You are on page 1of 96

ESI_Onishchik_titelei.qxd 30.01.

2004 9:33 Uhr Seite 1

S E
M M
E S

S E
M
E S
M
E M S
ESI_Onishchik_titelei.qxd 30.01.2004 9:33 Uhr Seite 2

ESI Lectures in Mathematics and Physics

Editors
Joachim Schwermer (Institute for Mathematics, University of Vienna)
Jakob Yngvason (Institute for Theoretical Physics, University of Vienna)

The Erwin Schrdinger International Institute for Mathematical Physics


Boltzmanngasse 9
A-1090 Wien
Austria

The Erwin Schrdinger International Institute for Mathematical Phyiscs is a meeting place for leading experts
in mathematical physics and mathematics, nurturing the development and exchange of ideas in the interna-
tional community, particularly stimulating intellectual exchange between scientists from Eastern Europe and
the rest of the world.

The purpose of the series ESI Lectures in Mathematics and Physics is to make selected texts arising from its
research programme better known to a wider community and easily available to a worldwide audience. It
publishes lecture notes on courses given by internationally renowned experts on highly active research topics.
In order to make the series attractive to graduate students as well as researchers, special emphasis is given to
concise and lively presentations with a level and focus appropriate to a student's background and at prices
commensurate with a student's means.
ESI_Onishchik_titelei.qxd 30.01.2004 9:33 Uhr Seite 3

Arkady L. Onishchik

Lectures on
Real Semisimple
Lie Algebras
and Their
Representations

S E
M M
E S

S E
M
E S
M
European Mathematical Society
ESI_Onishchik_titelei.qxd 30.01.2004 9:33 Uhr Seite 4

Author:

Arkady L. Onishchik
Faculty of Mathematics
Yaroslavl State University
Sovetskaya 14
150 000 Yaroslavl
Russia

2000 Mathematics Subject Classification (primary; secondary): 17-01, 17B10, 17B20; 22E46

Bibliographic information published by Die Deutsche Bibliothek

Die Deutsche Bibliothek lists this publication in the Deutsche Nationalbibliografie; detailed bibliographic data
are available in the Internet at http://dnb.ddb.de.

ISBN 3-03719-002-7

This work is subject to copyright. All rights are reserved, whether the whole or part of the material is
concerned, specifically the rights of translation, reprinting, re-use of illustrations, recitation, broadcasting,
reproduction on microfilms or in other ways, and storage in data banks. For any kind of use permission
of the copyright owner must be obtained.

2004 European Mathematical Society

Contact address:

European Mathematical Society Publishing House


Seminar for Applied Mathematics
ETH-Zentrum FLI C1
CH-8092 Zrich
Switzerland

Phone: +41 (0)1 632 34 36


Email: info@ems-ph.org
Homepage: www.ems-ph.org

Printed on acid-free paper produced from chlorine-free pulp. TCF


Printed in Germany

987654321
Contents

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii
1. Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
2. Complexication and real forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3. Real forms and involutive automorphisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4. Automorphisms of real semisimple Lie algebras . . . . . . . . . . . . . . . . . . . . . . . . . 27
5. Cartan decompositions and maximal compact subgroups . . . . . . . . . . . . . . . . 36
6. Homomorphisms and involutions of complex
semisimple Lie algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
7. Inclusions between real forms under an
irreducible representation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
8. Real representations of real semisimple Lie algebras . . . . . . . . . . . . . . . . . . . . . 64
9. Appendix on Satake diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
Introduction
These notes reproduce the lectures which I gave in Masaryk University (Brno,
Czech Republic) and in E. Schrodinger Institute for Mathematical Physics (Vi-
enna, Austria) during the autumn semesters 2001 and 2002, respectively. The
main goal was an exposition of the theory of nite dimensional representations of
real semisimple Lie algebras.
The foundation of this theory was laid in E. Cartan [3]. Iwahori [13] gave an
updated exposition of the E. Cartans work (see also [10], Ch. 7). This theory
reduces the classication of irreducible real representations of a real Lie algebra g0
to description of the so-called self-conjugate irreducible complex representations
of this algebra and to calculation of an invariant of such a representation (with
values 1) which is called the index. No general method for solving any of these two
problems were given in [10, 13], but they were reduced to the case when g0 is simple
and the highest weight of an irreducible complex representation is fundamental.
A complete case-by-case classication for all simple real Lie algebras g0 was given
(without proof) in the tables of Tits [24].
The aim proclaimed by E. Cartan in [2, 3] was to nd all irreducible subalgebras
of the linear Lie algebras gln (C) and gln (R). As a continuation of this line, one
should consider the works of Maltsev [17] and Dynkin [6] on semisimple subalgebras
of simple complex Lie algebras, also based on the representation theory. A similar
problem for real simple Lie algebras was studied in the paper of Karpelevich [15].
The results of this paper solve actually the problems on the self-conjugate complex
representations and the index mentioned above. Note that the paper appeared
before the publication of [13], but, unfortunately, still is not widely known. Our
goal is to give a simplied (and somewhat extended and corrected) exposition of
the main part of [15] and to relate it to the theory developped in [10, 13].
The rst section contains some classical facts from the theory of Lie groups,
Lie algebras and their representations, including the structure theory of complex
semisimple Lie algebras, given without proofs. The necessary notation is also xed
there. In the main body of the lecture notes (2 8) the exposition is given with
detailed proofs.
2 deals with general facts on complexication and realication, real and com-
plex structures and real forms. In particular, a general description of simple real
Lie algebras in terms of simple complex ones is given and two main examples of real
forms (normal and compact) in a complex semisimple Lie algebra are constructed.
3 is devoted to the correspondence between real forms of a complex semisimple
Lie algebra g and involutive automorphisms of g discovered by E. Cartan [4]
which is the main tool in the subsequent study of real Lie algebras and their
representations. Instead of Riemannian geometry exploited in [4], we use the
elementary techniques of Hermitian vector spaces, as in [11], Ch. III or [19],
Ch. 5.
In 4, necessary facts about automorphisms of complex semisimple Lie algebras
are given. As an example, we classify involutive automorphisms (and hence real
forms) of the Lie algebra sln (C). At the same time, we do not give such a classi-
viii Introduction

cation for other simple Lie algebras, referring to [11] or [19]. We also construct
here the so-called principal three-dimensional subalgebra of a complex semisimple
Lie algebra g and use it for studying the Weyl involution of g which corresponds
to the normal real form of this algebra.
In 5, we study the Cartan decompositions of a real semisimple Lie algebra g0
and of the corresponding adjoint linear group Int g0 . Then we prove the conjugacy
of maximal compact subgroups of Int g0 (the elementary proof borrowed from [19]
makes use of some properties of convex functions, instead of Riemannian geometry
exploited in the original proof of E. Cartan).
6 is devoted to the following problem. Suppose a homomorphism f of one
complex semisimple Lie algebra g into another h be given. Choosing real forms
g0 g and h0 h, one asks, when f (g0 ) h0 . We prove the result of [15]
which claims that this inclusion is equivalent to the relation f =  f if the classes
of conjugate real forms and conjugate automorphisms are considered. Thus, the
problem is reduced to that of extension of involutive automorphisms by homo-
morphisms. To get this reduction, we use, as in 3, the elementary techniques
of Hermitian vector spaces. A more precise result is got in the case when f is a
so-called S-homomorphism (in [15], it was proved for an irreducible representation
into a classical linear Lie algebra h).
In 7, we study the extension problem for involutive automorphisms in the main
case, when f = is an irreducible representation g sln (C). First we consider
the relation =  in the case when  is an outer automorphism of sln (C), i.e.,

 (X) = BX  B 1 , X sln (C), where B  = B.

A condition of existence of such an extension in terms of the highest weight of is


found; it is expressed as invariance of the highest weight of under an involutive
automorphism s0 of the Dynkin diagram. An important problem is to calculate
the sign j = 1 in the above formula provided that this condition is satised. In
[15], a simple explicit formula for this Karpelevich index j in the case when is
an inner automorphism was proved. We generalize this formula to an arbitrary
involutive . Then we consider the case of an inner  , getting, in particular, two
formulas from [15] which express the signature of the invariant Hermitian form
through the character of . We do not reproduce the explicit formulas expressing
this signature through the highest weight of which were deduced in [15], using
cumbersome calculations. We also omit the study of the extension problem for
reducible representations g sl(V ) and for representations g so(V ) and g
sp(V ) which was made in [15], too.
In 8, we give a classication of irreducible real representations of real semi-
simple Lie algebras following the method exposed in [10, 13]. It turns out that
an irreducible complex representation 0 of a real semisimple Lie algebra g0 is
self-conjugate if and only if the corresponding involution of g0 (C) extends to an
outer involution of sln (C) by the complexied representation 0 (C). Moreover, the
Cartan index of 0 coincides with the Karpelevich index of 0 (C). This allows to
apply the results of 7 which leads to an explicit classication of irreducible real
representations in terms of highest weights.
Introduction ix

In 9, written by J. Silhan, a description of the symmetry s0 , introduced in


7, in terms of the Satake diagram of a real semisimple Lie algebra is given. This
allows to determine self-conjugate complex representations by means of the Satake
diagram.
We conclude with some tables, giving in particular the indices of irreducible
representations of simple complex Lie algebras.
I would like to express my deep gratitude to Masaryk University (Brno) and
E. Schrodinger Institute for Mathematical Physics (Vienna) for hospitality and,
in particular, to J. Slovak and P. Michor for inviting me and for their interest in
my lectures. I also thank J. Silhan who wrote the rst version of these lecture
notes and helped me very much in preparing the tables. My pleasant duty is to
thank the EMS Publishing House, and in particular Dr. Thomas Hintermann, for
including my work into one of lecture notes series.
This work was partially supported by Russian Foundation for Basic Research,
grant no. 01-01-00709, and by the grant no. 123.2003.01 for supporting scientic
schools.
1. Preliminaries
In this section, some necessary facts from the theory of Lie groups and Lie algebras
are formulated without proofs. We refer the reader for details to the corresponding
text-books (see, e.g., [10, 11, 16, 19, 22]).

I. Lie groups and Lie algebras


We consider here nite dimensional Lie algebras over the eld K = C or R. Let g
be such a Lie algebra. Any x g determines the linear operator ad x in g given
by
(ad x)y = [x, y] , y g .
The mapping ad : x  ad x is a homomorphism of Lie algebras g gl(g) called
the adjoint representation of g. Its image ad g is an ideal of the Lie algebra der g
of all derivations of g. The Lie algebra ad g is called the adjoint algebra or the
algebra of inner derivations of g.
Consider the linear algebraic group Aut g of all automorphisms of g. For any
x g we have exp(ad x) Aut g. The (normal) subgroup of Aut g generated by all
exp(ad x), x g, is called the group of inner automorphisms of g and is denoted
by Int g. The elements of Int g are called inner automorphisms of g, while the
elements of Aut g \ Int g are called outer automorphisms.
(I.1) The Lie subalgebras ad g der g are tangent Lie algebras of the linear
groups Int g Aut g.
Let G be a Lie group with tangent Lie algebra g. For any g G, let g denote
the inner automorphism h  ghg 1 of G generated by g. Then the automorphism

Ad g = de g Aut g

is dened. The mapping Ad : g  Ad g is a representation of G in the vector


space g called the adjoint representation of G. We have de Ad = ad. The image
Ad G is a normal subgroup of Aut g called the adjoint linear group of G.
(I.2) If G is connected, then Ad G = Int g and Ker Ad = Z(G) (the centre of
the group G).
A bilinear (or a sesquilinear) form b on a Lie algebra g is called invariant if all
ad x, x g, are skew-symmetric relative to b, i.e., if

b([x, u], v) = b(u, [x, v]) , x, u, v g .

This is equivalent to the property

b(u, v) = b(u, v) , Int g , u, v g .

The most important example of an invariant bilinear form is the Killing form
kg of g given by
kg (u, v) = tr((ad u)(ad v)) , u, v g . (1)
2 1. Preliminaries

(I.3) The Killing form kg is a symmetric invariant bilinear form on g satisfying

kg (u, v) = kg (u, v) , Aut g , u, v g .

(I.4) For any ideal h of g, we have kg |h = kh .


A Lie algebra g is called semisimple if g has no non-trivial solvable (or, which
is equivalent, commutative) ideal. A Lie group is called semisimple if its tangent
Lie algebra is semisimple. The following important Cartan criterion is valid:
(I.5) A Lie algebra g over K = C or R is semisimple if and only if its Killing
form kg is non-degenerate.
(I.6) A Lie algebra g over K = C or R is semisimple if and only if


k
g= gi ,
i=1

where gi are non-commutative simple ideals. This implies that [g, g] = g.


(I.7) If g is semisimple, then der g = ad g. This implies that Int g = (Aut g) .
A Lie algebra g is called reductive if

g = g0 z(g) ,

where g0 is semisimple and z(g) is the centre of g. This implies that [g, g] = g0 .
Let G be a compact Lie group. Sometimes we will use the invariant integration
of smooth (scalar or vector-valued) functions over G (see, e.g., [12], Ch. I, or [16],
 invariant integral of a function f is denoted by G f (g)dg; we
Ch. VIII). The
suppose that G dg = 1. The following invariance conditions are satised:
   
1
f (gh)dg = f (hg)dg = f (g )dg = f (g)dg , h G.
G G G G

(I.8) Let R : G GL(V ) be a linear representation of a compact Lie group G


in a real or complex vector space V . For any xed v V , the vector

v0 = (R(g)v)dg V
G

is invariant under R, i.e., satises R(g)v0 = v0 , g G.


Under a scalar product in a real or complex vector space V we mean a positive
denite symmetric bilinear (or Hermitian) form on V . The following corollary of
(I.8) is called Theorem of Weyl .
1. Preliminaries 3

(I.9) Let R : G GL(V ) be a linear representation of a compact Lie group G


in a real or complex vector space V . For any xed scalar product ( , ) on V , the
function ( , )0 given by

(u, v)0 = (R(g)u, R(g)v)dg , u, v V ,
G

is a scalar product on V , invariant under R, i.e., satisfying (R(g)u, R(g)v)0 =


(u, v)0 , u, v V, g G.

A real Lie algebra g is called compact if there exists an invariant scalar product
on g. Clearly, any subalgebra of a compact Lie algebra is compact, too. (I.9)
implies

(I.10) If G is a compact Lie group, then its Lie algebra g is compact .

We also formulate the following important properties of compact Lie algebras:

(I.11) Any compact Lie algebra is reductive.

(I.12). The Killing form kg of a compact Lie algebra g satises kg (x, x)


0, x g. A real Lie algebra g is compact semisimple if and only if kg is negative
denite.

(I.13) For any compact semisimple Lie algebra g, the Lie groups Aut g and Int g
are compact. For any compact Lie algebra g, there exists a compact Lie group G
such that g is the tangent Lie algebra of G. If g is compact semisimple, then each
connected Lie group with the tangent Lie algebra g is compact .

Let g be a real or complex Lie algebra and let Aut g be involutive, i.e.,
satises 2 = e. Then is diagonalizable with eigenvalues 1, and hence we have
the eigenspace decomposition

g = g+ g , where g = {x g | x = x} . (2)

(I.14) The decomposition (2) is a Z2 -grading, i.e.,

[g+ , g+ ] g+ , [g+ , g ] g , [g , g ] g+ .

In particular, g = g+ is a subalgebra of g. Conversely, for any Z2 -grading

g = kp, (3)

the linear transformation : x + y  x y, x k, y p, is an involutive


automorphism of g, and (3) is the corresponding eigenspace decomposition.
4 1. Preliminaries

II. Structure of complex semisimple Lie algebras


Let g be a complex semisimple Lie algebra. Choose a maximal toral subalgebra
t of g, i.e., a maximal subalgebra consisting of semisimple elements. It is neces-
sarily commutative, and g is decomposed into the direct sum of weight subspaces
corresponding to the adjoint representation of t in the vector space g (see (III.4)).
Then t coincides with the subspace corresponding to the weight 0 (i.e., with its
centralizer in g). The non-zero weights are called the roots of g; they form a -
nite subset t (the system of roots). Thus, we have the so-called root space
decomposition: 
g=t g , (4)

where g is the weight subspace (called the root space) corresponding to .


(II.1) Any two maximal toral subalgebras of g are conjugate by an inner auto-
morphism of g.
Hence the root space decomposition is determined uniquely, up to inner auto-
morphisms. The dimension dim t of any maximal toral subalgebra is called the
rank of g and is denoted by rk g.
Dene the following real vector subspace of t:

t(R) = {h t | (h) R for all } . (5)

Then the following properties hold:


(II.2) t = C ;
(II.3) t(R) is a real form of t;
(II.4) t(R) := R is a real form of t .
As this notation suggests, t(R) may be regarded as the real dual vector space
for t(R).
Remind (see (I.5)) that the Killing form k = kg is non-degenerate.
(II.5) The Killing form k satises k(x, y) = 0 for two weight vectors x, y g
corresponding to any weights , t such that + = 0. It follows that t and
the dierent vector subspaces g g , , are pairwise orthogonal, and the
restrictions k|t and k|(g g ), are non-degenerate. The restriction k|t(R) is
real and positive denite.
Thus, we have a natural structure of Euclidean space on t(R). We often write
(x, y) = k(x, y) for x, y t. It is also convenient to carry this bilinear form to the
dual space t . Consider the vector space isomorphism  u given by

(u , h) = (h) , h t ; (6)

it maps t(R) onto t(R). Now we dene a non-degenerate form on t by

(, ) := (u , u ) = (u ) = (u ) , , t .
1. Preliminaries 5

This form is real and positive denite on t(R) . We also dene the vector

2
h = u (7)
(, )

for any non-zero t(R) . In particular, the vectors h , , are called


coroots.
Clearly, h t(R), . Also, any is uniquely determined by its
restriction to t(R) and, hence, may be identied with this restriction.
(II.6) For any we have , but c
/ for any c C, c = 1.
2(,)
(II.7) (h ) = (,) Z for any , .
(II.8) dim g = 1 for any .
It follows from (II.8) that we get a basis of g if we take a basis of t and a
non-zero vector e g (a root vector ) for each root . By denition, we
have
(II.9) [h, e ] = (h)e , h t.
Other commutation relations are the following ones:
(II.10) Let , be such that + = 0. Then

N, e+ , N, = 0, if + ,
[e , e ] =
0 otherwise.

The simplest example of a non-commutative simple complex Lie algebra is g =


sl2 (C). Denote
     
1 0 0 1 0 0
H= , E= , F = .
0 1 0 0 1 0

Then t = HC is a maximal toral subalgebra of g. We have = {, }, where


(H) = 2. Hence t(R) = HR . We have

[H, E] = 2E , [H, F ] = 2F , [E, F ] = H .

Let again g be an arbitrary complex semisimple Lie algebra. A triple {e, h, f }


of elements of g is called an sl2 -triple if [h, e] = 2e, [h, f ] = 2f, [e, f ] = h and
h = 0. Such a triple spans the subalgebra e, h, f C of g which is isomorphic to
sl2 (C). An isomorphism is given by

E  e , H  h , F  f .

(II.11) Given a root vector e g , we can choose f = e g in such a


2
way that k(e , f ) = (,) , which implies [e , f ] = h . Then {e , h , f } is an
sl2 -triple.
6 1. Preliminaries

(II.12) Given an automorphism Aut g such that (t) = t, the transposed


transformation  of t satises

 () = ,
(t(R)) = t(R) ,
(g ) = g( )1 .

Any root determines the hyperplane P = Ker t(R). Let r denote


the orthogonal reection relative to P . It is given by

r (h) = h (h)h , h t(R) . (8)

Dually, we have the hyperplane L = { t(R) | (, ) = 0}. The orthogonal


reection relative to L coincides with the transposed transformation r and is
given by
2(, )
r () = , t(R) . (9)
(, )
The reections r are actually induced by inner automorphisms of g leaving t
invariant.
(II.13) Given a root , suppose that the root vectors e and f are chosen
as in (II.11). Then

r = |t(R) , where = exp ad (e f ) .
2

(II.14) The reections r (respectively, r ) generate a nite group W of or-


thogonal transformations of t(R) (respectively, a nite group W of orthogonal
transformations of t(R) ). The mapping w  (w )1 is an isomorphism of W
onto W . The system of roots is invariant under W .
Any of two groups dened in (II.14) is called the Weyl group of g.
An element h t(R) is called regular if (t) = 0 for all , and singular
otherwise. The connected components of the set t(R)reg of regular elements of t(R)
are called the Weyl chambers; they are open polyhedral cones. (II.14) implies that
the Weyl group W permutes the Weyl chambers.
(II.15) The natural action of W on the Weyl chambers is simply transitive.
The property (II.13) suggests another denition of the Weyl group. Let us
dene the subgroups N = { Int g | (t) = t} and T = {exp(ad h) | h t} of
Int g.
(II.16) If N maps a xed Weyl chamber onto itself, then T , and
|t = id. The restriction homomorphism  |t(R) maps N onto the Weyl
group W , and hence W
N/T .
Let us x a Weyl chamber D. Since D is connected, for any root we
have either (h) > 0 for all h D or (h) < 0 for all h D. In the rst case,
1. Preliminaries 7

is called positive (with respect to D), and in the second one it is called negative
(with respect to D). The sets of positive and negative roots are denoted by +
and = + respectively. A positive root is called simple (with respect to
D) if = + , where , + . Let + denote the subset of simple roots.
We will write = {1 , . . . , l }.
(II.17) Any + has the form = i1 + . . .+ ir , where i1 + . . .+ ik +
for each k = 1, . . . , r.
(II.18) The set is a basis of t(R) . In particular, l = dimR t(R) = dimC t is

the rank of g. Any can be uniquely written in the form = li=1 ki i ,
where ki Z and either all ki  0 (for ) or all ki  0 (for ).
+

The property (II.18) suggests to call the set of simple roots a base of the
system of roots .
(II.19) The group W acts simply transitively on the subsets of positive roots
and on the subsets of simple roots of . Any is simple with respect to an
appropriate Weyl chamber. Given a xed subset of simple roots , the Weyl
group W is generated by r , .
Let us x a Weyl chamber D and consider the corresponding subsets +
of positive and simple roots. Dene the linear form
1 
= . (10)
2 +

(II.20) If and + , then r () + , except the case = , when


r ()= . We have

r () = , (h ) = 1 f or any .

Using the system = {1 , . . . , l }, one can construct a system of generators


of the Lie algebra g in the following way. Denote

hi = hi , ei = ei , fi = fi , i = 1, . . . , l , (11)

where the root vectors ei and fi are chosen as in (II.11). Introducing the integers
2(i , j )
aij = i (hj ) =
(j , j )

(see (II.7)), we get a l l matrix A = (aij ) called the Cartan matrix .


(II.21) The elements hi , ei , fi , i = 1, . . . , l, form a system of generators of g
and satisfy the relations
[hi , hj ] = 0 ,
[hi , ej ] = aji ej , [hi , fj ] = aji fj , (12)
[ei , fi ] = hi , [ei , fj ] = 0 f or i = j .
8 1. Preliminaries

The set {hi , ei , fi | i = 1, . . . , l} is called the canonical system of generators of


g associated with t and .
The structure of g is completely determined by the Cartan matrix A. More
precisely, we have
(II.22) Suppose that two complex semisimple Lie algebras g and g with canonical
generator systems ei , fi , hi and ei , fi , hi , respectively, are given, and let A and A be
their Cartan matrices. If A = A, then there exists a unique isomorphism : g g
such that (ei ) = ei , (fi ) = fi and (hi ) = hi .
For the entries aij , i = j, of the Cartan matrix only the values 0, 1, 2, 3 are
possible. Moreover, the only possible values for mij = aij aji , i = j, are 0, 1, 2, 3,
and the corresponding values for the angle between i and j are ij = (1 n1ij ),
| |
where nij = 2, 3, 4, 6, respectively. We also have |ji | = mij , whenever mij = 0
and |j | |i |. It is usual to depict the system of simple roots (or the Cartan
matrix) by a graph, assigning to any i , i = 1, . . . , l, a vertex and joining the i-th
and j-th vertices by the edge of multiplicity mij (in the case mij = 0 the vertices
are not linked). If |j | > |i |, then the edge is oriented by an arrow pointing from
the j-th vertex towards the i-th one. This graph is called the Dynkin diagram
of the Lie algebra g. As (II.22) shows, the Dynkin diagram determines g up to
isomorphy.
Let gi , i = 1, 2, be two complex semisimple Lie algebras. Consider g = g1 g2 .
Any maximal toral subalgebra t of g is of the form t = t1 t2 , where ti is a
maximal toral subalgebra of gi , i = 1, 2, the summands being orthogonal with
respect to kg . Hence t = t1 t2 . Now, the system of roots of g decomposes
as = 1 2 , where i is the system of roots of gi relative to ti , i = 1, 2. It
follows that t (R) = t1 (R)t2 (R). Moreover, any Weyl chamber for g has the form
D = D1 D2 , where Di is a Weyl chamber for gi . Therefore we have = 1 2 ,
where i is the system of simple roots of gi corresponding to Di and (, ) = 0
for any 1 , 2 .
(II.23) A semisimple complex Lie algebra g is simple if and only if its Dynkin
diagram is connected. Generally, the decomposition of the Dynkin diagram into
connected components corresponds to the decomposition of g into the direct sum of
simple ideals (see (I.6)).
The Dynkin diagrams of simple complex Lie algebras are listed in Table 1.

