Sunteți pe pagina 1din 152

Lie groups

E.P. van den Ban


Lecture Notes, Spring 2010

Contents
1 Groups

2 Lie groups, definition and examples

3 Invariant vector fields and the exponential map

15

4 The Lie algebra of a Lie group

18

5 Commuting elements

22

6 Commutative Lie groups

25

7 Lie subgroups

28

8 Proof of the analytic subgroup theorem

32

9 Closed subgroups

37

10 The groups SU(2) and SO(3)

40

11 Group actions and orbit spaces

43

12 Smooth actions and principal fiber bundles

45

13 Proper free actions

49

14 Coset spaces

53

15 Orbits of smooth actions

54

16 Intermezzo: the Baire category theorem

56

17 Normal subgroups and ideals

58

18 Detour: actions of discrete groups

61

19 Densities and integration

62

20 Representations

69

21 Schur orthogonality

77

22 Characters

81

23 The Peter-Weyl theorem

85

24 Appendix: compact self-adjoint operators

87

25 Proof of the Peter-Weyl Theorem

90

26 Class functions

93

27 Abelian groups and Fourier series

94

28 The group SU(2)

96

29 Lie algebra representations

100

30 Representations of sl(2,C)

102

31 Roots and weights

105

32 Conjugacy of maximal tori

115

33 Automorphisms of a Lie algebra

116

34 The Killing form

117

35 Compact and reductive Lie algebras

118

36 Root systems for compact algebras

122

37 Weyls formulas

128

38 The classification of root systems


130
38.1 Cartan integers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
38.2 Fundamental and positive systems . . . . . . . . . . . . . . . . . . . . . . . . . 132
38.3 The rank two root systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
2

38.4 Weyl chambers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139


38.5 Dynkin diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145

1 Groups
The purpose of this section is to collect some basic facts about groups. We leave it to the reader
to prove the easy statements given in the text.
We recall that a group is a set G together with a map  W G  G ! G; .x; y/ 7! xy and an
element e D eG ; such that the following conditions are fulfilled
(a) .xy/z D x.yz/ for all x; y; z 2 GI

(b) xe D ex D x for all x 2 GI


(c) for every x 2 G there exists an element x

2 G such that xx

Dx

x D e:

Remark 1.1 Property (a) is called associativity of the group operation. The element e is called
the neutral element of the group.
The element x 1 is uniquely determined by the property (c); indeed, if x 2 G is given, and
y 2 G an element with xy D e; then x 1 .xy/ D x 1 e D x 1 ; hence x 1 D .x 1 x/y D ey D
y: The element x 1 is called the inverse of x:
Example 1.2 Let S be a set. Then Sym.S/; the set of bijections S ! S; equipped with
composition, is a group. The neutral element e equals IS ; the identity map S ! S; x 7! x:
If S D f1; : : : ; ng; then Sym.S/ equals Sn ; the group of permutations of n elements.
A group G is said to be commutative or abelian if xy D yx for all x; y 2 G: We recall that a
subgroup of G is a subset H  G such that
(a) eG 2 H I

(b) xy 2 H for all x 2 H and y 2 H I


(c) x

2 H for every x 2 H:

We note that a subgroup is a group of its own right. If G; H are groups, then a homomorphism
from G to H is defined to be a map ' W G ! H such that
(a) '.eG / D eH I

(b) '.xy/ D '.x/'.y/ for all x; y 2 G:


We note that the image im.'/ WD '.G/ is a subgroup of H: The kernel of '; defined by
ker ' WD '

.feH g/ D fx 2 G j '.x/ D eH g

is also readily seen to be a subgroup of G: A surjective group homomorphism is called an epimorphism. An injective group homomorphism is called a monomorphism. We recall that a group
homomorphism ' W G ! H is injective if and only if its kernel is trivial, i.e., ker ' D feG g: A
bijective group homomorphism is called an isomorphism. The inverse ' 1 of an isomorphism
' W G ! H is a group homomorphism from H to G: Two groups G1 and G2 are called isomorphic if there exists an isomorphism from G1 onto G2 :
If G is a group, then by an automorphism of G we mean an isomorphism of G onto itself.
The collection of such automorphisms, denoted Aut.G/; is a subgroup of Sym.G/:
4

Example 1.3 If G is a group and x 2 G; then the map lx W G ! G; y 7! xy; is called left
translation by x: We leave it to the reader to verify that x 7! lx is a group homomorphism from
G to Sym.G/:
Likewise, if x 2 G; then rx W G ! G; y 7! yx; is called right translation by x: We leave it
to the reader to verify that x 7! .rx / 1 is a group homomorphism from G to Sym.G/:
If x 2 G; then Cx W G ! G; y 7! xyx 1 is called conjugation by x: We note that Cx is an
automorphism of G; with inverse Cx 1 : The map C W x ! Cx is a group homomorphism from G
into Aut.G/: Its kernel is the subgroup of G consisting of the elements x 2 G with the property
that xyx 1 D y for all y 2 G; or, equivalently, that xy D yx for all y 2 G: Thus, the kernel of
C equals the center Z.G/ of G:
We end this preparatory section with the isomorphism theorem for groups. To start with we recall
that a relation on a set S is a subset R of the Cartesian product S  S: We agree to also write
xRy in stead of .x; y/ 2 R: A relation  on S is called an equivalence relation if the following
conditions are fulfilled, for all x; y; z 2 S;
(a) x  x (reflexivity);
(b) x  y ) y  x (symmetry);
(c) x  y ^ y  z ) x  z (transitivity).
If x 2 S; then the collection x WD fy 2 S j y  xg is called the equivalence class of x: The
collection of all equivalence classes is denoted by S=  :
A partition of a set S is a collection P of non-empty subsets of S with the following properties
(a) if A; B 2 P; then A \ B D ; or A D BI
(b) [A2P A D S:
If  is an equivalence relation on S then S=  is a partition of S: Conversely, if P is a
partition of S; we may define a relation P as follows: x P y if and only if there exists a
set A 2 P such that x and y both belong to A: One readily verifies that P is an equivalence
relation; moreover, S= P D P:
Equivalence relations naturally occur in the context of maps. If f W S ! T is a map between
sets, then the relation  on S defined by x  y () f .x/ D f .y/ is an equivalence relation.
If x 2 S and f .x/ D c; then the class x equals the fiber
f

.c/ WD f

.fcg/ D fy 2 S j f .y/ D cg:

Let  denote the natural map x 7! x from S onto S=  : Then there exists a unique map
fN W S=  ! T such that the following diagram commutes
f

! T
% fN

S
 #
S= 
5

We say that f factors through a map fN W S=  ! T: Note that fN.x/ D f .x/ for all x 2 S: The
map fN is injective, and has image equal to f .S/: Thus, if f is surjective, then fN is a bijection
from S=  onto T:
Partitions, hence equivalence relations, naturally occur in the context of subgroups. If K is a
subgroup of a group G; then for every x 2 G we define the right coset of x by xK WD lx .K/:
The collection of these cosets, called the right coset space, is a partition of G and denoted by
G=K: The associated equivalence relation is given by x  y () xK D yK; for all x; y 2 G:
The subgroup K is called a normal subgroup if xKx 1 D K; for every x 2 G: If K is a
normal subgroup then G=K carries a unique group structure for which the natural map  W G !
G=K; x 7! xK is a homomorphism. Accordingly, xK  yK D .x/.y/ D .xy/ D xyK:
Lemma 1.4 (The isomorphism theorem) Let f W G ! H be an epimorphism of groups. Then
K WD ker f is a normal subgroup of G: There exists a unique map fN W G=K ! H; such that
fN  D f: The factor map fN is an isomorphism of groups.
Proof: Let x 2 G and k 2 K: Then f .xkx 1 / D f .x/f .k/f .x/ 1 D f .x/eH f .x/ 1 D eH ;
hence xkx 1 2 ker f D K: It follows that xKx 1  K: Similarly it follows that x 1 Kx  K;
hence K  xKx 1 and we see that xKx 1 D K: It follows that K is normal.
Let x 2 G and write f .x/ D h: Then, for every y 2 G; we have yK D xK () f .y/ D
f .x/ () y 2 f 1 .h/: Hence G=K consists of the fibers of f: In the above we saw that there
exists a unique map fN W G=K ! H; such that fN  D f: The factor map is bijective, since f is
surjective. It remains to be checked that fN is a homomorphism. Now fN.eK/ D f .eG / D eH ;
since f is a homomorphism. Moreover, if x; y 2 G; then fN.xKyK/ D fN.xyK/ D f .xy/ D
f .x/f .y/: This completes the proof.


2 Lie groups, definition and examples


Definition 2.1 (Lie group) A Lie group is a smooth (i.e., C 1 ) manifold G equipped with a
group structure so that the maps  W .x; y/ 7! xy; G  G ! G and  W x 7! x 1 ; G ! G are
smooth.
Remark 2.2 For a Lie group, the group operation is usually denoted multiplicatively as above.
The neutral element is denoted by e D eG : Sometimes, if the group is commutative, i.e.,
.x; y/ D .y; x/ for all x; y 2 G; the group operation is denoted additively, .x; y/ 7! x C yI
in this case the neutral element is denoted by 0:
Example 2.3 We begin with a few easy examples of Lie groups.
(a) Rn together with addition C and the neutral element 0 is a Lie group.
(b) Cn ' R2n together with addition C and the neutral element 0 is a Lie group.
(c) R WD R n f0g is an open subset of R; hence a smooth manifold. Equipped with the
ordinary scalar multiplication and the neutral element 1; R is a Lie group. Similarly, RC WD
0; 1 together with scalar multiplication and 1 is a Lie group.
(d) C WD C n f0g is an open subset of C ' R2 ; hence a smooth manifold. Together with
complex scalar multiplication and 1; C is a Lie group.
6

If G1 and G2 are Lie groups, we may equip the product manifold G D G1  G2 with the
product group structure, i.e., .x1 ; x2 /.y1 ; y2 / WD .x1 y1 ; x2 y2 /; and eG D .eG1 ; eG2 /:
Lemma 2.4 Let G1 ; G2 be Lie groups. Then G WD G1  G2 ; equipped with the above manifold
and group structure, is a Lie group.
Proof: The multiplication map  W G  G ! G is given by ..x1 ; x2 /; .y1 ; y2 // D 1 
2 ..x1 ; y1 /; .x2 ; y2 /: Hence,  D .1  2 / .IG1  S  IG2 /; where S W G2  G1 ! G1  G2
is the switch map given by S.x2 ; y1 / D .y1 ; x2 /: It follows that  is the composition of smooth
maps, hence smooth.
The inversion map  of G is given by  D .1 ; 2 /; hence smooth.

Lemma 2.5 Let G be a Lie group, and let H  G be both a subgroup and a smooth submanifold. Then H is a Lie group.
Proof: Let  D G W G  G ! G be the multiplication map of G: Then the multiplication map
H of H is given by H D jH H : Since  is smooth and H  H a smooth submanifold of
G  G; the map H W H  H ! G is smooth. Since H is a subgroup, H maps into the smooth
submanifold H; hence is smooth as a map H  H ! H: Likewise, H D G jH is smooth as a
map H ! H:

Example 2.6
(a) The unit circle T WD fz 2 C j jzj D 1g is a smooth submanifold as well as a subgroup of
the Lie group C : Therefore it is a Lie group.
(b) The q-dimensional torus Tq is a Lie group.
So far, all of our examples of Lie groups were commutative. We shall formulate a result that
asserts that interesting connected Lie groups are not to be found among the commutative ones.
For this we need the concept of isomorphic Lie groups.
Definition 2.7 Let G and H be Lie groups.
(a) A Lie group homomorphism from G to H is a smooth map ' W G ! H that is a homomorphism of groups.
(b) An Lie group isomorphism from G onto H is a bijective Lie group homomorphism ' W
G ! H whose inverse is also a Lie group homomorphism.
(c) An automorphism of G is an isomorphism of G onto itself.
Remark 2.8 (a) If ' W G ! H is a Lie group isomorphism, then ' is smooth and bijective and
its inverse is smooth as well. Hence, ' is a diffeomorphism.
(b) The collection of Lie group automorphisms of G; equipped with composition, forms a
group, denoted Aut.G/:
7

We recall that a topological space X is said to be connected if ; and X are the only subsets
of X that are both open and closed. The space X is said to be arcwise connected if for each pair
of points a; b 2 X there exists a continous curve c W 0; 1 ! X with initial point a and end
point b; i.e., c.0/ D a and c.1/ D b: If X is a manifold then X is connected if and only if X is
arcwise connected.
We can now formulate the promised results about connected commutative Lie groups.
Theorem 2.9 Let G be a connected commutative Lie group. Then there exist integers p; q  0
such that G is isomorphic to Tp  Rq :
The proof of this theorem will be given at a later stage, when we have developed enough
technology. See Theorem 6.1.
A more interesting example is the following. In the sequel we will often discuss new general
concepts in the context of this important particular example.
Example 2.10 Let n be a positive integer, and let M.n; R/ be the set of real n  n matrices.
Equipped with entry wise addition and scalar multiplication, M.n; R/ is a linear space, which in
2
an obvious way may be identified with Rn : For A 2 M.n; R/ we denote by Aij the entry of A
in the i -th row and the j -th column. The maps ij W A 7! Aij may be viewed as a system of
(linear) coordinate functions on M.n; R/:
In terms of these coordinate functions, the determinant function det W M.n; R/ ! R is given
by
X
det D
sgn ./1.1/    n.n/ ;
2Sn

where Sn denotes the group of permutations of f1; : : : ; ng; and where sgn denotes the sign of a
permutation. It follows from this formula that det is smooth.
The set GL.n; R/ of invertible matrices in M.n; R/; equipped with the multiplication of matrices, is a group. As a set it is given by
GL.n; R/ D fA 2 M.n; R/ j detA 0g:
Thus, GL.n; R/ is the pre-image of the open subset R D R n f0g of R under det: As the latter
function is continuous, it follows that GL.n; R/ is an open subset of M.n; R/: As such, it may
be viewed as a smooth manifold of dimension n2 : In terms of the coordinate functions ij ; the
multiplication map  W GL.n; R/  GL.n; R/ ! GL.n; R/ is given by
kl ..A; B// D

n
X

ki .A/il .B/:

iD1

It follows that  is smooth. Given A 2 M.n; R/ we denote by AT the transpose of A: Moreover,


for 1  i; j  n we denote by Mij .A/ the matrix obtained from A by deleting the i -th row
iCj
and j -th column. The co-matrix of A is defined by Aco
detMij .AT /: Clearly, the
ij D . 1/
8

map A 7! Aco is a polynomial, hence smooth map from M.n; R/ to itself. By Cramers rule the
inversion  W GL.n; R/ ! GL.n; R/; A ! A 1 is given by
.A/ D .detA/

Aco :

It follows that  is smooth, and we see that GL.n; R/ is a Lie group.


Example 2.11 Let V be a real linear space of finite dimension n: Let v D .v1 ; : : : ; vn / be
an ordered basis of V: Then there is a unique linear isomorphism ev from Rn onto V; mapping
the j -th standard basis vector ej onto vj : If w is a second basis, then L WD ev 1 ew is a linear
isomorphism of Rn onto itself, hence a diffeomorphism. It follows that V has a unique structure
of smooth manifold such that the map ev is a diffeomorphism, for any choice of basis v:
We denote by End.V / the set of linear endomorphisms of V; i.e., linear maps of V into
itself. Equipped with pointwise addition and scalar multiplication, End.V / is a linear space. Let
v D .v1 ; : : : ; vn/ be an ordered basis of V: Given A 2 End.V /; we write mat.A/ D matv .A/
for the matrix of A with respect to v: The entries Aij of this matrix are determined by Avj D
P
n
iD1 Aij vi ; for all 1  j  n: As in Example 2.10 we denote by M.n; R/ the set of all real nn
matrices. Equipped with entry wise addition and scalar multiplication, M.n; R/ is a linear space.
Accordingly, mat is a linear isomorphism from End.V / onto M.n; R/: Via this map, composition
in End.V / corresponds with matrix multiplication in M.n; R/: More precisely,
mat.A B/ D mat.A/ mat.B/
for all A; B 2 End.V /:
We note that the matrix matv .A/ equals the matrix of ev 1 A ev with respect to the standard
basis of Rn : Let now w D .w1 ; : : : ; wn / be a second ordered basis of V and let S be the matrix
of the linear endomorphism L D ev 1 ew 2 End.Rn / with respect to the standard basis. Then
from ev 1 Aev D L ew 1 Aew L 1 it follows that
matv .A/ D S matw .A/ S

By conjugation invariance of determinant and trace, we find that


detmatv .A/ D detmatw A and

trmatv .A/ D trmatw A

for all A 2 End.V /: It follows that determinant and trace are independent of the choice of
basis. Hence, there exist unique maps det; tr W End.V / ! R such that detA D detmatA and
trA D trmatA for any choice of basis.
We denote by GL.V /; or also Aut.V /; the set of invertible elements of End.V /: Then GL.V /
is a group. Moreover, fix a basis of V; then the associated matrix map mat W End.V / ! M.n; R/
is a diffeomorphism, mapping GL.V / onto GL.n; R/: It follows that GL.V / is an open subset,
hence a submanifold of End.V /: Moreover, as mat restricts to a group isomorphism from GL.V /
onto GL.n; R/; it follows from the discussion in the previous example that GL.V / is a Lie group
and that mat is an isomorphism of Lie groups from GL.V / onto GL.n; R/:
9

Remark 2.12 In the above example we have distinguished between linear maps and their matrices with respect to a basis. In the particular situation that V D Rn ; we shall often use the map
mat D mate ; defined relative to the standard basis e of Rn to identify the linear space End.Rn /
with M.n; R/ and to identify the Lie group GL.Rn / with GL.n; R/:
We shall now discuss an important criterion for a subgroup of a Lie group G to be a Lie
group. In particular this criterion will have useful applications for G D GL.V /: We start with a
result that illustrates the idea of homogeneity.
Let G be a Lie group. If x 2 G; then the left translation lx W G ! G; see Example 1.3, is
given by y 7! .x; y/; hence smooth. The map lx is bijective with inverse lx 1 ; which is also
smooth. Therefore, lx is a diffeomorphism from G onto itself. Likewise, the right multiplication
map rx W y 7! yx is a diffeomorphism from G onto itself. Thus, for every pair of points
a; b 2 G both lba 1 and ra 1 b are diffeomorphisms of G mapping a onto b: This allows us
to compare structures on G at different points. As a first application of this idea we have the
following.
Lemma 2.13 Let G be a Lie group and H a subgroup. Let h 2 H be a given point (in the
applications h D e will be most important). Then the following assertions are equivalent.
(a) H is a submanifold of G at the point hI
(b) H is a submanifold of G:
Proof: Obviously, (b) implies (a). Assume (a). Let n be the dimension of G and let m be the
dimension of H at h: Then m  n: Moreover, there exists an open neighborhood U of h in
G and a diffeomorphism  of U onto an open subset of Rn such that .h/ D 0 and such that
.U \ H / D .U / \ .Rm  f0g/: Let k 2 H: Put a D kh 1 : Then la is a diffeomorphism of G
onto itself, mapping h onto k: We shall use this to show that H is a submanifold of dimension
m at the point k: Since a 2 H; the map la maps the subset H bijectively onto itself. The set
Uk WD la .U / is an open neighborhood of k in G: Moreover, k D  la 1 is a diffeomorphism
of Uk onto the open subset .U / of Rn : Finally,
k .Uk \ H / D k .la U \ la H / D k la .U \ H / D .U \ H / D .U / \ .Rm  f0g/:
This shows that H is a submanifold of dimension m at the point k: Since k was an arbitrary point
of H; assertion (b) follows.

Example 2.14 Let V be a finite dimensional real linear space. We define the special linear
group
SL.V / WD fA 2 GL.V / j detA D 1g:

Note that det is a group homomorphism from GL.V / to R : Moreover, SL.V / is the kernel of
det: In particular, SL.V / is a subgroup of GL.V /: We will show that SL.V / is a submanifold of
GL.V / of codimension 1: By Lemma 2.13 it suffices to do this at the element I D IV :
10

Since G WD GL.V / is an open subset of the linear space End.V / its tangent space TI G may
be identified with End.V /: The determinant function is smooth from G to R hence its tangent
map is a linear map from End.V / to R: In Lemma 2.15 below we show that this tangent map is
the trace tr W End.V / ! R; A 7! tr.A/: Clearly tr is a surjective linear map. This implies that det
is submersive at I: By the submersion theorem, it follows that SL.V / is a smooth codimension 1
submanifold at I:
Lemma 2.15 The function det W GL.V / ! R has tangent map at I given by TI det D tr W
End.V / ! R; A 7! trA:
Proof: Put G D GL.V /: In the discussion in Example 2.14 we saw that TI G D End.V / and,
similarly, T1 R D R: Thus TI det is a linear map End.V / ! R: Let H 2 End.V /: Then by the
chain rule,

d
det.I C tH /:
TI .det/.H / D
dt t D0
Fix a basis v1 ; : : : ; vn of V: We denote the matrix coefficients of a map A 2 End.V / with respect
to this basis by Aij ; for 1  i; j  n: Using the definition of the determinant, we obtain
det.I C tH / D 1 C t .H11 C    C Hnn / C t 2 R.t; H /;
where R is polynomial in t and the matrix coefficients Hij : Differentiating this expression with
respect to t and substituting t D 0 we obtain
TI .det/.H / D H11 C    C Hnn D trH:

We shall now formulate a result that allows us to give many examples of Lie groups. The
complete proof of this result will be given at a later stage. Of course we will make sure not to
use the result in the development of the theory until then.
Theorem 2.16 Let G be a Lie group and let H be a subgroup of G: Then the following assertions are equivalent.
(a) H is closed in the sense of topology.
(b) H is a submanifold.
Proof: For the moment we will only prove that (b) implies (a). Assume (b). Then there exists
an open neighborhood U of e in G such that U \ HN D U \ H: Let y 2 HN : Since ly is a
diffeomorphism from G onto itself, yU is an open neighborhood of y in G; hence yU \ H ;:
Select h 2 yU \ H: Then y 1 h 2 U: On the other hand, from y 2 HN ; h 2 H it follows that
y 1 h 2 HN : Hence, y 1 h 2 U \ HN D U \H; and we see that y 2 H: We conclude that HN  H:
Therefore, H is closed.


11

By a closed subgroup of a Lie group G we mean a subgroup that is closed in the sense of
topology.
Corollary 2.17 Let G be a Lie group. Then every closed subgroup of G is a Lie group.
Proof: Let H be a closed subgroup of G: Then H is a smooth submanifold of G, by Theorem
2.16. By Lemma 2.5 it follows that H is a Lie group.

Corollary 2.18 Let ' W G ! H be a homomorphism of Lie groups. Then the kernel of ' is a
closed subgroup of G: In particular, ker ' is a Lie group.
Proof: Put K D ker ': Then K is a subgroup of G: Now ' is continuous and feH g is a closed
subset of H: Hence, K D ' 1 .feH g/ is a closed subset of G: Now apply Corollary 2.17.

Remark 2.19 We may apply the above corollary in Example 2.14 as follows. The map det W
GL.V / ! R is a Lie group homomorphism. Therefore, its kernel SL.V / is a Lie group.
Example 2.20 Let now V be a complex linear space of finite complex dimension n: Then by
End.V / we denote the complex linear space of complex linear maps from V to itself, and by
GL.V / the group of invertible maps. The discussion of Examples 2.10 and 2.11 goes through
with everywhere R replaced by C: In particular, the determinant det is a complex polynomial map
End.V / ! C; hence continuous. Since C D C n f0g is open in C; the set GL.V / D det 1 .C /
is open in End.V /: As in Example 2.11 we now see that GL.V / is a Lie group.
The map det W GL.V / ! C is a Lie group homomorphism. Hence, by Corollary 2.18 its
kernel, SL.V / WD fA 2 GL.V / j detA D 1g; is a Lie group.
Finally, let v D .v1 ; : : : ; vn/ be a basis of V (over C). Then the associated matrix map
mat D matv is a complex linear isomorphism from End.V / onto the space M.n; C/ of complex
n  n matrices. It restricts to a Lie group isomorphism GL.V / ' GL.n; C/ and to a Lie group
isomorphism SL.V / ' SL.n; C/:
Another very useful application of Corollary 2.17 is the following. Let V be a finite dimensional real linear space, and let W V  V ! W be a bilinear map into a finite dimensional
real linear space W: For g 2 GL.V / we define the bilinear map g  W V  V ! W by
g  .u; v/ D .g 1 u; g 1 v/: From g1  .g2  / D .g1 g2 /  one readily deduces that the
stabilizer of in GL.V /;
GL.V / D fg 2 GL.V / j g  D g
is a subgroup of GL.V /: Similarly SL.V / WD SL.V / \ GL.V / is a subgroup.
Lemma 2.21 The groups GL.V / and SL.V / are closed subgroups of GL.V /: In particular,
they are Lie groups.

12

Proof: Define Cu;v D fg 2 GL.V / j .g 1 u; g 1 v/ D .u; v/g; for u; v 2 V: Then GL.V /


is the intersection of the sets Cu;v ; for all u; v 2 V: Thus, to establish closedness of this group,
it suffices to show that each of the sets Cu;v is closed in GL.V /: For this, we consider the function f W GL.V / ! W given by f .g/ D .g 1 u; g 1 v/: Then f D .; /; hence f is
continuous. Since f.u; v/g is a closed subset of W; it follows that Cu;v D f 1 .f.u; v/g/ is
closed in GL.V /: This establishes that GL.V / is a closed subgroup of GL.V /: By application
of Corollary 2.17 it follows that GL.V / is a Lie group.
Since SL.V / is a closed subgroup of GL.V / as well, it follows that SL.V / D SL.V / \
GL.V / is a closed subgroup, hence a Lie group.

By application of the above to particular bilinear forms, we obtain interesting Lie groups.
Example 2.22 (a) Take V D Rn and the standard inner product on Rn : Then GL.V / D
O.n/; the orthogonal group. Moreover, SL.V / D SO.n/; the special orthogonal group.
Example 2.23 Let n D p C q; with p; q positive integers and put V D Rn : Let be the
standard inner product of signature .p; q/; i.e.,
.x; y/ D

p
X

xi yi

iD1

n
X

xi yi :

iDpC1

Then GL.V / D O.p; q/ and SL.V / D SO.p; q/: In particular, we see that the Lorentz group
O.3; 1/ is a Lie group.
Example 2.24 Let V D R2n and let be the standard symplectic form given by
.x; y/ D

n
X

xi ynCi

iD1

n
X

xnCi yi :

iD1

Then GL.V / is the real symplectic group Sp.n; R/:


Example 2.25 Let V be a finite dimensional complex linear space, equipped with a complex
inner product : This inner product is not a complex bilinear form, since it is skew linear in its
second component (this will always be our convention with complex inner products). However,
as a map V  V ! C it is bilinear over RI in particular, it is continuous. As in the proof of
Lemma 2.21 we infer that the associated unitary group U.V / D GL.V / is a closed subgroup
of GL.V /; hence a Lie group. Likewise, the special unitary group SU.V / WD U.V / \ SL.V / is
a Lie group.
Via the standard basis of Cn we identify End.Cn / ' M.n; C/ and GL.Cn / ' GL.n; C/; see
also Remark 2.12. We equip Cn with the standard inner product given by
hz ; wi D

n
X
iD1

zi wN i

.z; w 2 Cn /:

The associated unitary group U.Cn / may be identified with the group U.n/ of unitary n  nmatrices. Similarly, SU.Cn / corresponds with the special unitary matrix group SU.n/:
13

Remark 2.26 It is possible to immediately apply Lemma 2.21 in the above example, in order
to conclude that U.n/ is closed. For this we observe that we may forget the complex structure of
V and view it as a real linear space. We write V.R/ for V viewed as a linear space. If n D dimC V
and if v1 ; : : : ; vn is a basis of V; then v1 ; iv1 ; : : : ; vn; ivn is a basis of the real linear space V.R/ :
In particular we see that dimR V.R/ D 2n: Any complex linear map T 2 End.V / may be viewed
as a real linear map from V to itself, hence as an element of End.V.R/ /; which we denote by
T.R/ : We note that T 7! T.R/ is a real linear embedding of End.V / into End.V.R/ /: Accordingly
we may view End.V / as a real linear subspace of End.V.R/ /: Let J denote multiplication by i;
viewed as a real linear endomorphism of V.R/ : We leave it to the reader to verify that
End.V / D fA 2 End.V.R/ / j A J D J Ag:
Accordingly,
GL.V / D fa 2 GL.V.R/ / j a J D J Ag:
From this one readily deduces that GL.V / is a closed subgroup of GL.V.R/ /: In the situation
of Example 2.25, H WD GL.V.R/ / is a closed subgroup of GL.V.R/ /; by Lemma 2.21. Hence
U.V / D GL.V / \ H is a closed subgroup as well.
We end this section with useful descriptions of the orthogonal, unitary and symplectic groups.
Example 2.27 For a matrix A 2 M.n; R/ we define its transpose At 2 M.n; R/ by .At /ij D
Aj i : Let D h  ;  i be the standard inner product on Rn : Then hAx ; yi D hx ; At yi: Let
a 2 GL.n; R/: Then for all x; y 2 Rn ;
a

 .x; y/ D hax ; ayi D hat ax ; yi:

Since O.n/ D GL.n; R/ ; we infer that


O.n/ D fa 2 GL.n; R/ j at a D I g:
Example 2.28 If A 2 M.n; C/ we denote its complex adjoint by .A /ij D ANj i : Let h  ;  i be
the complex standard inner product on Cn : Then hAx ; yi D hx ; A yi for all x; y 2 Cn : As in
the previous example we now deduce that
U.n/ D fa 2 GL.n; C/ j a a D I g:
Example 2.29 Let be the standard symplectic form on R2n ; see Example 2.24. Let J 2
M.2n; R/ be defined by


0 I
;
J D
I 0

where the indicated blocks are of size n  n:


Let h  ;  i denote the standard inner product on R2n : Then for all x; y 2 R2n ; we have
.x; y/ D hx ; Jyi: Let a 2 GL.n; R/; then
a

 .x; y/ D hax ; Jayi D hx ; at Jayi:


14

From this we see that Sp.n; R/ D GL.2n; R/ consists of all a 2 GL.2n; R/ with at Ja D J;
or, equivalently, with
.at / 1 D JaJ 1
(1)

This description motivates the following definition. The map A 7! At uniquely extends to
a complex linear endomorphism of M.2n; C/: This extension is given by the usual formula
.At /ij D Aj i : We now define Sp.n; C/ to be the collection of a 2 GL.2n; C/ satisfying condition (1). One readily verifies that Sp.n; C/ is a closed subgroup of GL.2n; C/ hence a Lie group.
We call it the complex symplectic group.
Note that GL.2n; R/ is a closed subgroup of GL.2n; C/ and that Sp.n; R/ D GL.2n; R/ \
Sp.n; C/:
Finally, we define the compact symplectic group by
Sp.n/ WD U.2n/ \ Sp.n; C/:

Clearly, this is a closed subgroup of GL.2n; C/; hence a Lie group.


Remark 2.30 In this section we have frequently used the following principle. If G is a Lie
group, and if H; K  G are closed subgroups, then H \ K is a closed subgroup, hence a Lie
group.

3 Invariant vector fields and the exponential map


If M is a manifold, we denote by V.M / the real linear space of smooth vector fields on M: A
vector field v 2 V.G/ is called left invariant, if .lx / v D v for all x 2 G; or, equivalently if
v.xy/ D Ty .lx / v.y/

.x; y 2 G/:

(2)

The collection of smooth left invariant vector fields is a linear subspace of V.G/; which we
denote by VL .G/: From the above equation with y D e we see that a left invariant vector field
is completely determined by its value v.e/ 2 Te G at e: Differently said, v 7! v.e/ defines an
injective linear map from VL .G/ into Te G: The next result asserts that this map is surjective as
well. If X 2 Te G; we define the vector field vX on G by
vX .x/ D Te .lx /X;

.x 2 G/:

(3)

Lemma 3.1 The map X 7! vX defines a linear isomorphism from Te G onto VL .G/: Its inverse
is given by v 7! v.e/:

Proof: From the fact that .x; y/ 7! lx .y/ is a smooth map G  G ! G; it follows by differentiation with respect to y at y D e in the direction of X 2 Te G that x 7! Te .lx /X is smooth as
a map G ! T G: This implies that vX is a smooth vector field on G: Hence X 7! vX defines a
real linear map Te G ! V.G/: We claim that it maps into VL .G/:
Fix X 2 Te G: Differentiating the relation lxy D lx ly and applying the chain rule we see
that Te .lxy / D Ty .lx /Te .ly /: Applying this to the definition of vX we see that vX satisfies (2),
hence is left invariant. This establishes the claim.
From vX .e/ D X we see that the map  W v 7! v.e/ from VL .G/ to Te G is not only injective,
but also surjective. Thus,  is a linear isomorphism, with inverse X 7! vX :

15

If X 2 Te G; we define X to be the maximal integral curve of vX with initial point e:


Lemma 3.2 Let X 2 Te G: Then the integral curve X has domain R: Moreover, we have
X .s C t / D X .s/X .t / for all s; t 2 R: Finally the map .t; X/ 7! X .t /; R  Te G ! G is
smooth.
Proof: Let be any integral curve for vX ; let y 2 G; and put 1 .t / D y.t /: Differentiating
this relation with respect to t we obtain:
d
d
1 .t / D T.t / ly .t / D T.t /ly vX ..t // D vX .1 .t //;
dt
dt
by left invariance of vX : Hence 1 is an integral curve for vX as well.
Let now I be the domain of X ; fix t1 2 I; and put x1 D X .t1 /: Then 1 .t / WD x1 X .t /
is an integral curve for vX with starting point x1 and domain I: On the other hand, the maximal
integral curve for vX with starting point x1 is given by 2 W t 7! X .t C t1 /: It has domain I t1 :
We infer that I  I t1 : It follows that s C t1 2 I for all s; t1 2 I: Hence, I D R:
Fix s 2 R; then by what we saw above c W t 7! X .s/X .t / is the maximal integral curve for
vX with initial pont X .s/: On the other hand, the same holds for d W t 7! X .s C t /: Hence, by
uniqueness of the maximal integral curve, c D d:
The final assertion is a consequence of the fact that the vector field vX depends linearly, hence
smoothly on the parameter X: Let 'X denote the flow of vX : Then it is a well known (local)
result that the map .X; t; x/ 7! 'X .t; x/ is smooth. In particular, .t; X/ 7! X .t / D 'X .t; e/ is
a smooth map R  Te G ! G:

Definition 3.3 Let G be a Lie group. The exponential map exp D expG W Te G ! G is defined
by
exp.X/ D X .1/
where X is defined as above; i.e., X is the maximal integral curve with initial point e of the
left invariant vector field vX on G determined by vX .e/ D X:
Example 3.4 We return to the example of the group GL.V /; with V a finite dimensional real
linear space. Its neutral element e equals I D IV : Since GL.V / is open in End.V /; we have
Te GL.V / D End.V /: If x 2 GL.V /; then lx is the restriction of the linear map Lx W A 7!
xA; End.V / ! End.V /; to GL.V /; hence Te .lx / D Lx ; and we see that for X 2 End.V /
the invariant vectorfield vX is given by vX .x/ D xX: Hence, the integral curve X satisfies the
equation:
d
.t / D .t /X:
dt
Since t 7! e tX is a solution to this equation with the same initial value, we must have that
X .t / D e tX : Thus in this case exp is the ordinary exponential map X 7! e X ; End.V / !
GL.V /:

16

Remark 3.5 In the above example we have used the exponential e A of an endomorphism A 2
End.V /: One way to define this exponential is precisely by the method of differential equations
just described. Another way is to introduce it by its power series
1
X
1 n
A :
e D
n
nD0
A

From the theory of power series it follows that A ! e A is a smooth map End.V / ! End.V /:
Moreover,
d tA
e D Ae tA D e tA A;
dt
by termwise differentiation of power series. By multiplication of power series we obtain
e X e Y D e XCY

if

X; Y 2 End.V /

commute, i.e.,

XY D YX:

(4)

Applying this with X D sA and Y D tA; we obtain e .sCt /A D e sA e tA ; for all A 2 End.V / and
s; t 2 R: This formula will be established in general in Lemma 3.6 (b) below.
Lemma 3.6 For all s; t 2 R; X 2 Te G we have
(a) exp.sX/ D X .s/:
(b) exp.s C t /X D exp sX exp tX:
Moreover, the map exp W Te G ! G is smooth and a local diffeomorphism at 0: Its tangent map
at the origin is given by T0 exp D ITe G :
Proof: Consider the curve c.t / D X .st /: Then c.0/ D e; and
d
c.t / D s P X .st / D s vX .X .st // D vsX .c.t //:
dt
Hence c is the maximal integral curve of vsX with initial point e; and we conclude that c.t / D
sX .t /: Now evaluate at t D 1 to obtain the equality.
Formula (b) is an immediate consequence of (a) and Lemma 3.2. Finally, from Lemma 3.2
we have that .t; X/ 7! X .t / is a smooth map R  Te G ! G: Substituting t D 1 we obtain
smoothness of exp : Moreover,
T0 .exp/X D

d
exp.tX/jt D0 D P X .0/ D vX .e/ D X:
dt

Hence T0 .exp/ D ITe X ; and from the inverse function theorem it follows that exp is a local
diffeomorphism at 0; i.e., there exists an open neighborhood U of 0 in Te G such that exp maps
U diffeomorphically onto an open neighborhood of e in G:


17

Definition 3.7 A smooth group homomorphism W .R; C/ ! G is called a one-parameter


subgroup of G:
Lemma 3.8 If X 2 Te G; then t 7! exp tX is a one-parameter subgroup of G: Moreover, all
one-parameter subgroups are obtained in this way. More precisely, let be a one-parameter
subgroup in G; and put X D .0/:
P
Then .t / D exp.tX/ .t 2 R/:
Proof: The first assertion follows from Lemma 3.2. Let W R ! G be a one-parameter
subgroup. Then .0/ D e; and
d
d
d
.t / D
.t C s/jsD0 D
.t /.s/jsD0 D Te .l.t / /.0/
P
D vX ..t //;
dt
ds
ds
hence is an integral curve for vX with initial point e: Hence D X by the uniqueness of
integral curves. Now apply Lemma 3.6.

