative definite. There remains to see (a). For a -invariant , such as ours is by construction, this is equivalent to either of
-
[20]
-
(22)
sections of the bundle of +i-eigenspaces of in are closed under Lie bracket (Frobenius-Newlander-Nirenberg N (57));
- (23)
We prove ((23)). First observe that if and are eigenvectors of for eigenvalues and , then the matrix is an eigenvector of for eigenvalue :
(24) |
It follows that is diagonalizable with spectrum , and so is with spectrum . And indeed explicitly is “diagonal” with eigenvectors the elementary matrices : in more detail, writing tangent vectors (16) as self-adjoint matrices with blocks in the shape (11), formula (24) gives
(25) |
(the general pattern should be clear, though we only write out the case where the partition (10) is into 3 segments ). Hence we obtain by definition of and that the latter is the same as (25) with each divided by its modulus, i.e.
(26) |
Thus we see that the +i-eigenvectors of , and likewise their preimages under (14) or , are the block upper triangular matrices — hence a Lie subalgebra in .∎
Remark 2
Remark 3
Remark 4
The idea of using the polar decomposition to produce (“tamed”) almost complex structures occurs in a general context in (W, 77, p. 8); its application to obtain this one seems new. Other, less direct descriptions of are found in (S, 54, §2), (B, 54, §4), (B, 58, 14.6), (G, 82, p. 522), (B, 87, 8.34), (V, 87, 5.8).
3.1 The case of Grassmannians
Let be the Grassmannian of complex -planes in , each identified with the self-adjoint projector upon it, i.e.
(28) |
Its dominant element is the highest weight of the fundamental -module .
Proposition 5
In this case we have so that the canonical structure of is simply
(29a) | |||||
(29b) | |||||
. | (29c) |
Proof
Deriving and reusing the relations gives . This implies and . So and hence its square root are the identity.∎
Remark 5
The Hermitian metric in (29c) can be seen as Kähler reduction of the flat metric on .222Alternatively on the dual , if we insist on obtaining (29a) and not its opposite . It would be interesting to know if the of Theorem 5.2 can be similarly obtained by (pseudo-)Kähler reduction. Indeed acts there by , preserving with moment map , and (29c) obtains on passing to the quotient (G, 73, §V.5), (T, 06, p. 240). E.g. for one recovers the Fubini-Study metric on projective space, i.e. (S, 05, §5)
(30) |
Formulas (29c) are emblematic of the explicitness we’d like to have in general.
4 The invariant complex structures classified: parabolic subalgebras
In this section we review the classification of complex structures which results from the principle: a -invariant structure on amounts to an -invariant , squaring to . The results are well-known except perhaps Theorem 4.3.
4.1 The decomposition of the isotropy representation
Let denote, for segments in the partition (10), the matrices (25) whose blocks all vanish except perhaps , i.e.
(31) |
and (resp. ) the intersection of (resp. ) with .
Theorem 4.1
The isotypic decomposition of the isotropy representation of in the complexified tangent space (15) at into inequivalent irreducibles is
(32) |
Consequently,
-
(a)
Every -invariant almost complex structure on is obtained by flipping the sign of (and hence ) on some summands in .
-
(b)
As coincides with on , each such flip affects its signature by turning a block of pluses into minuses.
-
(c)
If has segments, then admits different -invariant almost complex structures.
Proof
Using the notation of (11) and (25), one checks without trouble that the isotropy action of takes block of to
(33) |
So the are -invariant and the representation on each factors through the natural representation of on . As these are irreducible and different for different pairs , we obtain (32). Now is determined by its i-eigenspaces
(34) |
which are (complex conjugate) -invariant subspaces of (32), hence are each the sum of some (B, 12, Prop. 4.4d) — one per pair . So they can only differ from those of (26) by the indicated sign flips, and we obtain (a, b, c).∎
4.2 The invariant complex structures
There remains to characterize which of the almost complex structures of Theorem 4.1 are integrable.
Theorem 4.2
We have
if ; | (35a) | ||||
if ; | (35b) | ||||
if ; | (35c) | ||||
if ; . | (35d) |
Consequently,
-
(a)
An almost complex structure obtained as in Theorem 4.1a is integrable iff it respects the Chasles rule: if the sign is flipped on and , then it is also flipped on .
- (b)
Proof
4.3 Parabolic subalgebras with a given Levi component
Theorem 4.2b reduces the classification of invariant complex structures on to describing the set of parabolic subalgebras of whose Levi component is (see (11); this set is discussed in e.g. (A, 81, p. 8), (D, 11, §5)). We claim:
Theorem 4.3
is in natural bijection with the symmetric group .