III. Representations of Lie algebras


Let g be a nite dimensional Lie algebra over the eld K = C or R. We consider
linear representations of g in nite dimensional vector spaces over K, i.e., homo-
morphisms of Lie algebras : g gl(V ). The number dim = dim V is called the
dimension of .
Two representations : g gl(V ) and  : g gl(V  ) are called equivalent (or
isomorphic) if there exists an isomorphism of vector spaces f : V V  such that
f (x) =  (x)f, x g .
In this case we write  .
1. Preliminaries 9

If : g gl(V ) is a representation, then we can choose a basis in V and dene a


homomorphism : g gln (K), sending any x g to the matrix of (x) relative to
our basis. We will call a matrix form of ; it can be regarded as a representation
of g in the vector space K n equivalent to . Clearly, two representations are
equivalent if and only if they have the same matrix form in appropriate bases.
With any representation : g gl(V ) the dual (or contragredient ) representa-
tion : g gl(V ) is associated. It is dened in the dual space V by

( (x))(v) = ((x)v) , x g, V , v V ,

or, equivalently,
(x) = (x) , x g. (13)

(III.1) If we choose dual bases in V and V , then the corresponding matrix


forms and of the representations and are related by


(x) =  (x) , x g.

A vector subspace W V is called invariant under a representation : g


gl(V ) if (x)(W ) W, x g. Restricting the operators (x), x g, to an invari-
ant subspace W , we get a representation of g in W , called the subrepresentation
of .

r
Suppose that V = i=1 Vi , where Vi , i = 1, . . . , r, are vector subspaces, invari-
ant under a representation : g gl(V ), and let i denote the subrepresentation
in Vi . Then we say that is the sum of i and write = ri=1 i .
A representation : g gl(V ) is called irreducible if V does not content
any invariant subspace W, {0} = W = V . It is called completely reducible if
for any invariant subspace W V there exists an invariant W  V such that
V = W W .
r
(III.2) A representation is completely reducible if and only if = i=1 i ,
where i are irreducible. In this case i are uniquely determined by , up to
equivalence.
The representations i in (III.2) are called the irreducible components of .
(III.3) Any representation of a semisimple Lie algebra is completely reducible.
Let us now consider representations of a complex semisimple Lie algebra g.
Note that, by (I.6), we may always regard it as a homomomorphism of g into
[gl(V ), gl(V )] = sl(V ). Let t be a maximal toral subalgebra of g.
(III.4) For any representation : g gl(V ) we have the weight space decompo-
sition 
V = V ,

10 1. Preliminaries

where t (R) is the system of weights of and

V = {v V | (h)v = (h)v , h t} = {0}

are the corresponding weight subspaces.

A special case of the weight space decomposition (for = ad) is the root space
decomposition (4); we have ad = {0}.

(III.5) For any representation we have = .


2(,)
(III.6) (h ) = (,) Z for any , .

(III.7) (e )V V+ for any , .

(III.8) For any Aut g such that (t) = t, we have = ( )1 ( ). The


system is invariant under W .

Let us x a Weyl chamber D and consider the corresponding systems +


of simple and positive roots and the canonical generators (11) of g. A weight
vector v V, v = 0, is called a highest vector for if (e )v = 0 for any +
or, equivalently, if (ei )v = 0, i = 1, . . . , l. The weight of v is called a highest
weight of .

(III.9) If is irreducible, then contains a unique highest weight , and


dim V = 1. Two irreducible representations of g are equivalent if and only if they
have the same highest weight. An arbitrary representation is determined up to
equivalence by the system of highest weights of its irreducible components.

A linear form t (R) is called dominant if (hi ) 0, i = 1, . . . , l.

(III.10) A highest weight of any representation is dominant. For any dominant


t (R) such that i = (hi ) Z there exists an irreducible representation of g
with the highest weight .

Let { 1 , . . . , l } denote the base of t (R) dual to the base {h1 , . . . , hl } of t(R).
The linear forms i are called the fundamental weights. By (III.10), they are the
highest weights of certain irreducible representations of g which are called the basic
l
representations. For any t (R) we have = i=1 i i where i = (hi ).
The irreducible representation with the highest weight can be described by
writing coecients i over the corresponding vertices of the Dynkin diagram of g.
In this way we get the so-called Dynkin diagram of an irreducible representation.
Let : g gl(V ) be an irreducible representation, and be its highest weight.
Then the highest vector v V generates V over g. More precisely, let us denote

v = v ,
vi1 ,... ,ik = (fi1 ) . . . (fik )v , 1 i 1 , . . . , ik l .
1. Preliminaries 11

(III.11) The vectors vi1 ,... ,ik span the vector space V . The canonical generators
act on these vectors as follows:

(hi )v = i v ,
(hi )vi1 ,... ,ik = (i ai1 ,i . . . aik ,i )vi1 ,... ,ik ,
(fi )vi1 ,... ,ik = vi,i1 ,... ,ik ,
(ei )v = 0 ,
(ei )vi1 ,... ,ik = (ii1 (hi ) + (fi1 )(ei ))vi2 ,... ,ik .

Any may be written in the form

= i1 . . . ik ,

where k 0.
Let gi , i = 1, 2, be two complex semisimple Lie algebras. Consider g = g1 g2 .
If i : gi gl(Vi ), i = 1, 2, are linear representations, then one can dene the
representation = 1 2 : g gl(V1 V2 ) by

(x1 , x2 )(v1 v2 ) = 1 (x1 )(v1 )v2 +v1 2 (x2 )(v2 ), xi gi , vi Vi , i = 1, 2 .

The representation is called the tensor product of 1 and 2 .


(III.12) If 1 and 2 are irreducible, then = 1 2 is irreducible, too. Choose
a maximal toral subalgebra ti gi , i = 1, 2, and denote t = t1 t2 . The highest
weight of relative to t is (1 , 2 ), where i is the highest weight of i relative
to ti , i = 1, 2. Any irreducible representation of g = g1 g2 is equivalent to the
tensor product 1 2 , where i is an irreducible representation of gi , i = 1, 2.
2. Complexification and real forms
Given a real vector space V0 , we can construct its complexication V = V0 (C) as
the complex vector space V0 R C = V0 iV0 . Clearly, any basis of V0 is a basis
of V , and hence dimC V0 = dimR V . The mapping S : V V , dened by

S(u + iv) = u iv , u, v V ,

is called the complex conjugation.


On the other hand, any complex vector space V can be regarded as a real vector
space, which is called its realication and is denoted by VR . Clearly, dimR VR =
2 dimC V .
Consider again the complexication V = V0 (C) of a real vector space V0 .
Clearly, the complex conjugation S given by (1) is an antiautomorphism (anti-
linear automorphism) of V , i.e., an automorphism of the real vector space VR
satisfying S(cv) = cSv, c C, v V . Also S is involutive, i.e., satises S 2 = e.
For brevity, we will call such an antiautomorphism of any complex vector space
an antiinvolution.
Next, let us consider a complex vector space V . A real vector subspace V0 of
VR is called a real form of V if VR = V0 iV0 . One veries easily that this is
equivalent to the following property: the mapping V0 (C) V sending u + iv
V0 (C) to u + iv V, u, v V0 , is an isomorphism of complex vector spaces.
This isomorphism identies the complex conjugation dened in V0 (C) with an
antiinvolution S of V , and we have

V0 = V S = {v V | Sv = v} . (1)

Generally, we call a real structure in a complex vector space V any antiinvolution


S : V V . Obviously, real structures in V are in bijective correspondence with
real forms V0 V (a real form V0 denes the corresponding complex conjugation
S, while a real structure S denes the real form V S given by (1)).
A complex structure in a real vector space V0 is, by denition, an automorphism
J GL(V0 ) such that J 2 = e. A complex vector space V may be regarded as
the real vector space VR endowed with the complex structure J : v  iv, v
V . Conversely, if a complex structure J in a real vector space V0 is given, then
we may regard (V0 , J) as a complex vector space with the realication V0 , the
multiplication by complex scalars being given by (a + bi)v = av + bJv, a, b
R, v V0 .
For any complex vector space V , let us denote by V the complex vector space
which coincides with V as an additive group, but is endowed with the following
multiplication by complex scalars: cx = cx, x V = V, c C. In other words, if
J is the given complex structure in V , i.e., J : v  iv, v V , then V = (VR , J).
The vector space V is called the complex conjugate to V . Given a second complex
vector space W , we easily see that any linear mapping : V W is a linear
mapping V W as well. If is expressed by a matrix A relative to certain bases
of V and W , then the matrix of : V W relative the same bases is equal to A.
On the other hand, an antilinear mapping V W can be interpreted as a linear
mapping V W or V W , and vice versa.
2. Complexification and real forms 13

All these general notions can be introduced in the category of Lie algebras.
Given a real Lie algebra g0 , we endow the complexication g = g0 (C) of the vector
space g0 with the bracket extending the bracket of g0 , i.e., given by

[x1 + iy1 , x2 + iy2 ] = [x1 , y1 ] [y1 , y2 ] + i([x1 , y2 ] + [y1 , x2 ]), x1 , y1 , x2 , y2 g0 .

Clearly, g is a complex Lie algebra. On the other hand, for any complex Lie algebra
g, the realication gR of the vector space g is a real Lie algebra.
Now, if g = g0 (C) is the complexication of a real Lie algebra g0 , then the
complex conjugation , dened by

(x + iy) = x iy , x, y g0 , (2)

is an involutive antiautomorphism (antilinear automorphism) of the complex Lie


algebra g (we call it an antiinvolution). Next, let g be a complex Lie algebra. A
real subalgebra g0 of gR is called a real form of g if it is a real form of the complex
vector space g, i.e., if gR = g0 ig0 . In this case, g0 (C) is naturally identied
with g.
A real structure in a complex Lie algebra g is an antiinvolution : g g. The
complex conjugation (2) corresponding to a real form of g is a real structure in g.
Conversely, any real structure in g denes the real form g . In this way, we get
a bijection between real structures and real forms of g.
Let us consider two real forms g0 and h0 of complex Lie algebras g and h,
respectively. Clearly, any homomorphism h : g0 h0 extends uniquely to a
homomorphism h(C) : g h of complex Lie algebras.
The following proposition is the rst (trivial!) step in the classication of real
forms of a given complex Lie algebra up to isomorphy.
Proposition 1. Consider two real forms g0 , g1 of a complex Lie algebra g, and let
0 , 1 be the corresponding real structures. Then g0
g1 if and only if 1 =
0 1 for a certain Aut g.
Proof. By the above remark, any isomorphism h : g0 g1 extends to an auto-
morphism = h(C) of g. Clearly, the antiautomorphisms 0 and 1 coincide
with h on g0 . It follows that 0 = 1 , whence 1 = 0 1 . Conversely, sup-
pose that there exists Aut g satisfying 1 = 0 1 . One veries easily that
(g0 ) = g1 , i.e., (g0 ) = g1 . 
A complex structure in a real Lie algebra g0 is a complex structure J in the
vector space g0 satisfying

J[x, y] = [x, Jy] , x, y g0 . (3)

Clearly, any complex Lie algebra is endowed with the complex structure Jv = iv.
Conversely, if a complex structure J in a real Lie algebra g0 is given, then we may
regard (g0 , J) as a complex Lie algebra (with the same bracket); its realication
is g0 .
If g is a complex Lie algebra, then the complex conjugate vector space g, en-
dowed with the same bracket as g, is a complex Lie algebra as well. It is called
14 2. Complexification and real forms

the complex conjugate Lie algebra. Clearly, the mapping id : g g = g is an


antilinear isomorphism. Also, isomorphisms g g are antilinear automorphisms
of g. In particular, g
g, whenever g possesses a real form.
Sometimes we will also deal with the following notion. A quaternion structure in
a complex vector space V is an antiautomorphism J : V V satisfying J 2 = e.
Thus, J is a complex structure on VR anticommuting with the given complex
structure. Let us view the skew-eld H of quaternions as a right vector space over
C with the basis 1, j. Thus, a quaternion is an element of the form q = z + jw with
z, w C, and the following relations hold: j 2 = 1, jw = wj. If a quaternion
structure J in a complex vector space V is given, then V can be regarded as a
right vector space over H, such that

v(z + jw) = zv + w(Jv) , v V , z, w C .

Conversely, any quaternion vector space V is, in a natural way, a complex vector
space endowed with a quaternion structure J, and linear endomorphisms V V
are the endomorphisms of the complex vector space commuting with J.
Example 1. Here we describe certain real forms of the Lie algebras gln (C) and
sln (C). We shall see later that our list contains all the real forms of sln (C) up to
isomorphism.
1. An obvious real form of gln (C) is the subalgebra gln (R). The corresponding
real structure is given by

: X = (xij )  X = (xij ).

The restriction of to sln (C) gives the real form sln (R) of sln (C).
2. Let us consider a non-degenerate hermitian form h on Cn with signature
(p, q), p + q = n. We may assume that

h(z, w) = z1 w1 . . . zp wp + zp+1 wp+1 + . . . + zn wn , z, w Cn .

Consider the (real) Lie algebra

up,q = {X gln (C) | h(Xz, w) + h(z, Xw) = 0 for all z, w Cn } .

Since h(z, w) = z  Ip,q w, where


 
Ip 0
Ip,q = , (4)
0 Iq

one sees easily that

up,q = {X gln (C) | XIp,q + Ip,q X  = 0} .

In particular, u0,n = un is the algebra of skew-hermitian matrices. One deduces


that up,q = gln (C) , where

(X) = Ip,q X  Ip,q , X gln (C) .


2. Complexification and real forms 15

One proves easily that is a real structure, and hence up,q is a real form of gln (C).
The elements of up,q are block matrices of the form
 
A B
,
B  D
where A up and D uq .
Clearly, sup,q = up,q sln (C) is a real form of sln (C).
3. For n = 2m, one can construct a real form using quaternions. The mapping

z1
.
.
q1 z1 + jw1 .

. .. zm
Hm  q = .. = .
 C2m
w1
qm zm + jwm .
..
wm
determines an isomorphism between Hm and C2m , both understood as right vector
spaces over C. We will identify Hm and C2m by this isomorphism. Clearly, Hm is
a right vector space over H, and the right multiplication q  qj in Hm is identied
with the quaternion structure J in C2m given by

z1 w1
.. ..
. .

z wm
J m=
w1 z1
. .
.. ..
wm zm
or
J(v) = Sm v , v C2m ,
where  
0 Im
Sm = . (5)
Im 0
Consider the Lie algebra glm (H) of endomorphisms of the vector space Hm (or of
quaternion n n-matrices). Clearly, it is identied with the real subalgebra {X
gl2m (C) | XJ = JX} of gl2m (C). Now, we dene the mapping : gl2m (C)
gl2m (C) by (X) = JXJ 1 = Sm XSm . This is an antiinvolution determining
the real form gl2m (C) = glm (H) of the Lie algebra gl2m (C). The elements of this
real form are block matrices of the form
 
A B
,
B A
where A, B glm (C). A real form slm (H) of sl2m (C) is then formed by the above
matrices with the condition Re tr A = 0.
We want now to compare the Killing forms of a complex Lie algebra g, of a real
form of g and of the complex conjugate Lie algebra g.
16 2. Complexification and real forms

Proposition 2. Let g be a complex Lie algebra.


(i) If g0 is a real form of g, then kg |g0 = kg0 , and g is semisimple if and only if
g0 is semisimple.
(ii) kg (x, y) = kg (x, y), x, y g.
(iii) If is an antilinear automorphism of g, then kg (x, y) = kg (x, y), x, y g.
Proof. (i) Any basis of the real vector space g0 is a basis of the complex vector
space g. If x g0 , then ad x(g0 ) g0 , and hence the operator ad x in g is expressed
in such a basis by the same matrix as its restriction to g0 . This is also true for
(ad x)(ad y), where x, y g0 . It follows that trg ((ad x)(ad y)) = trg0 ((ad x)(ad y)).
Clearly, kg and kg0 have the same matrix in any basis of g0 . Hence, kg0 is non-
degenerate if and only if kg is non-degenerate, and our second assertion follows
from the Cartan criterion (I.5).
(ii) Take x, y g. By a remark above, the matrix of the operator (ad x)(ad y)
in g is complex conjugate to the matrix of this operator in g (with respect to the
same basis). This implies our assertion.
(iii) We may regard as an isomorphism of complex Lie algebras : g g.
By (I.3), we have kg (x, y) = kg (x, y), and our assertion follows from (ii). 
We see from Proposition 2 (i) that the complexication operation gives a corre-
spondence between real and complex semisimple Lie algebras. We will study now
the corresponding question for simple Lie algebras.
Let g be a real Lie algebra, and let z  z denote the complex conjugation in
its complexication g(C). For any m g(C), let us denote m its image under
the complex conjugation. The following properties follow immediately from the
denitions.
1. For any subalgebra h g its complexication h(C) is a subalgebra of g(C).
2. In the situation of assertion 1, h is an ideal of g if and only if h(C) is an
ideal of g(C).
3. A vector subspace b g(C) has the form b = a(C) for a subspace a g if
and only if b = b. In this case, a = b g.
4. For any subalgebra (or ideal) h g(C), the subset h is a subalgebra (respecti-
vely, an ideal) of g(C).
5. For any two vector subspaces a, b g, we have [a, b](C) = [a(C), b(C)].
Proposition 3. If g is complex Lie algebra, then gR (C)
g g.
Proof. Let us denote gdbl = g g. The mapping : gdbl gdbl given by
(x, y) = (y, x) is a real structure in gdbl , since (c(x, y)) = (cy, cx) = c(y, x) =
c(x, y), x, y g, c C. The corresponding real form is gd = {(x, x) | x g}.
The mapping (x, x)  x is an isomorphism of real Lie algebras gd gR , and
hence gdbl
gR (C). 
Theorem 1. Let g be a complex simple Lie algebra. Then any real form of g and
gR (whenever g is non-commutative) are real simple Lie algebras. Conversely, any
real simple Lie algebra is either a real form of a complex simple Lie algebra or is
isomorphic to gR , where g is a non-commutative complex simple Lie algebra.
2. Complexification and real forms 17

Proof. Let g0 = 0 be a real Lie algebra, and let g = g0 (C) be simple. Then for
any non-zero ideal a g0 we get the non-zero ideal a(C) g. Hence, a(C) = g,
and a = g0 . Now, let g be a complex simple Lie algebra, and let a gR be a real
ideal such that 0 = a  gR . Then we get the real ideal ia of gR and two complex
ideals a ia and a + ia of g. These ideals must satisfy a ia = 0 and a + ia = g.
Thus, gR = a ia, whence [a, ia] = 0. This implies that g is commutative.
Now consider a real simple Lie algebra g0 = 0 and suppose that g = g0 (C) is not
simple. We are going to prove that g0 admits a complex structure J, converting
g0 into a complex Lie algebra. Fix a complex ideal 0 = a  g. Then we get the
ideals a, b = a + a and c = a a of g. Since b = b, c = c, we have b = (b g0 )(C)
and c = (c g0 )(C). As g0 = 0 is simple, this implies b g0 = g0 and c g0 = 0.
It follows that g = b = a a. Now we dene the desired complex structure J
by J(x + y) = ix iy, x, y a. Clearly, J 2 = e, and we only have to verify
the condition (3). For z1 = x1 + y1 and z2 = x2 + y2 , where xi , yi a, we get
[z1 , z2 ] = [x1 , x2 ] + [y1 , y2 ], whence
[z1 , Jz2 ] = [x1 , ix2 ] [y1 , iy2 ] = i[x1 , x2 ] i[y1 , y2 ] = J[z1 , z2 ] .
It follows that g0 is the realication of the complex Lie algebra h = (g0 , J). Clearly,
h is simple. If it is commutative, then h
C. But in this case g0
R R is not
simple. 
Let g be a complex semisimple Lie algebra. We are going now to construct two
real forms of g using its canonical generators (see (1.11)). Consider the complex
conjugate Lie algebra g. Clearly, g is semisimple, and any maximal toral sub-
algebra t of g is a maximal toral subalgebra of g, too. Consider the root space
decomposition (1.4) of g with respect to t. Clearly, any root subspace g of g
is the root subspace of g corresponding to the root t . Thus, the system of
= { | }, where is the system of roots of g. It follows
roots of g is
that the roots of g determine the same real form t(R) of t as the roots of g do,
and |t(R) = |t(R), . Identifying with its restriction to t(R), we
get = . By Proposition 2 (ii) and (II.5), the Killing forms of g and of g in-
duce the same scalar product on t(R) and also on t(R) . Now, choosing a Weyl
chamber D of t(R) for g, we see that it is a Weyl chamber for g, too. Denoting
by = {1 , . . . , l } the corresponding system of simple roots for g (and for g), it
follows that the canonical system of generators {hi , ei , fi | i = 1, . . . , l} of g is at
the same time a canonical system of generators of g, with the same Cartan matrix
A. By (II.22), we get an isomorphism : g g such that
(hi ) = hi , (ei ) = ei , (fi ) = fi , i = 1, . . . , l . (6)
2 2
Then is an automorphism of g leaving invariant our generators. Thus, = id,
and so is a real structure in g. The corresponding real form g is the real
subalgebra of g generated by {hi , ei , fi | i = 1, . . . , l}. It is called the normal (or
split ) real form.
Example 2. As in Example 1, consider the Lie algebra g = sln (C). As t, we may
take the subalgebra of all diagonal matrices
H = diag(x1 , . . . , xn ) , x1 + . . . + xn = 0 .
In particular, rk sln (C) = n 1.
18 2. Complexification and real forms

Let Eij denote the matrix with (i, j)-entry 1, all other entries being 0. Then

[H, Eij ] = (xi xj )Eij , H t.

It follows that = {xi xj | i = j}, and exi xj = Eij . Also hxi xj = Eii Ejj .
Hence, t(R) is the set of all real diagonal matrices with trace 0.
Let us x the Weyl chamber

D = {diag(x1 , . . . , xn ) | x1 < . . . < xn } t(R) .

Then

+ = {xi xj | i < j} ,
= {1 , . . . , n1 } , where i = xi xi+1 , i = 1, . . . , n 1 .

We may take the following canonical generators:

hi = Eii Ei+1,i+1 , ei = Ei,i+1 , fi = Ei+1,i , i = 1, . . . , n 1 . (7)

It follows that the normal real form is sln (R), and the corresponding real structure
is given by (X) = X.
To get another real form, we rst construct an involutive automorphism of g
which will be important in the sequel. With the Weyl chamber D xed above we
can associate the subset D t(R) which is, clearly, a Weyl chamber, too. The
corresponding system of simple roots is = {1 , . . . , l }, and its Cartan
matrix coincides with that of . As a canonical system of generators of g, we
may choose hi = hi , fi , ei , i = 1, . . . , l. By (II.22), we get a unique
automorphism Aut g such that

(hi ) = hi , (ei ) = fi , (fi ) = ei , i = 1, . . . , l . (8)

Clearly, 2 = id. The automorphism is called the Weyl involution.


Now, = , since both sides are antiautomorphisms of g and coincide on
the canonical generators. It follows that

= = (9)

is a real structure in g, giving a real form g .


Example 3. Let us consider again the Lie algebra g = sln (C). In this case, the
Weyl involution is given by

(X) = X  , X sln (C) .

In fact, the mapping X  X  is an automorphism of g, and one sees from (8)


that it coincides with on the generators (7). Therefore,

(X) = X  , X sln (C) .

It follows that the corresponding real form g = sun .