We now come to a very important application.
Lemma 3.9 Let ' W G ! H be a homomorphism of Lie groups. Then the following diagram
commutes:
'
G
!
H
expG "
" expH
Te G

Te '

Te H

Proof: Let X 2 Te G: Then .t / D '.expG .tX// is a one-parameter subgroup of H: Differentiating at t D 0 we obtain .0/
P
D Te .'/T0 .expG /X D Te .'/X: Now apply the above lemma to
conclude that .t / D expH .t Te .'/X/: The result follows by specializing to t D 1:


4 The Lie algebra of a Lie group


In this section we assume that G is a Lie group. If x 2 G then the translation maps lx W y 7! xy
and rx W y 7! yx are diffeomorphisms from G onto itself. Therefore, so is the conjugation map
Cx D lx rx 1 W y 7! xyx 1 : The latter map fixes the neutral element eI therefore, its tangent
map at e is a linear automorphism of Te G: Thus, Te Cx 2 GL.Te G/:
Definition 4.1 If x 2 G we define Ad.x/ 2 GL.Te G/ by Ad.x/ WD Te Cx : The map Ad W G !
GL.Te G/ is called the adjoint representation of G in Te G:
Example 4.2 We return to the example of GL.V /; with V a finite dimensional real linear
space. Since GL.V / is an open subset of the linear space End.V / we may identify its tangent space at I with End.V /: If x 2 GL.V /; then Cx is the restriction of the linear map
Cx W A 7! xAx 1 ; End.V / ! End.V /: Hence Ad.x/ D Te .Cx / D Cx is conjugation by
x:

18

The above example suggests that Ad.x/ should be looked at as an action of x on Te G by


conjugation. The following result is consistent with this point of view.
Lemma 4.3 Let x 2 G; then for every X 2 Te G we have
x exp X x

D exp.Ad.x/ X/:

Proof: We note that Cx W G ! G is a Lie group homomorphism. Hence we may apply


Lemma 3.9 with H D G and ' D Cx : Since Te Cx D Ad.x/; we see that the following diagram
commutes:
Cx
G
!
G
exp "
" exp
Te G
The result follows.

Ad.x/

Te G


Lemma 4.4 The map Ad W G ! GL.Te G/ is a Lie group homomorphism.


Proof: From the fact that .x; y/ 7! xyx 1 is a smooth map G  G ! G it follows by differentiation with respect to y at y D e that x 7! Ad.x/ is a smooth map from G to End.Te G/: Since
GL.V / is open in End.Te G/ it follows that Ad W G ! GL.Te G/ is smooth.
From Ce D IG it follows that Ad.e/ D ITe G : Moreover, differentiating the relation Cxy D
Cx Cy at e; we find, by application of the chain rule, that Ad.xy/ D Ad.x/Ad.y/ for all x; y 2 G:

Since Ad.e/ D I D ITe G and TI GL.Te G/ D End.Te G/; we see that the tangent map of Ad
at e is a linear map Te G ! End.Te G/:
Definition 4.5 The linear map ad W Te G ! End.Te G/ is defined by
ad WD Te Ad:
We note that, by the chain rule, for all X 2 Te G;

d
ad.X/ D
Ad.exp tX/:
dt t D0
Lemma 4.6 For all X 2 Te G we have:

Ad.exp X/ D e adX :

Proof: In view of Lemma 4.4, we may apply Lemma 3.9 with H D GL.Te G/ and ' D Ad:
Since Te H D TI GL.Te G/ D End.Te G/; whereas expH is given by X 7! e X ; we see that the
following diagram commutes:

exp

G
"

Te G

Ad

ad

GL.Te G/
" e.  /

End.Te G/

The result follows.


19

Example 4.7 Let V be finite dimensional real linear space. Then for x 2 GL.V / the linear
map Ad.x/ W End.V / ! End.V / is given by Ad.x/Y D xY x 1 : Substituting x D e tX and
differentiating the resulting expression with respect to t at t D 0 we obtain:
d tX
e Ye tX t D0 D XY YX:
. adX/Y D
dt
Hence in this case . adX/Y is the commutator bracket of X and Y .
Motivated by the above example we introduce the following notation.
Definition 4.8 For X; Y 2 Te G we define the Lie bracket X; Y 2 Te G by
X; Y WD . adX/Y

Lemma 4.9 The map .X; Y / 7! X; Y is bilinear Te G  Te G ! Te G: Moreover, it is antisymmetric, i.e.,


X; Y D Y; X
.X; Y 2 Te G/:

Proof: The bilinearity is an immediate consequence of the fact that ad W Te G ! End.Te G/ is


linear. Let Z 2 Te G: Then for all s; t 2 R we have
exp.t Z/ D exp.sZ/ exp.t Z/ exp. sZ/ D exp.t Ad.exp.sZ// Z/;

by Lemmas 3.6 and 4.3. Differentiating this relation with respect to t at t D 0 we obtain:
Z D Ad.exp.sZ// Z

Differentiating this with respect to s at s D 0 we obtain:

.s 2 R/:

0 D ad.Z/T0 .exp/Z D ad.Z/Z D Z; Z:

Now substitute Z D X C Y and use the bilinarity to arrive at the desired conclusion.

Lemma 4.10 Let ' W G ! H be a homomorphism of Lie groups. Then


Te '.X; Y G / D Te 'X; Te 'Y H ;

.X; Y 2 Te G/:

(5)

H
Proof: One readily verifies that ' CxG D C'.x/
': Taking the tangent map of both sides of this
equation at e; we obtain that the following diagram commutes:

Te G
AdG .x/ "

Te G

Te '

Te '

Te H
" AdH .'.x//
Te H

Differentiating once more at x D e; in the direction of X 2 Te G; we obtain that the following


diagram commutes:
Te G
adG .X/ "

Te G

Te '

Te H
" adH .Te 'X/

Te '

Te H

We now agree to write X; Y D ad.X/Y: Then by applying Te ' adG X to Y 2 Te G the


commutativity of the above diagram yields (5).

20

Corollary 4.11 For all X; Y; Z 2 Te G;


X; Y ; Z D X; Y; Z

Y; X; Z:

(6)

Proof: Put ' D Ad and H D GL.Te G/: Then eH D I and TI H D End.Te G/: Moreover,
A; BH D AB BA for all A; B 2 End.Te G/: Applying Lemma 4.10 and using that  ;  G D
 ;  and Te ' D ad; we obtain
ad.X; Y / D adX; adY H D adX adY

adY adX:

Applying the latter relation to Z 2 Te G; we obtain (6).

Definition 4.12 A real Lie algebra is a real linear space a equipped with a bilinear map ;  W
a  a ! a; such that for all X; Y; Z 2 a we have:
(a) X; Y D Y; X (anti-symmetry);
(b) X; Y; Z C Y; Z; X C Z; X; Y D 0

(Jacobi identity).

Remark 4.13 Note that condition (a) may be replaced by the equivalent condition (a): X; X D
0 for all X 2 a: In view of the anti-symmetry (a), condition (b) may be replaced by the equivalent condition (6). We leave it to the reader to check that another equivalent form of the Jacobi
identity is given by the Leibniz type rule
X; Y; Z D X; Y ; Z C Y; X; Z:

(7)

Corollary 4.14 Let G be a Lie group. Then Te G equipped with the bilinear map .X; Y / 7!
X; Y WD . adX/Y is a Lie algebra.
Proof: The anti-linearity was established in Lemma 4.9. The Jacobi identity follows from (6)
combined with the anti-linearity.

Definition 4.15 Let a; b be Lie algebras. A Lie algebra homomorphism from a to b is a linear
map ' W a ! b such that
'.X; Y a / D '.X/; '.Y /b ;

for all X; Y 2 a:

From now on we will adopt the convention that Roman capitals denote Lie groups. The
corresponding Gothic lower case letters will denote the associated Lie algebras. If ' W G ! H
is a Lie group homomorphism then the associated tangent map Te ' will be denoted by ' : We
now have the following.
Lemma 4.16 Let ' W G ! H be a homomorphism of Lie groups. Then the associated tangent map ' W g ! h is a homomorphism of Lie algebras. Moreover, the following diagram
commutes:
'
G
!
H
expG "
" expH
g

'

Proof: The first assertion follows from Lemma 4.10, the second from Lemma 3.9.
21

Example 4.17 We consider the Lie group G D Rn : Its Lie algebra g D T0 Rn may be identified
with Rn : From the fact that G is commutative, it follows that Cx D IG ; for all x 2 G: Hence,
Ad.x/ D Ig ; for all x 2 G: It follows that ad.X/ D 0 for all X 2 g: Hence X; Y D 0 for all
X; Y 2 g:
Let X 2 g ' Rn : Then the associated one-parameter subgroup X is given by X .t / D tX:
Hence exp.X/ D X; for all X 2 g:
We consider the Lie group homomorphism ' D .'1 ; : : : ; 'n / W Rn ! Tn given by 'j .x/ D
e 2 ixj : One readily verifies that ' is a local diffeomorphism. Its kernel equals Zn : Hence, by
the isomorphism theorem for groups, the map ' factors through an isomorphism of groups 'N W
Rn =Zn ! Tn : Via this isomorphism we transfer the manifold structure of Tn to a manifold
structure on Rn =Zn : Thus, Rn =Zn becomes a Lie group, and 'N an isomorphism of Lie groups.
Note that the manifold structure on H WD Rn =Zn is the unique manifold structure for which the
canonical projection  W Rn ! Rn =Zn is a local diffeomorphism. The projection  is a Lie
group homomorphism. The associated homomorphism of Lie algebras  W g ! h is bijective,
since  is a local diffeomorphism. Hence,  is an isomorphism of Lie algebras. We adopt
the convention to identify h with g ' Rn via  : It then follows from Lemma 4.16 that the
exponential map expH W Rn ! H D Rn =Zn is given by expH .X/ D .X/ D X C Zn :

5 Commuting elements
In the following we assume that G is a Lie group with Lie algebra g: Two elements X; Y 2 g are
said to commute if X; Y D 0: The Lie algebra g is called commutative if every pair of elements
X; Y 2 g commutes.
Example 5.1 If G D GL.V /; with V a finite dimensional real or complex linear space, then
g D End.V /: In this case the Lie bracket of two elements A; B 2 End.V / equals the commutator
bracket A; B D AB BA: Hence, the assertion that A and B commute means that AB D BA;
as we are used to. In this case we know that the associated exponentials e A and e B commute as
linear maps, hence as elements of GI moreover, e A e B D e ACB : The following lemma generalizes
this fact to arbitrary Lie algebras.
Lemma 5.2 Let X; Y 2 g be commuting elements. Then the elements exp X and exp Y of G
commute. Moreover,
exp.X C Y / D exp X exp Y:
Proof: We will first show that x D exp X and y D exp Y commute. For this we observe that, by
Lemma 4.3, xyx 1 D exp.Ad.x/Y /: Now Ad.x/Y D e adX Y; by Lemma 4.6. Since ad.X/Y D
X; Y D 0; it follows that ad.X/n Y D 0 for all n  1: Hence, Ad.x/Y D e adX Y D Y:
Therefore, xyx 1 D y and we see that x and y commute.
For every s; t 2 R we have that sX; t Y D st X; Y D 0: Hence by the first part of this proof
the elements exp.sX/ and exp.t Y / commute for all s; t 2 R: Define the map W R ! G by
.t / D exp.tX/ exp.t Y /
22

.t 2 R/:

Then .0/ D e: Moreover, for s; t 2 R we have


.s C t / D exp.s C t /X exp.s C t /Y
D exp sX exp tX exp sY exp t Y
D exp sX exp sY exp tX exp t Y D .s/.t /:
It follows that is a one-parameter subgroup of G: Hence D Z with Z D 0 .0/; by Lemma
3.8. Now, by Lemma 5.3 below,
 
 
d
d
0
exp.tX/ exp.0/ C
exp.0/ exp.t Y / D X C Y:
.0/ D
dt t D0
dt t D0
From this it follows that .t / D Z .t / D exp.t Z/ D exp.t .X C Y //; for t 2 R: The desired
equality follows by substituting t D 1:

The following lemma gives a form of the chain rule for differentiation that has been used in
the above, and will often be useful to us.
Lemma 5.3 Let M be a smooth manifold, U a neighborhood of .0; 0/ in R2 and ' W U ! M
a map that is differentiable at .0; 0/: Then
 
 
 
d
d
d
'.t; t / D
'.t; 0/ C
'.0; t /:
dt t D0
dt t D0
dt t D0
Proof: Let D1 '.0; 0/ denote the tangent map of s 7! '.s; 0/ at zero. Similarly, let D2 '.0; 0/
denote the tangent map of s 7! '.0; s/ at zero. Then the tangent T.0;0/ ' W R2 ! T'.0;0/ M of '
at the origin is given by T.0;0/ '.X; Y / D D1 '.0; 0/X C D2 '.0; 0/Y; for .X; Y / 2 R2 :
Let d W R ! R2 be defined by d.t / D .t; t /: Then the tangent map of d at 0 is given by
T0 d W R 7! R2 ; X 7! .X; X/: By application of the chain rule, it follows that
.d=dt /t D0 '.t; t / D
D
D
D

.d=dt /t D0 '.d.t // D T0 .' d / 1


T.0;0/ ' T0 d 1 D T.0;0/ '.1; 1/
D1 '.0; 0/1 C D2 '.0; 0/1
.d=dt /t D0 '.t; 0/ C .d=dt /t D0 '.0; t /:


Definition 5.4 The subgroup Ge generated by the elements exp X; for X 2 g; is called the
component of the identity of G:
Remark 5.5 From this definition it follows that
Ge D fexp.X1 /    exp.Xk / j k  1; X1 ; : : : ; Xk 2 gg:
In general it is not true that Ge D exp.g/: Nevertheless, many properties of g can be lifted to
analogous properties of Ge : As we will see in this section, this is in particular true for the property
of commutativity.
23

By an open subgroup of a Lie group G we mean a subgroup H of G that is an open subset


of G in the sense of topology.
Lemma 5.6 Ge is an open subgroup of G:
Proof: Let a 2 Ge : Then there exists a positive integer k  1 and elements X1 ; : : : ; Xk 2 g such
that a D exp.X1 / : : : exp.Xk /: The map exp W g ! G is a local diffeomorphism at 0 hence there
exists an open neighborhood  of 0 in g such that exp is a diffeomorphism of  onto an open
subset of G: Since la is a diffeomorphism, it follows that la .exp.// is an open neighborhood of
a: We now observe that la .exp.// D fexp.X1 / : : : exp.Xk / exp.X/ j X 2 g  Ge : Hence a
is an interior point of Ge : It follows that Ge is open in G:

Lemma 5.7 Let H be an open subgroup of G: Then H is closed as well.
Proof: For all x; y 2 G we have xH D yH or xH \ yH D ;: Hence there exists a subset
S  G such that G is the disjoint union of the sets sH; s 2 S: The complement of H in G is the
disjoint union of the sets sH with s 2 S; s H: Being the union of open sets, this complement
is open. Hence H is closed.

Lemma 5.8 Ge equals the connected component of G containing e: In particular, G is connected if and only if Ge D G:
Proof: The set Ge is open and closed in G; hence a (disjoint) union of connected components. On the other hand Ge is arcwise connected. For let a 2 Ge ; then we may write
a D exp.X1 / : : : exp.Xk / with k  1 and X1 ; : : : ; Xk 2 g: It follows that c W 0; 1 ! G; t 7!
exp.tX1 / : : : exp.tXk / is a continuous curve with initial point c.0/ D e and end point c.1/ D a:
This establishes that Ge is arcwise connected, hence connected. Therefore Ge is a connected
component; it obviously contains e:

Lemma 5.9 Let G be a Lie group, x 2 G: Then the following assertions are equivalent.
(a) x commutes with Ge I

(b) Ad.x/ D I:
Proof: Assume (a). Then for every Y 2 g and t 2 R we have exp.t Y / 2 Ge ; hence
exp.t Ad.x/Y / D x exp t Y x

D exp t Y

Differentiating this expression at t D 0 we see that Ad.x/Y D Y: This holds for any Y 2 g;
hence (b).
For the converse implication, assume (b). If Y 2 g; then
x exp Y x

D exp Ad.x/Y D exp Y:

Hence x commutes with exp.g/: Since the latter set generates the subgroup Ge ; it follows that x
commutes with Ge :

24

Remark 5.10 Note that the point of the above proof is that one does not need exp W g ! G
to be surjective in order to derive properties of a connected Lie group G from properties of its
Lie algebra. It is often enough that G is generated by exp g: Another instance of this principle is
given by the following theorem.
Theorem 5.11 Let G be a Lie group. The following conditions are equivalent.
(a) The Lie algebra g is commutative.
(b) The group Ge is commutative.
In particular, if G is connected then g is commutative if and only if G is commutative.
Proof: Assume (a). Then X; Y D 0 for all X; Y 2 g: Hence exp X and exp Y commute for
all X; Y 2 g and it follows that Ge is commutative.
Conversely, assume (b). Let x 2 Ge : Then it follows by the previous lemma that Ad.x/ D I:
In particular this holds for x D exp.tX/; with X 2 g and t 2 R: It follows that e ad.tX/ D
Ad.exp.tX// D I: Differentiating at t D 0 we obtain ad.X/ D 0: Hence X; Y D 0 for all
X; Y 2 g and (a) follows.
Finally, if G is connected, then Ge D G and the last assertion follows.


6 Commutative Lie groups


From Example 4.17, we recall that the group Rp =Zp .p 2 N/ has a unique structure of manifold
which turns the natural projection  W Rp ! Rp =Zp into a local diffeomorphism. With this
structure of manifold, the group Rp =Zp is a commutative Lie group. It is isomorphic with the
p-dimensional torus Tp :
In this section we will prove the following classification of commutative Lie groups.
Theorem 6.1 Let G be a commutative connected Lie group. Then there exist p; q 2 N such that
G ' Tp  Rq :
Before we give the proof, we need to collect some results on discrete subgroups of a Lie group. A
subgroup H of a Lie group is called discrete if it is discrete as a topological space for the restriction topology. Equivalently, this means that for every h 2 H there exists an open neighborhood
U in G such that U \ H D fhg:
Proposition 6.2 Let G be a Lie group and H a subgroup. Then the following statements are
equivalent.
(a) There exists an open neighborhood U of e such that U \ H D feg:
(b) The group H is discrete.
(c) For every compact subset C  G; the intersection C \ H is finite.
(d) The group H is a closed Lie subgroup with Lie algebra f0g:
25

Proof: (a) ) (b): Let h 2 H: Then Uh D hU is an open neighborhood of h in G: Moreover,


Uh \ H D hU \ H D h.U \ h 1 H / D h.U \ H / D fhg:
(b) ) (c): We first prove that H is closed in G: Let U be an open neighborhood of e in
G such that U \ H D feg: Let g 2 G be a point in the closure of H: Then it suffices to show
that g 2 H: There exists a sequence fhj g in H converging to g: It follows that hj C1 hj 1 !
gg 1 D e; as j ! 1: Hence for j sufficiently large we have hj C1 hj 1 2 U \ H D feg; hence
hj D hj C1 : It follows that the sequence hj becomes stationary after a certain index; hence
hj D g for j sufficiently large and we conclude that g 2 H:
It follows from the above that the set H \ C is closed in C; hence compact. For h 2 H \ C
we select an open subset of Uh of G such that Uh \ H D fhg: Then fUh j h 2 H \ C g is an
open cover of H \ C which does not contain a proper subcover. By compactness of H \ C this
cover must therefore be finite, and we conclude that H \ C is finite.
(c) ) (d) Let g 2 G be a point in the closure of H: The point g has a compact neighborhood
C: Now g lies in the closure of H \ C I the latter set is finite, hence closed. Hence g 2 H \ C 
H and we conclude that the closure of H is contained in H: Therefore, H is closed.
It follows that H is a closed Lie subgroup. Its Lie algebra h consists of the X 2 g with
exp.RX/  H: Since exp W g ! G is a local diffeomorphism at 0, there exists an open neighborhood  of 0 if g such that exp is injective on : Let X 2 g n f0g: Then there exists an  > 0
such that ; X  : The curve c W ;  ! G; t 7! exp tX has compact image; this
image has a finite intersection with H: Hence ft 2 ;  j exp tX 2 H g is finite, and we see
that X h: It follows that h D f0g:
(d) ) (a) Assume (d). Then H is a closed smooth submanifold of X of dimension 0: By
definition this implies that there exists an open neighborhood U of e in G such that U \H D feg:
Hence (a).

Proof of Theorem 6.1: Assume that G is a connected Lie group that is commutative. Then
its Lie algebra g is commutative, i.e., X; Y D 0 for all X; Y 2 g: From this it follows that
exp X exp Y D exp.X C Y / for all X; Y 2 g: Therefore, the map exp W g ! G is a homomorphism of the Lie group .g; C; 0/ to G: It follows that exp.g/ is already a subgroup of G; hence
equals the subgroup Ge generated by it. Since G is connected, Ge D G; and it follows that
exp has image G; hence is a surjective Lie group homomorphism. Let be the closed subgroup
ker.exp/ of g: By the isomorphism theorem for groups we have G ' g= as groups.
Since exp is a local diffeomorphism at 0; there exists an open neighborhood  of 0 in g on
which exp is injective. In particular this implies that  \ D f0g: By Proposition 6.2 it follows
that is a discrete subgroup of g: In view of Lemma 6.4 below there exists a collection 1 ; : : : ; p
of linear independent elements in g such that D Z 1    Z p : We may extend the above
set to a basis 1 ; : : : ; n of gI here n D dimg D p C q for some q 2 N: Via the basis 1 ; : : : ; n
we obtain a linear isomorphism ' W g ! Rp  Rq : Let  D exp ' 1 ; then  W Rn ! G is
a surjective Lie group homomorphism, and a local diffeomorphism everywhere. Moreover, its
kernel equals './ D Zp f0g: It follows that  factors through a bijective group homomorphism
N W .R=Z/p  Rq ' Rn =.Zp  f0g/ ! G: The canonical map  W Rn ! .R=Z/p  Rq is a
local diffeomorphism. Moreover,  D N  is a local diffeomorphism as well. Hence N is a local
diffeomorphism. Since N is a bijective as well, we conclude that N is a diffeomorphism, hence an
26

isomorphism of Lie groups.

Lemma 6.3 Let ' W G ! H be a homomorphism of Lie groups. If ' is a local diffeomorphism
at e; then ' is a local diffeomorphism at every point of G:
Proof: We prove this by homogeneity. Let a 2 G: Then from '.ax/ D '.a/'.x/ we see that
' l D l'.a/ '; hence ' D l'.a/ ' la 1 : Now la and l'.a/ are diffeomorphisms. Since la 1
maps a to e; whereas ' is a local diffeomorphism at e it follows that l'.a/ ' la 1 is a local
diffeomorphism at aI hence ' is a local diffeomorphism at a:

Lemma 6.4 Let V be a finite dimensional real linear space. Let be a discrete subgroup
of V: Then there exists a collection of linearly independent elements 1 ; : : : ; p of V such that
D Z 1    Z p :
Proof: We prove the lemma by induction on the dimension of V:
First assume that dimV D 1: Via a choice of basis we may identify V with RI then becomes
a discrete subgroup of R: Suppose f0g: Then there exists an element a 2 n f0g: Passing
to a is necessary, we may assume that a > 0: Now \ 0; a is finite (cf. Prop. 6.2), hence
contains a smallest element : We note that \ 0; 1 D ;: Now  Z : On the other hand,
if g 2 ; then g Z would imply that g 2 m; m C 1 for a suitable m 2 Z: This would
imply that g m 2 \ 0; 1 D ;; contradiction. It follows that  Z : Hence D Z :
This completes the proof of the result for dimV D 1:
Now assume that dimV > 1 and that the result has been established for spaces of strictly
smaller dimension. If D f0g we may take p D 0 and we are done. Thus, assume that
2 n f0g: Then the intersection of R with is discrete in R and non-trivial, hence of the
form Z 1 by the first part of the proof. Select a linear subspace W of V such that R 1 W D V:
Let  denote the corresponding projection V ! W: Let C be a compact subset of W: Then


.C / D R 1 C C D .0; 1 1 C C / C Z 1 :

From this it follows that


C \ ./  .

.C / \ / D ..C C 0; 1 1 / \ C Z 1 /
D ..C C 0; 1 1 / \ /I

the latter set is finite by compactness of C C 0; 1 1 : Thus we see that ./ \ C is finite for
every compact subset of W: By Prop. 6.2 this implies that ./ is a discrete subgroup of W: By
the induction hypothesis there exist linearly independent elements N2 ; : : : ; Np of ./ such that
./ D Z N2    Z Np : Fix 2 ; : : : ; p 2 such that . j / D Nj : Then the elements 1 ; : : : ; p
are readily seen to be linear independent; moreover, D Z 1    Z p :


27

7 Lie subgroups
Definition 7.1 A Lie subgroup of a Lie group G is a subgroup H; equipped with the structure
of a Lie group, such that the inclusion map  W H ! G is a Lie group homomorphism.
The above definition allows examples of Lie subgroups that are not submanifolds. This is
already so if we restrict ourselves to one-parameter subgroups.
Lemma 7.2 Let G be a Lie group, and let X 2 g: The image of the one-parameter subgroup
X is a Lie subgroup of G:
Proof: The result is trivial for X D 0: Thus, assume that X 0: The map X W R ! G is a Lie
group homorphism. Its image H is a subgroup of G:
Assume first that X is injective. Then H has a unique structure of smooth manifold for
which the bijection X W R ! H is a diffeomorphism. Clearly, this structure turns H into a Lie
group and the inclusion map i W H ! G is a Lie group homomorphism.
0
Next, assume that X is not injective. As X
.0/ D X 0; the map t 7! X .t / D exp tX is
injective on a suitable open interval I containing 0: It follows that ker X is a discrete subgroup
of R: Hence ker X D Z for some 2 R: This implies that there exists a unique group
homomorphism N W R=Z ! R such that X D N pr: Since pr is a local diffeomorphism, the
map N is smooth, hence a Lie group homomorphism. Therefore, H D im N is compact. By
homogeneity, N is an injective immersion. This implies that N is an embedding of R=Z onto a
smooth submanifold of G: We conclude that H D im N is a smooth submanifold of G; hence a
Lie subgroup.

We will give an example of a one-parameter subgroup of T2 whose image is everywhere
dense. The following lemma is needed as a preliminary.
Lemma 7.3 Let S be an infinite subgroup of T: Then S is everywhere dense.
Proof: If the subgroup S were discrete at 1; it would be finite, by compactness of T: It follows
that there exists a sequence n in S n f1g such that n ! 1: We consider the surjective Lie group
homomorphism p W R ! T given by p.t / D e 2 it : Since p W R ! T is a local diffeomorphism
at 0; there exists a sequence sn in R n f0g such that p.sn / D n and sn ! 0:
Let  2 T: Fix x 2 R with p.x/ D : For each n there exists a unique kn 2 Z such that
x 2 kn sn ; .kn C 1/sn /: Thus, jkn sn xj < jsn j and it follows that kn sn ! x: Therefore, in T we
have nkn D p.kn sn / ! : Since nkn 2 S for every n; we conclude that x belongs to the closure
of S: Hence, S is dense.

Corollary 7.4 Let W R ! T2 be an injective one-parameter subgroup of T2 : Then the image
of is dense in T2 :

28

Proof: Let H denote the image of : For j D 1; 2; let pj W T2 ! T denote the projection onto
the j -th component and consider the one-parameter subgroup j WD pj W R ! T: Its kernel
j is an additive subgroup of R; hence either trivial or infinite. If j0 .0/ D 0 then j D R:
If j0 .0/ 0 then j is immersive at 0; hence everywhere by homogeneity. It follows that the
image of j is an open subgroup of T; hence equal to T; by connectedness of the latter group.
It follows that j is a local diffeomorphism from R onto T: On the other hand, j cannot be a
diffeomorphism, as T is compact and R is not. It follows that j is not trivial, hence infinite.
Thus, in all cases 1 and 2 are infinite additive subgroups of R:
As is injective, we observe that 1 \ 2 D ker D f0g: Hence, maps 2 injectively to
H \ .T  f1g/: Since 2 is infinite, it follows that H \ .T  f1g/ must be infinite. By Lemma
7.3 it follows that H \ .T  f1g/ is dense in T  f1g:
Likewise, H \ .f1g  T/ is dense in f1g  T: Let z D .s; t / 2 T: Then there exists a sequence
xn in H \ .T  f1g/ with limit .s; 1/: Similarly, there exists a sequence yn in H \ .f1g  T /
converging to .1; t /: It follows that xn yn is a sequence in H with limit z: Hence, H is dense. 
We finally come to our example.
Example 7.5 We consider the group G D R2 =Z2 : The canonical projection  W R2 ! R2 =Z2 is
a homomorphism of Lie groups. We recall from Example 4.17 that  is a local diffeomorphism.
Accordingly, we use its tangent map  to identify R2 D T0 R2 with g: Let X 2 R2 I then the
associated one-parameter subgroup D X in G is given by
.t / D tX C Z2 ;

.t 2 R/:

From Lemma 7.2 it follows that the image H of X is a Lie subgroup of G: If X D 0; then X is
constant, and its image is the trivial group. We now assume that X 0: If X1 ; X2 have a rational
ratio, and X1 0; then X2 D pX1 =q; with p; q 2 Z; q > 0: Hence qX1 1 X 2 Z2 ; and it follows
that X is not injective. In the proof of Lemma 7.2 we saw that H is a compact submanifold of
G; diffeomorphic to the circle. A similar assertion holds in case X1 =X2 2 Q:
If X1 ; X2 have an irrational ratio, then tX Z2 for all t 2 R; so that X is injective. From
Corollary 7.4 it follows that H is dense in G in this case.
Lemma 7.6 Let ' W H ! G be an injective homomorphism of Lie groups. Then ' is immersive
everywhere. In particular, the tangent map ' D Te ' W h ! g is injective.
Proof: We will first establish the last assertion. There exists an open neighborhood  of 0 in h
such that expH maps  diffeomorphically onto an open neighborhood of e in H: The following
diagram commutes:
'
H
!
G
expH "
" expG
h

'

Since expH is injective on ; it follows that ' expH is injective on I hence so is expG ' :
It follows that ' is injective on : Hence ker.' / \  D f0g: But ker.' / is a linear subspace
of hI it must be trivial, since its intersection with an open neighborhood of 0 is a point.
29

We have shown that ' is immersive at e: We may complete the proof by homogeneity. Let
h 2 H be arbitrary. Then l'.h/ ' lh 1 D ': Hence, by taking tangent maps at h it follows that
Th ' is injective.