Proof
We describe the construction of the bijection in general and illustrate it on the case where in (9) has eigenvalues with multiplicities , say :
(36) |
Let a permutation be given. Regard it as acting on the letters and rearrange the blocks of (9) accordingly, obtaining here e.g.
(37) |
Next, form the matrix whose columns are the standard basis vectors in the order that indices appear in : in our case
(38) |
This is by construction a (“uniform block”) permutation matrix such that A (08); T (61). Now let be the -conjugate of block upper triangular matrices of shape (37), i.e. (with both s and s denoting arbitrary entries)
(39) |
This is clearly a subalgebra of the form required by Theorem 4.2, i.e. obtained by sign flips from the block upper triangular decoration of (36) (see (26)).
Conversely, let be given — e.g. the one in (39). It is a parabolic containing (11), with half all off-diagonal blocks marked after Theorem 4.1a. Now collapse all blocks to size : becomes a Borel containing the diagonals. By (C, 57, Cor. 3), is conjugate to the upper triangular Borel by a unique permutation matrix , which is the one we attach to .
One checks without trouble that the maps and thus defined are each other’s inverse.∎
Remark 7
The cases and of Theorem 4.3 are due to Borel and Hirzebruch, who observed that all s are then related by the action of complex conjugation () or the Weyl group () (B, 58, 13.8), (B, 82, Exerc. 4.8e). But in general our bijection does not arise from a geometrical action of on . In fact, as stated in (B, 58, p. 506) and detailed in (N, 84, p. 44), any diffeomorphism transforming one invariant complex structure into another must come from the natural action, , of some belonging to the stabilizer of in the automorphism group
(40) |
Here is inner automorphisms and is the effect of complex conjugation; see (B, 82, Exerc. 4.3), (S, 01, Thm 1.5). As preserves , and preserves iff is in the normalizer , and () preserves any -invariant , we see that things boil down to an action of
(41) |
The Weyl-like quotient is computed in (M, 11, Cor. 12.11) and isomorphic to the subgroup of those that send each segment of the partition (10) to a same-sized segment, modulo the that take each segment to itself. When all segments have different sizes, that is trivial and so (41) is far from able to account for all structures.
Remark 8
Extending Theorem 4.3 to compact groups of other types seems challenging, which may explain its apparent absence from the literature. The role of should presumably be taken over by a putative Weyl “group” of either the quotient systems of (L, 04, 12.18) or the -root systems of A (86, 97, 98, 03); K (10) (their is our from (3, 12)). One would also need to generalize the rather mysterious (to us) map .
4.4 Example: The adjoint variety
|
|
|
|
|||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|
The orbit with dominant element (36) we have used as a running example is the adjoint variety , studied in (B, 58, 13.9), B (61); K (98); L (02); H (05). Table 1 traces the construction of the entire bijection in this case. Note how
Remark 9
A dominant with multiplicities can of course lead also to metrics of signature or on the “same” manifold : signature depends not only on but also on the chosen (or coadjoint orbit).
5 The invariant complex structures realized: eigenflag embeddings
Theorem 4.2 only spells out complex structures by giving the effect of at the base point . At any other point , computation of requires use of some , on whose nonuniqueness the outcome is known not to depend. Our goal below is a more tangible picture where can be explicit in terms of alone, as in (19, 27, 29a). We freely use the notation introduced in (9–10, 28–29c).
5.1 Maps to products of Grassmannians
A first idea is to note that spectral decomposition expresses each as a linear combination of eigenprojectors, , belonging to the (fixed) eigenvalues :
(43) |
(Lagrange interpolation (H, 71, §6.7)). So sending to embeds -equivariantly as a submanifold of a product of Grassmannians (28), hopefully pulling product structures back to useful ones on . Alas, Theorem 5.1 below dashes this hope: isn’t a complex submanifold of the product, so there is no complex structure to transport back. Fortunately, the same Theorem will also indicate the way out.
To state it, note that the are just a small part of ’s spectral measure which maps subsets of (or alternatively, of the spectrum ) to projectors
(44) |
with the property that (so the all commute). Thus, not only the singletons but any subfamily gives rise to a -equivariant map, , from to a product of Grassmannians.
Theorem 5.1
The image of this map is a complex submanifold of (for the product complex structure) iff is totally ordered by inclusion.
Proof
First note that as is transitive on , the map’s equivariance (visible on (43)) ensures that is an orbit of a smooth group action, hence as always an (“initial”) submanifold (H, 12, Prop. 10.1.14).