2. Complexification and real forms 19

Example 4. Due to Proposition 3, the existence of real structures on a complex


semisimple Lie algebra g implies that gR (C)
g g or, equivalently, that gR is
isomorphic to a real form of g g. If we x, e.g., the real structure on g dened
by (9) and identify g with g via x  (x), then the real structure given in the
proof of this proposition is identied with

(x, y) = ( (y), (x)) , x, y g . (10)

The real form (g g) = {(x, (x)) | x g} is isomorphic to gR by the projection


(x, (x))  x.
Example 5. Let g be the complex Lie algebra with basis e0 , e1 , e2 and commutation
relations [e0 , e1 ] = e1 , [e0 , e2 ] = ie2 , [e1 , e2 ] = 0. Prove that g possesses no real
forms.
3. Real forms and involutive automorphisms
By Proposition 2.1, the classication of real forms of a given complex Lie algebra
g up to isomorphy is equivalent to the classication of antiinvolutions in g up to
conjugacy by automorphisms of g. E. Cartan proved (see [4]) that for a semisimple
complex Lie algebra g, the word antiinvolutions in this trivial proposition may be
replaced by involutions, i.e., involutive automorphisms of g. This description of
real semisimple Lie algebras by involutive automorphisms of complex semisimple
Lie algebras is a powerful tool in the study and classication of real Lie algebras,
their subalgebras and homomorphisms. It is based on the existence of a compact
real form of g, i.e., of a real form which is a compact Lie algebra.
Let g be an arbitrary complex Lie algebra possessing real structures, and let
us x a real structure in g. Then there exists an obvious bijection between
real structures in g and certain automorphisms of this Lie algebra. Namely, we
assign to any real structure the automorphism = of g. We will say that
the automorphism corresponds to the real structure . In particular, to =
the identity automorphism e corresponds. Clearly, this correspondence depends
on the choice of . The following properties of this correspondence are deduced
directly.
Proposition 1.
(i) The correspondence  = is a bijection of the set of all antiinvolutions
of g onto the set of all Aut g satisfying

= 1 .

(ii) We have 2 = e if and only if = . In this case, the real form g is


invariant under and under .
(iii) If corresponds to and Aut g, then the automorphism  , corresponding
to 1 , has the form
 = ( )1 . 

Actually, the correspondence  is a special case of the description of the set


of k-forms of an K-object X dened over a eld k in terms of the Galois cohomology
set H 1 (, Aut X), where is the Galois group of the eld extension K k (see
[21], Ch. III). In our case, we deal with the cohomology set H 1 (Z2 , Aut g). The
assertion (i) of Proposition 1 means that the function 0  e, 1  is a 1-cocycle
of Z2 = {0, 1}, while (iii) expresses the cohomology relation between two cocycles,
and (ii) asserts that invariant cocycles correspond to involutive automorphisms.
Let now a complex semisimple Lie algebra g be given. In this case, a compact
real form always exists (see Theorem 1 below), and we choose the corresponding
real structure as the xed real structure in the correspondence described above.
E. Cartan proved that any real structure in g is conjugate to a real struc-
ture commuting with , which gives an involution = by Proposition 1 (ii)
(or, equivalently, that any Galois cohomology class contains an invariant cocycle).
Following [11], Ch. III (see also [19], 5.1), we use for constructing such a structure
a Hermitian scalar product in g associated with .
3. Real forms and involutive automorphisms 21

More generally, with any real structure on g we associate the function h :


g g C given by
h (x, y) = k(x, y) , x, y g , (1)

where k = kg is the Killing form of g. Clearly, it is linear in the rst argument


and antilinear in the second one.

Proposition 2.
(i) The function h is a Hermitian form coinciding with kg on the real form g .
(ii) h (x, y) = h (x, y), x, y g, for any Aut g satisfying = .
(iii) h ([x, u], v) = h (u, [x, v]) for x g , u, v g.
(iv) The real form g is compact if and only if the Hermitian form h is positive
denite.

Proof. (i) Using Proposition 2.2 (ii), we get

h (y, x) = k(y, x) = k(x, (y)) = h (x, y) ,

and so h is hermitian. Clearly, h (x, y) = k(x, y), whenever x, y g .


(ii) follows trivially from the invariance of k under (see (I.3)), and (iii) from
the invariance of k.
(iv) If h is positive denite, i.e., h (x, x) > 0 for any non-zero x g, then,
by (i), the restriction k|g is negative denite. But this restriction coincides with
kg (Proposition 2.2 (i)), and hence g is compact by (I.12). Conversely, if g is
compact, then, by (I.12), h is positive denite on the real form g . Since h is
hermitian, it is positive denite on g. 

Now we are going to show that each complex semisimple Lie algebra g has a
compact real form. More precisely, we prove that the real form g constructed in
2 is compact. We will use the notation of 2. The real structure was dened
by (2.9), in terms of the canonical generators ei , fi , hi corresponding to a simple
root system = {1 , . . . , l }. We note some properties of .

Proposition 3. We have |t(R) = id, (g ) = g , .

Proof. The rst relation follows from the denition of . To prove the second one,
we apply to the equation

[t, e ] = (h)e , h t(R) .

Since (h) R, we get

[h, (e )] = (h) (e ) , h t(R) .

By linearity, this is also true for all t t, whence (e ) g . 


22 3. Real forms and involutive automorphisms

Proposition 4. Any w W is of the form w = |t(R), where Int g has the


following properties: (t) = t, = .
Proof. By (II.17), any w W is a product of reections ri = ri for some simple
roots i . Now, by (II.12), ri = i |t(R), where i = i = exp(ad 2 (ei fi )).
Thus, we may choose as a product of i , and it is sucient to verify that
i = i . Since Aut gR , we have i 1 = exp(ad 2 (ei fi )) = i . 
Theorem 1. The real form g is compact.
Proof. Due to Proposition 2 (iv), it suces to show that the Hermitian form h is
positive denite. Proposition 3 and (II.4) imply that the root space decomposition
(1.4) is orthogonal with respect to h . Therefore we have to verify that h is
positive denite on any g , , and on t. By (II.11), we have h (ei , ei ) =
k(ei , fi ) = k(ei , fi ) = (i2,i ) > 0, i = 1, . . . , l. Let be an arbitrary
root. By (II.19), there exists w W such w () = i . Applying Proposition
4, we get Int g such that (t) = t, w = |t(R) and = . Now, by (II.11),
we have (gi ) = g . But leaves h invariant (Proposition 2 (iii)), and hence h
is positive denite on g . Finally, we see that h (h, h) = k(h, h) = k(h, h) > 0
for any non-zero h t(R) (see (II.4)). Since h is Hermitian, this is also true for
any non-zero h t. 
Example 1. For g = sln (C) we have (X) = X and (X) = X  , and hence
(X) = X  (see Example 2.3). It follows that sun is a compact real form
of sln (C). This is also evident from the fact that the corresponding Lie group
SUn SLn (C) is compact and from (I.10).
Now we discuss some well-known facts from Hermitian geometry which will be
useful in what follows. Let E be a nite dimensional vector space over the eld
K = C or R, endowed with a scalar product, i.e., a positive denite Hermitian (for
K = C) or symmetric bilinear (for K = R) form ( , ). Denote n = dimK E. For
any linear operator (endomorphism) gl(E), the adjoint operator is dened
by the condition
(x, y) = (x, y) , x, y E .
A linear operator is called self-adjoint if = (one also says that is Hermit-
ian (in the complex case) or symmetric (in the real one). Let S(E) gl(E) denote
the vector subspace of all self-adjoint operators. Self-adjoint operators can be also
described as linear operators admitting a system of eigenvectors with real eigen-
values which forms an orthonormal basis of E. This is equivalent to the following

k
property: E admits an orthogonal direct decomposition E = i=1 Ei , where
Ei is the eigenspace of corresponding to the eigenvalue i R. A self-adjoint
operator is said to be positive denite if (x, x) > 0 for all non-zero x E
or, equivalently, if all its eigenvalues are positive. Let P(E) GL(E) denote the
subset of all positive denite self-adjoint operators. It is open in S(E).

k
Take S(E) and let E = i=1 Ei be the corresponding eigenspace de-
composition. One sees easily that the operator exp admits the same eigenspace
decomposition, with the eigenvalues exp i > 0 instead of i . Thus, exp P(E),
and we have the mapping exp : S(E) P(E). This mapping is bijective. In
3. Real forms and involutive automorphisms 23

fact, we can dene the inverse mapping log sending a positive denite self-adjoint
operator with eigenvalues i > 0 to the linear operator admitting the same
eigenspace decomposition, but with the eigenvalues log i R.
Any P(E) can be included into a unique real one-parameter subgroup
(t), t R, of the Lie group GL(E), lying in P(E) and satisfying (1) = . In
fact, if (t) is a one-parameter subgroup with these properties, then  (0) S(E)
and = (1) = exp  (0). Thus,  (0) = log is uniquely determined by . On
the other hand, the formula (t) = exp(t log ) gives, clearly, the desired subgroup.
If t Z, then
(t) = t .
The same notation will be used for arbitrary t R. This is suggested by the
following fact: if i > 0 are the eigenvalues of , then the operator (t) can be
given by the same eigenspace decomposition as with the eigenvalues ti .
If , P(E) satisfy = 1 , where GL(E), then log = (log) 1 and
= t 1 for any t R. In the complex case the same is true for any invertible
t

antilinear operator in E. This can be proved directly, using the eigenspace


decompositions.
Lemma 1. Let g be a complex or real Lie algebra of nite dimension endowed by
a scalar product, and let Aut g P(g). Then t Aut g for all t R. If g is
semisimple, then t Int g for all t R.

k
Proof. Let g = i=1 gi be the eigenspace decomposition for , where i > 0 are
the eigenvalues. If x gi , y gj , then

[x, y] = [x, y] = [i x, j y] = i j [x, y] ,

whence [x, y] gi j . Therefore

t [x, y] = (i j )t [x, y] = [ti x, tj y] = [t x, t y]

for any t R. It follows that t Aut g. The last assertion follows from (I.7). 
A compact real structure in a complex Lie algebra is a real structure, whose
corresponding real form is compact. In this case, any conjugate real structure is
compact as well. From now on, we suppose that a compact real structure in a
given complex semisimple Lie algebra g (existing by Theorem 1) is xed. We will
study the correspondence  = between antiinvolutions and automorphisms
of g (see the beginning of the section). Let us regard g as the Hermitian vector
space with the scalar product ( , ) = h (see (1)).
Lemma 2.
(i) For any antiinvolution , we have = S(g), and thus 2 P(g)Aut g.
(ii) For any involution Aut g, we have = ( )2 P(g) Aut g.
Proof. (i) To prove that = , we get, using (I.3),

(x, y) = k(x, y) = k(x, 1 y) = k(x, y) = (x, y) .

for any x, y g. Clearly, 2 is an positive denite automorphism.


24 3. Real forms and involutive automorphisms

(ii) Again, using (I.3), we get

(x, y) = k(x, y) = k(x, 1 y) = k(x, y) = k(x, y) = (x, y) ,

for any x, y g. Thus, = . Now,

h (x, x) = k(x, x) = k( x, x) = k( x, x)
= k( x, x) = (( )x, ( )x) > 0 for any x = 0 . 

Given a complex Lie algebra g and two real structures , in g, we say that
real forms g , g are compatible if = .
Proposition 5. Let g be a complex semisimple Lie algebra. If two compact real
forms of g are compatible, then they coincide.
Proof. Let , be two compact real structures in g, such that = . We want
to prove that = . To do this, consider the involution = . We have to
show that = id (this is equivalent to = 1 = ). Since is an automorphism
and g is a real form of g, it is sucient to prove that x = x for each x g .
By Proposition 1 (ii), g is invariant under , and hence we have the eigenspace
decomposition g = g+ g , where g = {x g | (x) = x} (see (1.2)). To
show that the second summand is 0, we use the positive denite Hermitian forms
h and h given by (1). If there is a non-zero x g , then h (x, x) > 0 and
h (x, x) > 0. But h (x, x) = k(x, x), while h (x, x) = k(x, x) = k(x, x). This
gives a contradiction. 
Proposition 6. For any real structure in g, there exists Int g such that

 = 1 satises  =  , i.e., the real forms g = (g ) and g are
1
compatible. We may choose = 4 , where = ( )2 .
Proof. Consider the automorphism = Aut g. By Lemma 2, = 2
P(g) Aut g. By Lemma 1, we get the one-parameter subgroup t Int g, t R.
Using Proposition 1 (i), we see that = 1 . This implies, by Lemma 2, that
t = t , t R. Clearly, we also have 1 = , whence t 1 = t , t R.
We will search the desired inner automorphism in the form = t for a certain
t R. We can write

(1 ) = t t = t t = t t = 2t ,
(1 ) = t t = t t = t 1 t = 2t 1 .

These operators coincide if and only if 4t = 2 = 1 , i.e., if t = 14 . Thus,


1
= 4 is the desired automorphism. The last assertion follows from the property
t
of discussed before Lemma 1. 
Now we can conclude that any two compact real structures are conjugate by an
inner automorphism of g.
3. Real forms and involutive automorphisms 25

Corollary. For any two compact real structures , in g there exists Int g such
that = 1 . Thus, any complex semisimple Lie algebra admits a unique, up
to conjugacy by inner automorphisms, compact real structure. The automorphism
can be chosen in the form described in Proposition 6.
Proof. Apply Proposition 6 to the real structure and use Proposition 5. 
Now we prove a similar proposition, where instead of the antiinvolution an
involution is considered.
Proposition 7. Given an involution Aut g, there exists Int g such that
1
1 commutes with . We may choose = 4 , where = ( )2 .
Proof. The proof is similar to that of Proposition 6. One considers the auto-
morphism = ( )2 P(g) Aut g (see Lemma 2 (ii)). Then one veries that
1
= 4 is the desired inner automorphism. 
Now we dene the desired bijection between the conjugacy classes of antiin-
volutions and involutions of a complex semisimple Lie algebra g. As above, we
assume that a compact real structure is xed. For any real structure , we
can choose a real structure  = 1 such that  =  , where Int g is
the inner automorphism constructed in Proposition 6. Then we assign to the
automorphism =  , which is involutive due to Proposition 1 (ii).
Theorem 2. Let g be a complex semisimple Lie algebra. The mapping 
described above determines a bijection between the conjugacy classes of antiinvolu-
tions and involutions by inner automorphisms (or by automorphisms) of g. This
bijection does not depend on the choice of the compact real structure in g.
Proof. First we shall prove that our mapping actually determines a mapping of
conjugacy classes. Suppose that we have two antiinvolutions , 1 which are con-
jugate by an inner automorphism. Then the corresponding antiinvolutions  , 1
commuting with are also conjugate by an inner automorphism, i.e., 1 =  1 ,
where Int g. Clearly, 1 commutes with and 1 . By Corollary of Propo-
sition 6, there exists Int g commuting with 1 , such that 1 = 1 .
Then = 1 commutes with and satises 1 =  1 . Clearly, for the
corresponding involutions =  and 1 = 1 we have 1 = 1 . Thus, we
have a well dened mapping of conjugacy classes.
Now we show that this mapping is bijective. First we prove the surjectivity.
Let an involution Aut g be given. By Proposition 7, there exists Int g
such that  = 1 satises  =  . Then =  is an antiinvolution. Since
= , our mapping sends the conjugacy class of to that of  .
Next we show the injectivity. Suppose that two antiinvolutions 1 , 2 are given
and that 1 = 1 and 2 = 2 satisfy 2 = 1 1 for a certain Aut g.
Here i are antiinvolutions conjugate to i and commuting with . Clearly, we
may assume that i = i and i = i , i = 1, 2. Then 2 commutes with
and 1 . By Corollary of Proposition 6, there exists Int g commuting with
2 , such that 1 = 1 . Then = 1 commutes with and satises
2 = 1 1 . Clearly, 2 = 1 1 .
26 3. Real forms and involutive automorphisms

Finally, let us consider another compact real structure 1 in g and the corre-
sponding bijection between conjugacy classes of antiinvolutions and involutions.
We claim that it coincides with the original bijection determined by . In fact,
by Corollary of Proposition 6, we have 1 = 1 , where Int g. Let
   = , where  is conjugate to and commutes with , be the
mapping giving the original bijection. Then, clearly, the new bijection is deter-
mined by the mapping   1  1 = (  1 )1 = 1 . This implies
our assertion.
Clearly, the above argument also gives a bijection between the conjugacy classes
of antiinvolutions and involutions by arbitrary automorphisms of g. 
Example 2. By the bijection of Theorem 2, to the class of compact structures
the identical involution = e = id corresponds. Now, if is the normal real
structure given by (2.5), then the corresponding involutive automorphism is the
Weyl involution = given by (2.7).
Example 3. In Example 2.1 we described certain real structures in sln (C). All
these structures commute with the compact real structure (X) = X  (see
Example 1). We list the involutive automorphisms = corresponding to various
real structures .
For the normal real structure (X) = X we get (X) = (X) = X .
1
For (X) = Sm XSm , n = 2m, we get (X) = Sm X  Sm
1
= (AdSm )(X T ).

For (X) = Ip,q X Ip,q we get (X) = Ip,q XIp,q .
Example 4. Let g be a complex semisimple Lie algebra. Due to Example 2.4, we
may regard gR as the real form of g g corresponding to the antiinvolution
given by (2.9). Clearly, is a compact real structure on g g commuting
with . Thus, the involution, corresponding to the real form gR by Theorem 2, is
= ( ) : (x, y)  (y, x), x, y g.
4. Automorphisms of complex semisimple Lie algebras
We present here some main facts about the automorphism group Aut g of a com-
plex semisimple Lie algebra g. Suppose that a maximal toral subalgebra t of g
is chosen and let denote the corresponding system of roots. By (II.12), for
any Aut g leaving t invariant, the transformation  of t (R) maps onto
itself. Let us also choose a Weyl chamber in t(R), and let denote the
corresponding subset of simple roots. Consider the subgroups of Aut g dened by

Aut(g, t) = { Aut g | (t) = t} ,


Aut(g, t, ) = { Aut(g, t) |  () = } .

Write = {1 , . . . , l } and let A = (aij ) denote the corresponding Cartan ma-


trix. Any bijection s : can be regarded as a permutation s Sl . The
bijection is called an automorphism of if s leaves invariant the matrix A, i.e.,
if as(i)s(j) = aij for all (i, j). Denote by Aut the group of automorphisms of
. If Aut(g, t, ), then |t(R) is an orthogonal transformation, and hence
( )1 induces an automorphism of . Clearly, the mapping :  ( )1 | is
a homomorphism of the group Aut(g, t, ) to Aut .
The homomorphism is surjective. Moreover, there exists a homomorphism
: Aut Aut(g, t, ) such that = id. To see this, consider the canonical
system of generators {hi , ei , fi | i = 1, . . . , l} given by (1.11). By (II.22), for any
s Aut there exists a unique automorphism s of g satisfying

s(ei ) = es(i) , s(fi ) = fs(i) , s(hi ) = hs(i) , i = 1, . . . , l . (1)

Clearly, s Aut(g, t, ). Also (s) = s. In fact, for any i, j we have

(s (i ))(hj ) = i (s(hj )) = i (hs(j) ) = ai,s(j) = as1 (i),j = s1 (i )(hj ) ,

whence s | = s1 . Now, st  = st, and hence we get the homomorphism :


s  s of Aut into Aut(g, t, ) with the desired property = id. This implies
that is surjective, is injective, and we have the semidirect decomposition
Aut(g, t, ) = Ker  (Aut ).
The automorphism (s) = s dened by (1) is called the diagram automorphism
of g corresponding to s Aut . In what follows, we denote by the same symbol s
the linear extension of s onto t(R) . The above argument shows that (s )1 = s
on t(R) . One also sees easily that any diagram automorphism commutes with
the real structures , and the Weyl involution associated with (see 2).
Proposition 1. The normal subgroup Ker Aut(g, t, ) coincides with T =
exp(ad t), and we have the semidirect decomposition

Aut(g, t, ) = T  (Aut ) . (2)

Proof. We only have to prove the assertion concerning Ker . Clearly, exp(ad h)
acts trivially on t for any h t. Conversely, suppose that Aut(g, t, ) satises
28 4. Automorphisms of complex semisimple Lie algebras

() = e. Then |t = id. By (II.12), (ei ) = ci ei , (fi ) = di fi , (hi ) = hi , where


ci , di C \ {0}, i = 1, . . . , l. The relation [ei , fi ] = hi implies di = c1 i , i =
1, . . . , l. Choosing h0 t in such a way that i (h0 ) = log ci , i = 1, . . . , l, we see
that and exp(ad h0 ) coincide on the generators ei , fi . Hence = exp(ad h0 ) T .

We are now going to extend this semidirect decomposition to the entire group
Aut g.
Theorem 1. Let g be a complex semisimple Lie algebra. Then

Aut g = Int g  (Aut ) , (3)

and the corresponding projection Aut g Aut coincides with on Aut(g, t, ).


Proof. First we prove that Aut g = (Int g) Aut(g, t, ). If Aut g, then (t) is
a maximal toral subalgebra of g, and, by (II.1), there exists Int g such that
(t) = t. By (II.12), permutes the Weyl chambers in t(R). Let D be the
Weyl chamber corresponding to . Then, by (II.15), we have w((D)) = D
for an element w W . By (II.13), w = |t(R) for an automorphism Int g.
Then (D) = D, and hence 0 = Aut(g, t, ). We got the desired
decomposition = ()1 0 , where ()1 Int g.
Since T Int g, this decomposition together with (2) gives the decomposition

Aut g = (Int g)(Aut ) .

Then (II.16) implies that Int g (Aut ) = {e}. Hence, the latter decomposition
is semidirect. 
Corollary 1. We have
Aut g/ Int g
Aut . 

In other words, the group Out g = Aut g/ Int g is isomorphic to the group of all
symmetries of the Dynkin diagram of g. For the simple Lie algebras this group is
indicated in Table 1. For the non-simple Lie algebras, there are also symmetries
which permute isomorphic simple components of g.
Corollary 2. There is the semidirect decomposition

Aut(g, t) = N  (Aut ) ,

where N = Aut(g, t) Int g. 


Example 1. Consider the Lie algebra g = sln (C). We want to use Theorem 1 for
the classication of involutive automorphisms of g. The result is that any such
automorphism is conjugate in Aut g (by an inner automorphism) to one of the
automorphisms listed in Example 3.3.
If n 3, then we see from Table 1 that Aut
Z2 . The automorphism
: X  X  is in this case an outer one. In fact, if is an inner automorphism,
then = Ad C for a certain C SLn (C) (see (I.2)). Then X  = CXC 1 , X
4. Automorphisms of complex semisimple Lie algebras 29

sln (C). But, e.g., X = diag(2, 0, . . . , 0, 1, 1) cannot satisfy to such an equation,


since X and X  have dierent eigenvalues.
Let us classify rst the outer involutive automorphisms. By Theorem 1, any
outer automorphism has the form = , where Int g. Writing = Ad C,
where C SLn (C), we have

(X) = CX  C 1 , X sln (C) .

The relation 2 = id is equivalent to the condition

X = (C(C  )1 )X(C(C  )1 )1 , X sln (C) .

Thus, C(C  )1 should commute with all X sln (C). This is true precisely in the
case, when C(C  )1 = In , where C , which can be rewritten as C = C  .
Transposing this relation, we get C  = C. It follows that C = 2 C, whence
= 1. Thus, we get two dierent cases.
I. If = 1, then we have C  = C. The reduction theory of symmetric bilinear
forms implies that C = U U  , where the matrix U can be chosen from SLn (C).
Then

(X) = U U  X  (U  )1 U 1 = (Ad U )(Ad U )1 (X) , X sln (C) .

Thus, = (Ad U )(Ad U )1 is conjugate to by an inner automorphism of g.


II. If = 1, then we have C  = C and n = 2m. Using the reduction
theory of skew-symmetric bilinear forms, we see that C = U Sm U  , where U
SLn (C) and Sm is given by (2.5). It follows that = (Ad U )(Ad Sm )(Ad U )1 is
conjugate to (Ad Sm ) by an inner automorphism of g.
Note that these two cases give two dierent conjugacy classes of automorphisms.
In fact, in the case I we have g
son (C), while in the case II g
spn (C), and
these Lie algebras are non-isomorphic for any even n 2.
Next, we consider the case
III. is an inner involutive automorphism. Suppose that

(X) = CXC 1 , X sln (C) ,

where C SLn (C). The condition 2 = id is equivalent to C 2 X(C 2 )1 = X, X


sln (C), whence C 2 = In , where C . Multiplying C by an appropriate scalar,
we can assume that C 2 = In , but C / SLn (C). Then C is diagonalizable with
eigenvalues 1, i.e., C = U Ip,q U 1 , where U GLn (C) and Ip,q is given by (2.4).
This implies that = (Ad U ) Ad Ip,q (Ad U )1 is conjugate to Ad Ip,q in the group
Aut g. By considering the subalgebras g , one sees easily that Ad Ip,q , p q, are
pairwise non-conjugate.
Now let us consider the case n = 2. Then all the automorphisms are inner ones,
and so we have only the case III. One sees that is either trivial, or conjugate to
Ad I1,1 . Thus, we have classied all involutive automorphisms up to conjugacy in
Aut g.
30 4. Automorphisms of complex semisimple Lie algebras

Using this classication, we can apply Theorem 3.2 to describe all the real forms
of sln (C). We see that any real form is isomorphic to one of the forms described
in Example 2.1. In the case n 3 all these forms are pairwise non-isomorphic.
In the case n = 2 all the non-compact real forms are isomorphic: sl2 (R)
su1,1 ,
while sl1 (H) = su2 .
Now we illustrate Theorem 1 by determining the corresponding decomposition
of the Weyl involution Aut g, dened by (2.8) in terms of the canonical
generators {hi , ei , fi | i = 1, . . . , l}. Here we use the notation of the beginning
of this section. Clearly, Aut(g, t), |t(R) = id, and so (D) = D is the
Weyl chamber opposite to D. By (II.15), there exists a unique w0 W such
that w0 (D) = D. Since w02 (D) = D, we have w02 = e. Thus, (w0 )(D) = D,
and hence w0 induces an involutive automorphism Aut . Then we get
the diagram automorphism which is involutive as well. Following the proof of
Theorem 1, we can write
= ,
where N satises |t(R) = w0 . Our goal is to describe the automorphism
explicitly.
Consider the element 
h= h t(R) . (4)
+

Lemma 1. We have (h) = 2 for any . In particular, h D.