In the following we assume that H is a Lie subgroup of G: The inclusion map is denoted by
 W H ! G: As usual we denote the Lie algebras of these Lie groups by h and g; respectively.
The following result is an immediate consequence of the above lemma.
Corollary 7.6 The tangent map  WD Te  W h ! g is injective.
We recall that  is a homomorphism of Lie algebras. Thus, via the embedding  the Lie algebra
h may be identified with a Lie subalgebra of g; i.e., a linear subspace that is closed under the Lie
bracket. We will make this identification from now on. Note that after this identification the map
 of the above diagram becomes the inclusion map.
Lemma 7.7 As a subalgebra of g; the Lie algebra of H is given by:
h D fX 2 g j 8t 2 R W expG .tX/ 2 H g:
Proof: We denote the set on the right-hand side of the above equation by V:
Let X 2 h: Then expG .tX/ D .expH tX/ by commutativity of the above diagram with
' D : Hence, expG .tX/ 2 .H / D H for all t 2 R: This shows that h  V:
To prove the converse inclusion, let X 2 g; and assume that X h: We consider the map
' W R  h ! G defined by
'.t; Y / D exp.tX/ exp.Y /:
The tangent map of ' at .0; 0/ is the linear map T.0;0/ ' W R  h ! g given by
T.0;0/ ' W .; Y / 7! X C Y:
Since X h; its kernel is trivial. By the immersion theorem there exists a constant  > 0 and
an open neighbourhood  of 0 in h; such that ' maps ;   injectively into G: Shrinking
 if necessary, we may in addition assume that expH maps  diffeomorphically onto an open
neighborhood U of e in H:
The map m W .x; y/ 7! x 1 y; H  H ! H is continuous, and maps .e; e/ to e: Since U
is an open neighborhood of e in H; there exists an open neighborhood U0 of e in H such that
m.U0  U0 /  U; or, written differently,
U0 1 U0  U:
Since H is a union of countably many compact sets, there exists a countable collection fhj j j 2
Ng  H such that the open sets hj U0 cover H: For every j 2 N we define
Tj D ft 2 R j exp tX 2 hj U0 g:
Let now j 2 N be fixed for the moment, and assume that s; t 2 Tj ; js t j < : Then it
follows from the definition of Tj that exp.t s/X D exp. sX/ exp.tX/ 2 U0 1 U0  U:
30

Hence exp.t s/X D exp Y for a unique Y 2 ; and we see that '.t s; 0/ D '.0; Y /: By
injectivity of ' on ;   it follows that Y D 0 and s D t: From the above we conclude
that different elements s; t 2 Tj satify js t j  : Hence Tj is countable.
The union of countably many countable sets is countable. Hence the union of the sets Tj is
properly contained in R and we see that there exists a t 2 R such that t Tj for all j 2 N: This
implies that exp tX [j 2N hj U0 D H: Hence X V: Thus we see that g n h  g n V and it
follows that V  h:

Example 7.8 Let V be a finite dimensional linear space (with k D R or C). In Example
2.14 we saw that SL.V / is a submanifold of GL.V /; hence a Lie subgroup. The Lie algebra of
GL.V / is equal to gl.V / D End.V /; equipped with the commutator brackets. We recall from
Example 2.14 that det W GL.V / ! k is a submersion at I: Hence the tangent space sl.V / of
SL.V / D det 1 .1/ at I is equal to ker.TI det/ D ker tr: We conclude that the Lie algebra of
SL.V / is given by
sl.V / D fX 2 End.V / j trX D 0gI
(8)
in particular, it is a subalgebra of gl.V /: The validity of (8) may also be derived by using the
methods of this section, as follows.
If X 2 sl.V /; then by Lemma 7.7, exp.tX/ 2 SL.V / for all t 2 R; hence

d
d
tX
det.e / D
1 D 0:
trX D
dt t D0
dt t D0

It follows that sl.V / is contained in the set on the right-hand side of (8).
For the converse inclusion, let X 2 End.V /; and assume that trX D 0: Then for every t 2 R
we have dete tX D e tr.tX/ D 1; hence exp tX D e tX 2 SL.V /: Using Lemma 7.7 we conclude
that X 2 sl.V /:

Example 7.9 We consider the subgroup O.n/ of GL.n; R/ consisting of real n  n matrices x
with x t x D I: Being a closed subgroup, O.n/ is a Lie subgroup. We claim that its Lie algebra is
given by
o.n/ D fX 2 M.n; R/ j X t D Xg;
(9)
the space of anti-symmetric n  n matrices. Indeed, let X 2 o.n/: Then by Lemma 7.7, exp sX 2
O.n/; for all s 2 R: Hence,
t
I D .e sX /t e sX D e sX e sX :

Differentiating with respect to s at s D 0 we obtain X t C X D 0; hence X belongs to the set on


the right-ghand side of (9).
For the converse inclusion, assume that X 2 M.n; R/ and X t D X: Then, for every s 2 R;
t

.e sX /t e sX D e sX e sX D e

sX sX

D I:

Hence exp sX 2 O.n/ for all s 2 R; and it follows that X 2 o.n/:


If X 2 o.n/ then its diagonal elements are zero. Hence trX D 0 and we conclude that
X 2 sl.n; R/: Therefore, o.n/  sl.n; R/: It follows that exp.o.n//  SL.n; R/; hence O.n/e 
31

SL.n; R/: We conclude that O.n/e  SO.n/  O.n/: Since SO.n/ is connected, see exercises,
it follows that
O.n/e D SO.n/:

The determinant det W O.n/ ! R has image f 1; 1g and kernel SO.n/; hence induces a group
isomorphism O.n/=O.n/e ' f 1; 1g: It follows that O.n/ consists of two connected components, O.n/e and xO.n/e ; where x is any orthogonal matrix with determinant 1: Of course,
one may take x to be the diagonal matrix with 1 in the bottom diagonal entry, and 1 in the
remaining diagonal entries, i.e., x is the reflection in the hyperplane xn D 0:
Lemma 7.10 Let G be a Lie group and H  G a subgroup. Then H allows at most one
structure of Lie subgroup.


Proof: See exercises.

We now come to a result that is the main motivation for allowing Lie subgroups that are not
closed.
Theorem 7.11 Let G be a Lie group with Lie algebra g: If h  g is a Lie subalgebra, then
the subgroup hexp hi generated by exp h has a unique structure of Lie subgroup. Moreover, the
map h 7! hexp hi is a bijection from the collection of Lie subalgebras of g onto the collection of
connected Lie subgroups of G:


Proof: See next section.

Remark 7.12 In the literature, the group hexp hi is usually called the analytic subgroup of G
with Lie algebra h:

8 Proof of the analytic subgroup theorem


The proof of Theorem 7.11 will be based on the following result. Throughout this section we
assume that G is a Lie group and that h is a subalgebra of its Lie algebra g:
Lemma 8.1 There exists an open neighborhood  of 0 in g such that M D exp.h \ / is a
submanifold of G with tangent space equal to
Tm M D Te .lm /h;

(10)

for every m 2 M: If  is any such neighborhood, then also Tm M D Te .rm /h for all m 2 M:
In the literature one usually proves this result by using the Frobenius integrability theorem
for subbundles of the tangent bundle. We will first recall this proof, and then give an independent
proof based on a calculation of the derivative of the exponential map.

32

Proof: We consider the subbundle S of T G given by Sx D Te .lx /: Then for all X 2 h the
left invariant vector field vX is a section of S: We note that vX ; vY D vX;Y (see one of the
exercises). Hence for all X; Y 2 h the Lie bracket of vX and vY defines a section of S as well.
Let now X1 ; : : : ; Xk and put vj D vXj :
P
P
Let now ;  be any pair of smooth sections of SI then  D kiD1  i vi and  D kjD1 j vj
for uniquely defined smooth functions  i and j on G: Since
X
;  D
 i j vi ; vj C  i vi .j /vj C j vj . i /vi ;
i;j

it follows that ;  is a section of S: By the Frobenius integrability theorem it follows that the
bundle S is integrable. In particular, there exists a k-dimensional submanifold N of G containing
e; such that Tx N D Sx for all x 2 N:
For X 2 h; the vector field vX is everywhere tangent to N; hence restricts to a smooth vector
N
field vX
on N: By smooth parameter dependence of this vector field on X there exists an open
neighborhood U of 0 in h and a positive constant > 0 such that for every X 2 U the integral
N
curve X of vX
with initial point e is defined on I WD ; : This integral curve is also an
integral curve for vX ; hence equals X W t 7! exp tX on I : It follows that exp U  N: We
may now select an open neighborhood  of 0 in g such that exp is a diffeomorphism from 
onto an open subset of G and such that  \ h is contained in U: Then M WD exp. \ h/ is
a k-dimensional hence open submanifold of N: It follows that Tx M D Tx N D Te .lx /h for all
x 2 M:
For the last assertion, we note that rm D rm lm1 lm so that rm D lm Cm 1 and Te .rm / D
Te .lm /Ad.m 1 /; and it suffices to show that Ad.m 1 / leaves h invariant. Write m D exp X with
X 2  \ h: Then
Ad.m 1 / D Ad.exp. X// D e adX
and the result follows, since ad.X/ leaves the closed subspace h invariant.

We shall now give a different proof of Lemma 8.1. The following result plays a crucial role.
Lemma 8.2 Let X 2 g: Then
TX exp D Te .lexp X /
D Te .rexp X /

s adX

ds

e s adX ds:
0

Proof: For X; Y 2 g; we define


F .X; Y / D Te .lexp X /

TX .exp/Y

2g

and note that, by the chain rule,


F .X; Y / D Texp X .lexp.

@
exp. X/ exp.X C t Y /:
X/ /TX .exp/Y D
@t t D0
33

From this it follows by interchanging partial derivatives, that

@
@
@
F .sX; sY / D
exp. sX/ exp.sX C t sY /:
@s
@t
@s
t D0

Now,
@
exp. sX/ exp.sX C t sY / D
@s

@
exp. . s C /X/ exp..s C /.X C t Y //
@ D0

@
D
exp. sX/ exp. X/ exp..X C t Y // exp.sX C t sY /
@ D0

@
exp. X/ exp.X C  t Y /
D Te .lexp. sX/ rexp.sXCst Y / /
@
D0

D Te .lexp.

and we conclude that

@
@
Te .lexp.
F .sX; sY / D
@s
@t t D0

@
Te .lexp.
D
@t t D0

sX/ rexp.sXCst Y / /.t Y /;

sX/ rexp.sXCst Y / /.t Y /

@
Te .lexp.
sX/ rexp.sXCst Y / /.0/ C
@t t D0

D Ad.exp. sX//Y D e

It follows that

s adX

sX/ rexp.sX/ /.t Y

Y:

Z 1
@
F .X; Y / D
F .sX; sY / ds D
e s adX Y ds;
0 @s
0
whence the first identity. The second identity may be obtained in a similar manner. It can also be
derived from the first as follows. We have Te .lexp X / D Te .rexp X/ Ad.exp X/; hence
Z 1
Z 1
s adX
adX
Te .lexp X /
e
ds D Te .rexp X / e
e s adX ds
0
0
Z 1
D Te .rexp X /
e .1 s/ adX ds
0
Z 1
D Te .rexp X /
e s adX ds:
1

Remark 8.3 The integral in the above expression may be expressed as a power series as follows.
Let V be a finite dimensional linear space, and A 2 End.V /: Then using the power series
expansion for e sA ; we obtain
Z 1
1 Z 1
X
1 n n
sA
e ds D
A s ds
n
0
0
nD0
D

1
X

1
An :
.n
C
1/
nD0

34

For obvious reasons, the sum of the latter series is also denoted by .e A

I /=A:

Alternative proof of Lemma 8.1: Let  be an open neighborhood of 0 in g such that exp j is
a diffeomorphism onto an open subset of G: Then M WD exp.h \ / is a smooth submanifold
of G of dimension dimM D dimh: For (10), we put m D exp X; with X 2 h \ : Since h is a
subalgebra, .e adX I /= adX leaves h invariant. Hence


I e adX
h  Te .lm /h:
Tm M D TX .exp/h D Te .lm /
adX
Equality follows for dimensional reasons. The identity with Te rm is proved in a similar manner.

We shall now proceed with the proof of Theorem 7.11, starting with the the result of Lemma
8.1, with  and M as given there.
Lemma 8.4 Let C be a compact subset of M: Then there exists an open neighborhood U of 0
in g such that m exp.h \ U / is open in M for all m 2 C: In particular, C exp.h \ U / is an open
neighborhood of C in M:
Proof: For every X 2 h; we denote by X W R  G ! G the flow of the left invariant
vectorfield vX : We recall that X .t; x/ D x exp tX; for all X 2 h; t 2 R and x 2 G: For fixed
X 2 h; x 2 G; the map t 7! X .x; t / is the maximal integral curve of vX with initial point x:
Since the vector field vX is everywhere tangent to M; it follows that vX jM is a vector field
on M: For every X 2 h; and m 2 M; we denote by t 7! 'X .t; m/ the maximal integral curve of
vX jM in M; with initial point m: By the theory of systems of ordinary differential equations with
parameter dependence, it follows that .X; t; m/ 7! 'X .t; m/ is smooth on its domain, which is
an open subset D  h  R  M containing h  f0g  M: Clearly, t 7! 'X .t; m/ is also an integral
curve of vX in G with initial point m: Therefore, 'X .t; m/ D X .t; m/ for all .X; t; m/ 2 D: We
conclude that X .t; m/ 2 M for all .X; t; m/ 2 D: Using that sX .t; m/ D X .st; m/ and that
C is compact, we now readily deduce that there exists an open neighborhood U0 of 0 in h such
that m exp.tX/ D X .t; m/ 2 M; for all X 2 U0 ; t 2 0; 1 and m 2 C: We may now select an
open neighborhood U of 0 in g such that exp jU is a diffeomorphism. Moreover, replacing U by a
smaller subset if necessary, we may in addition assume that h \ U  U0 : Then, for every m 2 C;
the map X 7! m exp X is an injective immersion of h \ U into M: Since dimM D dimh; the
map is a diffeomorphism onto an open subset of M: The final assertion follows as C exp.h \ U /
is the union of the open sets m exp.h \ U /; for m 2 C:

Corollary 8.5 Let M  G be as in Lemma 8.1. Then for every x1 ; x2 2 G; the intersection
x1 M \ x2 M is open in both x1 M and x2 M:
Proof: Let y 2 x1 M \ x2 M: Then by Lemma 8.4, there exists an open neighborhood U of 0 in
g; such that the sets xj 1 y exp.U \ h/ are open in M; for j D 1; 2: It follows that y exp.U \ h/
is open in both x1 M and x2 M:

35

Proof of Theorem 7.11. Let H be the group generated by exp h: We will first equip H with the
structure of a manifold.
We fix  and M as in Lemma 8.1. Replacing  by a smaller neighborhood if necessary, we
may assume that exp j is a diffeomorphism of  onto an open subset of G: Then exp restricts
to a diffeomorphism of 0 WD  \ h onto the submanifold M of G: Accordingly, its inverse
 W M ! 0 is a diffeomorphism of manifolds as well.
Since M  H; it follows that H is covered by the submanifolds hM of G; where h 2 H:
We equip H with the finest topology that makes the inclusions hM ,! H continuous. Then
by definition a subset U  H is open if and only if U \ hM is open in hM for every h 2 H:
We note that by Corollary 8.5, each set hM , for h 2 H; is open in H: For each open subset
O  G and each h 2 H; the set O \ hM is open in hM I hence O \ H is open in H: It follows
that the inclusion map H ,! G is continuous. Since G is Hausdorff, it now follows that H;
with the defined topology, is Hausdorff. For each h 2 H; the map h D  lh 1 W hM ! 0
is a diffeomorphism. This automatically implies that the transition maps are smooth. Hence
fh j h 2 H g is an atlas.
Fix a compact neighborhood C0 of 0 in  \ h: Then C D exp C0 is a compact neighborhood
of e in M: It follows that C is compact in H: Since h is the union of the sets nC0 for n 2 N; it
follows that exp h is the union of the sets fc n j c 2 C g; for n 2 N: One now readily sees that
H is the union of the sets C n ; for n 2 N: Each of the sets C n is compact, being the image of
the compact Cartesian product C      C (n factors) under the continuous multiplication map.
Hence the manifold H is a countable union of compact subsets, which in turn implies that its
topology has a countable basis.
We will finish the proof by showing that H with the manifold structure just defined is a Lie
group. If h 2 H then the map lh W H ! H is a diffeomorphism by definition of the atlas. We
will first show that right multiplication rh W H ! H is a diffeomorphism as well.
If X 2 h then the linear endomorphism Ad.exp X/ W g ! g equals e adX hence leaves h
invariant. Since H is generated by elements of the form exp X with X 2 h; it follows that
for every h 2 H the linear endomorphism Ad.h/ of g leaves h invariant. Fix h 2 H: Then
there exists an open neighborhood O of 0 in   g such that Ad.h 1 /.O/   hence also
Ad.h 1 /.h \ O/  h \ : From exp Xh D h exp Ad.h 1 /X we now see that
h rh D Ad.h/

e

on

exp.h \ O/:

This implies that rh W exp.h \ O/ ! M is smooth. Hence, rh W H ! H is smooth at e: By


left homogeneity it follows that rh W H ! H is smooth everywhere. Since rh is bijective with
inverse rh 1 ; it follows that rh is a diffeomorphism from H to itself.
We will finish by showing that the multiplication map H W H  H ! H; .h; h0 / 7! hh0
and the inversion map H W H ! H; h 7! h 1 are both smooth. If h1 ; h2 2 H then H .lh1 
rh2 / D lh1 rh2 H : Since lh1 and rh2 are diffeomorphisms, smoothness of H at .h1 ; h2 / follows
from smoothness of H at .e; e/: Thus, it suffices to show smoothness of H at .e; e/: From
H lh D rh 1 H ; we see that it also suffices to prove smoothness of the inversion map H at e:
Fix an open neighborhood Ne of e in M such that NN e is a compact subset of M: Then by
Lemma 8.4, there exists an open neighborhood U of 0 in g such that Ne exp.h \ U /  M:
36

Replacing U by its intersection with ; we see that N0 D exp.h \ U / is an open neighborhood


of e in M and that Ne N0  M: It follows that the smooth map G maps Ne  N0 into M; hence
its restriction G jNe N0 ; which equals H jNe N0 ; maps Ne  N0 smoothly into the smooth
submanifold M of G: This implies that H is smooth in an open neighborhood of .e; e/ in
H  H:
For the inversion, we note that 1 WD  \ . / is an open neighborhood of 0 in g that
is stable under reflection in the origin. It follows that G maps the open neighborhood N1 WD
exp.1 \ h/ of e in M into itself. Hence its restriction to N1 ; which equals H jN1 ; maps N1
smoothly into the smooth manifold M: It follows that H is smooth in an open neighborhood of
e:


9 Closed subgroups
Theorem 9.1 Let H be a subgroup of a Lie group G: Then the following assertions are equivalent:
(a) H is a C 1 -submanifold of G at the point eI
(b) H is a C 1 -submanifold of GI
(c) H is a closed subset of G:
Note that condition (b) implies that H is a Lie subgroup of G: Indeed, the map H W H !
H; h 7! h 1 is the restriction of the smooth map G to the smooth manifold H; hence smooth.
Similarly H is the restriction of G to the smooth submanifold H H of G G; hence smooth.
In the proof of the theorem we will need the following result. If G is a Lie group we shall
use the notation log for the map G ! g; defined on a sufficiently small neighborhood U of e;
that inverts the exponential map, i.e., exp log D I on U:
Lemma 9.2 Let X; Y 2 g: Then
X C Y D lim n log.exp.n 1 X/ exp.n 1 Y //:
n!1

Proof: Being the local inverse to exp; the map log is a local diffeomorphism at e: Its tangent
map at e is given by Te log D .T0 exp/ 1 D Ig :
The map
W g  g ! G; .X; Y / 7! exp X exp Y has tangent map at .0; 0/ given by
T.0;0/ W .X; Y / 7! X C Y: The composition log is well defined on a sufficiently small
neighborhood of .0; 0/ in g  g: Moreover, by the chain rule its derivative at .0; 0/ is given by
.X; Y / 7! X C Y:
It follows that, for .X; Y / 2 g  g sufficiently close to .0; 0/;
log.exp X exp Y / D X C Y C .X; Y /;

37

(11)

where .X; Y / D o.kXk C kY k/ as .X; Y / ! .0; 0/ (here k  k is any choice of norm on g).
Hence
n log.exp.n 1 X/ exp.n 1 Y // D nn 1 X C n 1 Y C .n 1 X; n 1 Y /
D X C Y C n o.n 1 / ! X C Y
.n ! 1/:

Proof of Theorem 9.1. Let n D dimG: We first show that .a/ ) .b/: Let k be the dimension of
H at e: By the assumption there exists an open neighborhood U of e in G and a diffeomorphism
 onto an open subset of Rn ; such that .U \ H / D .U / \ Rk  f0g: Let now h 2 H: Then
Uh WD hU D lh .U / is an open neighborhood of h in G; and h WD  lh 1 is a diffeomorphism
from Uh onto .U /: Moreover, since hU \ H D h.U \ H /; it follows that h .Uh \ H / D
.U \ H / D .U / \ Rk  f0g: It follows that H is a C 1 -submanifold at any of its points.
Next we show that .b/ ) .c/: Assume (b). Then there exists an open neighborhood U of e
in G such that U \ HN D U \ H: Let y 2 HN : Then yU is an open neighborhood of y 2 HN in G;
hence there exists a h 2 yU \ H: Hence y 1 h 2 U: On the other hand, from y 2 HN ; h 2 H it
follows that y 1 h 2 HN : Hence y 1 h 2 U \ HN D U \ H; and we see that y 2 H: We conclude
that HN  H; hence H is closed.
Finally, we show that .c/ ) .a/: We call an element X 2 g tangential to H if there exist
sequences Xn 2 g; n 2 R such that limn!1 Xn D 0; exp Xn 2 H; and limn!1 n Xn D X:
Let T be the set of X 2 g that are tangential to H: From the definition it is obvious that for all
X 2 T we have RX  T:
We claim that for every X 2 T we have exp X 2 H: Indeed, let X 2 T; and let Xn ; n be
sequences as above. If X D 0 then obviously exp X 2 H: If X 0; then jn j ! 1: Choose
mn 2 Z such that n 2 mn ; mn C 1: Then mn ! 1 hence
jn =mn

1j  1=jmn j ! 0

and it follows that mn =n ! 1: Thus, mn Xn D .mn =n /n Xn ! X: Hence


exp X D lim .exp Xn /mn 2 HN  H:
n!1

We also claim that T is a linear subspace of g: Let X; Y 2 T: Put Xn D n1 X and Yn 2 n1 Y:


Then Xn ; Yn ! 0 and Xn ; Yn 2 T; hence exp Xn ; exp Yn 2 H: For n 2 N sufficiently large we
may define Zn D log.exp Xn exp Yn /: Then exp Zn D exp.Xn / exp.Yn / 2 H: Moreover, by (11)
we have Zn ! 0 and nZn ! X C Y: It follows from this that X C Y 2 T:
We will finish the proof by showing that H is a C 1 submanifold at the point e: Fix a linear
subspace S  g such that
g D S T:
Then ' W .X; Y / 7! exp X exp Y is smooth as a map S  T ! G and has tangent map at .0; 0/
given by T.0;0/ ' W .; / 7!  C ; S  T ! g: This tangent map is bijective, and it follows that
' is a local diffeomorphism at .0; 0/: Hence, there exist open neighborhoods S and T of the
38

origins in S and T respectively such that ' is a diffeomorphism from S  T onto an open
neighborhood U of e in G:
We will finish the proof by establishing the claim that for S and T sufficiently small, we
have
'.f0g  T / D U \ H:

Assume the latter claim to be false. Fix decreasing sequences of neighborhoods kS and kT of
the origins in S and T ; respectively, with kS  kT ! f0g: By the latter assertion we mean
that for every open neighborhood O of .0; 0/ in S  T; there exists a k such that kS  kT  S:
By the assumed falseness of the claim, we may select hk 2 '.kS  kT / \ H such that hk
'.f0g  kT /; for all k:
There exist unique Xk 2 kS and Yk 2 kT such that hk D '.Xk ; Yk / D exp Xk exp Yk :
From the above it follows that Xk is a sequence in S n f0g converging to 0: Moreover, from
exp Xk D hk exp. Yk / we see that exp Xk 2 H for all k: Fix a norm k  k on S: Then the
sequence Xk =kXk k is contained in the closed unit ball in S; which is compact. Passing to a
suitable subsequence we may arrange that kXk k 1 Xk converges to an element X 2 S of norm
1: Applying the definition of T with n D kXk k 1 ; we see that also X 2 T: This contradicts the
assumption that S \ T D f0g:

Corollary 9.3 Let G; H be Lie groups, and let ' W G ! H be a continuous homomorphism of
groups. Then ' is a C 1 -map (hence a homomorphism of Lie groups).
Proof: Let D f.x; '.x// j x 2 Gg be the graph of ': Then obviously is a subgroup of the
Lie group G  H: From the continuity of ' it follows that is closed. Indeed, let .g; h/ belong
to the closure of in G  H: Let n D .gn ; hn / be a sequence in converging to .g; h/: Then
gn ! g and hn ! h as n ! 1: Note that hn D '.gn /: By the continuity of ' it follows that
hn D '.gn / ! '.g/: Hence '.g/ D h and we see that .g; h/ 2 : Hence is closed.
It follows that is a C 1 -submanifold of G H: Let p1 W G H ! G and p2 W G H ! H
the natural projection maps. Then p D p1 j is a smooth map from the Lie group onto G: Note
that p is a bijective Lie group homomorphism with inverse p 1 W g 7! .g; '.g//: Thus p 1 is
continuous. By the lemma below p is a diffeomorphism, hence p 1 W G ! is C 1 : It follows
that ' D p2 p 1 is a C 1 -map.

Lemma 9.4 Let G; H be Lie groups, and p W G ! H a bijective Lie homomorphism. If p is a
homeomorphism (i.e., p 1 is continuous), then p is a diffeomorphism (i.e., p 1 is C 1 ).
Proof: Consider the commutative diagram

expG

G
"
g

p

H
"

expH

where p D Te p: Fix open neighborhoods G ; H of the origins in g; h; respectively, such that


expG jG ; expH jH are diffeomorphisms onto open subsets UG of G and UH of H respectively.
39

Replacing G by a smaller neigborhood if necessary we may assume that p.UG /  UH (use


continuity of p). Since p is a homeomorphism, p.UG / is an open subset of UH ; containing e:
Thus 0H WD .expH / 1 .p.UG //\H is an open neighborhood of 0 in h; contained in H : Note
that expH is a diffeomorphism from 0H onto UH0 D p.UG /  UH : From the commutativity of
the diagram and the bijectivity of expG W G ! UG ; expH W 0H ! UH0 and p W UG ! UH0
it follows that p is a bijection of G onto 0H : It follows from this that p is a bijective linear
map. Its inverse p 1 is linear, hence C 1 : Lifting via the exponential maps we see that p 1 maps
UH0 smoothly onto UG I it follows that p 1 is C 1 at e: By homogeneity it follows that p 1 is C 1
everywhere. Indeed, let h 2 H: Then p 1 D lp.h/ p 1 lh 1 ; since p 1 is a homomorphism.
But lh 1 maps hUH0 smoothly onto UH0 I hence p 1 is C 1 on hUH0 :


10 The groups SU(2) and SO(3)


We recall that SU.2/ is the closed subgroup of matrices x 2 SL.2; C/ satisfying x  x D I: Here
x  denotes the Hermitian adjoint of x: By a simple calculation we find that SU.2/ consists of all
matrices



N
;
(12)
N

with .; / 2 C2 ; jj2 C jj2 D 1:


Let j W C2 ! M.2; C/ be the map assigning to any point .; / the matrix given in (12).
Then j is an injective real linear map from C2 ' R4 into M.2; C/ ' C4 ' R8 : In particular,
it follows that j is an embedding. Hence, the restriction of j to the unit sphere S  C2 ' R4
is an embedding of S onto a compact submanifold of M.2; C/ ' R8 : On the other hand, it
follows from the above that j.S/ D SU.2/: Hence, as a manifold, SU.2/ is diffeomorphic to the
3-dimensional sphere. In particular, SU.2/ is a compact and connected Lie group.
By a calculation which is completely analogous to the calculation in Example 7.9 we find
that the Lie algebra su.2/ of SU.2/ is the algebra of X 2 M.2; C/ with
X D

trX D 0:

X;

From this one sees that as a real linear space su.2/ is generated by the elements






0 i
0 1
i
0
:
; r3 D
; r2 D
r1 D
i 0
1 0
0
i
Note that rj D i j ; where 1 ; 2 ; 3 are the famous Pauli spin matrices. One readily verifies
that r21 D r22 D r23 D I and r1 r2 D r2 r1 D r3 ; and r2 r3 D r3 r2 D r1 :
Remark 10.1 One often sees the notation i D r1 ; j D r2 ; k D r3 : Indeed, the real linear span
H D RI Ri Rj Rk is a realization of the quaternion algebra. The latter is the unique (up to
isomorphism) associative R algebra with unit, on the generators i; j; k; subject to the above well
known quaternionic relations.
40

It follows from the above product rules that the commutator brackets are given by
r1 ; r2 D 2r3 ;

r2 ; r3 D 2r1 ;

From this it follows that the endomorphisms adrj


with respect to the basis r1 ; r2 ; r3 W
0
1
0
0 0
0
0
@
A
@
2 ; mat adr2 D
0
mat adr1 D 0 0
0 2
0
2

The above elements belong to

r3 ; r1 D 2r2 :

2 End.su.2// have the following matrices


1
0
0 2
0
A
@
0 0 ; mat adr3 D 2
0 0
0

so.3/ D fX 2 M.3; R/ j X  D

1
2 0
0 0 A:
0 0

Xg;

the Lie algebra of the group SO.3/:


If a 2 R3 ; then the exterior product map X 7! a  X; R3 ! R3 has matrix
0
1
0
a3
a2
0
a1 A
Ra D @ a3
a2
a1
0
with respect to the standard basis e1 ; e2 ; e3 of R3 : Clearly Ra 2 so.3/:

Lemma 10.2 Let t 2 R: Then exp tRa is the rotation with axis a and angle t jaj:
Proof: Let r 2 SO.3/: Then one readily verifies that Ra D r Rr
exp tRa D r exptRr

1a

1a

; and hence

Selecting r such that r 1 a D jaje1 ; we see that we may reduce to the case that a D jaje1 : In that
case one readily computes that:
0
1
1
0
0
sin t jaj A :
exp tRa D @ 0 cos t jaj
0 sin t jaj
cos t jaj


Write Rj D Rej ; for j D 1; 2; 3: Then by the above formulas for mat ad.rj / we have
mat ad.rj / D 2Rj

.j D 1; 2; 3/:

(13)

We now define the map ' W SU.2/ ! GL.3; R/ by '.x/ D matAd.x/; the matrix being taken
with respect to the basis r1 ; r2 ; r3 : Then ' is a homomorphism of Lie groups.

41

Proposition 10.3 The map ' W SU.2/ ! GL.3; R/; x 7! matAd.x/ is a surjective group
homomorphism onto SO.3/; and induces an isomorphism:
SU.2/=fI g ' SO.3/:
Proof: From
'.exp X/ D mate adX D e mat adX

we see that ' maps SU.2/e into SO.3/: Since SU.2/ is obviously connected, we have SU.2/ D
SU.2/e ; so that ' is a Lie group homomorphism from SU.2/ to SO.3/: The tangent map of
' is given by ' W X 7! mat adX: It maps the basis frj g of su.2/ onto the basis f2Rj g of
so.3/; hence is a linear isomorphism. It follows that ' is a local diffeomorphism at I; hence
its image im' contains an open neighborhood of I in SO.3/: By homogeneity, im' is an open
connected subgroup of SO.3/; and we see that im' D SO.3/e : As SO.3/ is connected, it follows
that im' D SO.3/: From this we conclude that ' W SU.2/ ! SO.3/ is a surjective group
homomorphism. Hence SO.3/ ' SU.2/= ker ': The kernel of ' may be computed as follows.
If x 2 ker '; then Ad.x/ D I: Hence xrj D rj x for j D 1; 2; 3: From this one sees that
x 2 f I; I g: Hence ker ' D f I; I g:


It is of particular interest to understand the restriction of ' to one-parameter subgroups of


SU.2/: We first consider the one-parameter group W t 7! exp.t r1 /: Its image T in SU.2/
consists of the matrices
 it

e
0
ut D
;
.t 2 R/:
0 e it
Obviously, T is the circle group. The image of ut under ' is given by
'.ut / D '.e t r1 / D e ' .t r1 / D e 2tR1 :
By a simple calculation, we deduce that, for  2 R;
0
1
1
0
0
cos  sin  A ;
R WD e R1 D @ 0
0
sin  cos 

the rotation with angle  around the x1 -axis. Let D be the group consisting of these rotations.
Then ' maps T onto D: Moreover, from '.ut / D R2t we see that ' restricts to a double covering
from T onto D:
More generally, if X is any element of su.2/; different from 0; there exists a x 2 SU.2/ such
that Ad.x 1 /X D x 1 Xx D r1 ; for some  > 0: It follows that the one-parameter subgroup
X has image exp RX D expAd.x/Rr1 D xT x 1 in SU.2/: The image of xT x 1 under '
equals rDr 1 ; with r D '.x/: Moreover, the following diagram commutes:
T
#

Cx
'

! xT x 1
# '

Cr

! rDr

The horizontal arrows being diffeomorphisms, it follows that 'jxT x


xT x 1 onto rDr 1 :
42

is a double covering from

11 Group actions and orbit spaces


Definition 11.1 Let M be a set and G a group. A (left) action of G on M is a map W G M !
M such that
(a) .g1 ; .g2 ; m// D .g1 g2 ; m/
(b) .e; m/ D m

.m 2 M; g1 ; g2 2 G/I

.m 2 M /:

Instead of the cumbersome notation we usually exploit the notation g  m or gm for .g; m/:
Then the above rules (a) and (b) become: g1  .g2  m/ D .g1 g2 /  m; and e  m D m:
If g 2 G; then we sometimes use the notation g for the map m 7! .g; m/ D gm; M !
M: From (a) and (b) we see that g is a bijection with inverse map equal to g 1 : Let Sym.M /
denote the set of bijections from M onto itself. Then Sym.M /; equipped with the composition
of maps, is a group. According to (a) and (b) the map W g 7! g is a group homomorphism of
G into Sym.M /: Conversely, any group homomorphism G ! Sym.M / comes from a unique
left action of G on M in the above fashion.
Let M1 ; M2 be two sets equipped with (left) G-actions. A map ' W M1 ! M2 is said to
intertwine the G-actions, or to be equivariant, if '.gm/ D g'.m/ for all m 2 M1 and g 2 G:
Remark 11.2 Similarly, a right action of a group G on a set M is defined to be a map W
M  G ! M; .m; g/ 7! mg; such that meG D m and .mg1 /g2 D m.g1 g2 / for all m 2 M and
g1 ; g2 2 G: Notice that these requirements on are equivalent to the requirement that the map
_ W G  M ! M defined by _ .g; m/ D mg 1 is a left action. Thus, all results for left actions
have natural counterparts for right actions.
Our goal is to study smooth actions of a Lie group on a manifold. As a first step we concentrate on continuous actions. This is most naturally done for topological groups.
Definition 11.3 A topological group is a group G equipped with a topology such that the multiplication map  W G  G ! G; .x; y/ 7! xy and the inversion map  W G ! G; x 7! x 1 are
continuous.
Note that a Lie group is in particular a topological group.
Definition 11.4 Let G be a topological group. By a continuous right action of G on a topological space M we mean an action W M  G ! M that is continuous as a map between
topological spaces. A (right) G-space is a topological space equipped with a continuous (right)
G-action.
We assume that H is a topological group and that M is a topological space equipped with
a continuous right action of H: Given h 2 H we denote by h the map M ! M given by
m 7! mh: Then h is continuous and so is its inverse h 1 : Therefore, h is a homeomorphism
of M onto itself.
43

Sets of the form mH .m 2 M / are called orbits for the action : Note that for two orbits
m1 H; m2 H either m1 H D m2 H or m1 H \ m2 H D ;: Thus, the orbits constitue a partition
of M: The set of all orbits, called the orbit space, is denoted by M=H: The canonical projection
M ! M=H; m 7! mH is denoted by :
The orbit space X D M=H is equipped with the quotient topology. This is the finest topology
for which the map  W M ! M=H is continuous. Thus, a subset O of X is open if and only if
its preimage  1 .O/ is open in M:
In general this topology need not be Hausdorff even if M is Hausdorff. We will return to this
issue later.
The following result is useful, but particular for group actions. It is not true for quotient
topologies in general.
Lemma 11.5 The natural map  W M ! M=H is open.
Proof: Let U  M be open and put O D .U /: Then the preimage  1 .O/ equals the union
of the sets U h D h .U /; which are open in M: It follows that  1 .O/ is open, hence O is open
by definition of the quotient topology.