Assume that is totally ordered by inclusion. Then a tuple in is a member iff it satisfies
(45) |
for all pairs in (the reverse order follows by taking adjoints); and a tangent vector is in iff we also have the derived relation
(46) |
Assume (46). Multiplying it on the left by gives and hence
(47) | ||||
Thus we see that also satisfies (46). This confirms that the product complex structure preserves .
Conversely, assume that is not totally ordered. So there are such that and . Pick and and nonzero eigenvectors for eigenvalues of ; thus we have
(48) |
Now put and consider the image of . By equivariance and (48), its components in and are respectively
(49) |
They (of course) satisfy the relation which any tangent vector to must, as one sees by deriving . On the other hand, we claim that and fail that relation. Indeed (29a) gives
(50) |
whence (using (49))
(51) | ||||
Thus the product complex structure fails to preserve , as claimed.∎
5.2 The eigenflag embeddings
Choosing in Theorem 5.1, we obtain our main result which provides
- •
-
•
for other , explicit models of with every pseudo-Kähler structure:
Theorem 5.2
Let give rise to complex structure and metric (Theorems 4.2, 4.3) and write where is the partition (10). Then the coadjoint orbit with pseudo-Kähler structure is isomorphic to the orbit of in endowed with the product complex structure and the metric and -form
(52) |
where is the Grassmannian (28, 29c) and we set . The (moment) map from to and inverse map from to are respectively, with defined by (43, 44),
(53) |
Proof
Formula (52) defines on the product a -form which is clearly symplectic and -invariant with moment map given by (53). Its restriction to is a priori presymplectic with moment map still given by (53). Equivariance ensures that maps onto a coadjoint orbit, which is since summation by parts gives (37).
An easy dimension count, or indeed the explicit inverse in (53), then shows that is a diffeomorphism which is symplectic by (S, 70, 11.17). There remains to see that the derivative of maps (the +i-eigenspace of) the product complex structure at to (the +i-eigenspace of) at the base point . But this boils down to the observation that linear combination takes the block upper triangular matrices in to block upper triangular matrices in (39).∎
Remark 10
It seems natural to refer to as an eigenflag of the corresponding matrix . Thus we have as many “eigenflag embeddings” of as there are orderings of its eigenvalues, and each induces a different complex structure. Note that by the observation made before (45), is algebraic in with equations .
5.3 Example: The adjoint variety (continued)
Table 2 details all embeddings when is the adjoint variety (§4.4) with ; the singleton could of course be mostly omitted from the notation. Taking the last row as an example, the signature metric is
(54) |
and gives with the product complex structure .
|
|
manifold |
|
||||||
---|---|---|---|---|---|---|---|---|---|
Acknowledgements.
We wish to thank Arnaud Beauville, Ivan Penkov, Jacqueline Rey-Glardon, Loren Spice and Alan Weinstein for very helpful indications.References
- \hyper@normalise
- A (08) Marcelo Aguiar and Rosa C. Orellana, The Hopf algebra of uniform block permutations. J. Algebraic Combin. 28 (2008) 115–138. \hyper@normalise
- A (86) Dmitri V. Alekseevskii and Askold M. Perelomov, Invariant Kähler-Einstein metrics on compact homogeneous spaces. Funktsional. Anal. i Prilozhen. 20 (1986), no. 3, 1–16, 96. (Translation: Funct. Anal. Appl. 20 (1986) 171–182.) \hyper@normalise
- A (97) Dmitri V. Alekseevskii, Flag manifolds. Zbornik Radova (N.S.) 6(14) (1997) 3–35. \hyper@normalise
- A (98) , Isotropy representation of flag manifolds. Rend. Circ. Mat. Palermo (2) Suppl. 54 (1998) 13–24. \hyper@normalise
- A (81) James Arthur, The trace formula in invariant form. Ann. of Math. (2) 114 (1981) 1–74. \hyper@normalise