Proof. The element h can be regarded as dual to the linear form 2 (see (1.10)),
and our assertion is similar to (II.20). To prove it, one rst deduces from (1.7),
(1.8) and (1.9) the relation

r (h ) = hr () , ,

(we remind that r () , due to (II.14)). Then we see that


 
r (h) = r (h ) = hr () .
+ +

If , then, by (II.20), r () + , except of the case = , when r () =


. This implies that r (h) = h 2h , and our assertion follows from (1.8). 
Clearly, we have

l
h= ri hi , (5)
i=1

where ri N. The numbers ri are important invariants of the Lie algebra g. For
simple complex Lie algebras, they are given in Table 4.
We also dene the elements e, f g by


l

l

e= ri ei , f= ri fi . (6)
i=1 i=1
4. Automorphisms of complex semisimple Lie algebras 31

Then {e, h, f } is an sl2 -triple. In fact, Lemma 1 implies easily the relations [h, ei ] =
2ei and [h, fi ] = 2fi , whence [h, e] = 2e and [h, f ] = 2f . Also [e, f ] = h, due to
(1.12). It follows that the correspondence
     
0 1 0 0 1 0
E=  e , F =  f , H=  h (7)
0 0 1 0 0 1

determines a homomorphism q : sl2 (C) g which maps sl2 (C) isomorphically


onto s. The subalgebra s is called the principal three-dimensional subalgebra of g.
We shall prove now that s is pointwise invariant under any diagram automor-
phism.
Proposition 2. For any s Aut , we have s(x) = x, x s.
Proof. One sees easily that any s Aut permutes the positive roots and that
s(h ) = h(s )1 () = hs() , + . It follows that s(h) = h. By (1) and (5), we
have
l l
s(h) = ri hs(i) = ri hi = h .
i=1 i=1

This implies the relations

rs(i) = ri , i = 1, . . . , l , s Aut . (8)

Using (1), (6) and (8), we get s(e) = e, s(f ) = f . Thus, s|s = id. 
In the Lie group SL2 (C), take the element
 
0 1
g0 = exp (E F ) = .
2 1 0

One easily veries that (Ad g0 )(X) = g0 Xg01 = X  , X sl2 (C). Thus, 0 =
Ad g0 = exp(ad 2 (EF )) is the Weyl involution of sl2 (C) (see Example 2.3). Using
the principal three-dimensional subalgebra, we can dene a similar automorphism
Int g by

= exp(ad (e f )) . (9)
2
To relate it to , we need the following simple lemma, which will be used later in
various situations and therefore is formulated in a general form.
Lemma 2. Suppose a homomorphism f : h g of arbitrary Lie algebras be given.
Then for any Int h we have f = f , where Int g. If = exp(ad x), x
h, then we may take = exp(ad f (x)). If we have a homomorphism F : H G of
the corresponding connected Lie groups such that de F = f and if = Ad g0 , g0
H, then we may take = Ad F (g0 ).
Proof. If = Ad g0 , where g0 H, then we may write F (g0 (h)) = F (g0 hg01 ) =
F (g0 )F (h)F (g0 )1 = F (g0 ) (F (h)), h G, i.e., F g0 = F (g0 ) F . Dierentiating
this relation, we get f = (Ad F (g0 ))f , and so we may set = Ad F (g). In
32 4. Automorphisms of complex semisimple Lie algebras

the case when = exp(ad x) = Ad(exp x), x h, we have  = Ad F (exp x) =


exp(ad f (x)). Note that the Lie group homomorphism always exists if we choose
H to be simply connected. 
Returning to our situation, we denote by G a connected Lie group with Lie
algebra g. Since SL2 (C) is simply connected, we may consider a Lie group homo-
morphism Q : SL2 (C) G such that de Q = q. Let us denote

g = Q(g0 ) = exp( (e f )) G . (10)
2

Due to the denition (9), = Ad g. Clearly, Lemma 2 implies the relation

q0 = q . (11)

Proposition 3. The automorphism given by (9) and the element g G dened


by (10) possess the following properties:
(i) (h) = h, (e) = f , (f ) = e.
(ii) (t(R)) = t(R), |t(R) = w0 .
(iii) 2 = e, g 2 = exp(ih) Z(G).
(iv) s = s for any s Aut .
(v) = = .
Proof. (i) follows immediately from (11).
To prove (ii), we note that, by Lemma 1, h is a regular element of t(R), and hence
t = zg (h). On sees from (i) that leaves t and hence t(R) invariant. By (II.16),
|t(R) W . Since h D, we have (D) = D, and therefore |t(R) = w0 .
Clearly, g02 = I2 = exp(iH). Applying Q, we get g 2 = exp(iq(H)) =
exp(ih). It follows from Lemma 1 that Ad g 2 (ei ) = eii (h) ei = e2i ei = ei , i =
1, . . . , l. Similarly, Ad g 2 (fi ) = fi . It follows that 2 = Ad g 2 = e. Therefore
g 2 Z(G) (see (I.2)), and (iii) is proved.
If s Aut , then, by Proposition 2,

ss1 = exp(ad s(e f )) = exp(ad (e f )) = ,
2 2

and (iv) is proved.


Now we prove (v). It follows from (ii) that  | = , i.e.,  (i ) =
(i) , i = 1, . . . , l. By (II.12), this implies that (gi ) = g(i) . Therefore
(ei ) = ci f(i) and (fi ) = di e(i) , where ci , di C \ {0}. Also (hi ) = h(i) .
Setting s = in (8), we get

r(i) = ri , i = 1, . . . , l . (12)

Since (e) = f , we have

l

l

l

( ri ei ) = ri ci f(i) = ri fi .
i=1 i=1 i=1
4. Automorphisms of complex semisimple Lie algebras 33

It follows from (12) that ci = 1, i = 1, . . . , l. Similarly, (f ) = e implies


di = 1, i = 1, . . . , l. Thus,

(hi ) = h(i) , (ei ) = f(i) , (fi ) = e(i) , i = 1, . . . , l .

Now it is easy to see that coincides with = on the canonical generators.



The involution Aut introduced above is important for the theory of
representations of semisimple Lie algebras. We are going now to describe the
explicit form of this involution.
Proposition 4.
(i) If g is a non-commutative simple complex Lie algebra, then is non-trivial
precisely in the cases g = Al , l 2; D2m+1 , m 1; E6 . In these cases,
is the only

s non-trivial symmetry of the Dynkin diagram.
(ii) If g = i=1 gi , where gi are simple, then Aut induces the s correspond-
ing involution i on each component of the decomposition = i=1 i , where
i is the system of simple roots of gi , i = 1, . . . , s (see (II.23)).
(iii) lies in the centre Z(Aut ).
Proof. (i) We have only to investigate the cases g = Al , l 2, g = Dl , l 3,
and g = E6 , since for other simple g the Dynkin diagram has no symmetries (see
Table 1).
1) g = Al = sll+1 (C), l 2.
We take t as in Example 2.2 and use the notation of this example. The Weyl
group W is the group of all permutations of the diagonal elements xi of a matrix
H = diag(x1 , . . . , xl+1 ) t. Clearly,

w0 (diag(x1 , . . . , xl+1 )) = diag(xl+1 , . . . , x1 ) .

This implies that


(i ) = li+1 , i = 1, . . . , l .
2) g = Dl = so2l (C), l 3. This Lie algebra is the algebra of all block matrices
of the form
 
X Y
, where X gll (C) , Y  = Y , Z  = Z .
Z X 

As a maximal toral subalgebra t, we may choose the subalgebra of diagonal ma-


trices
H = diag(x1 , . . . , xl , x1 , . . . , xl ) .
The system of roots is = {xi xj | i = j}. We choose the Weyl chamber

D = {H | x1 > > xl1 > |xl |} .

Then

= {1 , . . . , l } , where i = xi xi+1 , i = 1, . . . , l 1 , l = xl1 + xl .


34 4. Automorphisms of complex semisimple Lie algebras

The Weyl group is generated by permutations of the diagonal elements xi and by


the transformations of the form xi  xi , xj  xj , i = j; xk  xk , k = i, j.
We shall consider separately the cases of even and odd l.
If l = 2m, then w0 = e W transforms D into D. Thus, = id.
If l = 2m + 1, then w0 W is given by

w0 (diag(x1 , . . . , xl , x1 , . . . , xl ))
= diag(x1 , . . . , xl1 , xl , x1 , . . . , xl1 , xl ) .

It follows that (i ) = i , i = 1, . . . , l 2, (l1 ) = l .


3) g = E6 . We refer to [1], Ch. VI, 4, for a proof of the relation e / W which
implies = id.
(ii) Any Weyl chamber for g has the form D = D1 . . . Ds , where Di is a
Weyl chamber for gi , and w0 = (w0 )1 . . . (w0 )s , where (w0 )i is the element of
the Weyl group of gi mapping Di onto Di . Thus,  = 1 . . . s , which
implies our assertion.
(iii) For any s Aut , we have = ss1 Int g, where N satises
|t(R) = w0 (e.g., is given by (9)). By (II.16), w = |t(R) = (s )1 w0 s W .
But clearly w(D) = D. Therefore w = w0 , whence = w0 = s(w0 )s1 =
ss1 . 
Proposition 1 and Theorem 1 are rst steps in the study of automorphisms of
complex semisimple Lie algebras which leads, in particular, to the classication of
involutive automorphisms up to conjugacy in Aut g and hence to the classication
of real semisimple Lie algebras (see, e.g., [11], [19]). We will not give the details
of this classication and only prove an old result of Gantmacher [8] establishing
the so called canonical presentation of an involutive automorphism. We formulate
it in a modern form (it is actually a special case of Theorem 4.4.3 of [19]). The
proof will use the following remark.
Remark 1. Suppose a compact real form u of a complex semisimple Lie algebra g be
given. Then we can construct a maximal toral subalgebra t of g in the following
way: we choose an arbitrary maximal commutative subalgebra t0 of u and set
t = t0 (C). By Proposition 3.1, (ii), the operators ad u, u u, in g are skew-
Hermitian with respect to a scalar product, and hence diagonalizable with pure
imaginary eigenvalues. It follows that t is a toral subalgebra. This is a maximal
commutative subalgebra of g, since it coincides with the centralizer zg (t0 ). We also
note that t(R) = it0 . In fact, all roots of g with respect to t are pure imaginary
on t0 and real on it0 .
Theorem 2. Let an involutive automorphism Aut g be given. Then
(i) There exist a maximal toral subalgebra t g and a subset of simple roots
of the corresponding system of roots such that Aut(g, t, ).
(ii) is conjugate by an inner automorphism to an automorphism of the form
1 = s = s, where s Aut and = exp(ad 2 it), the element t t
satisfying s(t) = t and (t) 2Z, .
(iii) Any involutive automorphism 1 of the form described in (ii) commutes with
the real structures , and the Weyl involution associated with (see 2).
4. Automorphisms of complex semisimple Lie algebras 35

Proof. (i) First we show the existence of a maximal toral subalgebra t g which
is invariant under . By Proposition 3.7, there exists a -invariant compact form
u of g. Since is involutive, we have the eigenspace decomposition u = u+ u ,
where u corresponds to the eigenvalue 1 (see (I.14)). Let a denote a maximal
commutative subalgebra of u+ . Then the centralizer

t = zg (a) = {x g | [x, a] = 0}

is the desired maximal toral subalgebra. Let us denote

t0 = t u = zu (a) .

It suces to show that t0 is commutative. In fact, it is easy to verify that t = t0 (C).


If t0 is commutative, then, clearly, it is a maximal commutative subalgebra of u,
and due to Remark 1, t is a maximal toral subalgebra of g. Obviously, (t) = t.
By (I.11), t0 = z s, where z and s are respectively the centre and the commu-
tator subalgebra of t0 . Clearly, (t0 ) = t0 . Due to maximality of a, this implies
that the corresponding eigenspace decomposition has the form t0 = a b, where
b u and a z. Since this is a Z2 -grading (see I.14), we have s = [b, b] a. But
s is semisimple, and hence s = 0. Thus, t0 is commutative, and t is a -invariant
maximal toral subalgebra. By Remark 1, t(R) = it0 .
Next, we want to prove that leaves invariant a Weyl chamber in t(R). To do
this, we note that ia contains a regular element. In fact, let denote the system
of roots relative to t. If all the elements of ia are singular, then there exists
such that (a) = 0. Hence, e , e zg (a) = t, which contradicts to the fact
that t is commutative (see (II.11)). Thus, there exists a Weyl chamber D t(R)
such that D ia = . Further, any element of this intersection is invariant under
. It follows that (D) = D, and hence the corresponding system of simple roots
satises  () = . Thus, Aut(g, t, ).
(ii) Applying Proposition 1, we see that = s, where s Aut , s2 = id, and
= exp(ad x), where x t. We also have |t = s|t. Consider the corresponding
eigenspace decomposition t = t+ t . Then x t decomposes as x = 12 (x +
s(x)) + 12 (x s(x)). Denoting u = 12 (x + s(x)) t+ and v = 12 x, we can write
x = u + s(v) v. Then ad x = ad u + ad s(v) + ad(v) = ad u + s(ad v)s + ad(v),
and so = exp(ad u)s exp(ad v)s(exp(ad v))1 . Denoting = exp(ad u), we have
ss = exp(ad(s(u))) = . Therefore, = exp(ad v)(s)(exp(ad v))1 is conjugate
to 1 = s = s.
Since and s commute, we have 2 = e. It follows that

2 (e ) = exp(ad 2u)(e ) = e(2u) e = e

for all , whence (2u) 2iZ, . Denoting u = 2 it, we get (t)


2Z, , and 1 is of the desired form.
(iii) Since , , commute with s, we only have to prove that they commute
with . Clearly, u = u, whence = . Now, u = u, whence = 1 =
. Finally, = , since = . 
5. Cartan decompositions and maximal compact subgroups
Here we return to the correspondence between involutions (involutive automor-
phisms) and real forms of a complex semisimple Lie algebra established in 3. Let
a complex semisimple Lie algebra g and a compact real structure in g be given,
and let u = g denote the corresponding compact real form of g. Then any invo-
lution Aut g commuting with determines a real structure = . It is easy
to understand, how to get the real form g0 = g directly from u and .
The involution determines the eigenspace decomposition g = g+ g . Since
commutes with and , both real forms u and g0 are stable under . Thus, we
have the eigenspace decompositions

g0 = (g0 )+ (g0 ) , where (g0 ) = g g0 ,


(1)
u = u+ u , where u = g u .

Let us denote
k = (g0 )+ , p = (g0 ) .
Since = coincides with on g0 , we have k = u+ , p = iu . Thus, the
decompositions (1) have the form

g0 = k p , (2)
u = k ip . (3)

These decompositions are Z2 -gradings (see (I.14)), i.e.,

[k, k] k , [k, p] p , [p, p] k . (4)

Denote by k the Killing form kg and its restriction to g0 which coincides with kg0
by Proposition 2.2, and consider the real bilinear form

b (x, y) = k(x, y) , x, y g0 . (5)

Clearly, b is the restriction of the positive denite Hermitian form h and hence
is a scalar product (a positive denite symmetric bilinear form). It follows that

k(x, x) < 0 for x k , x = 0 ; k(y, y) > 0 for y p , y = 0 . (6)

Now let us start with a real semisimple Lie algebra g0 . A direct sum decompo-
sition
g0 = k p (7)
is called a Cartan decomposition if it is a Z2 -grading and if the Killing form k = kg0
satises (6). As we have seen above, the decomposition (2) of g0 constructed with
the help of a compact real form u of the complexication g = g0 (C), compatible
with g0 , is a Cartan decomposition. We will prove that any Cartan decomposition
of g0 can by obtained in this way.
5. Cartan decompositions and maximal compact subgroups 37

Suppose that a Cartan decomposition (7) is given. By (I.14), the involutive


transformation of g0 dened by
(x + y) = x y , x k, y p. (8)
is an automorphism of g0 . Extend to an (involutive) automorphism of g = g0 (C)
denoted by the same symbol. Clearly, = , where is the complex conjugation
in g relative g0 . Therefore = is an antiinvolution commuting with . We
claim that it is compact. In fact, for any x k, y p we have
h (x + y, x + y) = k(x + y, (x + y)) = k(x + y, (x + y))
= k(x + y, x y) = k(x, x) + k(y, y) .
Due to (6), this value is positive whenever x+ y = 0. Since h is a Hermitian form,
it follows that it is positive denite on g = g0 (C). Further, if (x + y) = x + y for
x k(C), y p(C), then (x y) = x + y, whence x k, y ip. Thus, u = g is
described by (3), and therefore k = u+ , p = iu .
Theorem 1. Each Cartan decomposition of a real semisimple Lie algebra g0 has
the form (2), where k = u+ , p = iu , for a compact real form u of g0 (C), com-
patible with g0 . Any two Cartan decompositions of g0 are conjugate by an inner
automorphism of g0 .
Proof. The rst assertion has been already proved. Suppose that two Cartan
decompositions of g0 are given, and let and 1 be the compact real structures
on g = g0 (C), commuting with the complex conjugation relative to g0 and
determining our decompositions. By Corollary of Proposition 3.6, there exists
Int g such that 1 = 1 and = . Then leaves g0 invariant and
maps g onto g1 . It follows that transforms the rst Cartan decomposition
1
into the second one. Actually, this automorphism is expressed as = 4 , where
{ | t R} is a 1-parameter subgroup of Int g consisting of positive denite
t

symmetric operators with respect to h . We have t = exp(t ad z), where z g,


and the condition = implies (ad z) = ad((z)) = ad z. Thus, z g0 ,
and |g0 is the inner automorphism exp( 14 ad z) of g0 . 
Example 1. Example 4.1 allows to write down the Cartan decompositions g0 = kp
for all real forms g0 of sln (C).
(i) For g0 = sln (R):
k = {X sln (R) | X  = X} = son , p = {X sln (R) | X  = X} .
(ii) For g0 = slm (H), n = 2m:
k = {X slm (H) | X  = X} = spm , p = {X slm (H) | X  = X} .
(iii) For g0 = sup,q :
  
X 0
k= | X up , Y uq , tr X + tr Y = 0 ,
0 Y
  
0 Z
p= | Z is a (p q)-matrix .
Z  0
38 5. Cartan decompositions and maximal compact subgroups

Example 2. Let g be a complex semisimple Lie algebra. Let us nd a Cartan


decomposition of the real Lie algebra gR . Consider gR as a real form of g g
(see Examples 2.4 and 3.4). Then one should take the compact real structure
on g g, where is a compact real structure on g, and the involution
: (x, y)  (y, x), x, y g. Clearly, we get

k = {(x, x) | x u} , p = {(ix, ix) | x u} .

The projection (x, (x))  x maps k onto u and p onto iu, and thus we get the
Cartan decomposition in the form

gR = u iu . (9)

Let us prove some properties of the Cartan decompositions which will be useful
in what follows. We remind that a Cartan decomposition of a real semisimple
Lie algebra g0 gives rise to the scalar product (5). We regard g0 as the euclidean
vector space with this scalar product.
Proposition 1.
(i) = ( )1 for any Aut g0 . In particular, = .
(ii) ad (x) = (ad x) for any x g0 .
Proof.

(i) (1 x, y) = k(1 x, y)
= k(x, y) = (x, y)

for any x, y g0 . Applying this to = , we see that = .

(ii) ((ad x)y, z) = k((ad x)y, z)


= k(y, (ad x)z)
= (y, (ad x)z) = (y, (ad (x))z)

for any x, y, z g0 . 
Now we go over to the multiplicative Cartan decompositions of real semisimple
Lie groups. For simplicity, we will consider only the automorphism groups of real
semisimple Lie algebras, referring to [11, 19] for the general case. The basic fact
here is the well-known polar decomposition theorem for linear operators in an
euclidean vector space E. As in 3, we denote by S(E) and P(E) respectively
the vector space of symmetric linear operators and its subset of positive denite
operators. We remind that we have the bijective mapping exp : S(E) P(E) and
denote by log its inverse. Let also O(E) denote the orthogonal group of E.
Lemma 1. The mapping exp : S(E) P(E) is real bianalytic, and thus log is an
analytic mapping.
Proof. Since exp is analytic, it suces to prove that d0 exp : S(E) S(E) is
an isomorphism for any 0 S(E). Take S(E) such that (d0 exp) = 0.
5. Cartan decompositions and maximal compact subgroups 39

Denoting (t) = 0 + t and (t) = exp (t), we have (t)(t) = (t)(t) for all
t R. Dierentiating this relation at t = 0, we get
exp 0 + 0  (0) =  (0)0 + (exp 0 ).
Since  (0) = (d0 exp) = 0, this implies that commutes with exp 0 , and
hence with 0 . Therefore (t) = (exp 0 )(exp t), whence 0 =  (0) = (exp 0 ),
and so = 0. 
Proposition 2. Let E be an euclidean vector space. Then the mapping : O(E)
P(E) GL(E) given by (k, p) = kp and the mapping (id exp) : O(E)
S(E) GL(E) are bianalytic.
Proof. By Lemma 1, exp : S(E) P(E) is bianalytic. We repeat the well-known
argument from linear algebra. Take u O(E) and p = exp s P(E), where s
S(E), and denote a = up = u exp s. Then a = pu1 , whence p2 = exp 2s = a a.
It follows that
1 1 1
s = log(a a), p = exp( log(a a)) , u = a exp( log(a a)) . (10)
2 2 2
Thus, p, s and u are expressed uniquely as analytic functions in a. 
Let now g0 be a real semisimple Lie algebra, endowed with a Cartan decom-
position (7). Consider the involutive automorphism of g0 given by (8) and the
bilinear form (5), where k is the Killing form of g0 . We regard g0 as the euclidean
vector space with the scalar product (x, y). Let us denote
K = Aut g0 O(g0 ) , P = Aut g0 P(g0 ) .
Clearly, K is a compact linear Lie group. By Proposition 1, K = { Aut g0 |
= }, and ad k ad g0 is the Lie subalgebra corresponding to K.
The multiplication mapping from Proposition 2 sends K P to Aut g0 . Now,
Proposition 1 (ii), implies that ad z S(g0 ), and hence exp(ad z) P for any
z p.
Theorem 2. The mappings : K P Aut g0 and (id exp ad) : K p
Aut g0 are bianalytic and induce bianalytic maps K P Int g0 and K p
Int g0 . We have
K = K Int g0 = { Int g0 | = } .