We denote by F .M / the complex linear space of functions M ! C: Let F .M /H denote the
subspace of F .M / consisting of functions g W M ! C that are H -invariant, i.e., g.mh/ D g.m/
for all m 2 M; h 2 H:
If f W M=H ! C is a function, then the pull-back of f by ; defined by   .f / WD f ; is
a function on M that is H -invariant, i.e., it belongs to F .M /H : One readily verifies that   is a
linear isomorphism from F .M=H / onto F .M /H :
Let C.M /H be the space C.M / \ F .M /H of continuous functions M ! C which are
H -invariant.
Lemma 11.6 The pull-back map   W f 7! f  maps C.M=H / bijectively onto C.M /H :
Proof: Obviously   maps C.M=H / injectively into C.M /H : It remains to establish surjectivity. Let f 2 C.M /H : Then f D   .g/ for a unique function g W M=H ! C: We must show
that g is continuous. Let  be an open subset of C: Then U D f 1 ./ is open in M: From the
H -invariance of f it follows that U is right H -invariant. Hence U D  1 ..U // and it follows
that .U / is open in M=H: But .U / D g 1 ./: Thus, g is continuous.

Remark 11.7 With exactly the same proof it follows: if X is an arbitrary topological space, then
  maps C.M=H; X/ (bijectively) onto C.M; X/H : In fact, the quotient topology on M=H is
uniquely characterized by this property for all X:
In what follows we shall mainly be interested in actions on locally compact Hausdorff spaces.
Recall that the topological space M is said to be Hausdorff if for each pair of distinct points
m1 ; m2 of M there exist open neighborhoods Uj of mj such that U1 \ U2 D ;: The space
M is said to be locally compact if each point in M has a compact neighborhood. Note that
in a Hausdorff space M each compact subset is closed. Moreover, if M is locally compact
Hausdorff, then for every point m 2 M and every open neighborhood U of m there exists a
compact neighborhood N of m contained in U:
44

Lemma 11.8 Let M be a locally compact Hausdorff space, equipped with a continuous right
action of a topological group H: Then the following assertions are equivalent.
(a) The orbit space M=H is Hausdorff.
(b) For each compact subset C  M the set CH is closed.
Proof: Assume (a) and let C  M be compact. Then .C / is compact. As M=H is Hausdorff,
it follows that .C / is closed. As CH D  1 ..C //; it follows that CH is closed.
Next, assume (b). From the fact that fmg is compact, for m 2 M; it follows that the orbit
mH is closed. Let x1 ; x2 2 X D M=H be distinct points. Select mj 2  1 .xj /: Then
xj D mj H; with m1 H \ m2 H D ;: The complement V of m2 H in M is open, right H invariant, and contains m1 H: Select an open neighborhood U1 of m1 in M such that UN 1 is
compact and contained in V: Then by (a), the set UN 1 H is closed and still contained in V: Its
complement V2 is open and containes m2 H: Hence, .V2 / is open in X and contains x2 :
On the other hand, V1 D U1 H is the union of the open sets U1 h; hence open in M: Moreover,
V1 contains m1 ; so that .V1 / is an open neighborhood of x1 in X: Clearly, the sets V1 and V2 are
right H -invariant and disjoint. It follows that the sets .V1 / and .V2 / are disjoint open subsets
of X containing the points x1 and x2 ; respectively. This establishes the Hausdorff property. 

12 Smooth actions and principal fiber bundles


Definition 12.1 Let M be a smooth manifold and H a Lie group. An action of H on M is said
to be smooth if the action map W M  H ! M; .m; h/ 7! mh is a C 1 map of manifolds.
In the rest of this section we will always assume that M is a smooth manifold on which H
has a smooth right action. We will first study smooth actions for which the quotient M=H allows
a natural structure of smooth manifold.
If  is a smooth manifold, then H has a right action on the manifold   H; given by
.x; g/  h D .x; gh/: We will say that such an action is of trivial principal fiber bundle (or trivial
PFB) type.1
More generally, the right action of H on a manifold M is called of trivial PFB type if there
exist a smooth manifold  and a diffeomorphism  W M !   H that intertwines the H actions. Such a map  is called a trivialization of the action. Note that dim D dimM dimH:
Definition 12.2 The right action of H on M is called of principal fiber bundle (PFB) type if the
following two conditions are fulfilled.
(a) Every point m of M possesses an open H -invariant neighborhood U such that the right
H -action on U is of trivial PFB type.
(b) If C is a compact subset of M; then CH is closed.
1

The terminology principal fiber bundle type is not standard, but used here for purposes of exposition.

45

In view of Lemma 11.8, the second condition is equivalent to the condition that the quotient
space M=H is Hausdorff.
We call the pair .U; / of condition (a) a local trivialization of the right H -space M at the
point m: Clearly, if the right H -space M is of PFB type, then there exists a collection f.U ;  / j
2 Ag of local trivializations such that the open sets U cover M: Such a covering is called a
trivializing covering.
Remark 12.3 If H is a closed subgroup of a Lie group G; then the map .h; g/ 7! gh; H G !
G defines a smooth right action of H on G: At a later stage we will see that this action is of PFBtype.
If the right H -action on M is of PFB type, then the quotient M=H admits a unique natural
structure of smooth manifolds. In order to understand the uniqueness, the following preliminary
result about submersions will prove to be very useful.
Lemma 12.4 Let X; Y; Z be smooth manifolds, and let  W X ! Y; ' W X ! Z and
Z be maps such that the following diagram commutes

If ' is smooth and  a submersion, then

X
#
Y

WY !

'

! Z
%

is smooth on .X/:

Proof: The map ; being a submersion, is open. In particular, .X/ is an open subset of Y: Let
y0 2 .X/: Fix x0 2 X such that .x0 / D y0 : Since  is a submersion, there exists an open
neighborhood U of x0 and a diffeomorphism ' W U ! .U /  F; with F a smooth manifold of
dimension dimX dimY; such that  D pr2 ': Here pr2 denotes the projection .U /  F ! F
onto the second component. Let b D pr2 '.x0 /: Then the smooth map  W .U / ! X defined
by .y/ D ' 1 .y; b/ satisfies   D I on .U / and .y0 / D x0 : In other words,  admits a
smooth locally defined section  with .y0 / D x0 : From this it follows that D   D
'  on .U /: Hence, is smooth on .U /:

Theorem 12.5 Let the right H action on M be of PFB type. Then M=H carries a unique
structure of C 1 -manifold (compatible with the topology) such that the canonical projection
 W M ! M=H is a smooth submersion.
If m 2 M; then the tangent map Tm  W Tm M ! T.m/ .M=H / has kernel Tm .mH /;
the tangent space of the orbit mH at m: Accordingly, it induces a linear isomorphism from
Tm M=Tm .mH / onto T.m/ .M=H /:
Finally,   W f 7! f  restricts to a bijective linear map from C 1 .M=H / onto C 1 .M /H :
Remark 12.6 It follows from the assertion on the tangent maps that the dimension of M=H
equals dimM dimH:
46

Proof: We will first show that the manifold structure, if it exists, is unique. Let Xj denote
M=H; equipped with a manifold structure labeled by j 2 f1; 2g and assume that the projection
map  W M ! M=H is submersive for both manifold structures. We will show the identity map
I W X1 ! X2 is smooth for the given manifold structures. The following diagram commutes
M


X1

&

X2

Since  W M ! X1 is a submersion and  W M ! X2 smooth, it follows by Lemma 12.4


that I W X1 ! X2 is smooth. By symmetry of the argument it follows that the inverse to I is
smooth as well. Hence, X1 and X2 are diffeomorphic manifolds. This establishes uniqueness of
the manifold structure.
We defer the treatment of existence until the end of the proof, and will first derive the other
assertions as consequences.
We first address the assertion about the tangent map of : Let m 2 M: Since  is a submersion, Tm  W Tm M ! T.m/ .M=H / is a surjective linear map, with kernel equal to the tangent
space of the fiber  1 ..m//: This fiber equals mH: Hence ker Tm  D Tm .mH /:
Finally, it is obvious that   restricts to a linear injection from C 1 .M=H / into C 1 .M /H :
Let g 2 C 1 .M /H : Then g D f  for a unique function f W M ! C: Since g is smooth
and  a smooth submersion, it follows by application of Lemma 12.4 that f is smooth. This
establishes the surjectivity, and hence the bijectivity of   :
We end the proof by establishing the existence of a manifold structure on X D M=H for
which  becomes a smooth submersion. First of all, X is a topological space, which is Hausdorff
because of Lemma 11.8.
Let f.U ;  /g2A be a trivializing covering of M as above. Thus,  is a diffeomorphism of
U onto   H which intertwines the right H -actions. Writing the manifolds  as unions of
charts, we see that we may replace the trivializing covering by one for which each  equals an
open subset of Rn : We write i for the injection x 7! .x; e/;  !   H and p for the
projection   H !  onto the first coordinate.
We will use the trivializing covering to define a smooth atlas of X: The map  W U !
  H is a diffeomorphism intertwining the H -actions, hence induces a homeomorphism  W
.U / ! .  H /=H: The projection map p induces a homeomorphism of the latter space
onto  ; by which we shall identify. Put V D .U /: Then the following diagram commutes:


U
#
V

!   H
# p

(14)

The sets V ; for 2 A; constitute an open covering of X; and the maps  W V !  are
homeomorphisms. We will show that the pairs .V ;  /; for 2 A; constitute a smooth atlas.
Put  D  .V \ V /: Then the transition map
 WD   1
47

is a homeomorphism from  onto  : We must show it is smooth.


The transition map  D   1 is a diffeomorphism from  H onto  H: Moreover,
the diagram
  H
p #


!   H
# p

commutes. As the vertical arrow represent smooth submersions, it follows by application of


Lemma 12.4 that  is smooth.
Let X be equipped with the structure of C 1 -manifold determined by the atlas defined above.
The map  maps U onto V : Moreover, from the commutativity of the diagram (14) we see that
jU corresponds via the horizontal diffeomorphisms  and  with the smooth projection p :
Hence  is smooth and submersive on each U I it follows that  is a smooth submersion.

The following terminology is standard in the literature, and explains the terminology PFB
type used so far. We assume that X is a smooth manifold.
Definition 12.7 A principal fiber bundle over X with structure group H is a pair .P; / consisting of a smooth right H -manifold P and a smooth map  W P ! X with the following property.
For every point x 2 X there exists an open neighborhood V of x in X and a diffeomorphism
 W  1 .V / ! V  H such that
(a)  D prV  on 

.V /; where prV denotes the projection V  H ! V I

(b)  intertwines the right H -actions.


The manifold P is called the total space, X is called the base space of the bundle. A map  as
above is called a local trivialization of the bundle.
The terminology action of PFB type is finally justified by the following result.
Lemma 12.8 Let H be a Lie group.
(a) If  W P ! X is a principal bundle with structure group H; then the right action of H
'
on P is of PFB-type. Moreover,  factors through a diffeomorphism P =H !X:
(b) Conversely, if M is a smooth manifold equipped with a smooth right-action of H that is
of PFB type, then  W M ! M=H is a principal fiber bundle with structure group H:
Proof: Assertion (a) is a straightforward consequence of Definition 12.7. Assertion (b) is easily
seen from the proof of Theorem 12.5.

Example 12.9 (Frame bundle of a vector bundle) Let V be a finite dimensional real vector space
of dimension k: Let Hom.Rk ; V / denote the linear space of linear maps Rk ! V: A frame in
V is defined to be an injective linear map f W Rk ! V: The set of frames, denoted F .V /, is
a dense open subset of Hom.Rk ; V /: Let e1 ; : : : ; ek be the standard basis of Rk : Then the map
48

f 7! .f .e1 /; : : : ; f .ek // is a bijection from F .V / onto the set of ordered bases of V: Thus, a
frame may be specified by giving an ordered basis of V:
The group H WD GL.k; R/ ' GL.Rk / acts on F .V / from the right; indeed the action is
given by .f; a/ 7! f a: This action is free and transitive; see the text preceding Theorem 13.5
and Proposition 15.5 for the definitions of these notions. Thus, for each f 2 F .V / the map
a 7! f a is a diffeomorphism from H onto F .V /:
Let now p W E ! M be a vector bundle of rank k over a smooth manifold M: For an open
subset U  M we write EU WD p 1 .U /: Then p W EU ! U is a vector bundle over U; called
the restriction of E to U: A trivialization of E over an open subset U  M is defined to be an
isomorphism  W EU ! U  Rk of vector bundles. For x 2 U we define the linear isomorphism
x W Ex ! Rk by  D .x; x / on Ex : Let RkM denote the trivial vector bundle M  Rk over M:
Then the vector bundle Hom.RkM ; E/ has fiber Hom.Rk ; Ex / at the point x 2 M: A trivialization
.U; / of E induces a trivialization  0 of Hom.Rk ; E/ given by x0 .Tx / D x Tx for Tx 2
Hom.Rk ; Ex /:
We define F .E/ to be the subset [x2M F .Ex / of Hom.RkM ; E/: This subset is readily seen
to be open; the natural map F .E/ ! M mapping F .E/x to x defines a sub fiber bundle of
Hom.RkM ; E/: A trivialization  of E over U induces a trivialization  00 of F .E/ over U given
by x00 .f / D x f for f 2 F .Ex /: The group H acts from the right on each fiber F .Ex /: By
looking at trivializations we see that these actions together constitute a smooth right action of H
on F .E/ which turns F .E/ into a principal fiber bundle with structure group H:

13 Proper free actions


In this section we discuss a useful criterion for smooth actions to be of PFB type.
We recall that a continuous map f W X ! Y between locally compact Hausdorff (topological) spaces X and Y is said to be proper if for every compact subset C  Y the preimage
f 1 .C / is compact.
For the moment we assume that M is a locally compact Hausdorff space equipped with a
continuous right action of a locally compact Hausdorff topological group H:
Definition 13.1 The action of H on M is called proper if .m; h/ 7! .m; mh/ is a proper map
M  H ! M  M:
Remark 13.2 Note that a continuous action of a compact (in particular of a finite) group is
always proper.
Lemma 13.3 The following conditions are equivalent.
(a) The action is proper.
(b) For every pair of compact subsets C1 ; C2  M the set HC1 ;C2 WD fh 2 H j C1 h\C2 ;g
is compact.

49

Proof: Let ' W M  H ! M  M; .m; h/ 7! .m; mh/: Assume (a) and let C1 ; C2  M be
compact sets. Then C1  C2 is compact, hence ' 1 .C1  C2 / is a compact subset of M  H:
Now
' 1 .C1  C2 / D f.m; h/ j m 2 C1 ; mh 2 C2 g;

hence HC1 ;C2 D p2 .' 1 .C1  C2 //; with p2 denoting the projection M  H ! H: It follows
that HC1 ;C2 is compact. Hence (b).
Now assume that (b) holds, and let C be a compact subset of M M: Then there exist compact
subsets C1 ; C2  M such that C  C1  C2 : Now ' 1 .C / is a closed subset of ' 1 .C1  C2 /;
hence it suffices to show that the latter set is compact. The latter set is clearly closed; moreover,
it is contained in C1  HC1 ;C2 ; hence compact.

Remark 13.4 We leave it to the reader to verify that condition (b) is equivalent to the condition
that fh 2 H j C h \ C ;g be compact, for any compact set C  M:
The action of H on M is called free if for all m 2 M; h 2 H we have mh D m ) h D e:
From now on we assume that H is a Lie group.
Theorem 13.5 Let M be a smooth manifold equipped with a smooth right H -action. Then the
following statements are equivalent.
(a) the action of H on M is proper and free;
(b) the action of H on M is of PFB type.
As a preparation for the proof we need the following lemma.
Lemma 13.6 Let M be a smooth right H -manifold. If C  H is compact, and m 2 M a point
such that m mC; then there exists an open neighborhood U of m in M such that U h \ U D ;
for all h 2 C:
Proof: Since mC is compact, there exist disjoint open neighborhoods 1 ; 2 of m and mC in
M: By continuity of the action and compactness of C there exists an open neighborhood U of m
in 1 such that UC  2 : It follows that UC \ U D ;:

The following lemma is the key to the proof of Theorem 13.5.

Lemma 13.7 (Slice Lemma). Let M be a smooth manifold equipped with a smooth right H
action which is proper and free. Then for each m 2 M there exists a smooth submanifold S of
M containing m such that the map .s; h/ 7! sh maps S  H diffeomorphically onto an open
H -invariant neighborhood of m in M:
Remark 13.8 The manifold S is called a slice for the H -action at the point m:

50

Proof: Fix m 2 M and define the map m W H ! M by h 7! mh: By freeness of the action,
this map is injective. We claim that its tangent map at e is an injective linear map h ! Tm M:
Given X 2 h we define the smooth vector field vX on M by

d
m exp tX:
vX .m/ D
dt t D0
By application of the chain rule we see that

vX .m/ D Te .m /.X/:
One readily sees that the integral curve of vX with initial point m is given by c W t 7! m exp tX:
From vX .m/ D 0 it follows that c is constant; by freeness of the action this implies that exp tX D
e for all t 2 R; hence X D 0: Thus vX .m/ D 0 ) X D 0 and it follows that the linear map
Te .m / has trivial kernel, hence is injective.
We now select a linear space s  Tm M such that s Te .m /.h/ D Tm M: Moreover, we
select a submanifold S 0 of M of dimension dimM dimH which has tangent space at m equal
to s: Consider the map ' W S 0  H ! M; .s; h/ 7! sh: Then T.m;e/ ' W s  h ! Tm M is
given by .X; Y / 7! X C Te .m /Y; hence bijective. Replacing S 0 by a neighborhood (in S 0 ) of
its point m we may as well assume that S 0 has compact closure and that there exists an open
neighborhood O of e in H such that ' maps S 0  O diffeomorphically onto an open subset of
M: In particular it follows that the tangent map T.s;e/ ' is injective for every s 2 S 0 : Using the
homogeneity h ' .I  rh 1 / D ' for all h 2 H we see that ' has bijective tangent map at
every point of S 0  H:
Let C D HSN 0 ;SN 0 : Then C is a compact subset of H: Hence C0 D C n O is a compact subset
of H; not containing e: Note that m mC0 by freeness of the action. Hence there exists an open
subset S of S 0 containing m such that S \ Sh D ; for all h 2 C0 (use Lemma 13.6).
We claim that the map ' is injective on S  H: Indeed, assume '.s1 ; h1 / D '.s2 ; h2 /;
for s1 ; s2 2 S; h1 ; h2 2 H: Then s2 D s1 .h1 h2 1 /; hence h1 h2 1 belongs to the compact set
C D HSN 0 ;SN 0 : From the definition of S it follows that h1 h2 1 2 C n C0  O: From the injectivity
of ' on S 0  O it now follows that s1 D s2 and h1 h2 1 D e: Hence ' is injective on S  H:
Since we established already that ' has a bijective tangent map at every point of S H it now
follows that ' is a diffeomorphism from S  H onto an open subset U of M: As '.m; e/ D m;
it follows that m 2 U: Moreover, ' intertwines the H -action on S  H with the H -action on U:
Therefore, U is H -invariant.

Proof of Theorem 13.5. (a) ) (b): Assume (a). We shall first prove that the first condition of
Definition 12.2 holds. Let m 2 M and let S be a slice through m as in the above lemma. Then
the map ' W S  H ! M given in the lemma is an H -equivariant diffeomorphism onto an H invariant open neighborhood U of m in M: It follows that the inverse map  D ' 1 W U ! S H
is a trivialization of the H -action on U:
We now turn to the second condition of Definition 12.2. Let C  M be compact and let x
be a point in the closure of CH: Fix a compact neighborhood C 0 of x in M: Then there exists a
sequence .xn /n1 in C 0 \ CH such that xn ! x as n ! 1: Write xn D cn hn ; with cn 2 C
51

and hn 2 H: Then hn is contained in HC;C 0 I the latter set is compact by condition (a). By
passing to subsequences if necessary, we arrive in the situation that the sequences .cn / and .hn /
are convergent, say with limits c 2 C and h 2 H; respectively. Now x D lim cn hn D ch 2 CH:
It follows CH contains its closure, hence is closed. This establishes the second condition of
Definition 12.2. Thus, (b) follows.
(b) ) (a): Assume (b) holds. To see that the action of H on M is free, let x 2 M; h 2 H
and assume that xh D x: There exists an H -invariant open neighborhood U of x on which the
H -action is of trivial PFB-type. Let  W U !   H be a trivialization of the action. Then from
.xh/ D .x/ and .xh/ D .x/h it follows that .x/h D .x/: Hence h D e: This establishes
freeness of the action.
To see that the action of H on M is proper, let C; C 0  M be compact subsets. Then
it suffices to show that HC;C 0 D fh 2 H j C h \ C 0 ;g is compact. For every x 2 C
there exists an H -invariant open neighborhood Ux of x on which the action is of trivial type.
Moreover, there exists a compact neighborhood Cx of x contained in Ux : The interiors of the
sets Cx form an open cover of C; hence contain a finite subcover, parametrized by finitely many
elements x1 ; : : : ; xn 2 M: Put Ci D Cxi ; then C  [niD1 Ci where Ci is contained in Uxi : One
easily verifies that H[i Ci ;C 0 D [i HCi ;C 0 : Therefore it suffices to prove that HC;C 0 is compact
under the assumption that C is contained in an H -invariant open set U on which the action is of
trivial type. Now CH is closed, hence C 00 D CH \C 0 is compact and contained in U: Moreover,
HC;C 0 D HC;C 00 : Thus, we may as well assume that C 0  U: Using a trivializing diffeomorphism
we see that we may as well assume that M is of the form  H: Let D and D 0 be the projections
of C and C 0 onto H; respectively. Then D and D 0 are compact. Moreover, HC;C 0 is a closed
subset of fh 2 H j Dh \ D 0 ;g D D 1 D 0 : The latter set is the image of the compact set
D  D 0 under the continuous map H  H ! H; .h1 ; h2 / 7! h1 1 h2 ; hence compact. It follows

that HC;C 0 is compact as well. This establishes (a).
Example 13.9 We return to the setting of Example 12.9, with p W E ! M a rank k-vector
bundle. The frame bundle  W F .E/ ! M is a principal fiber bundle with structure group H D
GL.k; R/: Thus, the action of H on F .E/ is proper and free, with quotient space F .E/=H ' M:
We observe that the bundle E can be retrieved from F .E/ as follows. The map ' W F .E/ 
Rk ! E defined by .f; v/ 7! f .v/ on F .E/x  Rk ; for x 2 M; is a surjective smooth map.
Using trivializations of E one sees that ' is a submersion. Two elements .f1 ; v1 / and .f2 ; v2 /
have the same image if and only if they belong to the same fiber F .E/x  Rk and there exists
a h 2 H such that .f2 ; v2 / D .f1 h; h 1 v1 /: Define the right action of H on F .E/  Rk by
.f; v/a D .f h; h 1 v/: Then it follows that the fibers of ' are precisely the orbits for the right
action of H on F .E/  Rk :
Via the projection q W F .E/  Rk ! F .E/ we view F .E/  Rk as a trivial vector bundle
over F .E/: The map q intertwines the given right actions of H: As the action of H on F .E/ is
proper and free, so is the action of H on F .E/  Rk (argument left to the reader). It follows
that the induced map qN W .F .E/  Rk /=H ! F .E/=H D M is smooth (show this). Using
trivializations of E, hence of F .E/; one readily checks that the projection qN defines a smooth
rank k vector bundle over M:
52

The map ' W F .E/Rk ! E defined above induces a smooth map 'N W .F .E/Rk /=H ! E
(give the argument). Again using trivializations of E one checks that 'N is an isomorphism
of vector bundles. Thus, the vector bundle qN W .F .E/  Rk /=GL.k; R/ ! M is naturally
isomorphic to E:

14 Coset spaces
We now consider a type of proper and free action that naturally occurs in many situations. Let
G be a Lie group and H a closed subgroup. The map .g; h/ 7! gh defines a smooth right action
of H on G: The associated orbit space is the coset space G=H; consisting of the right cosets
gH; g 2 G:
Lemma 14.1 Let H be a closed subgroup of the Lie group G: Then the right action of H on G
is proper and free.
Proof: It is clear that the action is free. To prove it is proper, let C1 ; C2 be compact subsets of
G: Then HC1 ;C2 D C1 1 C2 \ H: Now C1 1 C2 is the image of C1  C2 under the continuous map
.x; y/ 7! x 1 y; hence compact. Moreover, H is closed, hence HC1 ;C2 is compact.

Corollary 14.2 Let G be a Lie group and H a closed subgroup. Then the coset space G=H has
a unique structure of smooth manifold such that the canonical projection  W G ! G=H is a
smooth submersion. Relative to this manifold structure, the following hold.
(a) The map  W G ! G=H is a principal fiber bundle with structure group H:
(b) The left action of G on G=H given by .g; xH / 7! gxH is smooth.

Proof: From Lemma 14.1 and Theorem 13.5 it follows that the right action of H on G is of
PFB type. Hence, the first assertion is an immediate consequence of Theorem 12.5. Moreover,
assertion (a) follows from Lemma 12.8 (b). Finally, put X D G=H and let denote the action
map G  X ! X: Then the following diagram commutes:


G G
# I 
GX

G
#
X:

Since the vertical map on the left side of the diagram is a submersion, whereas  and  are
smooth, it follows that is smooth (see Lemma 12.4).

Corollary 14.3 Let G be a Lie group and H a closed subgroup. The tangent map Te  of
 W G ! G=H is surjective and has kernel equal to h:
Proof: This is an immediate consequence of the fact that  is a submersion with fiber 
H:

.eH / D


Remark 14.4 It follows from the above that the tangent map Te  induces a linear isomorphism
from g=h onto TeH .G=H /I we agree to identify the two spaces via this isomorphism from now
on. With this identification, Te  becomes identified with the canonical projection g ! g=h:
53

15 Orbits of smooth actions


In this section we assume that G is a Lie group and that M is a smooth manifold equipped with
a smooth left G-action :
Given X 2 g; we denote by X the smooth vector field on M defined by

d
.exp tX/ m
X .m/ WD
dt t D0

We leave it to the reader to verify that for every m 2 M; the curve t 7! .exp tX/ m is the maximal
integral curve of X with initial point m:
Lemma 15.1 The map X 7! X is a Lie algebra anti-homomorphism from g into the Lie
algebra V.M / of smooth vector fields on M:

Proof: Fix m 2 M; and let m W G ! M; g 7! gm: Then X .m/ D Te .m /X: It follows that
X 7! X .m/ is a linear map g ! Tm M: This shows that X 7! X is a linear map g ! V.M /:
It remains to be shown that X ; Y D Y;X ; for all X; Y 2 g: Since .t; m/ 7! .exp tX/ m D
exp tX .m/ is the flow of X ; the Lie bracket of the vector fields X and Y is given by

d
X ; Y .m/ D

Y .m/
dt t D0 exp tX

d
D
T.exp tX/ m .exp tX /Y ..exp tX/ m/
dt t D0

d
d
.exp tX/.exp sY /.exp tX/ m
D
dt t D0 ds sD0

d
d
D
.exp se t adX Y /m
dt t D0 ds sD0

d
D
e t adX Y .m/:
dt t D0
By linearity of Z 7! Z .m/ it follows from this that X ; Y .m/ D Z .m/; where Z D
.d=dt /e t adX Y jt D0 D X; Y :


Remark 15.2 Right multiplication x 7! rx defines a right action of G on itself. The associated
map g ! V.G/ is given by the map X 7! vX of Lemma 3.1 and defines a linear isomorphism of
g onto the space VL .G/ of left invariant vector fields on g: It follows from the above that VL .G/
is a Lie subalgebra of V.G/ and that X 7! vX is an isomorphism of Lie algebras from g onto
VL .G/:
If x 2 M; then the stabilizer Gx of x in G is defined by
Gx D fg 2 G j gx D xg:
Being the pre-image of x under the continuous map g 7! gx; the stabilizer is a closed subgroup
of G:
54

Lemma 15.3 Let x 2 M: The Lie algebra gx of Gx is given by


gx D fX 2 g j X .x/ D 0g:

(15)

Proof: Let gx denote the Lie algebra of Gx : Then for all t 2 R we have exp tX 2 Gx ; hence
.exp tX/x D x: Differentiating this expression with respect to t at t D 0 we see that X .x/ D 0:
It follows that gx is contained in the set on the right-hand side of (15).
To establish the converse inclusion, assume that X .x/ D 0: Then c.t / D .exp tX/ x is
the maximal integral curve of the vector field X with initial point x: On the other hand, since
X .x/ D 0; the constant curve d.t / D x is also an integral curve. It follows that exp tX x D
c.t / D d.t / D x; hence exp tX 2 Gx for all t 2 R: In view of Lemma 7.7 it now follows that
X 2 gx :

As Gx is a closed subgroup of G; it follows from Corollary 14.2 that the coset space G=Gx
has the structure of a smooth manifold. Moreover, let  W G ! G=Gx denote the canonical
projection. Then  is a submersion, and the tangent space of G=Gx at eN WD .e/ is given by
TeN .G=Gx / ' g= ker Te  D g=gx :
The map x W g 7! gx factors through a bijection N x of G=Gx onto the orbit Gx:
Lemma 15.4 The map N x W G=Gx ! M is a smooth immersion.
Proof: It follows from Corollary 14.2 that the natural projection  W G ! G=Gx is a smooth
submersion. Since x D N x ; it follows by application of Lemma 12.4 that N x is smooth.
From x D N x  it follows by taking tangent maps at e and application of the chain rule
that
Te x D TeN .N x / Te :
(16)
Now Te  is identified with the canonical projection g ! g=gx : Moreover, if X 2 g; then
Te .x /.X/ D d=dt x .exp tX/jt D0 D X .x/: Hence ker Te .x / D gx D ker Te : Combining
this with (16) we conclude that TeN N x is injective g=gx ! Tx M: Hence, N x is immersive at e:
N
We finish the proof by applying homogeneity. For g 2 G; let lg denote the left action of g on
G=Gx ; and let g denote the left action of g on M: Then the maps lg and g are diffeomorphisms
of G=Gx and M respectively, and
g N x lg

D N x :

By taking the tangent map of both sides at .g/ and applying the chain rule we may now conclude that N x is immersive at .g/:

The action of G on M is called transitive if it has only one orbit, namely the full manifold
M: In this case the G-manifold M is said to be a homogeneous space for G: The following result
asserts that all homogeneous spaces for G are of the form G=H with H a closed subgroup of G:
Proposition 15.5 Let the smooth action of G on M be transitive, and let x 2 M: Then the map
x W G ! M; g 7! gx induces a diffeomorphism G=Gx ' M:
55

Proof: The map N x W G=Gx ! M is a smooth immersion and a bijection. By Corollary 16.6
(see intermezzo on the Baire theorem) it must be a submersion at some point of G=Gx : By
homogeneity it must be a submersion everywhere. Hence N x is a local diffeomorphism. Since
N x is a bijection, we conclude that it is a diffeomorphism.

Example 15.6 Let n  0: The special orthogonal group SO.n C 1/ acts smoothly and naturally
on RnC1 : Let e1 be the first standard basis vector in RnC1 : Then the orbit SO.n C 1/e1 equals
the n-dimensional unit sphere S D S n in RnC1 : Since S is a smooth submanifold of RnC1 ; it
follows that the action of SO.n C 1/ on S is smooth and transitive. The stabilizer SO.n C 1/e1
equals the subgroup consisting of .n C 1/  .n C 1/ matrices of the form


1 0
; with
B 2 SO.n/:
0 B
It follows that S n is diffeomorphic to SO.n C 1/=SO.n C 1/e1 ' SO.n C 1/=SO.n/:
Example 15.7 Let n  0: We recall that n-dimensional real projective space P WD Pn .R/ is
defined to be the space of 1-dimensional linear subspaces of RnC1 : It has a structure of smooth
manifold, characterized by the requirement that the natural map  W RnC1 n f0g ! P; v 7! Rv
is a smooth submersion.
We consider the natural smooth action of G WD GL.n C 1; R/ on RnC1 n f0g given by
.g; x/ 7! gx: Then G maps fibers of  onto fibers, hence the given action induces an action
G  P ! P: Since  is a submersion, it follows by application of Lemma 12.4 that the action of
G on P is smooth. Let m 2 P be the line spanned by the first standard basis vector e1 of RnC1 :
Then Gm equals the group of invertible .n C 1/  .n C 1/ matrices with first column a multiple
of e1 : One readily sees that the action is transitive. Therefore, the induced map G=Gm ! P is a
diffeomorphism of manifolds.
We now consider the subgroup K D O.n C 1/ of G: One readily sees that K already acts
transitively on P: Hence the action induces a diffeomorphism from K=Km onto P: Here we note
that Km D K \ Gm consists of the matrices


a 0
;
0 B
with a D 1 and B 2 O.n/: Thus, Km ' O.1/  O.n/; and we see that
Pn .R/ ' O.n C 1/=.O.1/  O.n//:

16 Intermezzo: the Baire category theorem


Let X be a topological space. A subset A  X whose closure equals X is said to be dense.
Equivalently this means that A \ U ; for every non-empty open subset U of X:
56

If O1 ; O2 are two open dense subsets of X; then O1 \O2 is still open dense. Indeed, if U  X
is open non-empty, then U \ O1 is open non-empty by density of O1 : Hence, U \ O1 \ O2 is
open non-empty by density of O2 :
It follows that the intersection of finitely many open dense subsets of X is still open and
dense.
Definition 16.1 The topological space X is called a Baire space if every countable intersection
of open dense subsets is dense.
Remark 16.2 Let X be a topological space. A subset S  X is said to be nowhere dense in X
if its closure SN has empty interior. We leave it to the reader to verify that X is Baire if and only
if every countable union of nowhere dense subsets of X has empty interior.
We recall that a topological space X is said to be locally compact if every point p of X is
contained in a compact neighborhood C: If X is assumed to be locally compact and Hausdorff,
then it is known that for every point p 2 X and every neighborhood N of p there exists a
compact neighborhood C of p contained in N:
Theorem 16.3 (Baire category theorem) Let X be a Hausdorff topological space. Then X is a
Baire space as soon as one of the following two conditions if fulfilled.
(a) X is locally compact.
(b) There exists a complete metric on X that induces the topology of X:
Proof: Let fOk j k 2 Ng be a countable collection of open dense subsets of X: Let x0 2 X be
any point and let U0 be an open neighborhood of x0 : In case (a) we assume that U 0 is compact,
in case (b) assume that U 0 is contained in a ball of radius 1: It now suffices to show that U0 has
a non-empty intersection with \n2N On :
We will show inductively that we may select a sequence of non-empty open subsets Uk of X;
for k 2 N; with the property that U kC1  Ok \ Uk for all k 2 N: In case (b) we will show that
this can be done with the additional assumption that Uk is contained in a ball of radius 1=.k C 1/:
Suppose that U0 ; : : : ; Un have been selected. Since On is open dense, On \ Un ;: Select
a point xnC1 of the latter set, then in either of the cases (a) and (b) we may select an open
neighborhood UnC1 of xnC1 whose closure is contained in On \ Un : In case (b) we may select
UnC1 with the additional property that it is contained in the ball of radius 1=.n C 2/ around xnC1 :
The sequence .U n / is a descending sequence of non-empty closed subsets of the subset U0 :
At the end of the proof we will show that its intersection is non-empty. Since obviously \n Un
is contained in U0 \ \n2N On ; it then follows that the latter intersection is non-empty.
Thus, it remains to show that the intersection of the sets U n is non-empty. In case (b) this
follows from the lemma below. In case (a), the sequence .U n / is a decreasing sequence of closed
subsets of the compact set U 0 : Since each finite intersection contains a set Um ; it is non-empty.
Hence, by compactness, the intersection is non-empty.