- A (03) Andreas Arvanitoyeorgos, An introduction to Lie groups and the geometry of homogeneous spaces, Student Mathematical Library, vol. 22. Amer. Math. Soc., Providence, RI, 2003. \hyper@normalise
- B (87) Arthur L. Besse, Einstein Manifolds. Springer-Verlag, Berlin, 1987. \hyper@normalise
- B (61) William M. Boothby, Homogeneous complex contact manifolds. Proc. Symp. Pure Math. 3 (1961) 144–154. \hyper@normalise
- B (53) Armand Borel, Sur la cohomologie des espaces fibrés principaux et des espaces homogènes de groupes de Lie compacts. Ann. of Math. (2) 57 (1953) 115–207. \hyper@normalise
- B (54) , Kählerian coset spaces of semisimple Lie groups. Proc. Nat. Acad. Sci. U. S. A. 40 (1954) 1147–1151. \hyper@normalise
- B (58) Armand Borel and Friedrich Hirzebruch, Characteristic classes and homogeneous spaces. I. Amer. J. Math. 80 (1958) 458–538. \hyper@normalise
- B (75) Nicolas Bourbaki, Groupes et algèbres de Lie. Chapitres 7 et 8. Hermann, Paris, 1975. \hyper@normalise
- B (82) , Groupes et algèbres de Lie. Chapitre 9. Groupes de Lie réels compacts. Masson, Paris, 1982. \hyper@normalise
- B (12) , Algèbre. Chapitre 8. Modules et anneaux semi-simples. Springer-Verlag, Berlin, 2012. \hyper@normalise
- C (29) Élie Cartan, Sur les invariants intégraux de certains espaces homogènes clos et les propriétés topologiques de ces espaces. Ann. Soc. Polon. Math. 8 (1929) 181–225. \hyper@normalise
- C (57) Claude Chevalley, Le normalisateur d’un groupe de Borel. In Séminaire Claude Chevalley. Classification des groupes de Lie algébriques, Vol. 1, chap. 9, pp. 1–8. Secrétariat mathématique, Paris, 1957. \hyper@normalise
- D (11) Elizabeth Dan-Cohen and Ivan Penkov, Levi components of parabolic subalgebras of finitary Lie algebras. In Jeffrey Adams, Bong Lian, and Siddhartha Sahi (Eds.), Representation Theory and Mathematical Physics (New Haven, October 24–27, 2009), Contemp. Math., vol. 557, pp. 129–149. Amer. Math. Soc., Providence, RI, 2011. \hyper@normalise
- E (34) Charles Ehresmann, Sur la topologie de certains espaces homogènes. Ann. of Math. (2) 35 (1934) 396–443. \hyper@normalise
- F (49) Hans Freudenthal, La géométrie énumérative. In Topologie algébrique (Paris, 26 Juin – 2 Juillet 1947), Colloques Internationaux du Centre National de la Recherche Scientifique, vol. 12, pp. 17–33. C. N. R. S., Paris, 1949. \hyper@normalise
- F (69) Hans Freudenthal and Hendrik de Vries, Linear Lie Groups. Academic Press, New York-London, 1969. \hyper@normalise
- F (55) Alfred Frölicher, Zur Differentialgeometrie der komplexen Strukturen. Math. Ann. 129 (1955) 50–95. \hyper@normalise
- G (60) Roger Godement, Groupes linéaires algébriques sur un corps parfait. In Séminaire Bourbaki, Vol. 6, Exp. No. 206, pp. 1–22. Secrétariat mathématique, Paris, 1960. \hyper@normalise
- G (73) Werner Greub, Stephen Halperin, and Ray Vanstone, Connections, curvature, and cohomology. Vol. II: Lie groups, principal bundles, and characteristic classes. Academic Press, New York-London, 1973. \hyper@normalise
- G (82) Victor Guillemin and Shlomo Sternberg, Geometric quantization and multiplicities of group representations. Invent. Math. 67 (1982) 515–538. \hyper@normalise
- H (12) Joachim Hilgert and Karl-Hermann Neeb, Structure and Geometry of Lie Groups. Springer, New York, 2012. \hyper@normalise
- H (05) Friedrich Hirzebruch, The projective tangent bundles of a complex three-fold. Pure Appl. Math. Q. 1 (2005) 441–448. \hyper@normalise
- H (71) Kenneth Hoffman and Ray A. Kunze, Linear Algebra. 2nd ed. Prentice-Hall Inc., Englewood Cliffs, N.J., 1971. \hyper@normalise
- K (98) Hajime Kaji, Secant varieties of adjoint varieties. Matemática Contemporânea 14 (1998) 75–87. \hyper@normalise
- K (10) Bertram Kostant, Root systems for Levi factors and Borel-de Siebenthal theory. In H. E. A. (Eddy) Campbell, Aloysius G. Helminck, Hanspeter Kraft, and David Wehlau (Eds.), Symmetry and spaces, pp. 129–152. Birkhäuser, Boston, 2010. \hyper@normalise