Proof. To prove our claim concerning Aut g0 , it suces to verify that for any
a Aut g0 the operators s S(g0 ) and p P(g0 ) expressed by (10) belong to
ad p and Aut g0 , respectively. By Proposition 1 (i), we see that a Aut g0 , and
therefore p2 Aut g0 . By Lemma 3.1, pt Int g0 . Hence s = log p = ad z for a
certain z g0 . Since (ad z) = ad z, we see from Proposition 1 (ii) that (z) = z,
i.e., z p.
Since the mapping (id exp ad) : K p Aut g0 is a homeomorphism, it
should map the connected components of K p onto those of Aut g0 . It follows
that K p is mapped onto Int g0 . Hence (K P ) = Int g0 . The decomposition
Int g0 = K P implies the last assertion of the theorem. 
40 5. Cartan decompositions and maximal compact subgroups

Corollary. The subgroups K and K are maximal compact subgroups of the Lie
groups Aut g0 and Int g0 , respectively.
Proof. Suppose that we have a linear group L such that K  L Aut g0 . By
Theorem 2, L = K(L P ), where L P = {e}. Take p L P, p = e. Then the
sequence (p, p2 , . . . ) in L has no limit points, due to Lemma 1, since the sequence
(log p, 2 log p, . . . ) is unbounded. Therefore L is non-compact. The same argument
applies to the subgroup K of Int g0 . 
Our next aim is to prove that any two maximal compact subgroups of the group
Aut g0 or Int g0 are conjugate. Note that Theorem 1 implies the following weaker
assertion: any two maximal compact subgroups of Int g0 corresponding to Cartan
decompositions are conjugate in Int g0 . But it is not clear, whether each maximal
compact subgroup of Int g0 corresponds to a Cartan decomposition.
Theorem 3. For any compact subgroup L of Aut g0 or Int g0 , there exists q P
such that qLq 1 K. Any two maximal compact subgroups in Aut g0 or Int g0
are conjugate by an element of Int g0 .
The idea of the proof is as follows. We dene an analytic action  T of the
linear Lie group G = Aut g0 on the manifold P given by

T (p) = p , G, pP. (11)

Note that p P = G P(g0 ). In fact, G by Proposition 1 (i), (p ) =


p , and

(p (x), x) = (p( (x)), (x)) > 0 , x g0 \ {0} ,

since p is positive denite. Also T = T T , Te = id. This action is transitive. In


1 1
fact, any p P may be written as p = p 2 ep 2 = T 12 (e). The stabilizer of the point
p
e P is K, and hence the homogeneous space P of the group G can be identied
with G/K (or with G /K , where the identity component G = Int g0 ). The
crucial fact is that any compact subgroup L G leaves invariant a point q P . If
it is proved, then one sees easily that T 1 1
1 LT 1 (e) = e, and so T 1 LT 1 K. In
2 2
q2 q q2 q
the original Cartans proof of the conjugacy theorem (see, e.g., [11]), the existence
of a xed point was shown by a nice geometrical argument, using the G-invariant
Riemannian metric on the symmetric space P = G/K. The proof which follows
(see [19]) is quite elementary.
As suggested above, we have to nd a point in P which is xed under the
subgroup L by the action (11). This point will be constructed as a minimum
point of an L-invariant function on P . This function possesses a certain convexity
property which implies the uniqueness of its minimum point. Thus, it should be
xed under L.
Let us consider the more general situation, when an arbitrary euclidean vector
space E of dimension n is given. We dene the following analytic function on
P(E) P(E):
r(, ) = tr( 1 ) , , P(E) . (12)
5. Cartan decompositions and maximal compact subgroups 41

Further, for any compact subset P(E), we dene the function

() = max r(, ) , P(E) . (13)


One proves easily that it is continuous.


Let us give an explicit expression of r. For a xed P(E), choose an or-
thonormal basis v1 , . . . , vn of E such that (vi ) = i vi , i = 1, . . . , n. For any
P(E), we have 1 P(E); let 1 be expressed by the matrix (gij ) relative
to our basis. Then i > 0, gii > 0, i = 1, . . . , n, and we have


n
r(, ) = i gii , P(E) . (14)
i=1

We use this formula for establishing some properties of functions r and .


Let  denote the usual norm of a linear operator . For P(E), we have
 = maxi i , where i , i = 1, . . . , n, are the eigenvalues of .
Lemma 2. Let us x a compact subset P(E). Then there exists b > 0 such
that
() b , P(E) . (15)

Proof. We x P(E) and use the expression (14). Since the set of all orthonor-
mal bases (i.e., O(E)) and are compact, there exists a constant b > 0 such that
gii b, i = 1, . . . , n, for all and all orthonormal bases of E. Then


n
r(, ) b i b , P(E) , .
i=1

Clearly, this implies (15). 


Lemma 3. For any compact subset P(E), the function attains its minimum
on each subset F P(E) which is closed in S(E).
Proof. Take any 0 F . Then the set B = { F | () (0 )} is compact.
In fact, B is closed in S(E), since is continuous, and is contained in the ball
 1b (0 ), due to Lemma 2. Let 1 B be a minimum point of on B.
Since 0 B, we have (1 ) (0 ) < () for all F \ B. Thus, 1 is the
desired minimum point. 
Now we recall the following classic denition. A function f (t), t R, is called
strictly convex if for any a, b R, a < b, the graph of f between a and b lies under
the line segment linking the point (a, f (a)) with (b, f (b)), i.e., if

f (b) f (a) bt ta
f (t) < f (a) + (t a) = f (a) + f (b) , a < t < b. (16)
ba ba ba

It is well known that a smooth function f satisfying f  (t) > 0 for all t R is
strictly convex.
42 5. Cartan decompositions and maximal compact subgroups

Lemma 4. Let F (t, s) be a continuous function on R , where is a compact


space, and dene f (t) = maxs F (t, s), t R. If F (t, s) is a strictly convex
function of t for any xed s , then f is strictly convex.
Proof. For any t R, choose a point s(t) such that F (t, s(t)) F (t, s), s .
Then for a < t < b we get, using (16),

bt ta bt ta
f (t) = F (t, s(t)) < F (a, s(t)) + F (b, s(t)) f (a) + f (b) .
ba ba ba ba

We need this lemma to establish the following important fact.
Lemma 5. Choose an P(E) and a compact P(E). Then the function

f (t) = ( t ), t R,

is strictly convex.
Proof. Using an appropriate orthonormal basis of E, we can write, due to (14),


n 
n
r( t , ) = gii ti = gii et log i ,
i=1 i=1

where i > 0 are the eigenvalues of . Since gii > 0, the analytic function
t  r( t , ) is strictly convex for any xed P(E). By Lemma 4, f (t) =
max r( t , ) is strictly convex, too. 
Proof of Theorem 3. We apply the above considerations to the euclidean space
g0 with the scalar product (5). The set P = exp(ad p) is closed in Int g0 due to
Theorem 2. Note that tr ad x = 0 for any x g0 , since ad g0
g0 coincides with its
commutator subalgebra. Thus, ad g0 sl(g0 ). Using this inclusion, one deduces
that Int g0 SL(g0 ) is closed in gl(g0 ). It follows that P is closed in S(g0 ). Now we
consider the function r on P P which is the restriction of (12) and the function
on P given by (13), where is a compact subset of P . Applying Lemma 3 to
the subset F = P , we see that there exists 0 P such that (0 ) () for
all P . We want to prove that 0 is the only minimum point of .
One sees immediately from (11) that r is invariant under the action  T of
G = Aut g0 on P , i.e.,

r(T (), T ()) = r(, ) , G.

It follows that
(T ()) = T1 () () , G. (17)
Since the action is transitive, we may assume that 0 = e. Suppose that P is
another minimum point of . Consider the function

f (t) = ( t ) , t R.
5. Cartan decompositions and maximal compact subgroups 43

Clearly, 0 and 1 are two minimum points of f . But f is strictly convex, due to
Lemma 5, and hence f (t) < f (0) = f (1) for 0 < t < 1. This gives a contradiction.
Now, take as the orbit L(e) = { | L} of L under the action (11).
This is an L-invariant compact subset of P . By (17), the corresponding function
is L-invariant. Thus, L(0 ) = 0 , and the existence of the point xed under L
is proved.
As we saw above, this implies that qLq 1 K for a certain q P G =
Int g0 . Now, if L is a maximal compact subgroup of G, then qLq 1 is maximal
compact as well, and hence qLq 1 = K. If L is a maximal compact subgroup of
G , then qLq 1 K G = K , whence qLq 1 = K . 
Example 3. Let g be a complex semisimple Lie algebra. By Example 2, any Cartan
decomposition of the real Lie algebra gR has the form (9), where u is a compact real
form of g. Consider the Lie group G = Int g = Int gR . Theorem 2 implies that the
connected Lie subgroup U G corresponding to the subalgebra ad u ad gR is a
maximal compact subgroup of G. Applying Theorem 3, we see that any maximal
compact subgroup of G corresponds to a compact real form of g.
To nish, we note that the following generalization of Theorem 3 is true: any two
maximal compact subgroups of a Lie group G with a nite number of connected
components are conjugate by an element of G (see, e.g., [9]).
6. Homomorphisms and involutions of complex
semisimple Lie algebras
Here we start to study the behavior of real forms of complex semisimple Lie al-
gebras under homomorphisms of these algebras. More precisely, suppose a homo-
morphism f : g h of complex semisimple Lie algebras and a real form g0 of g be
xed. We would like to know, for which real forms h0 of h the inclusion f (g0 ) h0
holds. We will prove a theorem due to Karpelevich [15], giving an answer in terms
of the involutive automorphisms corresponding to the given real forms.
First we will make some preliminary remarks.
Let a homomorphism f : g h of arbitrary complex Lie algebras be given.

Let and  be real structures in g and h, respectively. Then f (g ) h if and

only if f = f . In fact, if the latter relation holds, then we get immediately

f (x) =  (f (x)) for any x g . Conversely, suppose that f (g ) h . Then f

and f coincide on the real form g . Since both mappings are antilinear, they
should coincide on the entire vector space g.
Generally, we say that a mapping a : g g extends by f to a mapping a : h
h, whenever f a = a f . In this case, we write a f a . The subscript f may be
omitted if it is clear which homomorphism is considered.
We have proved above the following
Proposition 1. Let f : g h be a homomorphism of complex semisimple Lie
algebras and let and  be real structures in g and h, respectively. Then f (g )

h if and only if f  . 
Now we note certain simple properties of the extension relation, assuming that
f : g h is xed.
Proposition 2.
(i) If a a and b b , then aa bb .
1
(ii) If a a and a and a are invertible, then a1 a .
(iii) An operator a in h extends idg , i.e., idg a if and only if a |f (g) = id.
 

(iv) For any Int g, there exists  Int h such that  . If = exp(ad x),
where x g, then we may take  = exp(ad f (x)).
(v) If idg , where P(h) relative to a compact real structure in h, then
idg t , t R.
Proof. The assertions (i)(iii) are trivial. Let us prove (iv). Let G and H be
connected Lie groups with tangent Lie algebras g and h such that there exists a
homomorphism F : G H satisfying de F = f (it always exist, if G is simply
connected). By (I.2), Int g = Ad G and Int h = Ad H. If = Ad g, where g G,
then we may write F (g (h)) = F (ghg 1 ) = F (g)F (h)F (g)1 = F (g) (F (h)), h
G, i.e., F g = F (g) F . Dierentiating this relation, we get f = (Ad F (g))f , and
so we may set  = Ad F (g). In the case when = exp(ad x) = Ad(exp x), x g,
we have  = Ad F (exp x) = exp(ad f (x)).
To prove (v), we note that, by (iii), idg means |f (g) = id. Thus, f (g) lies
in the eigenspace of corresponding to the eigenvalue 1. Clearly, all t , t R,
have the only eigenvalue 1 in this space. Hence idg t , t R. 
6. Homomorphisms and involutions 45

Now we will consider a homomorphism f : g h of complex semisimple Lie


algebras. Let a real structure on g extend to a real structure  on h, i.e., assume

that f (g ) h . Then, by Proposition 2 (iv), for any Int g there exists 
1 
Int h such that    f 1 or, equivalently, f ((g ))  (h ). So we may
consider the extension relation between conjugacy classes of real structures, where
the conjugacy by inner automorphisms is meant, and we see that it corresponds to
the inclusion relation between conjugacy classes of real forms in g and h. Our next
aim is to replace in this correspondence the conjugacy classes of real structures by
the conjugacy classes of involutions, using Theorem 3.2. The crucial step is the
following proposition giving a compact extension of any compact real structure.
Proposition 3. Let f : g h be a homomorphism of complex semisimple Lie
algebras.
(i) If a compact real structure on g is given, then there exists a compact real
structure  on h such that f  .
(ii) If  is another compact real structure on h extending , then there exists
Int h such that  =  1 and e f .
Proof. (i) Let us denote by u the compact real form g of g. Consider the simply
connected Lie group G with tangent Lie algebra g. There exists a homomorphism
F : G H = Int h such that de F = ad f . The connected Lie subgroup U G
corresponding to the subalgebra u is compact (see (I.13)), and hence the subgroup
F (U ) H is compact as well. Therefore F (U ) is contained in a maximal compact
subgroup V of H. By Example 5.3, any maximal compact subgroup of H corre-
sponds to a compact real form v of h. Since F (U ) V , we have ad f (u) ad v,
whence f (u) v. By Proposition 1, extends to the compact structure  on h
corresponding to v.
(ii) Let us apply Corollary of Proposition 3.6 to the compact real structures 
1
and  . It claims that  =  1 , and = 4 , where = (   )2 . Clearly,
e = ( ) f . By Proposition 2 (v), this yields e f .
2

Remark 1. If f = is a linear representation, i.e., h = sl(W ), then for proving (i)
you can use, instead of Theorem 5.3, Theorem of Weyl (I.9) claiming that there
exists a Hermitian scalar product in W , invariant under a given compact linear
group. The real form v (u) will consist of all skew-Hermitian operators, relative
to a scalar product in W invariant under R(U ), with zero trace. Here R is the
representation of the Lie group G such that de R = .
Let us x a compact real structure in g and a compact real structure  in
h such that f  . Consider the correspondence between antiinvolutions and
involutions in g dened in 3. It assigns to any real structure in g an involutive
automorphism = 1 Aut g, where 1 is a real structure commuting with
and having the form 1 = 1 , where Int g. We will say that is an
involution corresponding to (or to the real form g ), though is determined
only up to conjugacy by inner automorphisms. As was shown in Theorem 3.2, this
correspondence gives rise to a bijection between antiinvolutions and involutions.
Using the compact real structure  , we can dene a similar correspondence for h.
We will now prove the following theorem (see [15]).
46 6. Homomorphisms and involutions

Theorem 1. Let f : g h be a homomorphism of complex semisimple Lie algebras,


and let two real forms g0 g and h0 h be given. We also suppose that two
compact real structures in g and  in h, such that f  , are xed. Let
Aut g denote an involution corresponding to g0 . If f (g0 ) h0 , then extends
by f to an involution  Aut h corresponding to h0 . Conversely, if extends
to an involution  Aut h corresponding to h0 , then f (g0 ) is contained in a real
form of h which is conjugate to h0 by an inner automorphism.

Proof. Let us denote by and  the real structures determining g0 and h0 , re-
spectively. By Proposition 6.3, there exists such an Int g that 1 = 1
commutes with . Then = 1 is an involution corresponding to g0 .
First suppose that f (g0 ) h0 . Then  . By Proposition 2 (iv),  ,
where Int h, and hence 1    1 . Therefore we may assume that 1 = ,
and = . Consider now = (   )2 Aut h P(h) (see Lemma 3.2 (i)).
1
By Proposition 3.6, 1 =  1 , where = 4 , commutes with  , and then
 = 1  is an involution corresponding to h0 . But by Proposition 2 (i) e = 2 ,
and, by Proposition 2 (v), this yields e . Using Proposition 2 again, we see
that 1 , and hence  .
Conversely, suppose that extends to an involution  corresponding to h0 .
Consider = (  )2 Aut h P(h) (see Lemma 3.2 (ii)). By Proposition 3.7,
1
1 =  1 , where = 4 , commutes with  . By Proposition 2 (i), e =
2 , which implies that e . It follows that 1 , and hence 1 1 =
1  . By Proposition 2 (iv),  , where Int h. Therefore  1 1  ,
1 
and hence f (g0 )  (h1 ). On the other hand,  corresponds to both real

structures  and 1 , and Theorem 3.2 implies that h1 is conjugate to h0 by an
inner automorphism. 

Corollary 1. Let f : g0 h0 be a homomorphism of real semisimple Lie algebras


and let a Cartan decomposition g0 = k p be given. Then there exists a Cartan
decomposition h0 = k p such that f (k) k and f (p) p .

Proof. Consider the complexication of our homomorphism f (C) : g h, where


g = g0 (C), h = h0 (C), and denote by ,  the corresponding complex conjugations
in g and h, respectively. By Theorem 5.1, the given decomposition of g0 is the
eigenspace decomposition for an involution of g, commuting with and restricted
to g0 , and = is a compact real structure. By Proposition 3 (i), we have
f (C)  for a certain compact real structure  in h. Applying the rst assertion
of Theorem 1, we get an involution  and an automorphism Int h such that
 = (  1 )  , e f (C) , f (C)  and f (C)  . Then we have f (C) 1 =
1  and f (C) 1 = 1  . The desired Cartan decomposition of h0 may be
constructed as the eigenspace decomposition for 1 |h0 . 
Corollary 1 (often referred to as the canonical embedding property) was rst
proved in [14] using methods of Riemannian geometry and is equivalent to the
following fact. Let G0 be a connected semisimple Lie subgroup of a connected
semisimple real Lie group H0 and let K  denote the connected Lie subgroup of H0
corresponding to the compact part k of a Cartan decomposition of its Lie algebra
6. Homomorphisms and involutions 47

h0 = k p . Then G0 has a totally geodesic orbit in the Riemannian symmetric


space H0 /K  . It was also proved independently by Mostow [18].
The above results were specied in [15] in an important case, when f is given
by an irreducible linear representation of g. This means that h gl(V ) is a linear
complex Lie algebra and f : g h gl(V ) is an irreducible linear representation
of g in the complex vector space V . A natural generalization of this situation gives
the following denition. We say that a homomorphism f : g h of complex Lie
algebras is an S-homomorphism if for any Int h which satises e f (or,
which is the same, is identical on f (g)), we have = e.
Lemma 1. Suppose that f : g h gl(V ) is an irreducible linear representation
of g in the complex vector space V . Then f is an S-homomorphism. Conversely,
if a linear representation f : g gl(V ) of a semisimple complex Lie algebra g is
an S-homomorphism into gl(V ) or sl(V ), then f is irreducible.
Proof. Suppose that an inner automorphism Int h is identical on f (g). We
have (X) = (Ad C)X = CXC 1 , X h, for an element C of the connected Lie
subgroup H GL(V ) corresponding to h. By our assumption, CXC 1 = X for
any X g. Since f is irreducible, C = e, where C , by the Schur Lemma.
Hence = e.
If g is semisimple and a linear representation f : g gl(V ) is reducible, then f
leaves invariant a non-trivial direct sum decomposition V = V1 V2 . Constructing
the linear operator a in V such that a(v) = v, v V1 and a(v) = v, v V2 , we
get a non-trivial = Ad a Int gl(V ) such that |f (g) = id. Clearly, there exists
c C such that ca SL(V ), and = Ad(ca). 
Proposition 4. Let f : g h be an S-homomorphism of complex semisimple Lie
algebras, and let be a compact real structure on g. Then there exists a unique
compact real structure  on h such that  . If a real structure  on h extends
a real structure on g satisfying = , then   =   , and the involution
 =   extends the involution = .
Proof. By Proposition 3, there exists a compact real structure  on h which
extends . By the same proposition, any compact real structure  with the same
property has the form  =  1 , where Int h extends e. In our case = e,
and hence  =  .
Suppose now that a real structure  extends a real structure on g such that
= . An argument from the proof of Theorem 1 shows that  commutes
with a compact real structure  which extends . But  is the only compact real
structure with this property, and hence  =  . 
This proposition implies, in particular, that if an S-homomorphism is given,
then for any compact real form u of g there exists a unique compact real form v
of h such that f (u) v. We give now a strong generalization of this fact.
Let g be a complex semisimple Lie algebra and 1 , 2 Aut g. If 1 , 2 are con-
jugate by an inner automorphism Int g, then they belong to the same coset of
the group Aut g modulo Int g. In fact, we have 2 = 1 1 = 1 (11 1 )1
1 Int g. We will say that two real forms (or real structures) of g are of the
same kind if the corresponding involutions belong to the same coset modulo Int g
48 6. Homomorphisms and involutions

or, equivalently, determine the same automorphism of the Dynkin diagram of g


(see 4). The above remark shows that this denition is correct, and any two
real forms conjugate by an inner automorphism are of the same kind. The real
forms corresponding to inner involutions will be said to be of the rst kind . If
Aut g/ Int g
Z2 , then the real forms corresponding to outer involutions will be
said to be of the second kind . Thus, any real form of a simple complex Lie algebra
g, except of g
so8 , is of the rst or of the second kind. The following theorem
was proved in [15] in the case, when g is simple, non-isomorphic to so8 , and f is
given by an irreducible representation.
Theorem 2. Suppose that f : g h is an S-homomorphism. Let g0 be a real form
of g. Assume that we have two real forms h0 , h1 of h such that f (g0 ) h0 and
f (g0 ) h1 . If h0 and h1 are of the same kind, then h0 = h1 . This is true, in
particular, if h0 and h1 are conjugate by an inner automorphism of h.
Proof. Let , 1 and 2 denote the real structures determining g0 , h1 and h2 ,
respectively. Then 1 and 2 . The involution corresponding to g0 is
dened by = , where is a compact real structure on g such that = .
By the proof of Theorem 1, the involutions i which extend and correspond to
hi , i = 1, 2, are given by 1 = 1  and 2 = 2  , where  is the only compact
structure on h which satises  (see Proposition 4). By Proposition 2 (i),
e = 2 1 2 . But 1 2 Int h, since 1 and 2 belong to the same coset modulo
Int h. Therefore 1 2 = e, and 1 = 2 , whence 1 = 2 . 
The name S-homomorphism is suggested by the notion of S-subalgebra in-
troduced by Dynkin in [6]. A subalgebra f of a complex semisimple Lie algebra h is
called regular if f is normalized by a maximal toral subalgebra of h. In particular,
any subalgebra of maximal rank (i.e., containing a maximal toral subalgebra of
h) is regular. A subalgebra f h is called an R-subalgebra if it is contained in a
proper regular subalgebra, and an S-subalgebra otherwise.
Proposition 5. A homomorphism f : g h of complex semisimple Lie algebras
is an S-homomorphism if and only if f (S) is an S-subalgebra in h, i.e., is not
contained in any proper regular subalgebra.
Proof. First we prove that for any semisimple R-subalgebra f there exists a non-
trivial Int h such that |f = id. If f is an R-subalgebra, then f f, where
f is a maximal regular subalgebra. It follows (see [6]) that f is either a maximal
parabolic or a maximal semisimple subalgebra of maximal rank. In the rst case,
f lies in a Levi subalgebra of f, and hence its centralizer in h is non-trivial. In the
second case, one proves that f = h for a non-trivial Int h of a nite order (see
[9], Ch. 6).
Conversely, suppose that for a semisimple subalgebra f h there exists a non-
trivial Int h such that |f = id. Then f h  h. If the automorphism is
semisimple, then = exp(ad h), where h = 0 lies in a maximal toral subalgebra t
of h. Then t h , and hence h is regular, and f is an R-subalgebra. In the general
case, the Jordan decomposition in the algebraic group Aut h gives = s exp ad y,
where s is a semisimple inner automorphism, and y h is a nilpotent element of
h. If s = e, then h hs  h, and by the above argument f is an R-subalgebra.
6. Homomorphisms and involutions 49

If s = e, then h = zh (y) is the centralizer of a nilpotent element y = 0. By


Morozovs theorem, we can include y into a sl2 -triple {e0 = y, h0 , f0 } spanning
a simple three-dimensional subalgebra s h. It is known (see [9], 6.2) that the
centralizer zh (s) is a maximal reductive subalgebra of zh (y). By Maltsevs theorem,
we may assume that f zh (s). Then f zh (h0 ). But h0 is semisimple, and, by
the above argument, zh (h0 ) is regular. Thus, f is an R-subalgebra. 
Example 1. Suppose that h gl(V ) is a semisimple complex linear Lie algebra.
If a subalgebra f h is irreducible, then it is an S-subalgebra of h, see Lemma 1.
By the same lemma, the converse is true if h = sl(V ). One also can prove that
any semisimple S-subalgebra of so(V ), dim V odd, or of sp(V ), is irreducible. But
this is false for h = so(V ), dim V even. For example, the subalgebra
 so2r+1  (C)
X 0
so2s+1 (C) h = so2(r+s+1) (C), r, s 1, consisting of matrices , is an
0 Y
S-subalgebra. This gives an example of an S-homomorphism into h = so2m (C)
which is a reducible linear representation.
7. Inclusions between real forms under an
irreducible representation
In this section, the following problem will be studied. Let : g sl(V ) be a linear
representation of a complex semisimple Lie algebra g in a complex vector space V
of dimension n, and let an involution Aut g be given. We want to know, to
which involutions  Aut sl(V ) does extend by . As Theorem 6.1 shows, this
problem is equivalent to that of nding inclusions between real forms of complex
Lie algebras g and sl(V )
sln (C) determined by . We mainly will consider the
case, when is irreducible. Then, by Theorem 6.2, the extending involution  is
unique in its coset modulo inner automorphisms.
Let us choose a maximal toral subalgebra t of a given complex semisimple Lie
algebra g and a Weyl chamber D t (R). As usually, we denote by + the
corresponding subsets of simple and positive roots and by hi , ei , fi , i = 1, . . . , l,
the corresponding canonical generators of g. Let also , and = be dened
by (2.6), (2.8) and (2.9). For g = sln (C), we denote these (anti)involutions by
0 , 0 , 0 , respectively, and dene them as in Examples 2.2 and 2.3:

0 (X) = X , 0 (X) = X  , 0 (X) = X  . (1)

Irreducible representations of g will be described by their highest weights rela-


tive to t and D (see (III.9)). The following simple lemma will be useful.
Lemma 1. Let be an irreducible representation of a complex semisimple Lie al-
gebra g. Then, for any s Aut , we have s = s( ). The highest weight of
the representation s is s(), where is the highest weight of .
Proof. The rst assertion follows immediately from (III.8), since s = s1 (see
4). Moreover, any weight vector v of is a weight vector of s corresponding
to the weight s() s . In particular, consider the highest vector v of . Then
(4.1) implies that v is a highest vector of s. Thus, s() is the highest weight of
s. 
In our study of the extension problem, we will also use the representation R :
G SLn (C) of the simply connected Lie group G with Lie algebra g, such that
de R = .
Two inclusions between real forms are easy to establish. Due to Proposition
6.3, for any compact real form u of g there exists a compact real form u of sl(V )
such that (u) u . By Remark 6.1, u is the subalgebra of all skew-Hermitian
operators with zero trace with respect to a scalar product in V invariant under
R(U ), where U is the connected Lie subgroup of G corresponding to u. Also,
by (III.11), V admits a basis consisting of weight vectors such that the images
of canonical generators (hi ), (ei ), (fi ), i = 1, . . . , l, are expressed by real
matrices (to get such a basis, you may choose any maximal linearly independent
subset of the set of vectors vi1 ,... ,ik ). Then the corresponding matrix form of
satises (g0 ) sln (R), where g0 is the normal real form of g. Now we are going
to present these facts in a more precise form.
Clearly, we may replace a given representation : g sl(V ) by an equivalent
one. Usually we will replace by an equivalent representation in Cn , i.e., by its
7. Inclusions between real forms 51

matrix form : g sln (C) relative to a basis of V . If we change the basis, then
the new matrix form will be x  C (x)C 1 = (Ad C)(x), x g, where the
matrix C GLn (C) is the transition matrix from the new basis to the old one.
After choosing an appropriate basis in V , we will usually denote the matrix form
of a representation by the same symbol .
Proposition 1. Let : g sl(V ) be a linear representation of a complex semisimple
Lie algebra g. Then there exists a basis of V such that for the corresponding matrix
form : g sln (C) the following conditions are satised:

0 ; (2)
0 ; (3)
(x), x t , are diagonal matrices. (4)

Under the conditions (2) and (3) we have

0 . (5)

The conditions (2), (3) are equivalent, respectively, to the inclusions

(g ) sln (R) , (6)


(g ) sun .