57

Lemma 16.4 Let .X; d / be a complete metric space and let Ck be a decreasing sequence of
non-empty closed subsets of X whose diameters d.Ck / tend to zero. Then \k2N Ck consists of
precisely one point.
Proof: The condition about the diameter means that we may select a ball of radius rk containing
the set Ck ; for k 2 N; such that rk ! 0 as k ! 1: For each k we may select xk 2 Ck : Then
d.xm ; xn / < 2rk for all m; n  k; hence .xn / is a Cauchy sequence. By completeness of the
metric, the sequence .xn / has a limit x:
Fix k 2 N: Let > 0; then there exists n  k such that xn 2 B.xI /: Hence B.xI / \
Ck ;: It follows that x belongs to the closure of Ck ; hence to Ck ; for every k 2 N: Hence
x 2 \k2N Ck : If y is a second point in the intersection, then for every k; both x; y belong to Ck ;
hence d.x; y/ < 2rk : It follows that d.x; y/ D 0; hence x D y:

A useful application of the above is the following result.
Proposition 16.5 Let X be a manifold of dimension n: Let fYk j k 2 Ng be a countable
collection of submanifolds of X of dimension strictly smaller than n: Then the union [k2N Yk has
empty interior.
Proof: Since X is locally compact, it is a Baire space. Fix k 2 N: If y 2 Yk ; then by the
definition of submanifold, there exists an open neighborhood Uy of y in Yk such that Uy is
nowhere dense in X: By the second countability assumption for manifolds, it follows that Yk
can be covered with countably many neighborhoods Uk;j that are nowhere dense in X: Thus, the
union [k2N Yk is the countable union of the sets Uk;j : Since all of them are nowhere dense, their
union has empty interior.

Corollary 16.6 Let X and Y be smooth manifolds with dimX < dimY: Let ' W X ! Y be a
smooth immersion. Then '.X/ has empty interior.
Proof: Put d D dimX: For every x 2 X there exists an open neighborhood Ux of x in X such
that '.Ux / is a smooth submanifold of Y of dimension d: By the second countability assumption
there exists a countable covering of X by open subsets Uk of X such that '.Uk / is a smooth
submanifold of Y of dimension d: It follows that '.X/ D [k2N '.Uk / has empty interior.


17 Normal subgroups and ideals


If G is a Lie group and H a closed subgroup, then the coset space G=H is a smooth manifold in
a natural way. If H is a normal subgroup, i.e., gHg 1 D H for all g 2 G; then G=H is a group
as well. The following result asserts that these structures are compatible and turn G=H into a
Lie group.
Proposition 17.1 Let G be a Lie group and H a closed normal subgroup. Then G=H has a
unique structure of Lie group such that the canonical map  W G ! G=H is a homomorphism
of Lie groups.
58

Proof: We equip G=H with the unique manifold structure for which  is a submersion. Since
H is normal, G=H has a unique group structure such that  is a group homomorphism. Let N
denote the multiplication map of the quotient group G=H: Then the following diagram commutes


GG
# 


N

G=H  G=H

G
#

! G=H

Since  and  are smooth, so is  : Since the left vertical map is a submersion, it follows from
Lemma 12.4 that N is smooth. In a similar fashion it follows that the inversion map of G=H is
smooth. Hence G=H is a Lie group, and  is a Lie group homomorphism.
Suppose that G=H is equipped with a second structure of Lie group such that  W G ! G=H
is a Lie group homomorphism. We shall denote G=H; equipped with this structure of Lie group,
by .G=H /0 : The identity map I W G=H ! .G=H /0 clearly is an injective homomorphism of
groups. Since  is a submersion, it follows by application of Lemma 12.4 that I is smooth,
hence a Lie group homomorphism. Since I is injective, it follows by Lemma 7.6 that I is
immersive everywhere. Hence, by Lemma 17.2 below we see that I is a submersion. Thus, I
is a bijective local diffeomorphism, hence a diffeomorphism. Therefore, I is an isomorphism of
Lie groups, establishing the uniqueness.

Lemma 17.2 Let ' W G ! G 0 be an immersive homomorphism of Lie groups. Then ' is a
submersion if and only if '.G/ is an open subgroup of G 0 :
Remark 17.3 In Proposition 17.7 we will see that the assumption that ' be immersive is superfluous.
Proof: If ' is a submersion, then '.G/ is open in G 0 : Conversely, assume that '.G/ is open in
G 0 : Then it follows by Corollary 16.6 that dimG D dimG 0 : Hence Te ' W Te G ! Te G 0 is an
injective linear map between spaces of equal dimension. Therefore, it is surjective as well. By
homogeneity it follows that ' is a submersion everywhere.

Theorem 17.4 (The isomorphism theorem for Lie groups). Let ' W G ! G 0 be a homomorphism of Lie groups. Then H WD ker ' is a closed normal subgroup of G: Moreover, the induced
homomorphism 'N W G=H ! G 0 is a smooth injective immersion. If ' is surjective, then 'N is an
isomorphism of Lie groups.
Proof: The following diagram is a commutative diagram of group homomorphisms
G
#
G=H

'

! G0
%'N

Moreover, since  is a submersion, whereas ' is smooth, it follows that 'N is smooth. Hence 'N is
an injective homomorphism of Lie groups. It follows by Lemma 7.6 that ' is an immersion. Now
assume that ' is surjective. Then 'N is surjective, and it follows by application of Lemma 17.2
that 'N is a submersion. We conclude that 'N is a local diffeomorphism, hence a diffeomorphism,
hence an isomorphism of Lie groups.

59

Example 17.5 The isomorphism of Proposition 10.3 is an isomorphism of Lie groups.


Example 17.6 Let G be a Lie group. Then Ad is a Lie group homomorphism from G into
GL.g/: It induces an injective Lie group homomorphism G= ker Ad ! GL.g/; realizing the
image Ad.G/ as a Lie subgroup of GL.g/: If G is connected, then ker Ad is the center Z.G/ of
G; see exercises. Consequently, Ad.G/ ' G=Z.G/ in this case.
Proposition 17.7 Let ' W G ! G 0 be a homomorphism of Lie groups. Then '.G/ is an open
subgroup of G 0 if and only if ' is submersive.
Proof: If ' is submersive, then '.G/ is open. Thus, it remains to prove the only if statement.
Let H be the kernel of ': Then by Theorem 17.4 it follows that the induced map 'N W G=H ! G 0
is an injective homomorphism of Lie groups. By application of Lemma 17.2 it follows that 'N is
a submersion. Since 'N D ' ; whereas  is surjective, we now deduce that ' is a submersion
everywhere.

We end this section with a discussion of the Lie algebra of a quotient of a Lie group by a
closed normal subgroup.
Definition 17.8 Let l be a Lie algebra. An ideal of l is by definition a linear subspace a of l such
that l; a  a; i.e. X; Y 2 a for all X 2 l; Y 2 a:
Remark 17.9 Note that an ideal is always a Lie subalgebra.
Lemma 17.10 (a) Let l be a Lie algebra, a  l an ideal. Then the quotient (linear) space
l=a has a unique structure of Lie algebra such that the canonical projection  W l ! l=a is a
homomorphism of Lie algebras.
(b) Let ' W l ! l0 be a homomorphism of Lie algebras, with kernel a: Then a is an ideal in l
and ' factors through an injective homomorphism of Lie algebras 'N W l=a ! l0 :


Proof: Left as an exercise for the reader.


Lemma 17.11 Let G be a Lie group and let h be a subalgebra of its Lie algebra g:
(a) h is an ideal if and only if h is invariant under Ad.Ge /:
Let H be a Lie subgroup of G with Lie algebra h:
(b) If H is normal in G; then h is an ideal in g:
(c) If h is an ideal in g; then He is normal in Ge :
Proof: Left to the reader. Use Lemmas 4.3 and 4.6.

60

Lemma 17.12 Let H be a closed subgroup of the Lie group G: If H is normal, then its Lie
algebra h is an ideal in g: Moreover, the tangent map at e of the canonical projection  W G !
G=H induces an isomorphism from the quotient Lie algebra g=h onto the Lie algebra of the Lie
group G=H:
Proof: Let l D T.e/ .G=H / be equipped with the Lie algebra structure induced by the Lie
group structure of G=H: The tangent map  of the canonical projection  W G ! G=H is a
Lie algebra homomorphism from g onto l: On the other hand, its kernel is h: Hence, by Lemma
17.10,  factors through a Lie algebra isomorphism from g=h onto l:

Remark 17.13 Accordingly, if G is a Lie group and H a closed normal subgroup, then we shall
identify the Lie algebra of G=H with g=h via the isomorphism described above. In this fashion,
 becomes the canonical projection g ! g=h:
Corollary 17.14 Let ' W G ! G 0 be a homomorphism of Lie groups, with kernel H:
(a) The induced map 'N W G=H ! G 0 is a homomorphism of Lie groups.
(b) Put ' D Te ': Then ker ' equals the Lie algebra h of H:

(c) The tangent map 'N D TeN .'/


N is the linear map g=h ! g0 induced by ' :

(d) If ' is surjective, then 'N and 'N are isomorphisms.


Proof: Assertion (a) follows by application of Theorem 17.4. Let  W G ! G=H be the
canonical projection. Then  is a homomorphism and a smooth submersion. By the preceding
remark, its tangent map  is identified with the natural projection g ! g=h: From 'N  D '
it follows by differentiation at e and application of the chain rule that 'N  D ' : Since 'N is
a smooth immersion by Theorem 17.4, it follows that ker ' D ker  D h: Hence, (b). From
'N  D ' we also deduce (c). If ' is surjective, then 'N is an isomorphism of Lie groups by
Theorem 17.4. Hence, 'N is an isomorphism of Lie algebras.


18 Detour: actions of discrete groups


Let H be a group (without additional structure) acting on a topological space M from the right
by continuous transformations. Equivalently, this means that the action map M  H ! M is
continuous relative to the discrete topology on H (the topology for which all subsets of H are
open).
The action of H on M is said to be properly discontinuous if for each m 2 M there exists an
open neighborhood U of m such that U h \ U D ; for all h 2 H n feg: This condition amounts
to saying that the action of H is locally of trivial PFB type relative to the discrete topology on
H: A third equivalent way of phrasing the condition is that the action of H is free and that the
canonical projection  W M ! M=H is a covering map.
Now assume that M is locally compact and Hausdorff. Then in the setting of an action by
diffeomorphisms on a smooth manifold M the following result may be viewed as a consequence
of Theorem 13.5. We will give a direct proof to cover the topological setting.
61

Lemma 18.1 Let M be a locally compact Hausdorff space equipped with a right H -action by
continuous transformations. Then the following assertions are equivalent.
(a) The action of H on M is continuous, proper and free for the discrete topology on H:
(b) The action of H on M is properly discontinuous and the associated quotient space M=H
is Hausdorff.
Proof: Assume (a), and fix m 2 M: Then there exists a compact neighborhood N of m: The set
HN;N of h 2 H with N h \ N ; is compact in H; hence finite. Put C D HN;N n feg: For
every h 2 C we may select an open neighborhood Uh 3 m such that Uh h \ Uh D ; (observe
that mh m by freeness and use continuity of the action). Let U be the intersection of the finite
collection of open sets Uh .h 2 C / with the interior of N: Then U is open and U h \ U D ; for
all h 2 H n feg: It follows that the action of H is properly discontinuous. By the same argument
as in the proof of Theorem 13.5 it follows that CH is closed for every compact subset C  M:
By Lemma 11.8 we conclude that G=H is Hausdorff.
Next assume (b), and let H be equipped with the discrete topology. We will first show that
the action map W M  H ! M is continuous. Let U  M be open. Then for each h 2 H the
set U h 1 is open in M: Hence U h 1  fhg is open in M  H: The preimage 1 .U / equals the
union of these sets for h 2 H; hence is open.
Now suppose that C1 ; C2 are compact subsets of M: We will show that the set
HC1 ;C2 D fh 2 H j C1 h \ C2 ;g
is finite hence compact; from this (a) will follow.
For every m 2 M there exists an open neighborhood Um of m such that Um \ Um h D ;
for h 2 H; h e: It follows that Um h1 \ Um h2 D ; for distinct h1 ; h2 2 H: Let Nm be a
compact neighborhood of m contained in Um : By compactness there exists a finite collection F
of points from C1 such that C1  [m2F Um : It follows that HC1 ;C2 is contained in the finite union
[m2F HNm ;C2 : Therefore, it suffices to show that the set Hm WD HNm ;C2 is finite.
It follows from Lemma 11.8 that Nm H is closed. The complement U0 WD M n Nm H is open
in M: The sets U0 and Um h; h 2 Hm ; form an open cover of C2 : Hence, there exists a finite
subset S  Hm such that C2 is contained in the union of U0 and [h2S Nm h: Let h 2 Hm : Then
Nm h \ C2 is non-empty; let c be one of its points. As c U0 there exists a h0 2 S such that
c 2 Um h0 : From c 2 Nm h \ Um h0  Um h \ Um h0 it follows that the intersection Um h \ Um h0 is
non-empty. Hence h D h0 and we see that h 2 S: We conclude that Hm is contained in S hence
is finite.


19 Densities and integration


If V is an n-dimensional real linear space, then a density on V is a map ! W V n ! C transforming
according to the rule:
T  ! WD ! T n D jdetT j !
62

.T 2 End.V //:

In these notes the (complex linear) space of densities on V is denoted by DV: A density ! 2 DV
is called positive if it is non-zero and has values in 0; 1: The set of such densities is denoted
by DC V: It is obviously non-empty.
Example 19.1 If ! is an element of ^n V  ; the space of alternating multilinear maps V n ! R;
then j!j is a positive density on V:
If ' is a linear isomorphism from V onto a real linear space W , then the map '  W ! 7! ! ' n
is a linear isomorphism DW ! DV of the associated spaces of densities. Indeed, if ! 2 DW;
T 2 End.V /; then
'  ! T n D ! ' T n D .! ' T ' 1 n / ' n
D jdet.' T ' 1 /j '  ! D jdetT j '  !:
Note that '  maps DC W onto DC V:
The space DV is one dimensional; in fact, if v1 ; : : : ; vn is a basis of V then the map ! 7!
!.v1 ; : : : ; vn/ is a linear isomorphism from DV onto C; mapping DC V onto 0; 1:
If X is a smooth manifold, then by Tx X we denote the tangent space of X at a point x: By
a well known procedure we may define the bundle DTX of densities on XI it is a complex line
bundle with fiber .DTX/x ' D.Tx X/: The space of continuous sections of DTX is denoted
by .DTX/I this space is called the space of continuous densities on X: The space of smooth
densities on X is denoted by 1 .DTX/: We also have the fiber bundle DC TX of positive
densities on X: Its fiber above x equals DC Tx X: Its continuous sections are called the positive
continuous densities on X:
If ' is a diffeomorphism of X onto a manifold Y; then we define the (pull-back) map '  W
.DT Y / ! .DTX/ by
.'  !/.x/ D D'.x/ !.'.x//:
Note that '  maps positive densities to positive densities.
Let e1 ; : : : ; en be the standard basis of Rn : The density  2 DRn given by .e1 ; : : : ; en/ D 1
is called the standard density on Rn : Let U  Rn be an open subset. Then by triviality of the
tangent bundle T U ' U  Rn ; the map f 7! f  defines a linear isomorphism from C 1 .U /
onto 1 .DT U /: If f belongs to the space Cc .U / of compactly supported continuous functions
U ! C; we define the integral
Z
Z
f  WD
f .x/ dx;
Rn

where dx denotes normalized Lebesgue measure. If ' is a diffeomorphism from U onto a second
open subset V  Rn ; then, for g 2 Cc .V /; we have '  .g/.x/ D g.'.x// jdetD'.x/j .'.x//:
Thus, by the substitution of variables theorem:
Z
Z

' !D
!
.! 2 c .DT V //:
(17)
U

63

Let now .; / be a coordinate chart of X: If ! is a continuous density on X with compact


support supp !  ; then we define
Z
Z
! WD
. 1 / !:
X

./

This definition is unambiguous, because if .0 ; 0 / is a second chart such that supp !  0 ; then
Z
Z
Z
0 1
0
1 
0 1
. / ! D
.  / . / ! D
. 1 / !
0 .0 /

./

./

by the substitution of variables theorem.


We can now define the integral of a compactly supported density on the manifold X as follows. Let f j 2 Ag be an open cover of the manifold X with coordinate neighborhoods.
Then there exists a partition of unity f j 2 Ag subordinate to this cover. We recall that the
Cc1 .X/ with 0   1: Moreover, the collection of supports fsupp g
are functions in P
is locally finite and 2A D 1 on X (note that the sum is finite at every point of X; by the
local finiteness of the collection of supports). Let ! 2 c .DTX/ be a continuous density on X
with compact support. Then we define
Z
XZ
!D
!:
X

2A

Just as in the theory of integration of differential forms one shows that this definition is independent of the particular choice of partition of unity. Note that integration of forms is oriented,
whereas the present integration
of densities is non-oriented.
R
We note that ! 7! X ! is a linear map c DTX ! C: Moreover, the following lemma is an
easy consequence of the definitions (reduction to charts et cetera).
Lemma 19.2 Let
R ! be a positive density
R on X: Then for every f 2 Cc .X/ with f  0 everywhere we have X f !  0: Moreover, X f ! D 0 ) f D 0:

Also, by a straightforward reduction to charts we can prove the following substitution of


variables theorem.

Proposition 19.3 Let ' W X ! Y be a diffeomorphism of C 1 -manifolds. Then for every


! 2 c .DT Y / we have:
Z
Z

!:
' !D
Y

We now turn to the situation that G is a Lie group acting smoothly from the left on a smooth
manifold M: If g 2 G; we write lg for the diffeomorphism M ! M; m 7! gm:
Definition 19.4 A density ! 2 .DTM / is said to be G-invariant if lg ! D ! for all g 2 G:
The space of G-invariant continuous densities on M is denoted by .DTM /G :
64

The following result will be very important for applications.


Lemma 19.5 Let ! be a G-invariant continuous density on M: Then for every f 2 Cc .M / and
all g 2 G we have:
Z
Z
M

lg .f / ! D

f !:

(18)

Here lg f WD f lg :

Proof: We note that by invariance of ! we have lg .f /! D lg .f /lg .!/ D lg .f !/: Now observe
that lg is a diffeomorphism of M and apply the substitution of variables theorem (Proposition
19.3) with ' D lg :

Lemma 19.6 Let G be a Lie group and let .DT G/G denote the space of left invariant continuous densities on G:
'

(a) The evaluation map  W ! 7! !.e/ defines a linear isomorphism .DT G/G ! Dg:

(b) A density ! 2 .DT G/G is positive if and only if !.e/ is positive.

Proof: The map  is linear. If ! is a left invariant continuous density on G; then !.g/ D
..lg 1 / !/.g/ D Te .lg / 1 !.e/ for all g 2 G: Hence,  has trivial kernel. On the other hand, if
!0 is a density in Dg then the formula
!.g/ WD Te .lg /

1

!0

(19)

defines a continuous density on G whose value at e is !0 : By application of the chain rule for
tangent maps it follows that this density is left G-invariant. Thus, (a) follows. Assertion (b)
follows from (19).

The following result is an immediate consequence of the above lemma.
Corollary 19.7 Every Lie group G has a left (resp. right) invariant positive density. Two such
densities differ by a positive factor.
R
If ! is a density on G; then the map Cc .G/ ! R; f 7! I.f / D G f ! is continuous linear,
hence aRRadon measure on G: For this reason we shall often write dx for an invariant density on
G; and G f .x/ dx for the associated invariant integral of a function f 2 Cc .G/: Note that in the
example of G D Rn with addition, dx is a (complex) multiple of Lebesgue measure. Positivity
then means that the multiple is positive, and invariance corresponds with translation invariance
of the Lebesge measure.
We now recollect some of the above results in the present notation. Let dx be a left invariant
positive density on G: (Analogous statements will be valid for right invariant positive densities.)
R
Proposition 19.8 The map f 7! I.f / D G f .x/ dx is a complex linear functional on Cc .G/:
It satisfies the following, for every f 2 Cc .G/:
65

(a) If f is real then so is I.f /I if f  0 then I.f /  0:


(b) If f  0 and I.f / D 0 then f D 0:
(c) For every y 2 G W

f .yx/ dx D

f .x/ dx:

(20)

Proof: Assertion (a) follows from the positivity of !: Assertion (b) is immediate from Lemma
19.2. Finally (c) is a reformulation of Lemma 19.5.

Remark 19.9 One can show that up to a positive factor the linear functional I is uniquely
determined by the requirement I 0 and the properties (a) and (c). In particular property (b) is
a consequence. For details we refer the reader to the book by Brocker and tom Dieck.
It follows from the proposition that the Radon measure associated with a left invariant density
is left invariant, non-trivial and positive.
In the literature a left G-invariant positive Radon measure on G is called a left Haar measure of G: The above statement about the uniqueness of I is referred to as uniqueness of the
Haar measure. More generally a left (resp. right) Haar measure exists (and is unique up to a
positive factor) for any locally compact topological group. Of course one cannot use the present
differential geometric method of proof to establish the existence and uniqueness result in that
generality.
Lemma 19.10 Let G be a compact Lie group. Then there exists a unique left invariant density
dx on G with
Z
dx D 1:
G

This density is positive.

Proof: Fix a positive Rdensity  on G: Then it follows from assertions (a) and (b) of Proposition
19.8 for f D 1; that G  equals a positive constant c > 0: The densitity dx D c 1  satifies
the above. This proves existence. If ! is a density with the same property, then ! D Cdx for a
constant C 2 C: Integration over G shows that C D 1: This establishes uniqueness.

Remark 19.11 The density of the above lemma is called the normalized left invariant density
of G: The associated Haar measure is called normalized Haar measure.
The following result expresses how left invariant densities behave under right translation.
Lemma 19.12 Let dx be a left invariant density on a Lie group G: Then for every g 2 G;
rg .dx/ D jdetAd.g/j

66

dx:

Proof: Without loss of generality we may assume that dx is non-zero. For g; h 2 G we have that
lh rg D rg lh ; hence rg lh D lh rg and we see that lh .rg .dx// D rg .lh dx/ D rg .dx/: It follows
that rg .dx/ is a left invariant positive density. This implies that rg .dx/ D c dx for a non-zero
constant c: Applying lg 1 to both sides of this equation we find Cg 1 dx D c dx: Evaluating
both sides of the latter identity in e we obtain
c dx.e/ D Te .Cg 1 / dx.e/ D Ad.g 1 / dx.e/ D jdetAd.g

It follows that c D jdetAd.g/j 1 :

/j dx.e/:


A Lie group G with jdetAd.g/j D 1 for all g 2 G is said to be unimodular. The following
result is an immediate consequence of the lemma.
Corollary 19.13 Let G be a unimodular Lie group. Then every left invariant density is also
right invariant.
Lemma 19.14 Let G be a compact Lie group. Then G is unimodular.
Remark 19.15 It follows that the normalized Haar measure of a compact Lie group is biinvariant.
Proof: The map x 7! jdetAd.x/j is a continuous group homomorphism from G into the group
.RC ; / of positive real numbers equipped with multiplication. Its image H is a compact subgroup of .RC ; /: Now apply the lemma below to conclude that H D f1g:

Lemma 19.16 The only compact subgroup of .RC ; / is f1g:
Proof: Let be a compact subgroup of .RC ; /: Then 1 2 : Assume contains an element
> 0 different from 1: Since contains both and 1 we may as well assume that > 1: The
sequence . n /n1 belongs to and is unbounded from above, contradicting the compactness of
: It follows that D f1g:


We shall now investigate the existence of invariant densities on homogeneous spaces for G:
According to Proposition 15.5 such a space is of the form X D G=H; with H a closed subgroup
of G: Here G acts on X by left translation. For g 2 G we write lg W X ! X; xH 7! gxH:
The tangent map at e of the canonical projection  W G ! G=H induces a linear isomorphism g=h ' TeH .G=H / by which we identify. If h 2 H; then Ch W G ! G; g 7! hgh 1
leaves H invariant. Differentiation at e gives that Ad.h/ leaves the subspace h of g invariant,
hence induces a linear automorphism A.h/ of the quotient space g=h: The following lemma will
be useful in the sequel.
Lemma 19.17 Let h 2 H: Then the tangent map of lh W G=H ! G=H; xH 7! hxH at e is
given by
TeH .lh / D A.h/:
Proof: Let h 2 H: Recall that Ch W G ! G; x 7! hxh 1 has tangent map Ad.h/ at e: We
note that  Ch D lh : Differentiating at e and applying the chain rule we find Te  Ad.h/ D
TeH .lh / Te : It follows from this that TeH .lh / is the endomorphism of g=h induced by Ad.h/:

67

The fiber of the bundle DT .G=H / over eH is identified with D.g=h/: Thus A.h/ is an
automorphism of the associated space of densities D.g=h/: Note that for ! 2 D.g=h/ we have:
A.h/ ! D jdet A.h/j ! D

jdet Ad.h/jg j
!:
jdet Ad.h/jh j

(21)

We write D.g=h/H for the linear space of densities ! on g=h satisfying A.h/ ! D !: Such
densities are called H -invariant. Since D.g=h/ is one dimensional, the space of H -invariant
densities is either 0 or 1 dimensional. In view of (21) the latter is the case if and only if
jdetAd.h/jg j D jdetAd.h/jh j for all h 2 H:
Lemma 19.18
(a) The evaluation map  W ! 7! !.eH / defines a bijection from .DT .G=H //G onto
D.g=h/H : This bijection maps positive densities onto positive densities.
(b) The space of G-invariant densities on G=H is at most one dimensional. It is one dimensional if and only if
jdet Ad.h/jg j D jdet Ad.h/jh j

.h 2 H /:

Proof: Clearly  is a linear map. Assume that ! is a G-invariant density on G=H: Then for
g 2 G we have: TeH .lg / !.gH / D lg .!/.eH / D .!/; hence
!.gH / D .TeH .lg / 1 / .!/ D A.g/ 1 .!/:
This shows that the map  has a trivial kernel, hence is injective, and that its image is contained
in D.g=h/H : To establish its surjectivity, let !0 2 D.g=h/H : Then for all h 2 H we have
.TeH .lh / 1 / !0 D A.h/ 1 !0 D !0 ;
hence we may define a density on G=H by
!.gH / D .TeH .lg / 1 / !0 :
Note that the right hand side of this equation stays the same if g is replaced by gh; h 2 H: Hence
the definition is unambiguous. One readily verifies that ! thus defined is smooth, G-invariant,
and has image !0 under : This proves (a); the statement about positivity is obvious from the
above.
From (a) it follows that the dimension of .DT .G=H //G equals dimD.g=h/I hence, it is at
most one. The final assertion now follows from what was said in the preceding text.

Corollary 19.19 Let G be a Lie group, H a compact subgroup. Then G=H has a G-invariant
positive density. Two such densities differ by a positive factor.

68

Proof: For h 2 H; we put


.h/ D

jdetAd.h/jg j
:
jdetAd.h/jh j

Clearly is a Lie group homomorphism from H to the group RC consisting of the positive real
numbers, equipped with multiplication. Thus, .H / is a compact subgroup of RC : In view of
Lemma 19.16 this implies that .H / D f1g: The result follows.

Example 19.20 As S n ' SO.n C 1/=SO.n/; see Example 15.6, it follows that S n has a unique
SO.nC1/-invariant density of total volume 1: Similarly, Pn .R/ has a unique SO.nC1/-invariant
density of total volume 1I see Example 15.7. Real projective space is non-orientable, so it does
not have a volume form, i.e., a nowhere vanishing exterior differential form of top degree. This
problem of possible non-orientability of homogeneous spaces has been our motivation in using
densities rather than forms to define invariant integration.

20 Representations
In this section G will always be a Lie group.
In the following we will give some of the basic definitions of representation theory with V a
complete locally convex space over C: Every Banach space is an example of such a space. Natural spaces of importance for analysis, like C.M /; Cc .M /; C 1 .M /; Cc1 .M /; with M a smooth
manifold, and also the spaces D 0 .M / and E 0 .M / of distributions and compactly supported distributions, respectively, are complete locally convex, but in general not Banach. Of course Hilbert
spaces are Banach spaces; thus, they are covered as well.
Definition 20.1 Let V be a locally convex space. A continuous representation  D .; V / of
G in V is a continuous left action  W G  V ! V; such that .x/ W v 7! .x/v D .x; v/ is
a linear endomorphism of V; for every x 2 G: The representation is called finite dimensional if
dimV < 1:
Remark 20.2 If G is just a group, and V just a linear space, one defines a representation of G
in V similarly, but without the requirement of continuity.
Example 20.3
(a) Let G  X ! X; .g; x/ 7! gx be a left action of G on a set X; and let F .X/ denote the
space of functions X ! C: Then the action naturally induces the representation L of G
on F .X/ given by
Lg '.x/ D '.g 1 x/;
for ' 2 F .X/; g 2 G and x 2 X:

69

(b) Let L be the action of G on F .G/ induced by the left action G  G ! G; .g; x/ 7! gx:
This is called the left regular representation of G: It is given by the formula Lg '.x/ D
'.g 1 x/; for x; g 2 G:

Similarly, the right multiplication of G on itself induces the right regular representation
of G on F .G/ given by
Rg '.x/ D '.xg/;

for ' 2 F .G/; g; x 2 G: These representations leave the subspace C.G/  F .G/ invariant. Similarly, if dx is a left or right invariant Haar measure on G; then the associated
space L2 .G/ of square integrable functions is invariant under both L and R: One can show
that the restrictions of L and R to L2 .G/ are continuous, see Proposition 20.10.
(c) The natural action of SU.2/ on C2 induces a representation  of SU.2/ on F .C2 / given
by
N 2 ; z1 C z2 /;
.x/'.z/ D '.x 1 z/ D '.z
N 1 C z
for ' 2 F .C2 /; z 2 C2 and

xD

N
N

2 SU.2/;

i.e., ; 2 C; and jj2 C jj2 D 1:


Lemma 20.4 Let .; V / be a finite dimensional representation of G: If  is continuous, then 
is smooth.
Proof: By finite dimensionality of V; the group GL.V / is a Lie group. The map  W x 7! .x/ is
a homomorphism from G to GL.V /: The hypothesis that the representation is continuous means
that the map .x; v/ 7! .x/v is continuous. By finite dimensionality of V this implies that
 W G ! GL.V / is continuous. By Corollary 9.3 it follows that  W G ! GL.V / is smooth.
This in turn implies that .g; v/ 7! .g/v is smooth G  V ! V:

In the setting of the above lemma, the tangent map of  W G ! GL.V / at e is a Lie algebra
homomorphism  W g ! End.V /; where the latter space is equipped with the commutator
bracket. This motivates the following definition.
Definition 20.5 By a representation of l in V we mean a Lie algebra homomorphism  W l !
End.V /; i.e.,  is a linear map such that for all X; Y 2 l we have:
.X; Y / D .X/.Y /

.Y /.X/:

A representation of l in V is also called a structure of l-module on V: Accordingly,  is often


suppressed in the notation, by writing .X/v D Xv; for X 2 l; v 2 V: With this notation, the
above rule becomes
X; Y v D XY v

YXv
70

.X; Y 2 l; v 2 V /:

Remark 20.6 Similarly, a complete locally convex space V; equipped with a continuous representation of a Lie group G; will sometimes be called a G-module
Example 20.7 Ad W G ! GL.g/ is a continuous representation of G in g: The associated
infinitesimal representation of g in End.g/ is given by .X; Y / 7! . adX/Y D X; Y :
Proposition 20.8 Let  be a representation of G in a Banach space V: Then the following
conditions are equivalent:
(a)  W G  V ! V is continuous.
(b) For every x 2 G the map .x/ is continuous, and for every v 2 V the map G ! V; x 7!
.x/v is continuous at e:
Proof: That (b) follows from (a) is obvious. We will establish the converse implication by
application of the Banach-Steinhaus (or uniform boundedness) theorem.
Assume (b). Fix x0 2 G: If v 2 V then .x/v D .x0 /.x0 1 x/vI using (b) we see that
x 7! .x/v is continuous at x0 :
Now fix v0 2 V: Select a compact neighborhood N of x0 in G: Then f.x/ j x 2 N g is a
collection of continuous linear maps V ! V: Moreover, for every v 2 V; the map x 7! k.x/vk
is continuous, hence bounded on N: By the uniform boundedness theorem it follows that the
collection of operator norms k.x/k; for x 2 N is bounded, say by a constant C > 0: It follows
that for x 2 N; v 2 V we have
k.x/v

.x0 /v0 k  k.x/v .x/v0 k C k.x/v0 .x0 /v0 k


 C kv v0 k C k.x/v0 .x0 /v0 k:

The second term on the right-hand side tends to 0 if x ! x0 ; by (b). Hence .x; v/ 7! .x/v is
continuous in .x0 ; v0 /:

Remark 20.9 The above proof is based on the principle of uniform boundedness, and readily
generalizes to the category of complete locally convex spaces for which this principle holds, the
so called barrelled spaces.
The following result is in particular of interest if X D G and dx a left invariant positive
density on X:
Proposition 20.10 Let X be a manifold equipped with a continuous left G-action. Let dx be a
G-invariant positive continuous density on X: Then the natural representation L of G in L2 .X/
is continuous.
Proof: In view of the previous proposition it suffices to show that for every ' 2 L2 .X/ the
map W x 7! Lx '; G ! L2 .X/ is continuous at e: Thus we must estimate the L2 -norm of the
function Lx ' ' as x ! e: Let  > 0: Then there exists a 2 Cc .X/ such that k'
k2 < 13 :
71

Let g 2 Cc .G/ be a non-negative function such that g D 1 on an open neigbourhood of supp :


Then for x sufficiently close to e we have g D 1 on suppLx : Thus for such x we have:
kLx '

2
 C kLx
3
2
D
 C k.Lx
3
2

 C kLx
3

'k2 

k2
/gk2
k1 kgk2 :

Fix a compact neighborhood N of supp : For x sufficiently close to e one has suppLx  N:
By uniform continuity of on N; it now follows that kLx
k1 kgk2 < 3 for x sufficiently
close to e:

Definition 20.11 Let  be a representation of G in a (complex) linear space V: By an invariant
subspace we mean a linear subspace W  V such that .x/W  W for every x 2 G:
A continuous representation  of G in a complete locally convex space V is called irreducible, if 0 and V are the only closed invariant subspaces of V:
Remark 20.12 Note that for a finite dimensional representation .; V / an invariant subspace is
automatically closed. Thus, such a representation is irreducible if the only invariant subspaces
are 0 and V:
Definition 20.13 By a unitary representation of G we will always mean a continuous representation  of G in a (complex) Hilbert space H; such that .x/ is unitary for every x 2 G:
Remark 20.14 Let V be a complex linear space. Then by a sesquilinear form on V we mean a
map W V  V ! C which is linear in the first variable, and conjugate linear in the second, i.e.,
N
.v; w C w 0 / D .v;
w/ C .v; w 0 / for all v; w; w 0 2 V;  2 C:
A Hermitian inner product on V is a sesquilinear form h ; i that is conjugate symmetric, i.e.
hv ; wi D hw ; vi; and positive definite, i.e., hv ; vi  0 and hv ; vi D 0 ) v D 0 for all v 2 V:
Finally, we recall that a complex Hilbert space is a complex linear space H equipped with a
Hermitian inner product h ; i; whose associated norm is complete.
Remark 20.15 According to the above definition, a continuous representation of G in H is
unitary if and only if
h.x/v ; wi D hv ; .x

/wi

.v; w 2 H; x 2 G/:

Definition 20.16 A continuous finite dimensional representation .; V / of G will be called


unitarizable if there exists a Hermitian inner product on V for which  is unitary.
Proposition 20.17 Let G be compact, and suppose that .; V / is a continuous finite dimensional representation of G: Then  is unitarizable.
72

Proof: Let dx denote right Haar measure on G; and fix any positive definite Hermitian inner
product h ; i1 on V: Then we define a new Hermitian pairing on V by
Z
hv ; wi D h.x/v ; .x/wi1 dx
.v; w 2 V /:
G

Notice that the integrand v;w .x/ D h.x/v ; .x/wi1 in the above equation is a continuous
function of x: We claim that the pairing thus defined is positive definite. RIndeed, if v 2 V
then the function v;v is continuous and positive on G: Hence hv ; vi D G v;v .x/ dx  0
by positivity of the measure. Also, if hv ; vi D 0; then v;v  0 by Lemma 19.2, and hence
hv ; vi D v;v .e/ D 0; and positive definiteness follows.
Finally we claim that  is unitary for the inner product thus defined. Indeed this follows from
the invariance of the measure. If y 2 G; and v; w 2 V; then
Z
Z
h.y/v ; .y/wi D
v;w .xy/dx D
v;w .x/dx D hv ; wi:
G


Lemma 20.18 Let .; H/ be a unitary representation of G: If H1 is an invariant subspace for
; then its orthocomplement H2 D H1? is a closed invariant subspace for : If H1 is closed, then
we have the direct sum H D H1 H2 of closed invariant subspaces.
Proof: Let v 2 H2 and let x 2 G: We will show that .x/v 2 H2 : If w 2 H1 ; then .x 1 /w
belongs to H1 as well, so that h.x/v ; wi D hv ; .x 1 /wi D 0: It follows that .x/v 2 H1? :

Corollary 20.19 Let .; V / be a continuous finite dimensional representation of G: If  is
unitarizable, then it decomposes as a finite direct sum of irreducibles; i.e., there exists a direct
sum decomposition V D 1j n Vj of V into invariant subspaces such that for every j the
representation j defined by j .x/ D .x/jVj is irreducible.
Proof: Fix an inner product for which  is unitary, and apply the above lemma repeatedly.