- L (02) Joseph M. Landsberg and Laurent Manivel, Construction and classification of complex simple Lie algebras via projective geometry. Selecta Math. (N.S.) 8 (2002) 137–159. \hyper@normalise
- L (04) Ottmar Loos and Erhard Neher, Locally Finite Root Systems. Mem. Amer. Math. Soc., vol. 171 (2004), x+214 pp. \hyper@normalise
- M (11) Gunter Malle and Donna Testerman, Linear Algebraic Groups and Finite Groups of Lie Type. Cambridge University Press, Cambridge, 2011. \hyper@normalise
- N (17) Georg Nawratil, Point-models for the set of oriented line-elements – a survey. Mechanism and Machine Theory 111 (2017) 118–134. \hyper@normalise
- N (57) August Newlander and Louis Nirenberg, Complex analytic coordinates in almost complex manifolds. Ann. of Math. (2) 65 (1957) 391–404. \hyper@normalise
- N (84) Musubi Nishiyama, Classification of invariant complex structures on irreducible compact simply connected coset spaces. Osaka J. Math. 21 (1984), no. 1, 39–58. \hyper@normalise
- S (08) René de Saussure, La Geometrio “Folietara”. Internacia Scienca Revuo 5 (1908) 165–172, 197–208, 236–243. (Translation: Mém. Soc. phys. hist. nat. Genève 36 (1910) 211–266.) \hyper@normalise
- S (54) Jean-Pierre Serre, Représentations linéaires et espaces homogènes kählériens des groupes de Lie compacts (d’après Armand Borel et André Weil). In Séminaire Bourbaki, Vol. 2, Exp. No. 100, pp. 1–8. Secrétariat mathématique, Paris, 1954. \hyper@normalise
- S (01) Krishnan Shankar, Isometry groups of homogeneous spaces with positive sectional curvature. Differential Geom. Appl. 14 (2001) 57–78. \hyper@normalise
- S (69) Jean de Siebenthal, Sur certains modules dans une algèbre de Lie semisimple. Comment. Math. Helv. 44 (1969) 1–44. \hyper@normalise
- S (70) Jean-Marie Souriau, Structure des systèmes dynamiques. Dunod, Paris, 1970. (Reprint: Éditions Jacques Gabay, Sceaux, 2008. Translation: Structure of Dynamical Systems. Birkhäuser, Boston, 1997.) \hyper@normalise
- S (05) Eduard Study, Kürzeste Wege im komplexen Gebiet. Math. Ann. 60 (1905) 321–378. \hyper@normalise
- T (61) Olga Taussky, Commutators of unitary matrices which commute with one factor. J. Math. Mech. 10 (1961) 175–178. \hyper@normalise
- T (06) Richard P. Thomas, Notes on GIT and symplectic reduction for bundles and varieties. In Shing-Tung Yau (Ed.), Surveys in Differential Geometry, Vol. X, pp. 221–273. Int. Press, Somerville, MA, 2006. \hyper@normalise
- T (14) Kostiantyn Timchenko, A Constructive Proof of the Borel-Weil Theorem for Classical Groups. Master’s thesis, Georgia Southern University, Statesboro, 2014. \hyper@normalise
- V (90) Èrnest B. Vinberg, Vladimir V. Gorbatsevich, and Arkadii L. Onishchik, Structure of Lie Groups and Lie Algebras, Itogi Nauki i Tekhniki, vol. 41. VINITI, Moscow, 1990. (Translation: Encycl. Math. Sci., vol. 41. Springer-Verlag, Berlin (1994).) \hyper@normalise
- V (87) David A. Vogan, Jr, Representations of reductive Lie groups. In Proc. Internat. Congr. Math. (Berkeley, 1986) Vol. 1, pp. 245–266. Amer. Math. Soc., Providence, RI, 1987. \hyper@normalise
- W (36) Bartel L. van der Waerden, Reihenentwicklungen und Überschiebungen in der Invariantentheorie, insbesondere im quaternären Gebiet. Math. Ann. 113 (1936) 14–35. \hyper@normalise
- W (54) Hsien Chung Wang, Closed manifolds with homogeneous complex structure. Amer. J. Math. 76 (1954) 1–32. \hyper@normalise
- W (77) Alan Weinstein, Lectures on Symplectic Manifolds, Regional Conference Series in Mathematics, vol. 29. Amer. Math. Soc., Providence, RI, 1977. \hyper@normalise
- W (11) Willem A. Wythoff, Review of La géométrie des “feuillets” S (08). Revue semestrielle des publications mathématiques (Amsterdam) 19 (1911) 110–111. \hyper@normalise
- Y (14) Takumi Yamada, Invariant pseudo-Kähler metrics on generalized flag manifolds. Differential Geom. Appl. 36 (2014) 44–55.