(7)

Proof. First note that the equivalence of (2) and (6) and of (3) and (7) follows
from Proposition 6.1. Clearly, (6) means that (g ) leaves invariant the real span
of the chosen basis, while (7) means that the basis is orthonormal with respect
to a Hermitian scalar product invariant under R(U ). Finally, the condition (4)
means that all the vectors of the basis are weight vectors with respect to t.
Due to (III.2) and (III.3), we may assume that is irreducible. As was men-
tioned above, it follows from (III.11) that V admits a basis in which all (x), x
g , are expressed by real matrices. Then the corresponding matrix form of sat-
ises (6), and hence (2). Now, it follows from Proposition 6.4 that there exists
a unique compact real structure  on sln (C) which extends . Since = ,
we see from the same proposition that 0  =  0 . But the standard compact
real structure 0 also commutes with 0 . By Proposition 3.6,  = 0 1 , where
Int sln (C) and 0 = 0 . Then the matrix representation 1 = 1 satises
both (2) and (3). In fact, we have

1 = 1 = 1 0 = 0 1 = 0 1 ,
1 = 1 = 1  = 0 1 = 0 1 .

Now, = Ad C, where C GLn (C), and hence 1 (x) = C 1 C is the matrix


form of corresponding a new basis of V .
We also have to satisfy the condition (4) (it was satised by our rst basis,
but may be lost by changing it). We see from (6) and (7) that in the constructed
basis we have i(x) sun , (x) sln (R) for any x t(R). This means that all
52 7. Inclusions between real forms

these (x) are symmetric real matrices. Since they form a commutative family,
there exists a matrix D On such that all the matrices from D(t(R))D1 are
real diagonal ones. Going over to the new basis with transition matrix D1 , we
complete the proof. 
It is clear (and was used at the end of the proof) that if we change a basis of
V satisfying the conditions (2) and (3) of Proposition 1, using a real orthogonal
transition matrix, then the new basis satises the same conditions. The following
proposition shows that this is the only way to get a new basis satisfying the
conditions (2) and (3), provided that is irreducible.
Proposition 2. Let : g sl(V ) be an irreducible linear representation of a com-
plex semisimple Lie algebra g. Suppose that we x a basis of V such that (2) and
(3) are satised. If we have another basis of V with the same property, then the
transition matrix is of the form C = cB, where c C and B On .
Proof. Let C denote the transition matrix to the new basis. Consider the repre-
sentation  = 1 : g sln (C), where = Ad C. Clearly, the relations  0
and  0 imply, respectively, 0 1 and 0 1 . It follows from
Proposition 6.4 that 0 1 = 0 . Thus,  = 0 1 commutes with 0 , and
we get an involution  =  0 . Clearly,  . Since  = 0 1 , (5) and
Theorem 6.2 imply that  = 0 . Hence,  = 0 . Thus, commutes with 0 , 0
and 0 .
This allows to deduce our assertion, using simple calculations with matrices.
For any X sln (C), we have

C XC 1 = CXC 1 = C X C 1 ,

whence C = aC, a C. Applying the complex conjugation, we get C = aC,


whence aa = 1, i.e., |a| = 1. Then C1 = a C GLn (R). Thus, we may assume
that C GLn (R). Exploiting 0 = 0 , we get
1
C(X  )C 1 = (CXC 1 ) = C  X C  ,

for any X sln (C). It follows that C  C = bIn , where b R. Clearly, b > 0, and
hence B = 1b C On . 
Before studying our main problem, we consider an application of Proposition 1
to dual representations.
Proposition 3. For any linear representation of a complex semisimple Lie algebra
g we have
,
where Aut is the automorphism of the Dynkin diagram of g described in
Proposition 4.4. If is irreducible with highest weight , then the highest weights
of is ().
Proof. The condition (5) means that ((x)) = (x) . Thus, in the basis, chosen
in Proposition 1, has the same matrix form as (see (III.1)). It follows that
7. Inclusions between real forms 53

. Now, by Proposition 4.3 (v), = , where = exp(ad 2 (e f )) (see


(4.9)). Using Proposition 6.2 (iv), we get
= = Ad R(g) ,
where g G is given by (4.10). By (5), this implies our rst assertion. The claim
about highest weights follows from Lemma 1. 
A representation is called self-dual if .
Corollary 1. An irreducible linear representation with highest weight is self-
dual if and only if () = . 
In the sequel we will consider linear representations in their matrix form, i.e.,
as homomorphisms : g sln (C). Let be an involution of g, and suppose that
there exists an involution  of sln (C) such that  . If is irreducible, then
only one outer  and only one inner  extending may exist, due to Theorem
6.2. Let g0 denote a real form of g corresponding to (see 3).
Suppose that  is an outer automorphism. As we saw in Example 4.1,  =
(Ad B)0 , where 0 is given by (1), and B GLn (C) satises B  = B. The
number 
+1 if B  = B ,
j(g, , ) =
1 if B  = B
is called the Karpelevich index of the pair (, ). If j(g, , ) = 1, then (g0 ) is
conjugate to a subalgebra of sln (R). If j(g, , ) = 1, then n = 2m is even, and
(g0 ) is conjugate to a subalgebra of slm (H).
If  is an inner automorphism, then, by Example 4.1,  is conjugate (by an
inner automorphism) to Ad Ip,q for p q, the matrix Ip,q being given by (1.4). In
this case, (g0 ) is conjugate to a subalgebra of sup,q . The number q p 0 is
called the signature of the pair (, ).
As usually, we consider the real forms of g and the corresponding involutions
up to conjugacy by inner automorphisms of g (see Theorem 3.2). Therefore we
may assume that the given involution is of the form described in Theorem 4.2
(ii). More precisely,
= s , s Aut , = exp(ad u) ,
(8)
where u = it t , s(t) = t , (t) 2Z for all .
2
In particular, is an involutive inner automorphism commuting with , and
(see Theorem 4.2 (iii)). Let us denote z = exp u G.
Proposition 4.
(i) The element z 2 = exp(it) lies in the centre Z(G) of G.
(ii) Let : g sln (C) be an irreducible representation with highest weight .
Then R(z)2 = ei(t) In .
(iii) If the condition (4) is satised, then
1
R(z) = e 2 i(t) A , (9)
where A is a diagonal matrix from On , A2 = In .
54 7. Inclusions between real forms

Proof. (i) We have Ad z 2 = (Ad z)2 = 2 = e, and hence z 2 Z(G), by (I.2).


(ii) By the Schur Lemma, (i) implies R(z)2 = R(z 2 ) = aIn , where a C.
Applying R(z)2 to the highest vector v of , we get

R(z)2 v = e2(u) v = ei(t) v .

Thus, a = ei(t) .
1
(iii) It follows from (4) that the matrix A = e 2 i(t) R(z) is diagonal. By (ii),
A2 = In . Therefore the diagonal elements of A are 1, and hence A On . 
We want now to nd the outer automorphisms of sln (C) to which extends
by an irreducible representation and to deduce a formula expressing the index
j(g, , ) in terms of the highest weight of . Assuming that is of the form (8),
we denote
s0 = s = s (10)
(see Proposition 4.4 (iii)). Clearly, s20 = e.
Theorem 1. Let : g sln (C) be an irreducible representation of a complex
semisimple Lie algebra g with highest weight , and let denote an involutive
automorphism of g of the form (8). An outer involutive automorphism  of sln (C)
satisfying  exists if and only if

s0 () = . (11)

Under this condition, we have

j(g, , ) = (1)(t+h) , (12)

where h t(R) is given by (4.4).


Proof. By Proposition 1, we may assume that satises the conditions (2) (4).
We will also use the notation of Proposition 4 and Proposition 4.3.
Let us suppose that =  for an outer involutive  Aut sln (C). By
Example 4.1,  = (Ad B)0 , where B SLn (C) satises B  = jB, j = j(g, , ).
Using Proposition 6.2 (iv), we get = exp(ad u)s = (Ad R(z))s, whence

(Ad R(z))s = (Ad B)0 . (13)

Thus, s 0 (see Proposition 3), whence s() = (), due to


Lemma 1. Thus, (11) holds.
Conversely, suppose that s() = (), whence s . Then

= (Ad R(z))s = 0 .

Hence there exists B GLn (C) satisfying = (Ad B)0 . We show that the
outer automorphism  = (Ad B)0 is involutive. Clearly, e = 2  . But
2

2 2
Int sln (C). Since is irreducible, = e (see Lemma 6.1).
7. Inclusions between real forms 55

To calculate the index, consider again the relation (13). Using Proposition 4
(iii), we may replace R(z) in this formula by a diagonal matrix A On such
that A2 = In . Clearly, s commutes with and , and hence the left side of (12)
satises the conditions (2) and (3). Clearly, the same is true for 0 . Therefore,
by Proposition 2, we may assume that B On . Under this assumption, we clearly
have
B 2 = jIn . (14)
Now, by (5), 0 = . Using Proposition 4 (iii) and Proposition 4.3 (v), we
get from (13) the following relation:

(Ad A)s = (Ad B) exp(ad (e f )) .
2

By Proposition 6.2 (iv), we can rewrite it in the following form:

s = Ad(ABR(g)) , (15)

where g G is given by (4.10).


We also note that R(g) = exp ( 2 (e f )) SOn . In fact, (e f ) = e f , (e
f ) = (e f ), and by (6) and (7), the matrix ( 2 (e f )) is real skew-symmetric.
Moreover, Proposition 4.3 (iii) implies the following relation:

R(g)2 = exp(i(h)) = (1)(h) In . (16)

In fact, since g 2 Z(G), the Schur Lemma implies that R(g)2 = R(g 2 ) is a scalar
matrix. Applying this operator to the highest vector v of , we get

R(g)2 v = exp(i(h))v = ei(h) v = (1)(h) v ,

since

l
(h) = ri i Z , (17)
i=1

by (4.5).
One more relation we can obtain if we apply the rst and the last terms of (12)
to u = 2 it g. We get (u) = (Ad B)((u)) = B(u)B 1 . By exponentiating,
one gets R(z) = BR(z)1 B 1 . Using (9), we get BAB 1 = ei(t) A. It follows
that ei(t) = 1. Hence (t) Z and

BAB 1 = (1)(t) A . (18)

Now we will restrict both sides of (15) to the principal 3-dimensional subalgebra
s g constructed in 4. By Proposition 4.2, any canonical automorphism s, s
Aut , acts on s trivially. It follows from (14) that the matrix

C = ABR(g)
56 7. Inclusions between real forms

commutes with any (x), x s. Hence C leaves invariant any weight subspace
of |s. The weights of |s are restrictions of the weights of to hR and may
be identied with the corresponding eigenvalues of the operator (h). Thus, for
l
any we identify |hR with the integer (h) = i=1 ri (hi ), where ri
are dened by (4.5). In particular, to the highest weight of there corresponds
the non-negative integer m = (h) given by (17). By (III.11), any other weight
of has the form = i1 . . . ik , k > 0, and, due to Lemma 4.1, the
corresponding weight of |s is identied with (h) = m 2k < m. Thus, the
highest vector v Cn is the only, up to a scalar factor, weight vector of |s
corresponding to the weight m. It follows that Cv = cv for a certain c C. We
claim that c2 = 1.
In fact, by the condition (4), the standard basis of Cn consists of weight vectors
of . But v , up to a complex factor, should enter into this basis, since dim V = 1.
Thus, we may assume that v is one of the basic vectors and, in particular, that
v Rn . Thus, v is a real eigenvector of the matrix ABR(g) On , and hence
c = 1.
Now we calculate the matrix C 2 . Note that C commutes with R(g), since g
belongs to the connected Lie subgroup of G corresponding to s. Using (18), (14)
and (16), we may write

C 2 = (AB)2 R(g)2 = (1)(t) B 2 R(g)2 = j(1)(t+h) In .

Comparing this with the above result, we see that j(1)(t+h) = 1, but this is
equivalent to (12). 
Remark 1. One sees from the proof of Theorem 1 that the index formula (12) can
also be written in another form. As we know, g 2 = exp(ih) and z 2 = exp(it)
belong to the centre Z(G) of G, and hence R(g 2 ) and R(z 2 ) are scalar matrices.
We proved, actually, that their possible values are In and that

R(g 2 ) = (1)(h) In , R(z 2 ) = (1)(t) In . (19)

Note that the integer (h) can be calculated by the formula (17). Then (12) can
be expressed in the following form:

R(g 2 z 2 ) = j(g, , )In .

In the following examples, some special cases will be considered.


Example 1. Let us consider the Lie algebra g = sll+1 (C), l > 1. First we suppose
that the involution is an outer one. Since is the only non-trivial automorphism
of (Proposition 4.4 (i)), we have s = , s0 = e, and the condition (11) is satised
for any highest weight . By Example 4.1, there are two conjugacy classes of real
forms of the second kind represented by sll+1 (R) and slm (H), 2m = l + 1, the
corresponding involutions being, respectively, 0 and (Ad S)0 , where S is the
matrix Sm given by (2.5) or any other non-singular skew-symmetric matrix.
In the case = 0 , we need not use Theorem 1 to calculate the index. In fact,
sll+1 (R) is the normal real form of g, and it was proved in Proposition 1 that any
7. Inclusions between real forms 57

representation , expressed in an appropriate basis, maps the normal real form


into sln (R). Thus, j(g, , ) = 1.
In the second case, we may set = . In fact, by Proposition 4.3, we have =
(Ad g 1 )0 , where g = exp( 2 (e f )). Clearly, g SO2m . One sees from (17) and
Table 4 that 1 (h) = 2m 1 for g = sl2m (C), and by (19) g 2 = I2m . It follows
that g and g 1 are skew-symmetric, and hence we may set = (Ad g 1 )0 = .
By (12), we get
j(g, , ) = (1)(h) .
Thus, maps slm (H) into sln (R) if (h) is even, and into sl n2 (H) if (h) is odd.
By Remark 1, equivalent conditions are R(I2m ) = In and R(I2m ) = In ,
respectively. From (17) and Table 4 one gets the explicit formula

j(g, , ) = (1)1 +3 +...+2m1 .

If g0 is a real form of the rst kind, then, by Example 4.1, g0 = sup,l+1p , =


Ad Ip,l+1p , where p = 0, . . . , l. Since s = e, the condition (11) is satised,
whenever i = l+1i , i = 1, . . . , l. In the notation of Example 2.2 we have
= exp(ad 2 it), where t t(R) is given by p (t) = 2, i (t) = 0 for i = p. Using
Table 4, one gets from (12) the following result:

1 if l = 2m,
j(g, , ) =
(1)(m+p)m if l = 2m 1.

Example 2. Suppose that g = so2l (C), l 4. We will use the notation of the
proof of Proposition 4.4. As is well known (see, e.g., [11] or [19]), there exist, up
to isomorphy, the following real forms of the rst kind:
1. g0 = so2k,2(lk) , = Ad diag(Ik , Ilk , Ik , Ilk ), where k = 0, . . . , l 2.
2. g0 = ul (H), = Ad Sl .
Since s = e, we see from Proposition 4.4 that for l = 2m the condition (11) is
always satised, while for l = 2m + 1 it is satised, whenever 2m = 2m+1 .
Now, in the case 1 = exp(ad 2 it), where t t(R) is given by k (t) = 2, i (t) =
0 for i = k. One gets from (12) and Table 4 the formula

(1)(m+k)(2m1 +2m ) if l = 2m,
j(g, , ) =
1 if l = 2m + 1.

In the case 2, is conjugate to exp(ad 2 it), where t t(R) is given by l (t) =


2, i (t) = 0 for i < l. Using again (12) and Table 4, we get

j(g, , ) = (1)1 +3 +...+2m1 .

One also has the following real forms of the second kind:
3. g0 = so2k+1,2(lk)1 , = Ad diag(Ik , Ilk , Ik+1 , Ilk1 ), where k =
0, . . . , l 2.
If l = 2m, then s0 = s = e, and (11) is satised, whenever 2m1 = 2m . If
l = 2m + 1, then s0 = e, and (11) is always satised.
58 7. Inclusions between real forms

To calculate the index, we should present in the form (8). It is easy to verify
that the non-trivial diagram automorphism s is given by s = Ad T , where the
matrix T O2l is determined by the transposition of two vectors el , e2l of the
standard basis of C2l . One also proves that is conjugate to exp(ad 2 it)s, where
t t(R) is given by k (t) = 2, i (t) = 0 for i = k. Then (12) and Table 4 imply
the following result:

1 if l = 2m,
j(g, , ) =
(1)(m+k)(2m +2m+1 ) if l = 2m + 1.

Example 3. Suppose that g = so2l+1 (C), l 2. This is the algebra of all block
matrices of the form

0 U V
V  X Y , X gll (C), Y  = Y, Z  = Z, U, V Cl .
U  Z X 

As a maximal toral subalgebra t, one may choose the subalgebra of diagonal ma-
trices
H = diag(0, x1 , . . . , xl , x1 , . . . , xl ).
The system of roots is = {xi xj (i = j), xi }. We will use the system of
simple roots

= {1 , . . . , l }, where i = xi xi+1 , i = 1, . . . , l 1, l = xl .

It is well known (see, e.g., [11] or [19]) that there exist, up to isomorphy, only the
following real forms (all are of the rst kind):
g0 = so2k,2(lk)+1 , = Ad diag(Ik , Ilk , Ik , Ilk+1 ), where k = 0, . . . , l.
The condition (11) is always satised.
We can write = exp(ad 2 it), where t t(R) is given by k (t) = 2, i (t) = 0
for i = k. Using (12) and Table 4, we get
1
j(g, , ) = (1)(k+ 2 l(l+1))l .

Example 4. Consider the case g = sp2l (C), l 2. This is the algebra of all block
matrices of the form
 
X Y
, where X gll (C), Y  = Y, Z  = Z.
Z X 

As a maximal toral subalgebra t one may choose the subalgebra of diagonal ma-
trices
H = diag(x1 , . . . , xl , x1 , . . . , xl ).
The system of roots is = {xi xj (i = j), 2xi }. We will use the system of
simple roots

= {1 , . . . , l }, where i = xi xi+1 , i = 1, . . . , l 1, l = 2xl .


7. Inclusions between real forms 59

It is well known (see, e.g., [11] or [19]) that there exist, up to isomorphy, only the
following real forms (all are of the rst kind):
1. g0 = spp,lp , = Ad diag(Ip , Ilp , Ip , Ilp ), p = 0, . . . , l 1.
2. g0 = sp(R), = 0 |g.
The condition (11) is always satised.
In the case 1, we have = exp(ad 2 it), where t t(R) is given by p (t) =
2, i (t) = 0 for i = p. Using (12) and Table 4, we get
1 +3 +...+2[ 1 (l+1)]1
j(g, , ) = (1) 2 .

In the case 2, is conjugate to exp(ad 2 it), where t t(R) is given by l (t) =


2, i (t) = 0 for i < l. It follows that

j(g, , ) = 1.

Example 5. Let G be a simply connected complex semisimple Lie group and g its
Lie algebra. Let be an irreducible representation of g satisfying the condition
(11). Due to Remark 1, the Karpelevich index of may be found by calculating
the value R(w) of the corresponding representation R of G on the element w =
(gz)2 Z(G). This leads to the following simple observation: if the order of Z(G)
is odd, then j(g, , ) = 1. In fact, in this case w should have an odd order, and
R(w)2 = In implies R(w) = In . The order of Z(G) is odd for the simple Lie
algebras sl2m+1 (C) (Z(G)
Z2m+1 ), E6 (Z(G)
Z3 ), G2 , F4 , E8 (Z(G) = {e})
and their direct sums. Thus, under the condition (11), always maps any real
form of each of these algebras into sln (R).
Example 6. Let g be a complex semisimple Lie algebra. Consider an irreducible
representation : gg sl(V ). Consider the involution : (x, y)  (y, x) of gg
which corresponds to the real form of this algebra isomorphic to gR (see Example
3.4). Using Theorem 1, one can nd the condition, under which extends to an
outer involutive automorphism of sl(V ), and calculate the index.
Any irreducible representation of g g has the form = 1 2 , where
i , i = 1, 2, are two irreducible representations of g (see (III.12)). If t is a maximal
toral subalgebra of g, then we choose t t as a maximal toral subalgebra of
g g. Then the highest weight of is = (1 , 2 ), where i is the highest
weight of i , i = 1, 2. Clearly, = s is a canonical automorphism determined
by the transposition s of two copies of the system of simple roots of g, and
s() = (2 , 1 ). By (10) and Proposition 4.4 (ii), s0 () = ((2 ), (1 )). It
follows that the condition (11) has the form 2 = (1 ) or, equivalently, 2 1.
Using (12), we see that j(g g, s, 1 1 ) = (1)
1 (h)+(1 )(h)
= (1)21 (h) = 1.
Thus, under the above condition, always maps our real form into sln (R).
Due to Theorem 2.1, a real semisimple Lie algebra g0 is simple precisely in the
following two cases: 1) g0 (C) is simple; 2) g0 = gR , where g is a simple complex Lie
algebra. Let be an irreducible representation of g0 (C). The second case is studied
in Example 6. In the rst case, the involution s0 and the index j(g, , ) can be
calculated using Examples 1 5. The results are presented in Table 5. Note that
60 7. Inclusions between real forms

so2k+1,2(lk)
so2(lk),2k+1 , and therefore the index formula for g0 = sop,2l+1p ,
established in Example 7.3 in the case of even p, can be applied to the case of odd
p. This gives the corresponding formula in Table 5.
The same problem for an arbitrary real semisimple Lie algebra is reduced to
the case of a simple g0 with the help of the following proposition.
Proposition 5. Let gk , k = 1, 2, be two complex semisimple Lie algebras, k an
involutive automorphism of gk and s0k Aut k the corresponding involution of
the system of simple roots k of gk (see (10)). Then the involution s0 induced by
s01 and s02 on the system of simple roots = 1 2 of g = g1 g2 corresponds
to the automorphism = 1 2 of g. If k is an irreducible linear representation
of gk , k = 1, 2, then

j(g, , 1 2 ) = j(g1 , 1 , 1 )j(g2 , 2 , 2 ).