Corollary 20.20 Let .; V / be a continuous finite dimensional representation of a compact Lie
group. Then every invariant subspace of V has a complementary invariant subspace. Moreover,
 admits a decomposition as a finite direct sum of irreducible representations.
Proof: By Proposition 20.17  is unitarizable. Now apply Lemma 20.18 and Corollary 20.19.
Definition 20.21 Let .; V / be a finite dimensional continuous representation of G: Then by a
matrix coefficient of  we mean any function m W G ! C of the form
m.x/ D mv; .x/ WD h.x/v ; i
with v 2 V and  2 V  :
73

Remark 20.22 Note that the map x 7! .x/ is smooth, so that every matrix coefficient belongs
to C 1 .G/:
If h ; i is a Hermitian inner product on V; then a matrix coefficient of  may also be characterized as a function of the form
m D mv;w W x 7! h.x/v ; wi;
with v; w 2 V; since w 7! h ; wi is a conjugate linear bijection from V onto V  :
Let now .; V / be a finite dimensional unitary representation of G; and fix an orthonormal
basis u1 ; : : : ; un of V: Then for every x 2 G we define the matrix M.x/ D Mu .x/ by
M.x/ij D mui ;uj .x/:
This is just the matrix of .x/ with respect to the basis u: Note that it is unitary. Note also
that M.xy/ D M.x/M.y/: Thus M is a continuous group homomorphism from G to the group
U.n/ of unitary n  n matrices.
Definition 20.23 If .j ; Vj /; for j D 1; 2; are continuous representations of G in complete
locally convex spaces, then a continuous linear map T W V1 ! V2 is said to be equivariant, or
intertwining if the following diagram commutes for every x 2 G W
V1
1 .x/ "
V1

! V2
" 2 .x/

! V2

The representations 1 and 2 are said to be equivalent if there exists a topological linear isomorphism T from V1 onto V2 which is equivariant.
If the above representations are finite dimensional, then one does not need to require T to be
continuous, since every linear map V1 ! V2 has this property. In the case of finite dimensional
representations we shall write HomG .V1 ; V2 / for the linear space of interwining linear maps
V1 ! V2 and EndG .V1 / for the space of intertwining linear endomorphisms of V1 :
If V is a complex linear space, we write End.V / for the space of linear maps from V to itself,
and GL.V / for the group of invertible elements in End.V /: If  is a representation of G in V ,
then we may define a representation Q of G in End.V / by
.g/A
Q
D .g/A.g/ 1 :
Note that if  is finite dimensional and continuous, then so is :
Q Note also that the space
End.V /G D fA 2 End.V / j .g/A
Q
D Ag
of G-invariants in V is just the space EndG .V / of G-equivariant linear maps V ! V:
74

Exercise 20.24 Let .j ; Vj /; for j D 1; 2; be two finite dimensional representations of G:


Show that 1 and 2 are equivalent if and only if there exist choices of bases for V1 and V2 ; such
that for the associated matrices one has:
mat1 .x/ D mat2 .x/:
Example 20.25 We recall that SU.2/ is the group of matrices of the form



N
gD

N
with ; 2 C and jj2 C jj2 D 1: The group SU.2/ acts on C2 in a natural way, and we have
the associated representation  on the space P .C/ of polynomial functions p W C2 ! C: It is
given by the formula
N 2 ; z1 C z2 /
.g/p.z/ D p.g 1 z/ D p.z
N 1 C z
The subspace Pn D Pn .C2 / of homogeneous polynomials of degree n is an invariant subspace
for : We write n for the restriction of  to Pn :
We will now discuss a result that will allow us to show that the representations n of the
above example are irreducible. We first need the following lemma from linear algebra.
Lemma 20.26 Let V be a finite dimensional complex linear space, and let A; B 2 End.V / be
such that AB D BA: Then A leaves ker B; imB and all the eigenspaces of B invariant.


Proof: Elementary, and left to the reader.


From now on all representations of G are assumed to be continuous.

Lemma 20.27 (Schurs lemma) Let .; V / be a finite dimensional representation of G: Then
the following holds.
(a) If  is irreducible then EndG .V / D C IV :

(b) Conversely, if  is unitarizable and EndG .V / D C IV ; then  is irreducible.

Proof: (a) Suppose that  is irreducible, and let A 2 End.V /G : Let  2 C be an eigenvalue
of A; and let E D ker.A I/ be the associated eigenspace. Note that for non-triviality of this
eigenspace we need V to be complex. For every x 2 G we have that .x/ commutes with A;
hence leaves E invariant. In view of the irreducibility of  it now follows that E D V; hence
A D I:
(b) By unitarizability of ; there exists a positive definite inner product h ; i for which 
is unitary.
Let 0 W  V be a G-invariant subspace. For the proof that  is irreducible it suffices
to show that we must have W D V: Let P be the orthogonal projection V ! W: Since W
and W ? are both G-invariant, we have, for g 2 G; that .g/P D .g/ D P .g/ on W; and
.g/P D 0 D P .g/ on W ? : Hence P 2 EndG .V /; and it follows that P D I for some
 2 C: Now P 0; hence  0: Also, P 2 D P; hence 2 D ; and we see that  D 1:
Therefore P D I; and W D V:

75

We will now apply the above lemma to prove the following.


Proposition 20.28 The representations .n ; Pn.C2 // of SU.2/; for n 2 N; are irreducible.
For the proof we will need compactness of SU.2/: In fact we have the following more general
result.
Exercise 20.29 For n  1; let M.n; R/ and M.n; C/ denote the linear spaces of n  n matrices
with entries in R and C respectively. Show that SU.n/ is a closed and bounded subset of M.n; C/:
Show that SO.n/ D SU.n/ \ M.n; R/: Finally show that the Lie groups SO.n/ and SU.n/ are
compact.
Proof of Proposition 20.28: Let n  0 be fixed, and put  D n and V D Pn .C2 /: Then n is
unitarizable, since SU.2/ is compact. Suppose that A 2 End.V / is equivariant. Then in view of
Lemma 20.27 (b) it suffices to show that A is a scalar.
For 0  k  n we define the polynomial pk 2 V by
pk .z/ D z1n

k k
z2 :

Then fpk j 0  k  ng is a basis for V: For ' 2 R we put


 i'


e
0
cos '
t' D
;
r' D
i'
0 e
sin '

sin '
cos '

Then
T D ft' j ' 2 Rg

and R D fr' j ' 2 Rg

are (closed) subgroups of SU.2/: One readily verifies that for 0  k  n and ' 2 R we have:
.t' /pk D e i.2k

n/'

pk :

Thus every pk is a joint eigenvector for T: Fix a ' such that the numbers e i.2k n/' are mutually
different. Then for every 0  k  n the space Cpk is eigenspace for .t' / with eigenvalue
e i.2k n/' : Since A and .t' / commute it follows that A leaves all the spaces Cpk invariant.
Hence there exist k 2 C such that
Apk D k pk ;

0  k  n:

Let E0 be the eigenspace of A with eigenvalue 0 : We will show that E0 D V; thereby completing the proof. The space E0 is SU.2/-invariant, and contains p0 : Hence it contains .r' /p0 for
every ' 2 R: By a straightforward computation one sees that
!
n
X
n
.r' /p0 .z1 ; z2 / D .cos ' z1 C sin ' z2 /n D
cosn k ' sink ' pk :
k
kD0

From this one sees by application of A and using the intertwining property, that
!
n
X
n
cosn k ' sink ' .0 k / pk D 0;
k
kD0

for all ' 2 R: By linear independence of the pk ; it follows that k D 0 ; for every 0  k  n:
Hence E0 D V:

76

We end this section with two useful consequences of Schurs lemma.


Lemma 20.30 Let .; V /, . 0 ; V 0 / be two irreducible finite dimensional representations of G:
If  and  0 are not equivalent, then every intertwining linear map T W V ! V 0 is trivial.
Proof: Let T be intertwining, and non-trivial. Then ker T  V is a proper G-invariant subspace.
Hence ker T D 0; and it follows that T is injective. Therefore its image imT is a non-trivial Ginvariant subspace of V 0 : It follows that imT D V 0 ; hence T is a bijection, contradicting the
inequivalence.

If .; V / is a representation for a group G; then a sesquilinear form on V is called equivariant if ..g/v; .g/w/ D .v; w/ for all v; w 2 V; g 2 G:
Lemma 20.31 Let .; V / be an irreducible finite dimensional unitary representation of a locally compact group G: Then the equivariant sesquilinear forms on V are precisely the maps
W V  V ! C of the form D h ; i;  2 C: Here h ; i denotes the (equivariant) inner
product of the Hilbert space V:
Proof: Let W V  V ! C be sesquilinear. Then for every w 2 V the map v 7! .v; w/ is a
linear functional on V: Hence there exists a unique A.w/ 2 V such that .v; w/ D hv; A.w/ ; :i
One readily verifies that A W V ! V is a linear map. Moreover, the equivariance of and h ; i
imply that A is equivariant. Since  is irreducible it follows by Schurs lemma that A D I for
some  2 C; whence the result.


21 Schur orthogonality
Assumption In the rest of these notes every finite dimensional representation of a Lie group will
be assumed to be continuous, unless specified otherwise.
In this section G will be a compact
Lie group, unless stated otherwise. Let dx be the unique
R
left invariant density on G with G dx D 1I for its existence, see Lemma 19.10. Then dx is
positive. By Remark 19.15, the density dx is right invariant as well.
If  is a finite dimensional irreducible unitary representation of G we write
C.G/

(22)

for the linear span of the space of matrix coefficients of : Notice that the space C.G/ does not
depend on the chosen (unitary) inner product on V: Thus, by Proposition 20.17 we may define
C.G/ for any irreducible finite dimensional (continous) representation  of G:
There is a nice way to express sums of matrix coefficients of a finite dimensional unitary
representation .; V / of G by means of the trace of a linear map. Let v; w 2 V: Then we shall
write Lv;w for the linear map V ! V given by
Lv;w .u/ D hu ; wiv:
77

One readily sees that


tr.Lv;w / D hv ; wi;

v; w 2 V:

(23)

Indeed both sides of the above equation are sesquilinear forms in .v; w/; so it suffices to check
the equation for v; w members of an orthonormal basis, which is easily done.
It follows from the above equation that
mv;w .x/ D tr..x/Lv;w /:
Hence every sum m of matrix coefficients is of the form m.x/ D tr..x/A/; with A 2 End.V /:
Conversely if fek j 1  k  ng is an orthonormal basis for V; then one readily sees that any
endomorphism A 2 End.V / may be expressed as
X
AD
hAej ; ei iLei ;ej :
1i;j n

Using this one may express every function of the form x 7! tr..x/A/ as a sum of matrix
coefficients.
We now define the linear map T W End.V / ! C.G/ by
T .A/.x/ D tr..x/A/;

x 2 G;

for every A 2 End.V /: Let  be irreducible, then it follows from the above discussion that T
maps V onto C.G/ : Define the representation    of G  G on End.V / by
   .x; y/A D .x/A.y/ 1 ;
for A 2 End.V / and x; y 2 G:
We define the representation R  L of G  G on C.G/ by
.R  L/.x; y/ WD Rx Ly D Ly Rx :
Lemma 21.1 Let .; V / be a finite dimensional irreducible representation of G: Then C.G/
is invariant under R  L: The map T W V ! C.G/ is surjective, and intertwines the representations    and R  L of G  G:
Proof: We first prove the equivariance of T W End.V / ! C.G/: Let A 2 End.V / and x; y 2 G;
then for all g 2 G;
T .   .x; y/A/.g/ D tr..g/.x/A.y

// D tr..y

gx/A/ D Rx Ly .T .A//.g/:

Note that it follows from this equivariance that the image of T is R  L-invariant. In an earlier
discussion we showed already that im.T / D C.G/ :

Corollary 21.2 If  and  0 are equivalent finite dimensional irreducible representations of G;
then C.G/ D C.G/ 0 :
78

Proof: Let V; V 0 be the associated representation spaces. Then by equivalence there exists a
linear isomorphism T W V ! V 0 such that T .x/ D  0 .x/ T for all x 2 G: Hence for
A 2 End.V / and x 2 G;
T 0 .TAT

/.x/ D tr. 0 .x/TAT

/ D tr.T

 0 .x/TA/ D tr..x/A/ D T .A/.x/:

Now use that T and T 0 have images C.G/ and C.G/ 0 ; respectively, by Lemma 21.1.

We now have the following.


Theorem 21.3 (Schur orthogonality). Let .; V / and . 0 ; V 0 / be two irreducible finite dimensional representations of G: Then the following holds.
(a) If  and  0 are not equivalent, then C.G/ ? C.G/ 0 (with respect to the Hilbert structure
of L2 .G/).
(b) Let V be equipped with an inner product for which  is unitary. If v; w; v 0; w 0 2 V; then
the L2 -inner product of the matrix coefficients mv;w and mv 0 ;w 0 is given by:
Z
(24)
mv;w .x/ mv 0 ;w 0 .x/ dx D dim./ 1 hv ; v 0 ihw ; w 0 i
G

Remark 21.4 The relations (24) are known as the Schur orthogonality relations. Of course the
assumption that dx is normalized is a necessary assumption for (24) to hold.
Proof: For w 2 V and w 0 2 V 0 we define the linear map Lw 0 ;w W V ! V 0 by Lw 0 ;w u D
hu ; wiw 0 : Consider the following linear map V ! V 0 ; defined by averaging,
Z
Iw 0 ;w D
 0 .x/ 1 Lw 0 ;w .x/ dx:
G

One readily verifies that

hIw 0 ;w v ; v 0 i D hmv;w ; mv 0 ;w 0 iL2 :

(25)

Moreover, by right invariance of the measure dx it readily follows that Iw 0 ;w is an intertwining


map from .V; / to .V 0 ;  0 /:
(a): If  and  0 are inequivalent then the intertwining map Iw 0 ;w is trivial by Lemma 20.30.
Now apply (25) to prove (a).
(b): Now assume V D V 0 : Then for all w; w 0 2 V we have Iw 0 ;w 2 EndG .V /; hence Iw 0 ;w
is a scalar. It follows that there exists a sesquilinear form on V such that
Iw 0 ;w D .w 0 ; w/ IV :

Applying the trace to both sides of the above equation we find that tr.Iw 0 ;w / D d .w 0 ; w/: Here
we have abbreviated d D dim./: On the other hand, since tr is linear,
Z
Z
1
tr.Iw 0 ;w / D
tr..x/ Lw 0 ;w .x// dx D
tr.Lw 0 ;w / dx D tr.Lw 0 ;w / D hw 0 ; wi:
G

Hence

Iw 0 ;w D .w 0 ; w/ IV D d 1 hw 0 ; wi IV :

Now apply (25) to prove (b).

79

Another way to formulate the orthogonality relations is the following (V is assumed to be


equipped with an inner product for which  is unitary). If A 2 End.V /, let A denote the
Hermitian adjoint of A: Then one readily verifies that
.A; B/ 7! hA ; Bi WD trB  A
defines a Hermitian inner product on End.V /: Moreover, the representation    is readily
seen to be unitary for this inner product.
Corollary 21.5 The map T WD
onto C.G/ :

d T is a unitary G-equivariant isomorphism from End.V /

Proof: We begin by establishing a few properties of the endomorphisms Lv;w ; for v; w 2 V:


From the definition one readily sees that, for v 0 ; w 0 2 V the adjoint of Lv 0 ;w 0 is given by
Lv 0 ;w 0 D Lw 0 ;v 0 :
Moreover, one also readily checks that
Lw 0 ;v 0 Lv;w D hv ; v 0 iLw 0 ;w :
From these two properties it follows in turn that
hLv;w ; Lv 0 ;w 0 i D tr.Lw 0 ;v 0 Lv;w / D hv ; v 0 ihw ; w 0 i:

(26)

Finally, we recall that


T .Lv;w /.x/ D tr..x/Lv;w / D mv;w .x/

.x 2 G/;

hence
h T .Lv;w / ; T .Lv 0 ;w 0 /iL2 D d hmv;w ; mv 0 ;w 0 iL2 :

(27)

From (26) and (27) we see that the Schur orthogonality relations may be reformulated as
h T .Lv;w / ; T .Lv 0 ;w 0 /iL2 D hLv;w ; Lv 0 ;w 0 i;

(28)

for all v; v 0 ; w; w 0 2 V: The maps Lv;w ; for v; w 2 V; span the space End.V /: Hence the Schur
orthogonality relations are equivalent to the assertion that T is an isometry from End.V / into
C.G/ : We proved already that T is surjective onto C.G/ I hence T is a unitary isomorphism. The equivariance of T has been established before.

Definition 21.6 Let .V; / be a finite dimensional representation of G: The function  W G !
C defined by
 .x/ D tr.x/;
.x 2 G/;
is called the character of :
80

Remark 21.7 Since the representation  is continuous, it is also smooth, hence  2 C 1 .G/:
Note that  is a sum of matrix coefficients of : Thus, if G is compact and  irreducible, then
 2 C.G/ :
Lemma 21.8 Let .; V / be an irreducible finite dimensional representation of G: Then  is
the unique conjugation invariant function in C.G/ with  .e/ D d : Its L2 -norm relative to
the normalized Haar measure is k k2 D 1:
Proof: We equip V with an inner product for which  is unitary, and define the associated inner
product on End.V / as above. Let ' 2 C.G/ : Then ' D T .A/ for a unique A 2 End.V /:
By equivariance of T ; the function ' is conjugation invariant if and only if A is G-intertwining,
which in turn is equivalent to A D cIV for a constant c 2 C: We observe that c D '.e/=d3=2 :
This implies that there exists
p a unique conjugation invariant function ' with '.e/ D d : For this
function we have c D 1= d and
p
'.x/ D T .cIV /.x/ D d tr..x/cIV / D tr.x/ D  .x/:
The assertion about the L2 -norm follows from

k k2 D tr.cI / cI D c 2 tr.I / D c 2 d D 1:




22 Characters
In this section we assume that G is a Lie group. We shall discuss properties of characters of finite
dimensional representations of G:
If V is a finite dimensional complex linear space, we write End.V / for the space of complex
linear maps from V to itself, and det D detV and tr D trV for the complex determinant and trace
functions End.V / ! C:
Lemma 22.1 Let T W V ! W be a linear isomorphism of finite dimensional linear spaces.
Then for every linear map A W V ! V;
detW .T A T

/ D detV A

and

trW .T A T

/ D trV A:

Let V be a finite dimensional linear space. Then for all A; B 2 End.V /;


tr.A B/ D tr.B A/:


Proof: Exercise for the reader.

81

The character  of a finite dimensional representation .; V / of G is defined as in Definition 21.6.


Lemma 22.2 Let ;  be finite dimensional representations of G: If  and  are equivalent,
their characters are equal:  D  :
Proof: Let T W V ! V be an equivariant linear isomorphism. Then .x/ D T .x/ T 1
for every x 2 G: The result now follows by application of Lemma 22.1.

Lemma 22.3 Let .; V / be a finite dimensional representation of G: Then, for all x; y 2 G;
 .xyx

/ D  .y/:


Proof: Exercise for the reader.

Definition 22.4 Let  be a representation of G in a finite dimensional complex linear space V:


We define the contragredient or dual of  to be the representation  _ of G in the dual linear
space V  given by
 _ .x/ D .x

1 

/ W v  7! v  .x

.x 2 G/:

/;

Lemma 22.5 Let .; V / be a finite dimensional representation of G:


(a) If  is continuous, then  _ is continuous as well.
(b) The character of  _ is given by
 _ .x/ D  .x

.x 2 G/:

Proof: Let v1 ; : : : ; vn be a basis for V and let v 1 ; : : : ; v n be the dual basis for V  ; i.e., v i .vj / D
ij : Then, for x 2 G; the matrix of  _ .x/ with respect to the basis v1 ; : : : ; vn is given by
 _ .x/ij D h.x

1  j

/ v ; vi i D hv j ; .x

/vi i D .x

/j i :

If  is continuous, then its matrix coefficients are continuous functions. Therefore, so are the
matrix coefficients of  _ ; and (a) follows. Assertion (b) follows from the above identity as well.

Characters of unitarizable representations have the following special property.
Lemma 22.6 Let  be a finite dimensional representation of G: If  is unitarizable, then
 .x

/ D  .x/;

.x 2 G/:


Proof: Exercise for the reader.


82

If .1 ; V1 / and .2 ; V2 / are two continuous representations of G; then we define the direct
sum representation  D 1 2 in the direct sum V D V1 V2 by
.x/.v1 ; v2 / D .1 .x/v1 ; 2 .x/v2 /

.v1 2 V1 ; v2 2 V2 ; x 2 G/:

Lemma 22.7 Let 1 ; 2 be finite dimensional representations of G: Then


1 2 D 1 C 2 :


Proof: Exercise for the reader.

If .1 ; V1 / and .2 ; V2 / are two finite dimensional representations of G; we define their tensor
product 1 2 to be the representation of G in the tensor product space V1 V2 given by
.1 2 /.x/ D 1 .x/ 2 .x/: Thus, for x 2 G; the linear endomorphism .1 2 /.x/ of
V1 V2 is determined by
.1 2 /.x/.v1 v2 / D 1 .x/v1 2 .x/v2 ;
for all v1 2 V1 ; v2 2 V2 :
Lemma 22.8 Let .1 ; V1 / and .2 ; V2 / be finite dimensional representations of G: Then the
character of their tensor product 1 2 is given by
1 2 D 1 2 :
Proof: Exercise for the reader. Establish, more generally, an identity of the form tr.A B/ D
tr.A/tr.B/; by choosing suitable bases.

Exercise 22.9 Recall the definition, for n 2 N; of the representation n of SU.2/ in the finite
dimensional space Pn .C2 / of homogeneous polynomial functions C2 ! C of degree n: Show
that the character n of n is completely determined by its restriction to T D ft' j ' 2 Rg: Hint:
use that every matrix in SU.2/ is conjugate to a matrix of T:
Show that:
sin.n C 1/'
;
n.t' / D
sin '
for ' 2 R: Here t' denotes the diagonal matrix with entries e i' and e

i'

Assumption: In the rest of this section we assume that the Lie group G is compact. We denote
by h  ;  i the L2 -inner product with respect to the normalized Haar measure dx on G:
Lemma 22.10 Let ;  be finite dimensional irreducible representations of G:
(a) If    then h ;  i D 1:
(b) If  6  then h ;  i D 0:
Proof: This follows easily from Theorem 21.3.
83

Let  be a finite dimensional representation of the compact group G: Then  is unitarizable,


and therefore
equivalent to a direct sum niD1 i of irreducible representations. It follows that
Pn
 D iD1 i : Using the lemma above we see that for every irreducible representation of G
we have
#fi j i  g D h ;  i:
(29)

In particular this number is independent of the particular decomposition of  into irreducibles.


For obvious reasons the number (29) is called the multiplicity of in : We shall also denote it
by m.; /:
b denote the set of equivalence classes of finite dimensional irreducible representations
Let G
b to indicate that is a representative for an
of G: Then by abuse of language we shall write 2 G
b (A better notation would perhaps be 2 G:/
b If 2 G
b and m 2 N; then we write
element of G:
m for (the equivalence class of) the direct sum of m copies of :
We have proved the folllowing lemma.
Lemma 22.11 Let  be a finite dimensional representation of the compact group G: Then
M

m.; /;
b
2G

b Any decomposition of  into irreducibles is


where m.; / D h ;  i 2 N; for every 2 G:
equivalent to the above one.

Exercise 22.12 This exercise is meant to illustrate that a decomposition of a representation


into irreducibles is not unique. Let 1 ; 2 be irreducible representations in V1 ; V2 respectively.
Assume that 1 ; 2 are equivalent, and let T W V1 ! V2 be an intertwining isomorphism.
Equip V D V1 V2 with the direct sum representation , and show that W1 D f.v; T v/ j
v 2 V1 g is an invariant subspace of V: Show that the restriction of  to W1 is irreducible, and
equivalent to 1 : Find a complementary invariant subspace W2 and show that the restriction of 
to this space is also equivalent to 1 :
The following result expresses that the character is a powerful invariant.
Corollary 22.13 Let ;  be two finite dimensional continuous representations of G: Then
   ()  D  :
Proof: The implication ) follows from Lemma 22.2. For the converse implication, assume
b we have m.; / D h ;  i D h ;  i D m.; /: Now
that  D  : Then for every 2 G
use the previous lemma.

Corollary 22.14 Let  be a finite dimensional representation of G: Then  is irreducible if and
only if its character  has L2 -norm one.
b form an orthonormal set in L2 .G/:
Proof: By Schur orthogonality,
the characters  ; for 2 G
P
2
2
It follows that k k D m.; / : The result now easily follows.

84

23 The Peter-Weyl theorem


b the set of (equivalence
In this section we assume that G is a compact Lie group. We denote by G
classes of) irreducible continuous finite dimensional representations of G:
Definition 23.1 We define the space R.G/ of representative functions to be the space of funcb
tions f W G ! C that may be written as a finite sum of functions f 2 C.G/ ; for 2 G:

Note that the space R.G/ is contained in C 1 .G/: Moreover, it is invariant under both the
left- and the right regular representations of G:
Exercise 23.2 Show that R.G/ is the linear span of the set of all matrix coefficients of finite
dimensional continuous representations of G: Hint: consider the decomposition of finite dimensional representations into irreducibles.
Proposition 23.3 The space of representative functions decomposes according to the algebraic
direct sum
M
R.G/ D
C.G/ :
b
2G

The summands are mutually orthogonal with respect to the L2 -inner product. Every summand
C.G/ is invariant under the representation R  L of G  G: Moreover, the restriction of R  L
to that summand is an irreducible representation of G  G:
Proof: The orthogonality of the summands follows from Schur orthogonality. It follows that the
above sum is direct.
The map T W End.V / ! C.G/ is bijective and intertwines  with R  L: Hence it
suffices to show that  is an irreducible representation of G  G:
By a straightforward computation one checks that
  .x; y/ D  .x/ .y/;
for .x; y/ 2 G  G: If dx and dy are normalized right Haar measure on G; then the product
measure dx dy is the normalized right Haar measure on G  G: Moreover, by Fubinis theorem,
Z
2
j .x/j2 j .y/j2 dx dy
k  kL2 .GG/ D

Z
ZGG
2
2
j .x/j j .y/j dx dy
D
G
2
k kL2 .G/ k k2L2 .G/
G

D 1;

since is an irreducible representation of G: It follows from Corollary 22.14 that  is an


irreducible representation of G  G:


85

The proof of the following result is based on the spectral theorem for compact self-adjoint
operators in a Hilbert space. It will be given in the next section.
Proposition 23.4 The space R.G/ is dense in L2 .G/:
Let H be a collection of Hilbert spaces, indexed by a set A: Then the algebraic direct sum
M
H
2A

P
P
is
Pa pre-Hilbert space when equipped with the direct sum inner product: h v ; w i D
hv ; w i: Its completion is called the Hilbert direct sum of the spaces H ; and denoted by
^
M

H :

(30)

2A

This completion may be realized as the space of sequences v D .v /2A with v 2 H and
X
kv k2 < 1:
kvk2 D
2A

Its inner product is given by


hv ; wi D

hv ; w i:

2A

If  is a unitary representation of G in H ; for every 2 A; then the direct sum of the 


extends to a unitary representation of G in (30). We call this representation the Hilbert sum of
the  :
Theorem 23.5 (The Peter-Weyl Theorem). The space L2 .G/ decomposes as the Hilbert sum
2

L .G/ D

^
M

C.G/ ;

b
2G

each of the summands being an irreducible invariant subspace for the representation R  L of
G  G:
Proof: This follows from Propositions 23.3 and 23.4.

Exercise 23.6 Fix, for every (equivalence class of an) ireducible unitary representation .; V /
an orthonormal basis e;1 ; : : : ; e;dim./ : Denote the matrix coefficient associated to e;i and e;j
by m;ij : Use Schur orthogonality and the Peter-Weyl theorem to show that the functions
p
b 1  i; j  dim./
dim./ m;ij
2 G;

constitute a complete orthonormal system for L2 .G/:


86

24 Appendix: compact self-adjoint operators


Definition 24.1 Let V; W be Banach spaces. A linear map T W V ! W is said to be compact if
the image T .B/ of the unit ball B D B.0I 1/  V has compact closure in W:
A compact operator T W V ! W is obviously bounded. The set of compact operators forms
a linear subspace of the space L.V; W / of bounded linear operators V ! W: The latter space is
a Banach space for the operator norm.
Lemma 24.2 Let V; W be Banach spaces, and let L.V; W / be the Banach space of bounded
linear operators V ! W; equipped with the operator norm. Then the subspace of compact
linear operators V ! W is closed in L.V; W /:
Proof: See a standard textbook on functional analysis.

Remark 24.3 A linear map T W V ! W is said to be of finite rank if its image T .V / is finite
dimensional. Clearly an operator of finite rank is compact. Thus, if Tj is a sequence of operators
in L.V; W / all of which are of finite rank, and if Tj ! T with respect to the operator norm, then
it follows from the above result that T is compact.
We recall that a bounded linear operator T from a complex Hilbert space H to itself is said
to self-adjoint if T  D T; or, equivalently, if hT v ; wi D hv ; T wi for all v; w 2 H:
We now recall the important spectral theorem for compact self-adjoint operators in Hilbert
space. It will play a crucial role in the proof of the Peter-Weyl theorem in the next section. For a
proof of the spectral theorem, we refer to a standard text book on functional analysis.
Theorem 24.4 Let T be a compact self-adjoint operator in the (complex) Hilbert space H: Then
there exists a discrete subset  R n f0g such that the following hold.
(a) For every  2 the associated eigenspace H of T in H is finite dimensional.
(b) If ;  2 ;   then H ? H :
(c) For every  2 ; let P denote the orthogonal projection H ! H : Then
X
T D
 P ;
2

the convergence being absolute with respect to the operator norm.


(d) The set is bounded in R and has 0 as its only limit point.
We will end this section by describing a nice class of compact self-adjoint operators in
L2 .G/; for G a compact Lie group. First we examine the space of compactly supported continuous functions on product space.

87

Let X; Y be locally compact topological Hausdorff spaces. If ' 2 C.X/; and


then we write ' for the continuous function on X  Y defined by:
'

2 C.Y /;

W .x; y/ 7! '.x/ .y/:

The linear span of such functions in C.X  Y / is denoted by C.X/ C.Y /: If ' 2 Cc .X/
and 2 Cc .Y / then ' is compactly supported. Hence the span Cc .X/ Cc .Y / of such
functions is a subspace of Cc .X  Y /:
Proposition 24.5 Let X; Y be locally compact Hausdorff spaces. Then for every open subset
U  X  Y with compact closure, every 2 Cc .U / and every  > 0; there exists a function
' 2 Cc .X/ Cc .Y / with supp'  U and supz2U j.z/ '.z/j < : In particular, the space
Cc .X/ Cc .Y / is dense in Cc .X  Y /:
Proof: Using Cc -partitions of unity for X and Y; we see that we may reduce to the case that
U D UX  UY ; with UX and UY open neighborhoods with compact closures in X and Y
respectively.
Fix 2 Cc .X  Y /; with K D supp  U: Then, by compactness, K  KX  KY for
compact subsets KX  UX and KY  UY . Let  > 0: Then by compactness there exists a finite
open covering fVj g of KX such that for every j and all x1 ; x2 2 Vj ; y 2 KY one has
.x1 ; y/

.x2 ; y/ < :

Without loss of generality we may assume that Vj  UX for all j: Select a partition of unity f'j g
which is subordinate to the covering fVj g; and fix for every j a point j 2 Vj : Let x 2 KX ; y 2
KY : If j is such that x 2 Vj ; then j.xj ; y/ .x; y/j < : It follows from this that
X
X
j
'j .x/.xj ; y/ .x; y/j D j
'j .x/.xj ; y/ 'j .x/.x; y/j
j


<

X
j

Hence, if we put

j .y/

D .xj ; y/; then


X
k
'j

'j .x/j.xj ; y/

.x; y/j

'j .x/ D :

k1 < :

Moreover, supp'j

 UX  UY  UX  KY  U:

Let now G be a Lie group. We fix a left invariant density dx on G and equip G  G with
the left invariant product of dx with itself. This product density, denoted dxdy; is determined
by the formula


Z Z
Z
Z Z
f .x; y/dy dx;
f .x; y/dx dy D
f .x; y/ dxdy D
GG

88

for f 2 Cc .X  Y /:
If K 2 Cc .G  G/; then we define the linear operator TK W Cc .G/ ! Cc .G/ by
Z
TK .'/.x/ D
K.x; y/'.y/dy:
G

For obvious reasons this is called an integral operator with kernel K:


Lemma 24.6 Let K 2 Cc .G G/: Then the operator TK extends uniquely to a bounded
linear endomorphism of L2 .G/ with operator norm kTK kop  kKk2 : Moreover, this extension
is compact.
Proof: Let ' 2 Cc .G/: Then
hTK .'/ ;

i D

TK .'/.x/ .x/ dx

Z Z
.x/ dx
K.x; y/'.y/ dy
D
G

D hK ; ' N iL2 .GG/


 kKkL2 .GG/ k' N kL2 .GG/ D kKk2 k'k2 k k2 :
Hence kTK 'k2  kKk2 k'k2 : This implies the first assertion, since Cc .G/ is dense in L2 .G/:
For the second assertion, note that by Proposition 24.5 there exists a sequence Kj in Cc .G/
Cc .G/ which converges to K with respect to the L2 -norm on G  G: It follows that
kTKj

TK kop  kKj

Kk2 ! 0:

Every operator TKj has a finite dimensional image hence is compact. The subspace of compact
endomorphisms of L2 .G/ is closed for the operator norm, by Lemma 24.2. Therefore, TK is
compact.

Let G be a Lie group, equipped with a left invariant density dx: If .; V / is a continuous
finite dimensional representation of G, then for f 2 Cc .G/ we define the linear operator .f / W
V ! V by
Z
.f /v D

f .x/.x/v dx:

Referring to integration with values in a Banach space, this definition actually makes sense if 
is a continuous representation in a Banach space; it is readily seen that then .f / is a continuous
linear operator. In particular, the definition may be applied to the regular representations L and
R of G in L2 .G/: Thus, for f 2 Cc .G/ and ' 2 L2 .G/;
Z
Z
R.f /'.x/ D
f .y/'.xy/ dy D
f .x 1 y/'.y/ dy
.x 2 G/:
(31)
G

Of course, this formula can also be used as the defining formula, without reference to Banachvalued integration.
89

Corollary 24.7 Assume that G is compact, and let f 2 C.G/: Then the operator R.f / W
L2 .G/ ! L2 .G/ is compact.
Proof: If ' 2 C.G/; then from (31) we see that R.f / D TK ; with K.x; y/ D f .x
result now follows by application of Lemma 24.6.

y/: The


Remark 24.8 Note that for this argument it is crucial that G is compact. For if not, and f 2
Cc .G/; then the associated integral kernel K need not be compactly supported.
The following lemmas will in particular be needed for the right regular representation R:
Lemma 24.9 Let .; H/ be a unitary representation of G in a Hilbert space. Let f 2 Cc .G/;
then
.f / D .f  /;
where f  .x/ D f .x

1 /:

Proof: Straightforward and left to the reader.

Lemma 24.10 Let  be a continuous representation of G in a Banach space V: If f 2 Cc .G/


is conjugation invariant, then .f / is intertwining.


Proof: Straightforward and left to the reader.

Corollary 24.11 Assume that G is compact, and let f 2 C.G/ be such that f  D f: Then
R.f / (and L.f / as well) is a compact self-adjoint operator. If, in addition, f is conjugation
invariant then R.f / is G-equivariant.
Proof: This follows by combining Corollary 24.11 and Lemmas 24.9 and 24.10.

25 Proof of the Peter-Weyl Theorem


In the beginning of this section we assume that G is any Lie group. At a later stage we will
restrict our attention to compact G: We assume that G is equipped with a positive left invariant
density dx:
Lemma 25.1 Let ' 2 Cc .G/: Then R.'/ maps L2 .G/ into C.G/:

90

Proof: Let x0 2 G and let  > 0: Since ' has compact support C WD supp'; it follows by the
principle of uniform continuity that there exists a compact neighborhood U of e in G such that
j'.u/ '.v/j < .2k1C k2 C 1/ 1 for all u; v 2 G with vu 1 2 U:
Let now f 2 L2 .G/: For x; y 2 G with x 2 x0 U we have .x0 1 y/.x 1 y/ 1 D x0 1 x 2 U;
hence
Z
jR.'/f .x/ R.'/f .x0 /j D j '.x 1 y/ '.x0 1 y/f .y/ dyj
Z G
Z

 jf .y/j dy D 
1xC [x0 C jf .y/j dy
xC [x0 C

  k1xC [yC k2 kf k2  2 k1C k2 kf k2  kf k2 :

From this we deduce that R.'/f is continuous in x0 :

Lemma 25.2 Let f 2 L2 .G/ and let  > 0: There exists an open neighborhood U of e in G
such that for all x 2 U we have kRx f f k2 <R : Moreover, if U is any neighborhood with this
property and if ' 2 Cc .U / satisfies '  0 and G '.x/ dx D 1; then
kR.'/f

f k2 < :

(32)

Proof: The first assertion follows from the continuity of the map x 7! Rx f; see Proposition
20.10. Let U; ' be as stated. Then, for all x 2 G;
Z
R.'/f .x/ f .x/ D
'.y/f .xy/ f .x/ dy:
G

Hence, for every g 2 L2 .G/ we have


jhR.'/f

f ; gij 
D


Z Z

ZG

ZG
G

ZG
G

'.y/ jf .xy/

f .x/j g.x/ dy dx

'.y/ jf .xy/

f .x/j g.x/ dx dy

'.y/ kRy f

f k2 kgk2 dy

  kgk2 :


From this the estimate (32) follows.


From now on we assume that the group G is compact.

Lemma 25.3 Let V be a finite dimensional right G-invariant subspace of L2 .G/: Then V 
R.G/:

91

Proof: Decomposing V into a direct sum of irreducible subspaces, we see that we may reduce
the case that V is irreducible. We claim that V consists of continuous functions. For this we
observe that C.G/ \ V is an invariant subspace. Hence it suffices to show that V contains a
non-trivial continuous function. Fix f 2 V n f0g and fix 0 <  < 1=2kf k2 : Choose U and '
as in Lemma 25.2. Then kR.'/f k > 1=2; hence R.'/f 0: From Lemma 25.1 it follows
that R.'/f 2 C.G/: Moreover, since V is right invariant, it follows that R.'/f 2 V: This
establishes the claim that V  C.G/:
Choose an orthonormal basis . i / of V: Then for f 2 V we have
X
Rx f D
hRx f ; i i i ;
i

hence by evaluation in e;
f .x/ D

X
i

hRx f ;

By definition of R.G/ it now follows that f 2 R.G/:

ii

i .e/:

Lemma 25.4 Let U be an open neighborhood of e in G: Then there exists a ' 2 Cc .U / such
that:
R
(a) '  0 and G '.x/ dx D 1I
(b) '  D 'I

(c) ' is conjugation invariant.


Proof: From the continuity of the map x 7! x 1 one sees that there exists a compact neighborhood V of e such that V  U and V 1  U: For every x 2 G there exist an open neighborhood
Nx of x and a compact neighborhood Vx of e in V such that zyz 1 2 V for all z 2 Nx ; y 2 Vx :
By compactness of G finitely many of the Nx cover G: Let  be the intersection of the corresponding Vx : Then  is a compact neighborhood of e and for all x 2 G and y 2  we have
xyx 1 2 V:
R
Now select 0 2 Cc ./ such that 0  0 and G 0 .x/ dx D 1: Define
Z
1
.x/ D
/ dy:
0 .yxy
G

Since .x; y/ 7! .yxy 1 / is a continuous function, it follows that is a continuous function.


Clearly  0: Moreover, by interchanging theRorder of integration, and using the fact that dx
is bi-invariant and normalized, we deduce that G .x/ dx D 1: If .x/ 0; then yxy 1 2
supp 0 for some y 2 G; hence x 2 [y2G y 1 y  V: It follows that supp  V: One now

readily verifies that the function ' D 12 . C  / satisfies all our requirements.
Corollary 25.5 Let f 2 L2 .G/; f 0: Then there exists a left and right G-equivariant
bounded linear operator T W L2 .G/ ! L2 .G/ with:
92

(a) Tf 0:
(b) T is self-adjoint and compact;
(c) T maps every right G-invariant closed subspace of L2 .G/ into itself.
Proof: Let  D 12 kf k2 ; and fix an open neighborhood U of e in G that satisfies the assertion
of Lemma 25.2 Let ' 2 Cc .U / be as in Lemma 25.4, and define T D R.'/: Then kTf
f k < ; hence (a). Moreover, every closed right invariant subspace V of L2 .G/ equipped
with the restriction of R is a continuous representation in a Banach space, hence invariant under
T D R.'/: This implies (c).
The operator T is left G-equivariant, since L and R commute. It is right G-equivariant
because ' is conjugation invariant, cf. Lemma 24.10. Finally (b) follows from Corollary 24.11.

Proof of Propostion 23.4. The space R.G/ is left and right G-invariant, and by unitarity so is its
orthocomplement V: Suppose that V contains a non-trivial element f: Let T be as in Corollary
25.5. Then T jV W V ! V is a non-trivial compact self-adjoint operator which is both left and
right G-equivariant. By the spectral theorem for compact self-adjoint operators, Theorem 24.4,
there exists a  2 R,  0; such that the eigenspace V D ker.T IV / is non-trivial. By
compactness of T the eigenspace V is finite dimensional, and by equivariance of T it is both left
and right G-invariant. By Lemma 25.3 it now follows that V  R.G/; contradiction. Therefore,
V must be trivial.


26 Class functions
By a class function on a compact Lie group G we mean a function f W G ! C that is conjugation
invariant, i.e., Lx Rx f D f for all x 2 G: The name class function comes from the fact that a
conjugation invariant function is constant on the conjugacy classes, hence may be viewed as a
function on the set of conjugacy classes.
The space C.G; class/ of continuous class functions is a closed subspace of C.G/ (with
respect to the sup norm). Its closure in L2 .G/ equals L2 .G; class/; the space of square integrable
class functions on G:
b we denote the orthogonal projection from L2 .G/ onto the finite dimensional subIf 2 G;
space C.G/ by
P W L2 .G/ ! C.G/
Note that P is equivariant for both the representations R and L of G: In particular, this implies
that P maps C.G; class/ into its intersection with C.G/ : Hence, by Lemma 21.8
P .C.G; class// D C.G/ \ C.G; class/ D C :
It follows from this that the space R.G; class/ D C.G; class/ \ R.G/ of representative class
b
functions is the linear span of the characters  ; 2 G:
93

b form a complete
Lemma 26.1 Let G be a compact Lie group. Then the characters  ; 2 G;
2
orthonormal system for L .G; class/:

Proof: By Schur orthogonality, the characters form an orthonormal system. To establish its
b
completeness, let f 2 L2 .G; class/ and assume that f ?  for all 2 G:
From P f 2 C.G/ D C ; we see that
P f D hP f ;  i  D hf ;  i  D 0:

Hence f ? R.G/: By the Peter-Weyl theorem, the latter implies that f D 0:

Corollary 26.2 Let f 2 L2 .G; class/: Then


X
f D
hf ;  i  ;
with convergence in the L2 -norm.

b
2G

27 Abelian groups and Fourier series


In this section we consider the special case that the compact Lie group G is commutative. If, in
addition, G is connected, then G ' Rn =Zn for some n 2 N; and we will see that the Peter-Weyl
theorem specializes to the theory of Fourier series.
By a multiplicative character of G we mean a continuous (hence smooth) group homomorphism  W G ! C ; where C D C n f0g is equipped with complex multiplication. By the lemma
below, if  is a multiplicative character, then j.x/j D 1; x 2 G:
Lemma 27.1 Let H be a compact subgroup of C : Then H  T:
Proof: By compactness, there exists a constant r > 0 such that r 1 < jzj < r for all z 2 H:
Let w 2 H; then applying the estimate to z D w n we obtain that r 1=n  jwj  r 1=n : Taking
the limit for n ! 1 we see that jwj D 1:

Lemma 27.2 Let G be a commutative compact Lie group. If .; V / is a finite dimensional
irreducible representation of G; then dimV D 1: Moreover, .x/ D  .x/IV : The map 7! 
b onto the set of multiplicative characters of G:
induces a bijection from G
Proof: If x 2 G; then .y/.x/ D .yx/ D .xy/ D .x/.y/ for all y 2 G; hence .x/ is
equivariant, and it follows that
.x/ D .x/I;
(33)

for some .x/ 2 C; by Schurs lemma. It follows from this that every linear subspace of V is
invariant. By irreducibility of this implies that the dimension of V must be one. From the fact
that is a representation it follows immediately that x 7! .x/ is a character. Applying the trace
94

b of
to (33) we see that  D  ; the character of : Thus 7!  induces a map from the space G
equivalence classes of finite dimensional irreducible representations to the set of multiplicative
characters of G: This map is injective by Corollary 22.13. If  is a multiplicative character then
(33) defines an irreducible representation of G in C; and  D  : Therefore the map !  is
surjective onto the set of multiplicative characters.

Corollary 27.3 Assume that G is a commutative compact Lie group. Then the set of multiplicab is a complete orthonormal system for L2 .G/:
tive characters  ; 2 G;
Proof: This follows immediately from the previous lemma combined with the theorem of Peter
and Weyl (Theorem 23.5).

by

b ! C of a function f 2 L2 .G/
In the present setting we define the Fourier transform fO W G
fO./ D hf ;  i:

b be equipped with the counting measure. Then the associated L2 -space is l 2 .G/;
b the space
Let G
P
2
b ! C such that
of functions ' W G
b j'./j < 1; equipped with the inner product:
2G
X
h' ; i WD
'./ ./:
b
2G

Corollary 27.4 (The Plancherel theorem). Let G be a commutative compact Lie group. Then
b Moreover, if f 2 L2 .G/;
the Fourier transform f 7! fO is an isometry from L2 .G/ onto l 2 .G/:
then
X
f D
fO./  ;
with convergence in the L2 -sense.

b
2G

Proof: Exercise for the reader.

If in addition it is assumed that the group G is connected, then G ' .R=Z/n for some n 2 N;
see Theorem 6.1. The purpose of the following exercise is to accordingly view the classical
theory of Fourier series as a special case of the Peter-Weyl theory.
Exercise 27.5 Let G D Rn =2Zn : If m 2 Zn ; show that
m W x 7! e i.mx/
defines a multiplicative character of G: (Here m  x D m1 x1 C    C mn xn :) Show that every
b ' Zn : Accordingly for f 2 L2 .G/ we view the
multiplicative character is of this form. Thus G
Fourier transform fO as a map Zn ! C:
Show that the normalized Haar integral of G is given by
Z 2
Z 2
1
:::
f .x1 ; : : : ; xn / dx1 : : : dxn :
I.f / D
.2/n 0
0
95

Show that for f 2 L2 .G/; m 2 Zn we have:


Z 2
Z 2
1
O
:::
f .x1 ; : : : ; xn / e
f .m/ D
.2/n 0
0
Moreover, show that we have the inversion formula
X
f .x/ D
fO.m/ e i.mx/
m2Zn

i.m1 x1 CCmn xn /

dx1 : : : dxn :

.x 2 Rn =2Zn /

in the L2 -sense.

28 The group SU(2)


In this section we assume that G is the compact Lie group SU.2/:
Recall the definition of the representation n of SU.2/ in the space Vn D Pn .C2 / of homogeneous polynomials of degree n from Section 20. In Proposition 20.28 it was shown that n is
irreducible. Moreover, the associated character is determined by the formula:
n .t' / D

sin.n C 1/'
sin '

.' 2 R/

(34)

(see Exercise 22.9). The purpose of this section is to prove the following result:
Proposition 28.1 Every finite dimensional irreducible representation of SU.2/ is equivalent to
n ; for some n 2 N:
We recall that every element of SU.2/ is conjugate to an element of T D ft' j ' 2 Rg: Therefore
a class function f on SU.2/ is completely determined by its restriction f jT to T: For every
' 2 R the diagonal matrices t' and t' 1 D t ' are conjugate. Therefore, the restriction f jT
is invariant under the substitution t 7! t 1 : Thus, if C.T /ev denotes the space of continuous
functions g W T ! C satisfying g.t 1 / D g.t / for all t 2 T; then restriction to T defines an
injective linear map r W C.G; class/ ! C.T /ev :
Lemma 28.2 The map r W C.G; class/ ! C.T /ev is bijective. Moreover, r is isometric, i.e., it
preserves the sup-norms.
Proof: That r is isometric follows from the observation that the set of values of a function
f 2 C.G; class/ is equal to the set of values of its restriction r.f /: Thus it remains to establish
the surjectivity of r: Let g 2 C.T /ev : Then g.t' / D g.e
Q i' / for a unique continuous function
1
gQ W T ! C satisfying g.z/
Q
D g.z
Q
/: Now g.z/
Q
D G.Rez/ for a unique continuous function
G W 1; 1 ! C: It follows that g.t' / D G.cos '/; for ' 2 R:
An element x 2 SU.2/ has two eigenvalues z.x/ and z.x/ 1 ; with jz.x/j D 1: Clearly
x 7! Rez.x/ is a well defined continous function on SU.2/:
Define f .x/ D G.Rez.x//: Then f is a well defined continuous class function. Moreover,
f .t' / D G.Ree i' / D g.t' /; hence r.f / D g:

96

Corollary 28.3 The linear span of the characters n; for n 2 N; is dense in C.G; class/:
Proof: By Lemma 28.2 it suffices to show that the linear
Pn span S of the functions n jT is dense
in C.T /ev : From formula (34) we see that n .t' / D kD0 e i.n 2k/' : Hence S equals the linear
span of the functions n W t' 7! e i n' C e i n' D 2 cos n', .n 2 N/: The latter span is dense in
C.T /ev ; by the classical theory of Fourier series.

Corollary 28.4 Let f 2 C.G; class/: If f ? n for all n 2 N; then f D 0:
b this follows from the Peter-Weyl theorem.
Remark 28.5 Once we know that the n exhaust G

Proof: We first note that, for g 2 C.G/;


kgk2L2

jg.x/j2 dx  kgk21 ;

where k  k1 denotes the sup norm. Using this estimate we see that the linear span of the
characters n is dense in C.G; class/ with respect to the L2 -norm. Thus, if f 2 C.G; class/ is
perpendicular to all n ; then it follows that f ? C.G; class/: In particular, kf k22 D hf ; f i D 0;
which implies that f D 0:

Corollary 28.6 Every finite dimensional irreducible representation of SU.2/ is equivalent to
one of the n ; n 2 N:
b such that is not equivalent to n ; for every
Proof: Suppose not. Then there exists a 2 G
n 2 N: Hence the class function  is perpendicular to n for every n 2 N: This implies that
 D 0: This is impossible, since  .e/ D dim./  1:

From the fact that every element of SU.2/ is conjugate to an element of T one might expect
that there should exist a Jacobian J W T ! 0; 1 such that for every continuous class function
f on SU.2/ we have
Z
Z
2

SU.2/

f .x/ dx D

f .t' / J.t' / d':

It is indeed possible to compute this Jacobian by a substitution of variables. However, we shall


obtain the above integration formula by other means.
Lemma 28.7 For every continuous class function f W SU.2/ ! C we have:
Z

SU.2/

f .x/ dx D

2

97

f .t' /

sin2 '
d':


(35)

Proof: Consider the linear map L which assigns to f 2 C.G; class/ the expression on the lefthand side minus the expression on the right-hand side of the above equation. Then we must show
that L is zero.
Obviously the linear functional L W C.G; class/ ! C is continuous with respect to the sup
norm. Hence by density of the span of the characters it suffices to show that L.n / D 0 for every
n 2 N: The function 0 is identically one; therefore left- and right-hand side of (35) both equal
1 if one substitutes f D 0 : Hence L.0 / D 0: On the other hand, if n  1; and f D n ; then
the left hand side of (35) equals hn ; 0 i D 0: The right hand side of (35) also equals 0; hence
L.n / D 0 for all n:

Corollary 28.8 Let f 2 C.G/: Then
Z

f .x/ dx D

2

f .xt' x
G

sin2 '
d'
/ dx


Remark 28.9 The interpretation of the above formula is that the integration over G D SU.2/
may be split into an integration over conjugacy classes, followed by an integration over the circle
group T:
Proof: Put
F .y/ D

f .xyx

/ dx:

Then by bi-invariance of the Haar measure, F is a continuous class function. Hence by the
previous result
Z
Z 2 Z
sin2 '
F .y/ dy D
f .xt' x 1 / dx
d':

G
0
G
On the other hand,
Z
Z Z
F .y/ dy D
f .xyx 1 / dx dy
G
ZG ZG
D
f .xyx 1 / dy dx;
G

by Fubinis theorem. By bi-invariance of the Haar measure, the inner integral is independent of
x: Therefore,
Z
Z Z
F .y/ dy D
f .y/ dy dx
G
G
G
Z
D
f .y/ dy:
G

This completes the proof.

98

We end this section with a description of all irreducible representations of SO.3/: From Section 10 we recall that there exists a surjective Lie group homomorphism ' W SU.2/ ! SO.3/
with kernel ker ' D f I; I g: Accordingly, SO.3/ ' SU.2/=fI g (Thm. 17.4).
Proposition 28.10 For k 2 N the representation 2k of SU.2/ factors through a representation
N 2k of SO.3/ ' SU.2/=fI g: The representations N 2k are mutually inequivalent and exhaust
\
SO.3/:
Proof: One readily verifies that 2k .x/ D I for x 2 fI g: Hence 2k factors through a
representation N 2k of SO.3/: Every invariant subspace of the representation space V2k of 2k is
2k .SU.2// invariant if and only if it is N 2k .SO.3// invariant. A non-trivial SO.3/-equivariant
map V2k ! V2l would also be SU.2/-equivariant. Hence the N 2k are mutually inequivalent.
\ assume that .; V / is an irreducible
Finally, to see that the representations N 2k exhaust SO.3/;

representation of SO.3/: Then '  WD  ' is an irreducible representation of SU.2/; hence
equivalent to some n ; n 2 N: From '   D I on ker ' it follows that n D I on fI g; hence
n is even.


99

29 Lie algebra representations


Let V be a finite dimensional complex linear space. If  is a continuous representation of G
in V; then  is a (smooth) Lie group homomorphism G ! GL.V /; in view of Corollary 9.3.
Accordingly, the tangent map  W g ! End.V / at e is a homomorphism of Lie algebras.
Thus,  is a representation of g in V: In other words, we see that a finite dimensional Gmodule V automatically is a g-module (see Remark 20.6 and the text preceding the remark for
the terminology used here).
By the chain rule one readily sees that

d
.exp tX/v;
(36)
 .X/v D
dt t D0
for v 2 V and X 2 g: On the other hand, it follows from Lemma 4.16 that for all X 2 g we
have:
.exp X/ D e  .X/ :
(37)

When G is connected this equation allows us to compare the G- and the g-module structures on
V: When there is no chance of confusion, we will omit the star in the notation of the representation
of g in V:
Lemma 29.1 Assume that G is connected, and let V; V 0 be two finite dimensional G-modules.
(a) Let W be a linear subspace of V: Then W is G-invariant if and only if W is g-invariant.
(b) The G-module V is irreducible if and only V is irreducible as a g-module.
(c) Let T W V ! V 0 be a linear map. Then T is G-equivariant if and only if T is g-equivariant.

(d) V and V 0 are isomorphic as G-modules if and only if they are isomorphic as g-modules.

Proof: Write  and  0 for the representations of G in V and V 0 respectively. As agreed, we


denote the associated representations of g in V and V 0 by the same symbols, i.e., we omit the
stars in the notation.
(a): If W is g-invariant, then it follows from (37) that W is invariant under the group Ge
which is generated by exp g: But Ge D G; since G is connected. The converse implication is
proved by differentiating .exp.tX// at t D 0:
(b): This is now an immediate consequence of (a).
(c): Suppose that T is g-equivariant. Then for all X 2 g we have:  0 .X/ T D T .X/;
hence  0 .X/n T D T .X/n for all n 2 N; and since T is continuous linear it follows that
0

e  .X/ T D T e .X/ :
From this it follows that  0 .x/ T D T .x/ for all x 2 exp g; and hence for x 2 Ge D G:
The reverse implication follows by a straightforward differentiation argument as in part (a) of
this proof.
(d): This follows immediately from (c).

100

Lemma 29.2 Let G be a connected compact Lie group, and let  be a representation of G in a
finite dimensional Hilbert space V: Then  is unitary if and only if
.X/ D

.X/

(38)

for all X 2 g:
Proof: We recall that  W G ! GL.V / is a Lie group homomorphism. Hence for all X 2 g; t 2
R we have:
.exp tX/ D e t .X/ :
If  is unitary, then .exp tX/ D .exp. tX//; hence


e t .X/ D e

t .X/

(39)

Differentiating this relation at t D 0 we find (38). Conversely, if (38) holds, then (39) holds for
all X; t and it follows that .x/ is unitary for x 2 exp g: This implies that .x/ is unitary for
x 2 Ge D G:

It will turn out to be convenient to extend representations of g to its complexification gC :
If E is a real linear space, its complexification EC is defined as the real linear space E R C;
equipped with the complex scalar multiplication .v z/ D v z: We embed E as a real
linear subspace of EC by the map v 7! v 1: Then EC D E iE as a real linear space. In terms
of this decomposition, the complex scalar multiplication is given in the obvious fashion. If g is a
real Lie algebra, then its complexification gC is equipped with the complex bilinear extension of
the Lie bracket. Thus, gC is a complex Lie algebra.
Any representation  of g in a complex vector space V has a unique extension to a (complex)
representation of gC in V I this extension, denoted C ; is given by
C .X C iY / D .X/ C i.Y /;
for X; Y 2 g:
Lemma 29.3 Let V; V 0 be g-modules, and let W  V a (complex) linear subspace, and T W
V ! V 0 a (complex) linear map.
(a) The space W is g-invariant if and only if it is gC -invariant.
(b) V is irreducible as a g-module if and only if it is so as a gC -module.
(c) T is g-equivariant if and only if it is gC -equivariant.
(d) V and V 0 are isomorphic as g-modules if and only if they are isomorphic as gC -modules.


Proof: Left to the reader.

101

Example 29.4 The Lie algebra su.2/ of SU.2/ consists of complex 22 matrices A 2 M.2; C/;
satisfying trA D 0 and A D A: It follows from this that i su.2/ is the real linear subspace
of M.2; C/ consisting of matrices A with trA D 0 and A D A: In particular, we see that
su.2/ \ i su.2/ D f0g: Therefore, the embedding su.2/ ,! M.2; C/ extends to a complex linear
embedding
j W su.2/C ,! M.2; C/:

Clearly, the image of j is contained in the Lie algebra of SL.2; C/; which is given by
sl.2; C/ D fA 2 M.2; C/ j trA D 0g:

On the other hand, if A 2 sl.2; C/; then 12 .A A / belongs to su.2/ and 21 .A C A / belongs
to i su.2/I summing these elements, we see that A 2 j.su.2/C /: Therefore, j is an isomorphism
from su.2/C onto sl.2; C/; via which we shall identify from now on.

30 Representations of sl(2,C)
It follows from the discussion in the previous section that the SU.2/-module Pn .C2 /; for n 2
N; carries a natural structure of sl.2; C/-module. The associated representation of sl.2; C/ in
Pn .C2 / equals .n /C ; the complexification of n : We shall now compute this structure in
terms of the basis p0 ; : : : ; pn of Pn .C2 / given by
pj .z/ D z1j z2n

.z 2 C2 /:

Let p 2 Pn .C2 /: Then we recall that, for x 2 SU.2/; n .x/p.z/ D p.x


follows from this that, for  2 su.2/;

d
t
p.e z/
n ./p.z/ D
;
dt
t D0

z/; z 2 C2 : It

hence, by the chain rule

n ./p.z/ D

@p
@p
.z/. z/1 C
.z/. z/2 :
@z1
@z2

The expression on the right-hand side is complex linear in I hence it also gives p D .n /C ./p
for  2 sl.2; C/: Thus, we obtain, for  2 sl.2; C/ and p 2 Pn .C2 /;
p D

.z/1

@
@
C .z/2
p:
@z1
@z2

(40)

We shall now compute the action of the basis H; X; Y of sl.2; C/ given by








0 0
0 1
1
0
:
; Y D
; XD
H D
1 0
0 0
0
1
By a straightforward computation we see that
H; X D 2X;

H; Y D
102

2Y;

X; Y D H:

(41)

Definition 30.1 Let l be a Lie algebra. By a standard sl.2/-triple in l we mean a collection of


linear independent elements H; X; Y 2 l satisfying the relations (41).
Remark 30.2 Let l be a complex Lie algebra. Then the complex linear span of an sl.2/-triple
in l is a Lie subalgebra isomorphic to sl.2; C/:
Substituting H; X and Y for  in (40), we obtain, for p 2 Pn .C2 /;
H p D z1

@
@
C z2
p;
@z1
@z2

XpD

z2

@
p;
@z1

YpD

z1

@
p:
@z2

(42)

By a straightforward computation we now see that the action of the triple H; X; Y on the basis
element pj is given by
Hpj D .n

2j /pj ;

Xpj D

jpj

1;

Ypj D .j

n/pj C1 :

For the matrices of the action of H; X; Y on Pn.C2 / relative to the basis p0 ; : : : ; pn we thus find
0
1
n
0
::: 0
:: C
B
: C
B 0 n 2
mat.H / D B :
C;
:
: : ::: A
@ ::
0 ::: :::
n
and
0

B
B
B
mat.X/ D B
B
@

0
0
::
:

1
0

0 :::
2 :::
:: ::
:
:
0
0 :::
:::

0
0
::
:

C
C
C
C
C
n A
0

B
B
B
mat.Y / D B
B
@

1
::: 0
:: C
n
0
: C
C
0 1 n 0
C:
::
: : : : :: C
:
: : A
:
0
:::
0
1 0
0

:::

These matrices will guide us through the proof of the following theorem.
Theorem 30.3 Every irreducible finite dimensional sl.2; C/-module is isomorphic to Pn .C2 /;
for some n 2 N:
Remark 30.4 From the above theorem we deduce again, using Lemmas 29.1 and 29.3, that
every irreducible continuous finite dimensional representation of SU.2/ is equivalent to n ; for
some n 2 N:
The proof of the the above theorem will be given in the rest of this section. Let V be an
irreducible finite dimensional sl.2; C/-module.
Given  2 C; we shall write V WD ker.H I /: This space is non-trivial if and only if  is
an eigenvalue for the action of H on V:
103

Lemma 30.5 Let  2 C: Then


XV  VC2 ;

Y V  V 2 :

Proof: Let v 2 V : Then HXv D XH v C H; Xv D Xv C 2Xv D . C 2/Xv; hence


Xv 2 VC2 : This proves the first inclusion. The second inclusion is proved in a similar manner.

By a primitive vector of V we mean a vector v 2 V n f0g with the property that Xv D 0: The
idea behind this definition is to get hold of the analogue of p0 2 Pn.C2 /:
Lemma 30.6 V contains a primitive vector that is an eigenvector for H:
Proof: Let  be an eigenvalue of the action of H on V: Fix an eigenvector w 2 V ; w 0
and consider the sequence of vectors wk ; k  0; defined by w0 D w and wkC1 D Xwk :
Then wk 2 VC2k : If all vectors wk were non-zero, then they would be eigenvectors for different
eigenvalues of H; hence they would be linear independent, contradicting the finite dimensionality
of V: It follows that there exists a largest k such that wk 0: The vector wk is primitive.

In the following we assume that v 2 V is a fixed primitive vector that is an eigenvector for H:
The associated eigenvalue is denoted by : We now consider the vectors vk defined by v0 D v
and vkC1 D Y vk : By a similar reasoning as in the above proof it follows that there exists a largest
number n such that vn 0:
Lemma 30.7
(a) The vectors vk D Y k v; 0  k  n; form a basis for V:
(b) The eigenvalue  equals n D dimV

1:

(c) For every 0  k  n;


H vk D .

Xvk D k.

2k/vk ;

k C 1/vk 1 :

(d) The primitive vectors in V are the non-zero multiples of v0 :


Proof: We first prove (c) for all k 2 N (but note that vk D 0 for k > n). It follows from repeated
application of Lemma 30.5 that vk 2 V 2k ; hence H vk D . 2k/vk : We prove the second
assertion of (c) by induction. Since v0 D v is primitive, the second assertion of (c) holds for
k D 0: Let now k > 0 and assume that the assertion has been established for strictly smaller
values of k: Then
Xvk D
D
D
D
D

XY vk 1
YXvk 1 C X; Y vk 1
YXvk 1 C H vk 1
.k 1/. .k 2//Y vk
k. k C 1/vk 1
104

C .

2.k

1//vk

and (c) follows.