Proof. Let tk be a maximal toral subalgebra of gk , k = 1, 2. We may suppose


that k has the form (8), i.e., k = exp(ad 2 itk )sk , where sk Aut k and tk
tk , k = 1, 2, satisfy the conditions of (8). Then t = t1 t2 is a maximal toral
subalgebra and = 1 2 is a system of simple roots of g. One veries easily
that = 1 2 = exp(ad 2 it)s, where t = (t1 , t2 ) t and s = s1 s2 Aut also
satisfy the conditions of (8). By Proposition 4.4 (ii) and (iii), = 1 2 , where k is
the corresponding involution of k , k = 1, 2, and s0 = s = s01 s02 . Clearly, (4.4)
implies that h = (h1 , h2 ), where hk tk (R) is the sum of the positive coroots of
gk , k = 1, 2. Now, if k is the highest weight of k , then, by (III.12), = (1 , 2 )
is the highest weight of = 1 2 . Using (12), we get

j(g, , ) = (1)(t+h) = (1)(1 ,2 )((t1 ,t2 )+(h1 ,h2 ))


= (1)1 (t1 +h1 )+2 (t2 +h2 ) = j(g1 , 1 , 1 )j(g2 , 2 , 2 ).


Next, we apply Theorem 1 to study of bilinear invariants of an irreducible
representation. Let : g gl(V ) be a linear representation of a complex Lie
algebra g and b a bilinear form in V . We say that b is invariant under if

b((x)u, v) + b(u, (x)v) = 0 , x g, u, v V . (20)

An example is the Killing form kg invariant under the adjoint representation = ad


(see (1.1)).
Bilinear forms on V are in a natural bijection with linear mappings : V V .
Namely, if : V V is a linear mapping, then the function

b(u, v) = (u)(v) , u, v V,

is a bilinear form on V , and any bilinear form is obtained in this way. Clearly, non-
degenerate bilinear forms correspond to isomorphisms V V . Further, bilinear
forms invariant under correspond to the mappings satisfying

(x) = (x) , x g.
7. Inclusions between real forms 61

In particular, a non-degenerate invariant bilinear form exists if and only if is


self-dual. The Schur Lemma implies that if is irreducible, then the invariant
bilinear form (if it exists) is unique, up to a complex factor. Moreover, any non-
zero invariant bilinear form is non-degenerate.
Choose a basis in V and denote by B the matrix of b in this basis. Replacing
the representation by its matrix form, we see that (19) is expressed by

B(x) + (x) B = 0 , x g. (21)

In the case when det B = 0, the relation (21) can also be written in the form

= (Ad B 1 )0 ,

which means that the automorphism  = (Ad B)0 extends e by . If g is semisim-


ple, then we may apply Theorem 1. From its proof we know that we can assume
B On and B  = jB, where j = j(g, e, ) = 1. Thus, the invariant form b
is symmetric or skew symmetric, depending on the value of j. Summarizing and
applying Theorem 1, we get the following assertion.

Theorem 2. If an irreducible representation of a complex Lie algebra g admits


a non-zero invariant bilinear form b, then b is non-degenerate and unique up to a
complex factor. Such a form exists if and only if is self-dual. If g is semisimple,
then this condition is equivalent to () = , where is the highest weight of .
Moreover,
b(v, u) = (1)(h) b(u, v) , u, v V , (22)

where h is given by (4.4). 

A representation of a Lie algebra g is called orthogonal if it admits a sym-


metric non-degenerate invariant bilinear form, and symplectic if it admits a skew-
symmetric non-degenerate invariant bilinear form. In the rst case, is equivalent
to a homomorphism g son (C), and in the second one to a homomorphism
g spn (C), n even. Theorem 2 allows to describe all orthogonal and symplectic
linear representations of complex semisimple Lie algebras.
Note that this description is due to E.B. Dynkin [5, 7]. A similar result for
Chevalley groups see in Steinberg [23], 12.
To conclude, we consider the case when the extending involution is inner. In
particular, we prove two formulas (see [15]) for the signature of (, ) which express
this signature through the character of the representation R. We remind that the
character of R is the function R : G C dened by

R (g) = tr R(g) , g G.

One of these formulas also exploits the invariant integration over a maximal com-
pact subgroup of G (see 1 I).
62 7. Inclusions between real forms

Theorem 3. Let : g sln (C) be an irreducible representation with highest weight


, and let an involution = s Aut g of the form (8) be given.
(i) We have  for an involution  Int sln (C) if and only if s() = . In
particular, any involution Int g extends to an involution  Int sln (C).
(ii) Under the condition of (i), the signature q p of the pair (, ) is determined
by 
(q p)2 = n R (g(g)1 )dg , (23)
U

where is the automorphism of G satisfying de = and U is a maximal


compact subgroup of G invariant under .
(iii) If = = exp(ad u) Int g, then the signature can also be expressed by the
formula
q p = |R (exp u)| . (24)

Proof. (i) Denoting = Ad z, where z = exp u G, we can write the extension


condition in the form

= (Ad R(z))s = (Ad B) , (25)

where B GLn (C). If such an extension exists, then s , whence s() = ,


by Lemma 1. Conversely, if s() = , then s , and there exists B GLn (C)
satisfying (25). Then  = Ad B extends . Now, e = 2  , whence  = e by
2 2

Lemma 6.1. In the case s = e, we may take = Ad R(z).
(ii) Suppose that (25) is satised. Then

R((g)) = BR(g)B 1 , g G. (26)

By Example 4.1, we may assume that B = CIp,q C 1 for a certain C GLn (C)
and p q. In particular, B 2 = In , and tr B = q p. Clearly, u = g is a compact
real form of g invariant under , and the corresponding maximal compact subgroup
U G is invariant under . Applying (I.8) to the representation g  Ad R(g) of
U in the vector space gln (C), we get the matrix

B0 = (R(g)BR(g)1 )dg gln (C) (27)
U

satisfying R(g)B0 R(g)1 = B0 . Clearly, R(g)|U is irreducible, and the Schur


Lemma implies B0 = In , where C. From (27) we see that

tr B0 = (tr B)dg = tr B = q p.
U

Therefore, B0 = qp
n In . Since B = B 1 , we get from (26)

B(R(g)BR(g)1 ) = R((g))R(g)1 = R((g)g 1 ) , g G.


7. Inclusions between real forms 63

The integration over U gives



qp
BB0 = B= R((g)g 1 )dg.
n U

Calculating the traces, we get (23).


(iii) Using the Schur Lemma, we get R(z) = cB = cCIp,q C 1 , where c C
(see the proof of (ii)). Comparing the determinants of both sides, we see that
|c| = 1, by Proposition 4 (iii). It follows that

q p = tr Ip,q = | tr R(z)| = |R (z)| = |R (exp u)| . 


Using tedious calculations with the Weyl character formula (see [22]) and in-
tegration, in [15] explicit formulas for the signature of irreducible representations
of classical Lie algebras were deduced from (23) and (24). We will not reproduce
them here.
8. Real representations of real semisimple Lie algebras
The main goal of this section is the classication of irreducible real representa-
tions of real semisimple Lie algebras. We start with general notions concerning
complexication, realication and complex conjugation of representations, which
are quite similar to those discussed in 2 in connection with Lie algebras. Certain
simple facts similar to Proposition 2.3 and Theorem 2.1 will be established.
Let g0 be an arbitrary real Lie algebra. In this section, representations : g0
gl(V0 ) of g0 in a real vector space V0 will be studied. We call them generally real
representations of g0 . Real homomorphisms : g0 gl(V ), where V is a complex
vector space, will also be considered; they are called complex representations of
the real Lie algebra g0 . If g is a complex Lie algebra, then its representations in a
complex vector space studied in 7 will also be called complex representations.
Respectively, we have two complexication operations applied to a real repre-
sentation : g0 gl(V0 ). Let us denote V = V0 (C) and g = g0 (C). First, we can
extend any (x), x g0 , to a complex linear operator in V , obtaining a complex
representation C : g0 gl(V ). Second, we can extend C to a homomorphism of
complex Lie algebras (C) : g gl(V ), i.e., to a complex representation of g.
Now, if we begin with a complex representation : g0 gl(V ) of a real Lie
algebra g0 , then we also can extend it to a complex representation (C) : g
gl(V ). On the other hand, we may regard as a representation of g0 in the real
vector space VR , and thus we get a real representation R : g0 gl(VR ). This is
the realication operation. Also, any complex representation : g gl(V ) of a
complex Lie algebra g gives rise to a real representation R : gR gl(VR ).
A real (complex) representation is called irreducible if the representation space
does not contain any non-zero proper real (respectively, complex) invariant vector
subspaces. The equivalence of two real (complex) representations is dened, using
real (respectively, complex) isomorphisms of representation spaces. It is easy to
establish certain implications between the irreducibility of , C , (C) and R and
between the equivalence of two representations 1 and 2 , of their complexications
and realications (see Tables 2 and 3).
A complex structure J in a real vector space V0 is said to be invariant under
a real representation of g0 in V0 if (x)J = J(x), x g0 . In this case, we
may regard as a complex representation of g0 in (V0 , J), and the original real
representation will be its realication.
Similarly, a real (or quaternion) structure S in a complex vector space V is
said to be invariant under a complex representation : g0 gl(V ) if (x)S =
S(x), x g0 . If a real structure S in V is invariant under , then the real form
V0 = V S of V is invariant, and the real subrepresentation 0 : g0 gl(V0 ) of
satises C 0 = . If a quaternion structure in V is invariant under , then V
admits a structure of the vector space over H, invariant under , which means
that determines a homomorphism g0 glm (H), where 2m = dimC V .
Let : g0 gl(V ) be a complex representation of a real Lie algebra g0 . Then
may be regarded as a complex representation of g0 in the complex conjugate vector
space V (see 2). We denote this representation by and call it the complex con-
jugate to the representation . We distinguish between these two representations,
8. Real representations of real semisimple Lie algebras 65

since they act in distinct complex vector spaces and, as we shall see later, are not
necessarily equivalent. At the same time, ()R = R .
Let us nd the matrix form of the complex conjugate representation. Choose a
basis v1 , . . . , vn of V and denote by C(x) = (cij (x)) the corresponding matrix of
(x), x g0 . Then (x) is expressed in the same basis, regarded as a basis of V ,
by the complex conjugate matrix C(x). Thus, (x) has the following matrix form:

= 0 , (1)

where by 0 the real structure X  X in the Lie algebra gln (C) is denoted.
Now we will give a description of irreducible real representations of a real Lie
algebra g0 in terms of its irreducible complex representations, using the notions
introduced above. The results and their proofs are similar to the description of
simple real Lie algebras in terms of simple complex ones given by Theorem 2.1.
Proposition 1. For any complex representation : g0 gl(V ), we have

(R )C + .

Proof. Consider the complex vector space W = V V and dene the mapping
S : W W by S(u, v) = (v, u), u, v V . One sees easily that S is a real structure
in W invariant under + . The corresponding real form W S coincides with Wd =
{(u, u) | u V }, and the projection (u, u)  u, u V , gives an isomorphism of
the real subrepresentation 0 induced on Wd onto R . Thus, (R )C C 0 = + .

Theorem 1. Any irreducible real representation : g0 gl(V0 ) of a real Lie algebra
g0 satises precisely one of the following two conditions:
(i) C is an irreducible complex representation;
(ii) = R , where  is an irreducible complex representation admitting no in-
variant real structures.
Conversely, any real representation satisfying (i) or (ii) is irreducible.
Proof. Suppose that is irreducible, but C is reducible. We want to prove that
V0 admits a complex structure invariant under . Fix a non-zero complex vector
subspace W  V = V0 (C) invariant under C . Using the complex conjugation with
respect to V0 , we can construct the complex vector subspaces W , Y = W + W
and Z = W W of V invariant under C . We have Y = Y, Z = Z, whence
Y = (Y V0 )(C) and Z = (Z V0 )(C). Since is irreducible, this implies
that Y V0 = V0 and Z V0 = {0}. It follows that V = Y = W W , and
V0 = {u + u | u W }. Then we can dene a complex structure J on V0 by
J(u + u) = i(u u), u W . Clearly, we have

(x)J(u + u) = (x)i(u u) = i(C (x)u C (x)u) = i(C (x)u C (x)u)


= J(C (x)u + C (x)u) = J(x)(u + u) , x g0 , u W .

This means that J is invariant under .


66 8. Real representations of real semisimple Lie algebras

It follows that may be regarded as a complex representation  in (V0 , J) and is


the realication of  . Clearly,  is irreducible (Table 3). If  admits an invariant
real structure S, then V0S is invariant under , which gives a contradiction. Thus,
the condition (ii) is satised.
By Proposition 1, (ii) implies that C is reducible, and hence (i) is not satised.
Conversely, if (i) is satised, then is irreducible, due to Table 2. Suppose
that (ii) is satised, i.e., = R , where  : g0 gl(V ) is an irreducible complex
representation. If is reducible, then there exists a non-zero real vector subspace
W  V which is invariant under . Then we can construct the -invariant real
vector subspace iW and the -invariant complex vector subspaces W (iW ) and
W + iW . Since  is irreducible, we can deduce that V = W (iW ). Thus, W is
an -invariant real form of V . Clearly, the corresponding real structure is invariant
under , which is a contradiction. 
Thus, all irreducible real representations of a given real Lie algebra g0 split
into two disjoint classes I and II which are characterized by the conditions (i) and
(ii) of Theorem 1, respectively. The following corollary describes these classes from
another point of view.
Corollary 1. The class I consists of all irreducible real representations which
admit no invariant complex structure. In this case, C is an irreducible complex
representation, admitting an invariant real structure.
The class II consists of all irreducible real representations which admit an
invariant complex structure, i.e., have the form = R , where  is an irreducible
complex representation. In this case,  admits no invariant real structures. 
As Theorem 1 shows, the following question is important for our study: which
irreducible complex representations : g0 gl(V ) admit invariant real structures?
By denition, an invariant real structure S : V V is an isomorphism V V ,
satisfying (x)S = S (x), x g0 , i.e., an isomorphism of onto the complex
conjugate representation . A complex representation of g0 is called self-conjugate
whenever . We see that any complex representation admitting an invariant
real structure is self-conjugate. The converse is not true, in general.
Proposition 2. Let : g0 gl(V ) be an irreducible complex representation of a
real Lie algebra g0 .
(i) If S is an antiautomorphism V V such that S(x) = (x)S, x g0 , then
S 2 = ce, where c R .
(ii) If S1 is another antiautomorphism V V satisfying S1 (x) = (x)S1 , x
g0 , then S1 = dS, where d C , and S12 = c1 e, where = sgn c = sgn c1
does not depend on the choice of S.
(iii) If S1 and S2 are two real structures
invariant under , then S1 = dS2 , where
d C, |d| = 1, and V S1 = d V S2 .
Proof. (i) Clearly, S 2 GL(V ) satises (x)S 2 = S 2 (x), x g0 . By the Schur
Lemma, we have S 2 = ce, where c C . For any v V , we have c(Sv) =
S 2 (Sv) = S(S 2 v) = S(cv) = c(Sv), and thus c = c. Hence c R.
(ii) If we have another antiautomorphism S1 : V V such that S1 (x) =
(x)S1 , x g0 , then S1 S 1 is an automorphism commuting with all (x), and
8. Real representations of real semisimple Lie algebras 67

the Schur Lemma implies that S1 S 1 = de, where d C . We get S1 = dS, and
hence S12 = (dS)(dS) = (dd)S 2 = (|d|2 c)e. Clearly, sgn(|d|2 c) = sgn c.
(iii) By
(ii), we 1have S1 = S22 = e implies |d| = 1. If x V S2 ,
= dS2 , and S12
then S1 ( d x) = d S1 x = d x, and thus d x V S1 .


Let : g0 gl(V ) be a self-conjugate irreducible complex representation.


We dene the Cartan index of as () = sgn c = 1, where c is dened by
the following condition: we have S 2 = ce, where S is an antiautomorphism of V
commuting with . This sign is uniquely determined, due to Proposition 2. Setting
S0 = 1 S, we get an antiautomorphism of V commuting with and satisfying
|c|
S02 = ()e. If () = 1, then S0 is a real structure in V invariant under . If
() = 1, then S0 is a quaternion structure in VR invariant under , and V admits
a structure of the vector space over H, invariant under . Thus, determines a
homomorphism g0 glm (H), where 2m = dimC V .
Summarizing, we see that an irreducible complex representation : g0 gl(V )
admits an invariant real structure if and only if is self-conjugate and its Cartan
index is equal to 1. In this case, the invariant real structure is determined uniquely,
up to a complex factor, and hence the real form V0 of V , invariant under , is
determined uniquely, up to multiplication by a complex number.
We can now classify the irreducible real representations described in Theorem
1 up to equivalence.
Theorem 2.
(i) Two irreducible real representations 1 and 2 of the class I are equivalent if
and only if C C
1 and 2 are equivalent.
(ii) Two irreducible representations 1 = (1 )R and 2 = (2 )R of the class II are
equivalent if and only if 1 2 or 1 2 .
(iii) A representation of the class I cannot be equivalent to one of the class II.
Proof. (i) The implication 1 2 C C
1 2 is obvious (see Table 2). Conversely,
suppose that 1 and 2 are of the class I and that C C
1 2 . Let : V1
V2 be an isomorphism of the representation spaces of C 1 and C
2 respectively,
satisfying C 1 (x) = C
2 (x), x g 0 . Denote by S i a real structure in Vi invariant
under C , i = 1, 2. Then S 
= 1
S 2 is an invariant real structure in V1 , and
i 1
 S1
Proposition
2 implies that S1 = dS1 , where |d| = 1, and V1 = d V1 . It follows
S1

that d (V1S1 ) = V2S2 . Therefore d determines an equivalence between 1


and 2 .
(ii) If 1 2 , then, by Proposition 1, 1 + 1 2 + 2 . Therefore 1 2
or 1 2 . Conversely, any of two equivalences 1 2 and 1 2 implies
1 = (1 )R (2 )R = 2 .
(iii) is evident. 
From now on, we assume that g0 is a real semisimple Lie algebras. Theorems 1
and 2 reduce the problem of classication of irreducible real representations of g0
to the following ones: to describe the complex conjugate representation in terms
of an irreducible complex representation and, in particular, to determine the
self-conjugate irreducible complex representations of g0 ; to calculate the Cartan
68 8. Real representations of real semisimple Lie algebras

index of a self-conjugate irreducible complex representation. We will see now that


these two problems can be solved if we use the results of 7.
We will regard g0 as a real form of the complex semisimple Lie algebra g =
g0 (C). The Cartan correspondence described in 3 assigns to this real form an
involution of g. For any irreducible complex representation 0 of g0 , we denote
by = 0 (C) the corresponding complex representation of g which is irreducible,
too, and thus is determined, up to equivalence, by its highest weight . We will
say that is the highest weight of 0 . Clearly, determines 0 up to equivalence,
as well. In the following theorem, we use the notation of 7. In particular, by s
one denotes the involutive automorphism of which corresponds to the involution
by Theorem 4.1. We also put s0 = s Aut . By Proposition 4.4 (iii), s0 = s
is involutive. The Cartan index of 0 will be denoted by (g0 , 0 ).
Theorem 3. Let 0 : g0 sl(V ) be an irreducible complex representation of a
real semisimple Lie algebra g0 , and let , denote the highest weights of the
representations 0 , 0 , respectively. Then = s0 (). In particular, 0 is self-
conjugate if and only if s0 () = . In this case, the Cartan index of 0 is expressed
by
(g0 , 0 ) = j(g, , ) ,
where = 0 (C).
Proof. Let denote the complex conjugation in g = g0 (C) with respect to g0 . By
denition, = , where is a compact real structure in g commuting with .
Due to Proposition 7.1, we may choose a basis of V such that the conditions (7.3)
and (7.5) are satised (note that the notation does not coincide with that from
Proposition 7.1!). We may replace by this matrix form in this basis, regarding it
as a homomorphism g sln (C). By (1), we have 0 = 0 0 . Denoting = 0 (C),
we have
= 0 . (2)
In fact, both sides of this equality are complex homomorphisms g sln (C) which
coincide on the real form g0 . Multiplying both sides of (2) by from the right, we
deduce
= 0 = 0 0 = 0 = ,
where 0 is given by (7.1). By Theorem 4.1, = s, where Int g and
s Aut . It follows that s , whence s() = (), due to Proposition 7.3.
Hence, = s0 ().
Now suppose that 0 is self-conjugate, i.e., that s() = (). Then we can
apply Theorem 7.1. In particular,  , where  is an outer automorphism of
sln (C), i.e., there exists a matrix B GLn (C) such that

= (Ad B)0 . (3)

As was shown in the proof of this theorem, we may assume that B On satises
(7.13). One deduces from (3) that

= (Ad B)0 0 = (Ad B)0 ,


8. Real representations of real semisimple Lie algebras 69

whence = (Ad B)0 . Restricting both sides to g0 , we get

0 = (Ad B)0 0 . (4)

Let S0 : (z1 , . . . , zn )  (z1 , . . . , zn ) denote the standard real structure in Cn .


Clearly, the real structure 0 in the Lie algebra gln (C) may be written as 0 (X) =
S0 XS0 , where X gln (C) is viewed as a linear operator in Cn . Dene the anti-
linear operator S = BS0 in Cn . Then (4) may be written as

0 (x) = S0 (x)S 1 , x g0 .

Since B is real, BS0 = S0 B, whence S 2 = B 2 . Therefore we get from (7.13)

S 2 = j(g, , )In .

Using Proposition 2, we see that (g0 , 0 ) = j(g, , ). 


Summarizing, we see that Theorems 1, 2, 3 and Theorem 7.1 give the following
classication of irreducible real representations of real semisimple Lie algebras.
Let g0 be a real form of a complex semisimple Lie algebra g, and let Aut g
be the corresponding involution written in the form (7.8). Suppose an irreducible
representation in a complex vector space V be given, and let be its highest
weight. Denote 0 = |g0 . Then 0 determines an irreducible real representation
of g0 , and here three dierent cases are to be distinguished.
1. The real case. If

s0 () = , (g0 , 0 ) = (1)(t+h) = 1 , (5)

then 0 leaves invariant a real form V0 of V and induces there an irreducible


representation 0 |V0 of g0 .
2. The quaternion case. If

s0 () = , (g0 , 0 ) = (1)(t+h) = 1 , (6)

then the realication (0 )R of 0 acting in VR is irreducible, and VR admits a


structure of quaternion vector space invariant under (0 )R .
3. The complex case. If
s0 () = , (7)
then the realication (0 )R of 0 acting in VR is irreducible, and 0 admits
neither real nor quaternion invariant structures.
All irreducible real representations of real semisimple Lie algebras can be ob-
tained in one of these ways. In the real case, we get a bijection between the
dominant weights satisfying (5) and the irreducible representations 0 |V0 of g0
regarded up to equivalence. In the quaternion case, a similar assertion is true. In
the complex case, two dominant weights 1 , 2 satisfying (7) determine equivalent
real representations (0 )R if and only if either 2 = 1 or 2 = s0 (1 ).
70 8. Real representations of real semisimple Lie algebras

Example 1. Let g0 be a compact semisimple Lie algebra. We may regard it as a


compact real form of a complex semisimple Lie algebra g, corresponding to the
trivial involution = e. Consider an irreducible representation : g sl(V ) with
highest weight . The condition (5) is () = , (1)(h) = 1, it means that
is orthogonal (see Theorem 7.2). In this case, we get an irreducible representation
g0 son sln (R), where n = dim V . The condition (6) means that is symplec-
tic, and we get an irreducible representation g0 spm sl2n (R), n = 2m. The
condition (7) means that is not self-dual. Then we get an irreducible represen-
tation g0 sun sl2n (R).
Example 2. Let g0 = gR be a realication of a complex semisimple Lie algebra
g. As in Examples 2.4 and 3.4, we will identify g0 with the real form of g g
corresponding to the involution s : (x, y)  (y, x). The embedding g0 g g has
the form x  (x, x), where is a xed compact real structure in g. To apply our
classication, consider an irreducible complex representation : g g sl(V ).
Using the notation of Example 7.6, we may write = 1 2 , where 1 , 2 are
two irreducible complex representations of g, and = (1 , 2 ), where i is the
highest weight of i , i = 1, 2. If 2 = (1 ) (or 2 1 ), then the Cartan
index is 1, by Example 7.6 and Theorem 3. Therefore, we get an irreducible
representation of g0 in a real form of V . These real irreducible representations
are in 1-1 correspondence with the dominant weights 1 (or 2 ). The quaternion
case is impossible. If 2 = (1 ) (or 2  1 ), then we get a real irreducible
representation of g0 in VR . These real irreducible representations are determined
by ordered pairs (1 , 2 ) of dominant weights of g, and the pairs (1 , 2 ) and
((2 ), (1 )) give equivalent representations.
We also note that in the case 2 1 the vector space V can be identied with
End V1 , where V1 is the space of the representation 1 . Let us replace 1 by its
matrix form such that (3.7) is satised. Then End V1 is identied with glm (C),
where m = dim 1 , and acts by the formula

(x, y)Z = 1 (x)Z Z1 (y) , x, y g .

Therefore

(x, x)Z = 1 (x)Z + Z1 (x) , x, y g .
It follows that the corresponding irreducible representation of g0 acts in the real
form V0 = {Z glm (C) | Z  = Z} consisting of Hermitian matrices.
Example 3. Let : g gl(V ) be an irreducible complex representation of a
complex semisimple Lie algebra g and dim V > 1. Then the real representation
R of gR is irreducible. In fact, in the opposite case one proves, as in Theorem
1, the existence of a real structure S in V , invariant under . For any x g,
we have (ix)S = i(x)S = iS(x). But, on the other hand, (ix)S = S(ix) =
S(i(x)) = iS(x). Thus, is trivial, and dim V = 1. This construction is
actually a special case of Example 2, since = 2 , where 2 is the trivial
representation of dimension 1.
In the case, when g0 is a real form of a simple complex Lie algebra, the involution
s0 giving the self-conjugacy condition of an irreducible complex representation 0
8. Real representations of real semisimple Lie algebras 71

of g0 is shown in Table 5. It can be also read from the Satake diagrams of g0


which are given in this Table, using Theorem 9.1 (see Appendix). Table 5 also
contains the values of the corresponding Cartan index. The calculation is based
on Examples 7.1 7.5. The index is expressed through the coordinates i of
the highest weight of 0 in the basis of fundamental weights, following the
numeration of simple roots admitted in [19]. The real forms g0 are denoted as
in [11] or [19]. Together with Example 2, this solves the classication problem
of irreducible real representations of simple real Lie algebras. The general case is
reduced to this one by the following proposition which is an immediate corollary
of Proposition 7.5 and Theorem 3.

s
Proposition 3. Consider the Lie algebra g0 = i=1 gi , where gi are non-commu-
tative simple real Lie algebras, and its representation 0 = 1 . . . s , where i
is an irreducible complex representation of gi . Then 0 0 if and only if i i
for each i = 1, . . . , s. Under this condition, we have

s
(g0 , 0 ) = (gi , i ).
i=1


In conclusion, we will deduce a formula reducing the calculation of the Cartan
index to the case when the given highest weight is a fundamental one. Let us x
g0 and regard (0 ) as a function () of the highest weight of = 0 (C). By
Theorem 3, this function is dened on the subsemigroup (g0 ) of the semigroup of
all dominant weights of g = g0 (C) given by the equation s0 () = . By the same
theorem, (0 ) coincides with the Karpelevich index j(g, , ) which is expressed
by (7.12). It follows that ( +  ) = ()( ), i.e., is a homomorphism of
(g0 ) into the group {1, 1}. We also note the following property of the index.
Lemma 1. For any dominant weight of g, we have ( + s0 ()) = 1.
Proof. We should prove that ( + s0 ())(t + h) is even. Using s(t) = t, we get

( + s0 ())(t + h) = (t +  t) + (h) + (s


0 h) .

By Proposition 4.2, s0 h = h. Since (h) Z, we have only to prove that (t +


 t) = (+())(t) 2Z. Let be an irreducible representation of g with highest
weight . Then () = w0 () (this is the so-called lowest weight of ).
It follows that + () = w0 () is a sum of simple roots of g. By (7.8), its
value on t is even. 
Clearly, the function is completely determined by its values on any generators
of the semigroup (g0 ). To construct such generators, one can use the fundamental
weights. Let 1 , . . . , l denote the fundamental weights corresponding to the
system of simple roots = {1 , . . . , l }. Then any dominant weight can be
written as
l
= i i ,
i=1
72 8. Real representations of real semisimple Lie algebras

where i = (hi ) are non-negative integers. The involution s0 permutes the


weights i . Therefore we can write them as

1 , . . . , r , r+1 , . . . , 2r , 2r+1 , . . . , l ,

where

s0 (i) = r + i , i = 1, . . . , r ; s0 (k) = k , k = 2r + 1, . . . , l .

Then the weights

i + r+i , i = 1, . . . , r ; k , k = 2r + 1, . . . , l ,

generate the semigroup (g0 ). By Lemma 1, ( i + r+i ) = 1, i = 1, . . . , r.


Therefore, is completely determined by the values ( k ), k = 2r + 1, . . . , l.
More precisely,
l 
() = ( k )k = ( i )i . (8)
k=2r+1 s0 (i)=i

This formula goes back to E. Cartan [3] and was proved by Iwahori [13] (see
also [10], Ch. 7). It was used in the tables of Tits [24], were the indices of
representations were given without proof. More precisely, these tables contain
a description of self-conjugate irreducible representations of real forms of simple
complex Lie algebras and of the real and quaternion fundamental weights.
The same information you can obtain from Table 5 below.
9. Appendix on Satake diagrams
Another way to describe the classication of real semisimple Lie algebras is pro-
vided by the Satake diagrams, originally introduced in [20]. The Satake diagram
of a real semisimple Lie algebra g0 has the same sets of vertices and edges as the
Dynkin diagram of its complexication g (although coloured in black or white)
and possibly some arrows, relating white vertices. It is natural to ask about the
relation between these data and the symmetries of the Dynkin diagram discussed
above. In this Appendix, we will show how to read the involution s0 = s (see
(7.10)) of the Dynkin diagram of g from the Satake diagram of its real form g0
which corresponds to an involutive automorphism of g inducing the symmetry
s Aut by Theorem 4.1. This involution appears in Theorems 7.1 and 8.3
and, in particular, allows to determine all self-conjugate complex irreducible rep-
resentation of the real Lie algebra g0 . The answer is formulated in Theorem 1
below.
First we recall the construction of Satake diagrams (see, e.g., [20] or [19], 5.4,
for more details). Let us suppose that g0 is a real semisimple Lie algebra with the
complexication g. The Cartan correspondence described in 3 implies that there
are the following data: a compact structure , a real structure and an involution
of g, such that = , g0 = g and , and commute pairwise. Thus,
(g0 ) = g0 , and there is the Cartan decomposition g0 = k p (see 5). Choose a
subalgebra a of g0 which lies in p and is maximal among all such subalgebras. Then
a is commutative and any maximal commutative subalgebra t0 g0 containing
a satises (t0 ) = t0 and has the form t0 = t+ 0 a, where t0 k. Moreover,
+

t = t0 (C) g is a maximal toral subalgebra of g which is invariant under , and


, and we have
0 a.
t(R) = it+
Let t be the system of roots of g relative to t. Since Aut(g, t), we have
an involution  : , due to (II.12). The roots from the subset

c = { |  () = } = { | | a = 0} (1)

are called compact roots, while the roots from nc = \ c are called non-
compact ones. The system of positive roots + can be chosen in such a way that
 () + for each non-compact + . To obtain such a system + , it
suces to take the set of roots which are positive with respect to the lexicographical
ordering in t(R) determined by a base of t0 such that its rst elements constitute
a base of a. Let + be the corresponding set of simple roots, and let us
denote


c = c , nc = nc ,
c = c , nc = nc .

Then, by the choice of + , we have

 (+
nc ) = nc . (2)

The following important property was proved in [20]:


74 9. Appendix on Satake diagrams

Lemma 1. There exists an involution : nc nc such that for any nc


we have 
 () = () c ,
c

where c are non-negative integers. 


The Satake diagram of g0 is dened as the Dynkin diagram corresponding to ,
where compact roots are denoted by black circles , non-compact roots by white
circles and any pair , () nc , such that = (), is linked by an arrow.
It is known that the Satake diagram determines g0 uniquely, up to isomorphy, and
that g0 is simple if and only if its Satake diagram is connected. Satake diagrams
for all real simple Lie algebras can be found, e.g., in [19], see also Table 5 below.
By Corollary 2 of Theorem 4.1, we have = s, where N = Aut(g, t)Int g
and s Aut . We will interprete s as a symmetry of the Dynkin diagram or of
the Satake diagram. Another important symmetry is the involution arising from
the similar decomposition of the Weyl involution , see Proposition 4.3. Our aim
is to describe the symmetry s0 = s of the Satake diagrams in terms of the data
given by this diagram, i.e., black and white vertices and the arrows.
The set c of compact roots can be interpreted as follows (see [20, 19]). Let m
g denote the centralizer of a in g. One can verify that m is reductive and that t+ =
t+
0 (C) is a maximal toral subalgebra of the semisimple subalgebra m0 = [m, m]. Any
c vanishes on a(C) and hence may be identied with its restriction to t+ . In
this way, c is identied with the root system of m0 relative to t+ and c with a
set of simple roots of m0 . Let Wc denote the subgroup of the Weyl group W of g
generated by the reections r , c . One deduces from (1.8) that any w Wc
acts trivially on a and leaves it+0 invariant. Moreover, restricting all w Wc to
it+
0 , we get an isomorphism of W c onto the Weyl group of m0 . As in 4, consider
the element w0c Wc such that (w0c ) (c ) = c and dene c = (w0c ) | c .
We will interprete c as an involutive symmetry of the black part of the Satake
diagram which is the Dynkin diagram of c . We shall also use the following
Lemma 2.
(i) For any w Wc we have w (+ +
nc ) = nc .
(ii) If w Wc and nc , then

w () = + k , (3)
c

where k are non-negative integers.


Proof. If c , nc , then by (II.9) we have

(, )
r () = 2 , (4)
(, )

 
where 2 (,)
(,) Z. Clearly, r () | a = 0, and hence r () nc . By (II.20),

r () + , whenever nc . Thus, the assertion (i) is proved for w = r . By
+
9. Appendix on Satake diagrams 75

(II.19), Wc is generated by reections r , c , which implies (i) in the general


form.
To prove (ii), suppose that c , nc . Then in (4) we have 2 (,)
(,)  0.
Thus, (3) is true for w = r . The general case can be proved by induction on the
number of reections r , c , entering into an expression of w Wc through
these generators. In fact, suppose that (3) holds for a certain w Wc and take
c . Then

(wr ) () = r (w()) = r () + k r ().
c

Clearly, r () c . Using (4), we get



(wr ) () = + k ,
c

where k Z. Then (II.18) implies that k  0, c . 


With the notation introduced above, we state the main result:
Theorem 1. The involution s0 = s of the Satake diagram leaves invariant both
the subset of white vertices and the subset of black ones. On the white vertices, it
induces the involution depicted by the arrows, while on the black ones it coincides
with the involution c . If g0 is simple, then c = id, except the following cases:
g0 = sun , suk,nk , sok,2nk (n k odd), and the compact form of E6 .
Proof. We start with improving the involution by composing it with an ap-
propriate inner automorphism of g in order to get 1 = Aut(g, t, ). Then
we will be able to see the symmetry s Aut induced by . We have to choose
N , and hence on the dual level we have to compose the transformation 
of the vector space t(R) with an element of the Weyl group W . Note that
 (w0c ) (+ ) = . In fact, by Lemma 2 (i), (w0c ) (+ +
nc ) = nc , and by deni-
tion (w0c ) (+c ) =
c . Now our assertion follows from (1) and (2). Clearly, it
can also be written as  (w0c ) ( ) = + , and therefore the mapping  (w0c ) w0
leaves + invariant. Now, the elements w0c , w0 W are induced by inner auto-
morphisms ac , a N respectively, and it follows that 1 = Aut(g, t, ),
where = aac N . Writing = s, where N, s Aut , we get 1 = 1 s,
where 1 = lies in the subgroup T = {exp ad h | h t} of Int g, due to
Proposition 4.1. Then for the transposed linear transformations of t(R) we have
s = 1 =  (w0c ) w0 =  ((w0c ) )(w0 ), whence s (w0 ) =  ((w0c ) ).
The left-hand side leaves invariant and induces s0 = s Aut . Therefore the
right-hand side also leaves invariant and
s0 = ( ((w0c ) )) | . (5)
The relation (5) gives the desired description of the action of s0 on the Satake
diagram. In fact, if nc , then, by Lemma 2 (ii),

(w0c ) () = k ,
c
76 9. Appendix on Satake diagrams

where k are non-negative integers. Applying  and using Lemma 1 and (1),
we get  
s0 () = () + c k ,
c c

where c and k are non-negative integers. But s0 () and () lie in nc ,


whence s0 () = (). If c , then (5) implies

s0 () = ( ((w0c ) ))() =  (c ()) = c (),

by (1).
The last assertion is easy to verify case-by-case using Table 5. 
Tables

Table 1
Dynkin diagrams and their automorphism groups
Type Complex Lie algebra Dynkin diagram Aut
A1 sl2 (C) {e}

Al , l 2 sll+1 (C) . . . Z2

Bl , l 2 so2l+1 (C) . . . > {e}

Cl , l 2 sp2l (C) . . . < {e}



D4 so8 (C) .. S3
..



Dl , l 5 so2l (C) . . . .. Z2
..


E6 Z2

E7 {e}


E8 {e}

F4 < {e}

G2 < {e}

Table 2
Relations between , C , (C) (V0 is over R)

: g0 gl(V0 ) C : g0 gl(V0 (C)) (C) : g gl(V0 (C))


irreducible = C irreducible (C) irreducible
1 2 = C C
1 2 1 (C) 2 (C)
78 Tables

Table 3
Relations between , R , (C) (V is over C)

R : g0 gl(VR ) : g0 gl(V ) (C) : g gl(V (C))


R irreducible = irreducible (C) irreducible
(1 )R (2 )R = 1 2 1 (C) 2 (C)

Table 4
Elements h t(R) of the principal sl2 subalgebra
Type Coecients ri
l 2(l1) 3(l2) (l1)2 l
Al . . .
2l 2(2l1) 3(2l2) (l1)(l+2) l(l+1)/2
Bl , l 2 . . . >
2l1 2(2l2) 3(2l3) (l1)(l+1) ll
Cl , l 2 . . . <

(l1)l/2
yy
2l2 2(2l3) 3(2l4) (l3)(l+2)
yyy
Dl , l 4 . . . GyG(l2)(l+1)
GG
GG
l(l1)/2
16 30 42 30 16
E6
22

27 52 75 96 66 34
E7
49

58 114 168 220 270 182 92


E8
136
22 42 30 16
F4 <
10 6
G2 <
Tables 79

Table 5
Indices of irreducible representations of simple complex Lie algebras
Real form Satake diagram with a weight s0 s Index
1 2 l1 l
sll+1 (R) e = e = e +1

slm (H) m
1 2 3 l1 l
e = e = e (1) i=1 2i1
l = 2m 1
1 2 p p+1
O O O
sup,l+1p = e = e e
1 p 2l
  
l l1 lp l+1p
l even +1
l = 2m 1 (1)(m+p)m
1 2 p1
sup,p O O O LLL
l = 2p 1 L = e = e
r p e +1
p2   rrr
2p 2p1 p+1

1 2 l1 l
sul+1 = e = e e
l even +1
l = 2m 1 (1)mm
sop,2l+1p 1 p p+1 l1 l
> e e e
1pl
l(l+1)
p = 2k (1)(k+ 2 )l
l(l+3)
p = 2k + 1 (1)(k+ 2 )l
1 2 l1 l l(l+1)
l
so2l+1 > e e e (1) 2

spp,lp  [ l+1
2
]
1 2 3 2p 2p+1 l1 l 2i1
< e e e (1) i=1
1 p l1
2
spp,p 1 2 3 2p2 2p1 2p p
2i1
< e e e (1) i=1
l = 2p
1 2 l1 l
sp2l (R) < e e e +1
 [ l+1
2
]
1 2 l1 l 2i1
spl < e e e (1) i=1

(continued)
80 Tables

Table 5 continued
Real form Satake diagram with a weight s0 s Index
l1
sop,2lp {{
{
1 2 p p+1
1pl2 EEl2
E
l
p, l even e e lp
e (1) 2 (l1 +l )
p, l odd = e = e
p even, l odd = e e
= e +1
p odd, l even e = e
O l1
1 2 l2zzz
sol1,l+1 FFF
 l
l even e = e
= e +1
l odd = e e
l1
1 2 l2zzz
sol,l FFF
l
l even e e
e +1
l odd = e = e
l1
1 2 l2zzz
so2l FFF
l
l
l even e e e (1) 2 (l1 +l )
l odd e = e
= e +1
l1
ul (H) 1 2 3 l3 l2zzz m
2i1
l = 2m FFF e e e (1) i=1

O l1
ul (H) 1 2 3 l3 l2zzz m
FFF = e = e e (1) i=1 2i1
l = 2m + 1
 l
(continued)
Tables 81

Table 5 continued
Real form Satake diagram with a weight s0 s Index
1 2 3 4 5

EI e = e = e +1
6

EII x w ' &


= e = e e +1
1 2 3 4 5

EIII x &
= e = e e +1
1 2 3 4 5

6
1 2 3 4 5

EIV e = e = e +1
6

1 2 3 4 5
compact
= e = e e +1
form of E6
6

1 2 3 4 5 6

EV e e e +1
7

1 2 3 4 5 6

EVI e e e (1)1 +3 +7
7

1 2 3 4 5 6

EVII e e e +1
7

1 2 3 4 5 6
compact
e e e (1)1 +3 +7
form of E7
7

(continued)
82 Tables

Table 5 continued
Real form Satake diagram with a weight s0 s Index
1 2 3 4 5 6 7

EVIII e e e +1
8

1 2 3 4 5 6 7

EIX e e e +1
8

1 2 3 4 5 6 7
compact
e e e +1
form of E8
8

1 2 3 4
FI < e e e +1

1 2 3 4
F II < e e e +1

compact 1 2 3 4
< e e e +1
form of F4
1 2
G < e e e +1

compact 1 2
< e e e +1
form of G2
References
[1] N. Bourbaki, Groupes et algebres de Lie, Ch. IVVI, Hermann, Paris, 1968.
[2] E. Cartan, Les groupes projectifs qui ne laissent invariante aucune multiplicite plane, Bull.
Soc. Math. 41 (1913), 5394.
[3] E. Cartan, Les groupes projectifs continus reels qui ne laissent invariante aucune multi-
plicite plane, J. Math. pures appl. 10 (1914), 149186.
[4] E. Cartan, Les groupes reels simples, nis et continus, Ann. Ec. Norm. Sup. 31 (1914),
263355.
[5] E.B. Dynkin, Certain properties of the system of weights of linear representations of
semisimple Lie groups, Dokl. Akad. Nauk SSSR 76 (1951), 221224. (Russian)
[6] E.B. Dynkin, Semisimple subalgebras of semisimple Lie algebras, Mat. Sb. 30 (1952),
349462 (Russian); English transl. in Amer. Math. Soc. Transl. (2) 6 (1957), 111244.
[7] E.B. Dynkin, A.L. Onishchik, Compact Lie groups in the large, Uspekhi Mat. Nauk 10
(1955), 374 (Russian); English transl. in Amer. Math. Soc. Transl. (2) 21 (1962), 119192.
[8] F.R. Gantmacher, On the classication of real simple Lie groups, Mat. Sb. 5 (1939), 217
250.
[9] V.V. Gorbatsevich, A.L. Onishchik, E.B. Vinberg, Structure of Lie groups and Lie algebras,
Lie Groups and Lie Algebras III. Enc. Math. Sci., vol. 41, Springer-Verlag, Berlin e.a., 1994.
[10] M. Goto, F.D. Grosshans, Semisimple Lie Algebras, Marcel Dekker, New York Basel,
1978.
[11] S. Helgason, Dierential Geometry, Lie Groups, and Symmetric Spaces, Academic Press,
New York e.a., 1978.
[12] S. Helgason, Groups and Geometric Analysis, Academic Press, New York e.a., 1984.
[13] N. Iwahori, On real irreducible representations of Lie algebras, Nagoya Math. J. 14 (1959),
5983.
[14] F.I. Karpelevich, Transitivity surfaces of a semisimple subgroup of the motion group of a
symmetric space, Dokl. Akad. Nauk SSSR 93 (1953), 401404. (Russian)
[15] F.I. Karpelevich, Simple subalgebras of real Lie algebras, Trudy Mosk. Mat. Obshch. 4
(1955), 3112. (Russian)
[16] A.W. Knapp, Lie Groups beyond an Introduction, Birkhauser, Boston, 1996.
[17] A.I. Maltsev, On semisimple subgroups of Lie groups, Izv. AN SSSR. Ser. mat. 8 (1944),
no. 4, 143174 (Russian); English transl. in Amer. Math. Soc. Transl. (1) 11 (1957), 172
213.
[18] G.D. Mostow, Some new decomposition theorems for semisimple Lie groups, Mem. Amer.
Math. Soc. 14 (1955), 3154.
[19] A.L. Onishchik, E.B. Vinberg, Lie Groups and Algebraic Groups, Springer-Verlag, Berlin
e.a., 1990.
[20] I. Satake, On representations and compactications of symmetric Riemannian spaces, Ann.
of Math. 71, No.2 (1960), 77110.
[21] J.-P. Serre, Cohomologie Galoisienne, Lect. Notes in Math., 5, Springer-Verlag, Berlin,
1964.
[22] J.-P. Serre, Complex Semisimple Lie Algebras, Springer-Verlag, Berlin e.a., 1987.
[23] R. Steinberg, Lectures on Chevalley Groups, Yale Univ., New Haven, 1968.
[24] J. Tits, Tabellen zu den einfachen Lieschen Gruppen und ihren Darstellungen, Lect. Notes
in Math., 40, Springer-Verlag, Berlin, 1967.
Subject Index
Algebra, adjoint 1
of inner derivations 1
antiautomorphism 12, 13
, involutive 12
antiinvolution 12, 13
automorphism, corresponding to a real structure 20
, diagram 27
, inner of a Lie algebra 1
, group 1
, outer of a Lie algebra 1
of a system of simple roots 27
Base of a system of roots 7
Cartan decomposition 36
index 67
matrix 7
character 61
complexification 12, 64
component of a representation, irreducible 9
conjugation, complex 12
coroot 5
Dynkin diagram 8
of an irreducible representation 10
Element, regular 6
, singular 6
embedding, canonical 46
extension of a mapping 44
Form, bilinear (sesquilinear) invariant on a Lie algebra 1
, invariant under a representation 60
, linear dominant 10
, real 12, 13
, normal (split) 17
, of the first kind 48
, of the second kind 48
forms, real compatible 24
, of the same kind 47
function, strictly convex, 41
Group, linear adjoint 1
Integral, invariant 2
involution 12
Karpelevich index 53
Killing form 1
Lie algebra, compact 3
, complex conjugate 14
, reductive 2
, semisimple 2
group, semisimple 2
Operator, adjoint 22
, Hermitian, 22
Subject Index 85

, positive definite, 22
, self-adjoint, 22
, symmetric, 22
Presentation of an automorphism, canonical 34
product, scalar 2
of representations, tensor 11
Rank of a semisimple Lie algebra 4
realification 12, 64
representation of a Lie algebra 8
, adjoint 1
, basic 10
, completely reducible 9
, complex 64
, conjugate 64
, dual (contragredient) 9
, irreducible 9, 64
, matrix form 9
, orthogonal 61
, real 64
, self-conjugate 66
, self-dual 53
, symplectic 61
group, adjoint, 1
representations, equivalent (isomorphic) 8, 64
root 4
, compact 73
, negative 7
, non-compact 73
, positive 7
, simple 7
space 4
decomposition 4
vector 5
R-subalgebra 48
S-homomorphism 47
S-subalgebra 48
Satake diagram 74
signature 53
sl2 -triple 5
structure, complex 12, 13
, invariant 64
, quaternion 14
, invariant 64
, real 12, 13
, compact 23
, invariant 64
subalgebra, maximal toral 4
, principal three-dimensional 31
, regular 48
subrepresentation 9
sum of representations 9
system of generators, canonical 4
roots 4
weights 10
86 Subject Index

Theorem of Weyl 2

Vector, highest 10
space, complex conjugate 12
subspace, invariant 9
Weight, fundamental 10
, highest 10, 68
, lowest 71
space decomposition 9
subspace 10
Weyl chamber 6
group 6
involution 18
Z2 -grading 3

You might also like