Let W be the linear span of the vectors vk ; for 0  k  n: Then by definition of the vectors
vk ; Y vk D vkC1 : Therefore, Y leaves W invariant. By (c), H and X leave W invariant as well.
It follows that W is a non-trivial invariant subspace of V; hence V D W by irreducibility. The
vectors vk ; for 0  k  n; must be linear independent since they are eigenvectors for H for
distinct eigenvalues; hence (a).
Finally, we have established the second assertion of (c) for all k  0; in particular for k D
n C 1: Now vnC1 D 0; hence 0 D .n C 1/. n/vn and since vn 0 it follows that  D n: This
establishes (b).
It follows from (a) and (c) that the only primitive vectors in V are non-zero multiples of v0 :

Corollary 30.8 Let V and V 0 be two irreducible finite dimensional sl.2; C/-modules. Then
V ' V 0 if and only if dimV D dimV 0 : Moreover, if v and v 0 are primitive vectors of V and V 0 ;
respectively, then there is a unique isomorphism T W V ! V 0 mapping v onto v 0 :
Proof: Clearly if V ' V 0 then V and V 0 have equal dimension. Conversely, assume that
dimV D dimV 0 D n and that v and v 0 are primitive vectors of V and V 0 respectively. Then by
the above lemma, the vectors vk D Y k v; 0  k  n form a basis of V: Similarly the vectors
vk0 D Y k v 0 ; 0  k  n form a basis of V 0 : Any intertwining operator T W V ! V 0 that maps v
onto v 0 must map the basis vk onto the basis vk0 ; hence is uniquely determined. Let T W V ! V 0
be the linear map determined by T vk D vk0 ; for 0  k  n: Then T is a linear bijection.
Moreover, by the above lemma we see that T intertwines the actions of H; X; Y on V and V 0 : It
follows that T is equivariant, hence V ' V 0 :

Completion of the proof of Theorem 30.3: The space Pn .C2 / is an irreducible sl.2; C/module, of dimension n C 1: Hence if V is an irreducible sl.2; C/-module of dimension m  1;
then V ' Pn .C2 /; with n D m 1:


31 Roots and weights


Let t be a finite dimensional commutative real Lie algebra, and let .; V / be a finite dimensional
representation of t in V:
Let tC denote the space of complex linear functionals on tC : Note that t ; the space of real
linear functionals on t may be identified with the space of  2 tC that are real valued on t: Thus,
t is viewed as a real linear subspace of tC : Accordingly i t equals the space of  2 tC such that
jt has values in i R:
If  2 tC ; then we define the following subspace of V W
\
V D
ker..H / .H /I /:
(43)
H 2t

105

In other words, V consists of the space of v 2 V such that .H /v D .H /v for all H 2 t: If
V 0; then  is called a weight of t in V; and V is called the associated weight space. The set
of weights of t in V is denoted by ./:
Lemma 31.1 Let T 2 End.V / be a -intertwining linear endomorphism, then T leaves V
invariant, for every  2 ./:
Proof: Let  2 ./: The endomorphism T commutes with .H / hence leaves the eigenspace
ker..H / .H // invariant, for every H 2 t: Hence T leaves the intersection V of all these
spaces invariant.

Lemma 31.2 The set ./ is a non-empty finite subset of tC : Assume that .X/ is diagonalizable for every X 2 t: Then
M
V D
V :
(44)
2./

Moreover, if W is a t-invariant subspace of V; then W is the direct sum of the spaces W \ V ;


for  2 ./:
Proof: Fix a basis X1 ; : : : ; Xn of t: The endomorphism .X1 / has at least one eigenvalue, say
1 ; with corresponding eigenspace E1  V: Since t is commutative, this eigenspace is invariant
under the action of t: Proceeding by induction on dimt; we obtain a sequence of non-trivial
subspaces En  En 1      E1 such that Xj acts by a scalar j on Ej ; for each 1  j  n:
Define  2 tC by .Xj / D j ; then En  V ; hence  2 ./: This establishes the first
assertion.
If .X/ diagonalizes, for every X 2 t; then, in particular, V admits a decomposition of
eigenspaces for the endomorphism .X1 /: Each of these eigenspaces is invariant under t: Therefore, by induction on dimt there exists a direct sum decomposition V D V1    VN such that
Xj acts by a scalar ij on Vi ; for all 1  i  N and 1  j  n: Let i 2 tC be defined by
i .Xj / D ij ; for 1  i  N: Then ./ D f1 ; : : : ; N g: Moreover, one readily verifies that,
for  2 ./; V D j WDj Vj : Hence, (44) follows.
For the final assertion, we observe that by finite dimensionality of V the set ./ is finite.
Hence, there exists a X0 2 t such that . /.X0 / 0 for all ;  2 ./ with  : For
 2 ./; let P W V ! V be the projection along the remaining summands in (44). We claim
that
Y
P D
..X0 / .X0 // 1 ..X0 / .X0 //:
2./nfg

Indeed this is readily checked on each of the summands V of the decomposition in (44), for
 2 ./:
It follows from the above formula for P that P .W /  W: Hence, P .W /  W \ V ; and
the final assertion follows.


106

Assumption: In the rest of this section we assume that G is a compact Lie group, with Lie
algebra g:
Definition 31.3 A torus in g is by definition a commutative subalgebra of g: A torus t  g is
called maximal if there exists no torus of g that properly contains t:
From now on we assume that t is a fixed maximal torus in g:
Lemma 31.4 The centralizer of t in g equals t:
Proof: Since t is abelian, it is contained in its centralizer. Conversely, assume that X 2 g
centralizes t: Then t0 D t C RX is a torus which contains t: Hence t0 D t by maximality, and we
see that X 2 t:

Let .; V / be a finite dimensional representation of gC ; the complexification of the Lie algebra gI i.e.,  is a complex Lie algebra homomorphism from gC into End.V / (the latter is the space
of complex linear endomorphisms equipped with the commutator Lie bracket). Alternatively we
will also say that V is a finite dimensional gC -module. We denote by . / D . ; t/ the set
of weights of the representation  D jt of t in V: If  2 tC ; then as before, V is defined as in
(43), with jt in place of : Thus
V D fv 2 V j .H /v D .H /v

for all

H 2 tg:

From Lemma 31.2 we see that . / is a non-empty finite subset of tC :
Let .; V / be a finite dimensional continuous representation of G: Then the map  W G !
GL.V / is a homomorphism of Lie groups. Let  D Te : Then  W g ! End.V / is a Lie
algebra homomorphism, or, differently said, a representation of g in V: The homomorphism
 has a unique extension to a complex Lie algebra homomorphism from gC into End.V / (we
recall that V is a complex linear space by assumption). This extension is called the induced
infinitesimal representation of gC in V:
Lemma 31.5 Let  be a finite dimensional continuous representation of G: Then . / is a
finite subset of i t : Moreover,
M
V D
V :
2. /

If V is equipped with a G-invariant inner product, then for all ;  2 . / with   we
have V ? V :
Proof: There exists a G-invariant inner product on V I assume such an inner product h ; i to be
fixed. Then  maps G into U.V /; the associated group of unitary transformations. It follows
that  maps g into the Lie algebra u.V / of U.V /; which is the subalgebra of anti-Hermitian
endomorphisms in End.V /: It follows that for X 2 g the endomorphism  .X/ is anti-Hermitian,
hence diagonalizable with imaginary eigenvalues. The proof is now completed by application of
Lemma 31.2.

107

If A 2 End.g/; then we denote by AC the complex linear extension of A to gC : Obviously the


map A 7! AC induces a real linear embedding of End.g/ into End.gC / WD EndC .gC /: Accordingly we shall view End.g/ as a real linear subspace of the complex linear space End.gC / from
now on. Thus, we may view Ad as a representation of G in the complexification gC of g: The
associated infinitesimal representation is the adjoint representation ad of gC in gC : The associated collection . ad/ of weights contains the weight 0: Indeed the associated weight space gC0
equals the centralizer of t in gC ; which in turn equals tC ; by Lemma 31.4. Hence:
gC0 D tC :
Definition 31.6 The weights of ad in gC different from 0 are called the roots of t in gC I the set
of these is denoted by R D R.gC ; t/: Given 2 R; the associated weight space gC is called a
root space.
It follows from the definitions that
gC D fX 2 gC j H; X D .H /X

for all

H 2 tg:

From Lemma 31.5 we now obtain the so called root space decomposition of gC ; relative to the
torus t:
Corollary 31.7 The collection R D R.gC ; t/ of roots is a finite subset of i t : Moreover, we have
the following direct sum of vector spaces:
M
gC D tC
gC :
(45)
2R

Example 31.8 The Lie algebra g D su.2/ has complexification sl.2; C/; consisting of all complex 2  2 matrices with trace zero. Let H; X; Y be the standard basis of sl.2; C/I i.e.






0 0
0 1
1
0
:
;
Y D
;
XD
H D
1 0
0 0
0
1
Now t D i RH is a maximal torus in su.2/: We recall that H; X D 2X; H; Y D 2Y;
X; Y D H: Thus, if we define 2 tC by .H / D 2; then R D R.gC ; t/ equals f; g:
Moreover, gC D CX and gC. / D CY:
We recall that, by definition, the center z D zg of g is the ideal ker adI i.e., it is the space of
X 2 g that commute with all Y 2 g:
Lemma 31.9 The center of g is contained in t and equals the intersection of the root hyperplanes:
\
zg D
ker :
2R

In particular, if zg D 0; then R spans the real linear space i t :


108

Proof: The center of g centralizes t in particular, hence is contained in t; by Lemma 31.4. Let
H 2 t and assume that H centralizes gI then H centralizes gC ; hence every root space of gC : This
implies that .H / D 0 for all 2 R: Conversely, if H 2 t is in the intersection of all the root
hyperplanes, then H centralizes tC and every root space gC : By the root space decomposition it
then follows that H 2 z: This establishes the characterization of the center.
If z D 0; then the root hyperplanes ker . 2 R/ have a zero intersection in t: This implies
that the set R  i t spans the real linear space i t :

Lemma 31.10 Let .; V / be a finite dimensional representation of gC : Then for all  2 ./
and all 2 R [ f0g we have:
.gC /V  VC :
In particular, if  C ./; then .gC / anihilates V :
Proof: Let X 2 gC and v 2 V : Then, for H 2 t;
.H /.X/v D .X/.H /v C .H /; .X/v
D .H /.X/v C .H; X/v D .H / C .H /.X/v:
Hence .X/v 2 VC : If  C is not a weight of ; then VC D 0 and it follows that
.X/v D 0:

Corollary 31.11 If ; 2 R [ f0g; then
gC ; gC  gC.C / :
In particular, if C R [ f0g; then gC and gC commute.
Proof: This follows from the previous lemma applied to the adjoint representation.

P We shall write ZR for the Z-linear span of R; i.e., the Z-module of elements of the form
2R n ; with n 2 Z:
In the following corollary we do not assume that  comes from a representation of G:
Corollary 31.12 Let .; V / be a finite dimensional representation of gC : Then
M
W WD
V

(46)

2./

is a non-trivial gC -submodule. If  is irreducible, then W D V: Moreover, if ;  2 ./; then


  2 ZR:
Proof: By Lemma 31.2 the set ./ is non-empty and finite, and therefore W is a non-trivial
subspace of V: From Lemma 31.10 we see that W is gC -invariant. If  is irreducible, then
W D V: To establish the last assertion we define an equivalence relation on ./ by  
 ()   2 ZR: If S is a class for ; then VS D 2S V is a non-trivial gC -invariant
subspace of V; by Lemma 31.10. Hence VS D V and it follows that S D ./:

109

Remark 31.13 If g has trivial center, then the above result actually holds for every finite dimensional V -module. To see that a condition like this is necessary, consider g D R; the Lie algebra
of the circle. Define a representation of g in V D C2 by


0 x
:
.x/ D
0 0
Then ./ D f0g; but V0 D C  f0g is not all of V:
Note that this does not contradict the conclusion of Lemma 31.5, since  is not associated
with a continuous representation of the circle group in C2 :
Lemma 31.14 Let t be a maximal torus in g; and R the associated collection of roots. If 2 R
then 2 R:
Proof: Let  be the conjugation of gC with respect to the real form g: That is: .X C iY / D
X iY for all X; Y 2 g: One readily checks that  is an automorphism of gC ; considered as a
real Lie algebra (by forgetting the complex linear structure). Let 2 R; and let X 2 gC : Then
for every H 2 t;
H; .X/ D H; X D ..H /X/ D .H /.X/ D

.H /.X/:

For the latter equation we used that has imaginary values on t: It follows that 2 R and that
 maps gC into gC (in fact is a bijection between these root spaces; why?).

We recall that we identify i t with the real linear subspace of tC consisting of  such that jt
has values in i RI the latter condition is equivalent to saying that jit is real valued. One readily
verifies that the restriction map  7! jit defines a real linear isomorphism from i t onto the
real linear dual .i t/ : In the following we shall use this isomorphism to identify i t with .i t/ :
Now R is a finite subset of .i t/ n f0g: Hence the complement of the hyperplanes ker  i t;
for 2 R is a finite union of connected components, which are all convex. These components
are called the Weyl chambers associated with R: Let C be a fixed chamber. By definition every
root is either positive or negative on C: We define the system of positive roots RC WD RC .C/
associated with C by
RC D f 2 R j > 0 on Cg:
By what we said above, for every 2 R; we have that either or
both. It follows that
R D RC [ . RC / (disjoint union).

belongs to RC ; but not


(47)

We writePNRC for the subset of ZR consisting of the elements that can be written as a sum
of the form 2RC n ; with n 2 N:

Lemma 31.15 NRC \ . NRC / D 0:

Proof: Let  2 NRC : Then   0 on C; the chamber corresponding to RC : If also  2 NRC ;


then   0 on C as well. Hence  D 0 on C: Since C is a non-empty open subset of i t ; this
implies that  D 0:

110

Lemma 31.16 The spaces


gC
C WD

gC WD

gC ;

2RC

gC

2 RC

are ad.t/-stable subalgebras of gC : Moreover,


gC D gC
C tC gC :
Proof: Let ; 2 RC and assume that gC ; gC 0: Then C 2 R [ f0g; and C > 0
C
on C: This implies that C 2 RC ; hence gC.C /  gC
C : It follows that gC is a subalgebra. For
similar reasons gC is a subalgebra. Both subalgebras are ad.t/ stable, since root spaces are. The
direct sum decomposition is an immediate consequence of (45) and (47).

We are now able to define the notion of a highest weight vector for a finite dimensional gC module, relative to the system of positive roots RC : This is the appropriate generalization of the
notion of a primitive vector for sl.2; C/:
Definition 31.17 Let V be a finite dimensional gC -module. Then a highest weight vector of V
is by definition a non-trivial vector v 2 V such that
(a) tC v  CvI

(b) Xv D 0 for all X 2 gC


C:
Lemma 31.18 Let V be a finite dimensional gC -module. Then V has a highest weight vector.
Proof: We define the gC -submodule W of V as the sum of the tC -weight spaces, see Corollary
31.12.
Let C be the positive chamber determining RC : Fix X 2 C: Then .X/ > 0 for all 2 RC :
We may select 0 2 ./ such that the real part of .X/ is maximal. Then 0 C ./ for
all 2 RC : By Lemma 31.10 this implies that  .gC /V  V0 C D 0 for all 2 RC : Hence
gC

C annihilates V0 : Thus, every non-zero vector of V0 is a highest weight vector.
Definition 31.19 Let V be a finite dimensional gC -module. A vector v 2 V is called cyclic if it
generates the gC -module V; i.e., V is the smallest gC -submodule containing v:
Obviously, if V is irreducible, then every non-trivial vector is cyclic.
Proposition 31.20 Let V be a finite dimensional gC -module and let v 2 V be a cyclic highest
weight vector.
(a) There exists a (unique)  2 .V / such that v 2 V : Moreover, V D Cv:
(b) The space V is equal to the span of the vectors v and .X1 /    .Xn /v; with n 2 N and
Xj 2 gC ; for 1  j  n:
(c) Every weight  2 .V / is of the form 

; with  2 NRC :

111

(d) The module V has a unique maximal proper submodule W:


(e) The module V has a unique non-trivial irreducible quotient.
Proof: The first assertion of (a) follows from the definition of highest weight vector. We define
an increasing sequence of linear subspaces of V inductively by V0 D Cv and VnC1 D Vn C
.gC /Vn : Let W be the union of the spaces Vn : We claim that W is an invariant subspace of
V: To establish the claim, we note that by definition we have .gC /Vn  VnC1 I hence W is gC
invariant. The space V0 is t- and gC
C -invariant; by induction we will show that the same holds for
Vn : Assume that Vn is t- and gC
-invariant,
and let v 2 Vn ; Y 2 gC : Then for H in t we have
C
H Y v D Y H v C H; Y v: Now v 2 Vn and by the inductive hypothesis it follows that H v 2 Vn :
Hence Y H v 2 VnC1 : Also H; Y 2 gC and it follows that H; Y v 2 VnC1 : We conclude that
H Y v 2 VnC1 : It follows from this that
.t/.gC /Vn  VnC1 :
Hence VnC1 is t-invariant.
Let now v 2 Vn ; Y 2 gC and X 2 gC
C : Then XY v D YXv C X; Y v: Now Xv 2 Vn
by the induction hypothesis and we see that YXv 2 VnC1 : Also, X; Y 2 gC : By the induction
hypothesis it follows that gC Vn  VnC1 : Hence X; Y v 2 VnC1 : We conclude that XY v 2 VnC1 :
It follows from this that
.gC
C /.gC /Vn  VnC1 :

Hence VnC1 is gC
C -invariant. This establishes the claim that W is a gC -invariant subspace of V:
Since W contains the cyclic vector v; it follows that W D V: Hence, (b) follows. Let
w D .Y1 /    .Yn /v; with n 2 N; Yj 2 gC. j / ; j 2 RC : Then w belongs to the weight
space V  ; where  D 1 C    C n 2 NRC : Since v and such elements w span W D V;
we conclude that every weight  in .V / is of the form   with  2 NRC : This establishes
(c). Moreover, it follows from the above description that V equals the vector sum of Cv and
V ; where V denotes the sum of the weight spaces V with  2 .V / n fg: This implies that
V D Cv; whence the second assertion of (a).
We now turn to assertion (d). Let U be a submodule of V: In particular, U is a tC -invariant
subspace. Let .U / be the collection of  2 .V / for which U WD U \ V 0: In view of
Lemma 31.2, U is the direct sum of the spaces U ; for  2 .U /: If U is a proper submodule,
then U D 0; hence .U /  .V / n fg: It follows that the vector sum W of all proper
submodules satisfies .W /  .V / n fg hence is still proper. Therefore, V has W as unique
maximal submodule.
The final assertion (e) is equivalent to (d). To see this, let p W V ! V 0 be a surjective gC module homomorphism onto a non-trivial gC -module. Then U 7! p 1 .U / defines a bijection
from the collection of proper submodules of V 0 onto the collection of proper submodules of
V containing ker p: It follows that V 0 is irreducible if and only if ker p is a proper maximal
submodule of V: The equivalence of (d) and (e) now readily follows.

Corollary 31.21 Let V be a finite dimensional irreducible gC -module. Then V has a highest
weight vector v, which is unique up to a scalar factor. Let  be the weight of v: Then assertions
(a) - (c) of Proposition 31.20 are valid.
112

Proof: It follows from Lemma 31.18 that V has a highest weight vector. Let v be any highest
weight vector in V and let  be its weight. By irreducibility of V; the vector v is cyclic. Hence
all assertions of Proposition 31.20 are valid.
Let w be a second highest weight vector and let  be its weight. Then all assertions of
Proposition 31.20 are valid. Hence  2  NRC and  2  NRC ; from which   2
NRC \ . NRC / D f0g: It follows that  D I hence w 2 V D Cv:

Remark 31.22 For obvious reasons the above weight  is called the highest weight of the
irreducible gC -module V; relative to the choice RC of positive roots.
The following theorem is the first step towards the classification of all finite dimensional
irreducible representations of gC :
Theorem 31.23 Let V and V 0 be irreducible gC -modules. If V and V 0 have the same highest
weight (relative to RC ), then V and V 0 are isomorphic (i.e., the associated gC -representations
are equivalent).
Proof: We denote the highest weight by  and fix associated highest weight vectors v 2 V n f0g
and v 0 2 V0 n f0g: We consider the direct sum gC -module V V 0 and denote by W the smallest
gC -submodule containing the vector w WD .v; v 0/: Then w is a cyclic weight vector of W; of
weight :
Let p W V V 0 ! V be the projection onto the first component, and p 0 W V V 0 ! V 0 the
projection onto the second. Then p and p 0 are gC -module homomorphisms. Since p.w/ D v; it
follows that pjW is surjective onto V: Similarly, p 0 jW is surjective onto V 0 : It follows that V; V 0
are both irreducible quotients of W; hence isomorphic by Proposition 31.20 (e).

Remark 31.24 In the above proof it is easy to deduce that in fact W is irreducible, and pjW and
p 0 jW are isomorphisms from W onto V and V 0 ; respectively.

113

Changes to the 2009 version


 The formulation of Lemma 7.3 has been changed.
 Proof of Lemma 7.3 has been revised.
 Proof of Corollary 7.4 has been revised.

114

32 Conjugacy of maximal tori


We retain the notation of the previous section. In this section we shall investigate to what extent
the collection R D R.gC / depends on the choice of the maximal torus t: An element X 2 tC
will be called regular if .X/ 0 for all 2 R: The set of regular elements in t and tC will be
reg
denoted by treg and tC ; respectively. Since R is finite, treg is an open dense subset of tI similarly
reg
tC is an open dense subset of tC :
Lemma 32.1 Let t be a maximal torus in g; and let X 2 t: Then the following statements are
equivalent.
(a) X 2 treg I
(b) ker. ad.X// D tI

(c) with respect to any G-invariant inner product on g we have t D im. ad.X//? I

(d) with respect to some G-invariant inner product on g we have t D im. ad.X//? I
Proof: Assume (a), and
P let Y 2 g commute with X: In the complexification of g we may
decompose Y D Y0 C 2R Y ; with Y0 2 tC and Y 2 gC for 2 R: Then
X
0 D X; Y D
.X/Y :
2R

Since X is regular, .X/ 0 for all ; and it follows that Y D 0 for all 2 R: Hence
Y 2 g \ tC D t: This implies ker. ad.X//  tI the converse inclusion is obvious, hence (b)
follows.
Next, we assume that (b) holds. Since ad.X/ is anti-symmetric with respect to any invariant
inner product, it follows that im. ad.X//? D ker. ad.X//: The latter space equals t by (b). Hence
(c) follows.
That (c) implies (d) is obvious. Now assume that (d) holds. Then it follows that ad.X/
induces a linear automorphism of g=t: All eigenvalues of a linear automorphism must be different
from zero, hence .X/ 0 for all 2 R:

If g 2 G; then Ad.g/ is an automorphism of the Lie algebra gI hence Ad.g/t is a maximal
torus in g: The following result asserts that all maximal tori of g arise in this way.
Lemma 32.2 Let t; t0 be two maximal tori in g: Then there exists a g 2 G such that
t0 D Ad.g/t:
Proof: By the method of averaging over G we see that there exists a G-invariant positive definite
inner product on gI select such an inner product h ; i: Moreover, select regular elements X 2 t
and Y 2 t0 : Then by Lemma 32.1 we see that t equals the centralizer of X in g: We consider the
smooth function f W G ! R given by
f .x/ D hAd.x/X ; Y i:
115

By compactness of G; the continuous function f attains a minimal value at a point x0 2 G: It


follows that for every Z 2 g the function t 7! f .x0 exp t Z/ has a minimum at t D 0; hence

d
f .x0 exp t Z/ D hAd.x0 /Z; X ; Y i D h ad.X/.Z/ ; Ad.x0 / 1 Y i:
0D

dt t D0

By Lemma 32.1 we see that ad.X/ maps g onto t? : Hence Ad.x0 / 1 Y 2 .t? /? D t: It follows
from this that the maximal torus t00 D Ad.x0 /t contains Y I obviously t00 is contained in the
centralizer of Y; which equals t0 ; by Lemma 32.1. By maximality of t00 it follows that t0 D t00 D
Ad.x0 /t:

If g 2 G; then Ad.g/ is an automorphism of the Lie algebra g: More generally we now
consider an automorphism ' of the Lie algebra gI its complex linear extension, also denoted by
' is an automorphism of the complex Lie algebra gC : If t is a maximal torus, then t0 D '.t/ is a
maximal torus as well. The map tC ! t0C given by  7!  ' 1 is a linear isomorphism, which
we again denote by ': With this notation we have:
Lemma 32.3 Let ' be an automorphism of the Lie algebra g: If t is a maximal torus in g; then
t0 D '.t/ is a maximal torus in g as well. Moreover, the induced linear isomorphism ' W tC ! t0C
maps R D R.gC ; t/ bijectively onto R0 D R.gC ; t0 /: Finally, if 2 R; then
'.gC / D gC'./ :
Proof: Let 2 R and let X 2 gC : Then, for every H 0 2 t0 ;
H 0 ; '.X/ D '.'

.H 0 /; X/ D '..'

.H 0 /X/ D './.H 0 /'.X/:

From this we see that './ 2 R0 and '.gC /  gC'./ : The proof is completed by applying the
same reasoning to the inverse of ':

Corollary 32.4 Let R; R0 be the collections of roots associated with two maximal tori t; t0 of g:
Then there exists a bijective linear map from i t onto i t0  which maps R onto R0 :
Proof: By Lemma 32.2 there exists a g 2 G such that Ad.g/t D t0 : The map ' D Ad.g/ is
an automorphism of g: By Lemma 32.3 the induced isomorphism from tC onto t0C satisfies all
requirements.


33 Automorphisms of a Lie algebra


In this section we assume that g is a finite dimensional real or complex Lie algebra. We denote
by Aut.g/ the group of automorphisms of the Lie algebra g: This is clearly a subgroup of GL.g/:
In fact, Aut.g/ is the intersection, for X; Y 2 g; of the subsets AX;Y consisting of ' 2 GL.g/
with '.X; Y / '.X/; '.Y / D 0: All of these subsets are closed, hence Aut.g/ is a closed
116

subgroup of GL.g/: Its Lie algebra is a sub Lie algebra of End.g/; equipped with the commutator
bracket.
A derivation of g is by definition a linear map D 2 End.g/ such that
D.X; Y / D D.X/; Y C X; D.Y /

.X; Y 2 g/:

One readily sees that the space Der.g/ of all derivations of g is a Lie subalgebra of End.g/:
Proposition 33.1 Der.g/ is the Lie algebra of Aut.g/:
Proof: Let D be an element in the Lie algebra of Aut.g/: Then exp.tD/ 2 Aut.g/ for all t 2 R:
Let X; Y 2 g; then it follows that
e tD X; Y D e tD X; e tD Y :
Differentiating this expression with respect to t at t D 0 we obtain that DX; Y D DX; Y C
X; DY I hence D is a derivation. It follows that the Lie algebra of Aut.g/ is contained in Der.g/:
To prove the converse inclusion, let D 2 Der.g/; and let X; Y 2 g: Consider the function
c W R ! g defined by
c.t / D e tD e tD X; e tD Y :
Then c is differentiable, and its derivative is given by
c 0 .t / D .@t e tD /e tD X; e tD Y C e tD .@t e tD X/; e tD Y C e tD e tD X; .@t e tD Y /
D
e tD D.e tD X; e tD Y / C e tD De tD X; e tD Y C e tD e tD X; De tD Y
D 0:
Hence c is constantly equal to c.0/ D X; Y : It follows from this that e tD 2 Aut.g/ for all
t 2 R; hence D belongs to the Lie algebra of Aut.g/:

Corollary 33.2 The homomorphism ad maps g into Der.g/: If X 2 g; then e adX is an automorphism of the Lie algebra g:
Proof: The first assertion follows from the Jacobi identity. The last statement is now a consequence of the above lemma; indeed e D 2 Aut.g/; for D 2 Der.g/:


The subgroup of Aut.g/ generated by e ad.X/ ; X 2 g is called the group of interior automorphisms of g; notation: Int.g/: Its Lie algebra equals ad.g/; see Section 7.

34 The Killing form


Let g be a finite dimensional Lie algebra over K D R; C: Its Killing form is by definition the
bilinear form g  g ! K defined by
B.X; Y / D tr. ad.X/ ad.Y //;
117

.X; Y 2 g/:

Lemma 34.1 The Killing form is symmetric. Morever, if ' 2 Aut.g/; then
B.'.X/; '.Y // D B.X; Y /

.X; Y 2 g/:

(48)

.X; Y; Z 2 g/:

(49)

Finally,
B.Z; X; Y / D

B.X; Z; Y /

Proof: If A; B are endomorphisms of a linear space, it is well known that AB BA has trace 0:
Hence tr. ad.X/ ad.Y // D tr. ad.Y / ad.X//; for X; Y 2 g; and the symmetry of B follows.
If ' is a Lie algebra automorphism of g; then it follows that ' adX D ad.'.X// ': Hence
ad.'.X// D ' ad.X/ ' 1 : Using this and conjugation invariance of the trace (48) follows.
Let t 2 R; then e t adZ 2 Aut.g/I thus (48) holds with e t adZ inserted for ': Differentiation of
the resulting identity with respect to t at t D 0 yields (49).
The latter identity can also be derived algebraically, as follows. We note that ad.Z; X/ D
ad.Z/ ad.X/ ad.X/ ad.Z/; hence
B.Z; X; Y / D tr. ad.Z/ ad.X/ ad.Y // tr. ad.X/ ad.Z/ ad.Y //
D tr. ad.X/ ad.Y / ad.Z// tr. ad.Y / ad.Z/ ad.X//
D tr . ad.X/ ad.Y; Z// D B.X; Z; Y /:

The identity (49) is known as invariance of the Killing form. If v is a linear subspace of g;
then by v? we shall denote its orthocomplement with respect to B; i.e., the collection of Y 2 g
such that B.X; Y / D 0 for all X 2 v: Note that from the invariance of B the following lemma is
an immediate consequence.
Lemma 34.2 Let v  g: If v is an ideal, then so is v? :

35 Compact and reductive Lie algebras


Throughout this section g will be a real finite dimensional Lie algebra.
The algebra g is called compact if it is isomorphic to the Lie algebra of a compact Lie group.
The purpose of this section is to derive useful criteria for a Lie algebra to be compact. Our
starting point is the following result.
Let B denote the Killing form of g: We recall from Lemma 34.1 that B is a symmetric bilinear
form. Since ad is a Lie algebra homomorphism, its kernel z WD ker ad is an ideal in g: This ideal
is called the center of g:
Lemma 35.1 Let g be compact. Then the Killing form B is negative semi-definite. Moreover,
g? D z:

118

Proof: We may assume that g is the Lie algebra of the compact group G: The representation
Ad of G in gC is unitarizable, hence there exists a positive definite inner product on gC that is
Ad.G/-invariant. With respect to this inner product we have Ad.G/  U.gC /:
Since ad is the infinitesimal representation of g in gC associated with Ad; it follows that
ad.g/  u.gC /: Hence, ad.X/ is an anti-symmetric Hermitian endomorphism of gC ; for X 2 g:
This implies that adX has a an orthonormal basis of eigenvectors and imaginary eigenvalues.
Hence ad.X/2 has the same orthonormal basis of eigenvectors with eigenvalues  0: It follows
that B.X; X/ D tr ad.X/2  0: Hence, B is negative semi-definite. Moreover, if B.X; X/ D 0
then tr ad.X/2 D 0 and it follows that all eigenvalues of ad.X/2 ; hence of ad.X/ are zero.
Hence, ad.X/ D 0: This shows that g?  z: The converse inclusion is obvious.

If v; w are subspaces of g; then by v; w we denote the subspace of g spanned by the elements
X; Y ; where X 2 v; Y 2 w: If v; w are ideals, then v; w is an ideal of g: Indeed, this follows
by a straightforward application of the Jacobi identity. In particular
Dg WD g; g
is an ideal of g; called the commutator ideal.
If a; b are ideals of g; then a C b is an ideal of g as well. Two ideals a and b of g are said to
be complementary if
gDab
(50)
as linear spaces.
Lemma 35.2 If a; b are ideals of g; then a; b  a \ b: If a and b are complementary, then
a; b D 0: In that case, (50) is a direct sum of Lie algebras.
Proof: Since a is an ideal, a; b D b; a  a: Similarly, a; b  b: Hence a; b  a \ b: The
last two assertions now readily follow.

Lemma 35.3 Let g be a compact Lie algebra. Then every ideal of g has a complementary ideal.
Proof: As in the proof of Lemma 35.1 there exists a positive definite inner product h  ;  i on
g with respect to which Ad.G/  O.g/: It follows that ad.g/  o.g/; or, equivalently, that
hX; U ; V i D hU ; X; V i; for all X; U; V 2 g: By this property, if a  g is an ideal, then
a? is an ideal; moreover, g D a a? :

Lemma 35.4 Let g have the property that every ideal has a complementary ideal. Then
g D z Dg:
Proof: The ideal Dg has a complementary ideal, say a: Since obviously g; a  Dg; we have
g; a  a \ Dg D 0; from which we conclude that a  z: It follows that g D z C Dg:
The ideal z has a complementary ideal, say b: Thus, g D z b: Now Dg D g; g 
g; z C g; b  b; from which we conclude that z \ Dg D 0:

119

Theorem 35.5 The following assertions are equivalent.


(a) g is compact
(b) g D z Dg and B is negative definite on Dg:

(c) There exists a subalgebra g0  g such that g D z C g0 and such that B is negative definite
on g0 :

Finally, if g0 is as in (c) then g0 D Dg:


Proof: First, assume that (a) is valid. Then g D z Dg by Lemma 35.4. By Lemma 35.1 it
follows that B < 0 on Dg: Hence (b). The implication (b) ) (c) is obvious. The implication (c)
) (a) and the final assertion will be established in the following lemma.

Lemma 35.6 Let the Killing form of g be negative definite. Then Aut.g/ is compact. Moreover,
ad is a Lie algebra isomorphism from g onto Der.g/: In particular it follows that g is compact
and has trivial center.
Proof: Let O.g/ denote the group of invertible transformations of g that are orthogonal relative
to the positive definite inner product B: Then O.g/ is compact. From (48) it follows that the
closed subgroup Aut.g/ of GL.g/ is contained in the compact group O.g/; hence is compact.
By definition of the Killing form, ker ad  g? I since B is non-degenerate, it follows that ad
is an injective Lie algebra homomorphism. It follows from the Jacobi identity that ad maps g
into Der.g/:
If D 2 Der.g/; then for X 2 g we have that D ad.X/.Y / D ad.DX/Y C ad.X/ D.Y /
for Y 2 g: Hence
D; ad.X/ D ad.DX/:
(51)

It follows that ad.g/ is an ideal in Der.g/:


Now Der.g/ is the Lie algebra of the compact group Aut.g/; see Proposition 33.1. It follows
that Der.g/ is compact. By application of Lemma 35.3 it follows that ad.g/ has a complementary
ideal b in Der.g/: Let D 2 b: Then D commutes with ad.g/; hence from (51) we see that
ad.DX/ D 0; whence DX D 0 for all X 2 g: Hence D D 0: We conclude that b D 0; hence
ad.g/ D Der.g/:
It follows that ad is an isomorphism from g onto Der.g/I the latter is the Lie algebra of the
compact group Aut.g/: Hence g is compact.


Completion of the proof of Theorem 35.5: Assume that (c) holds. Then g0 has negative
definite Killing form, hence is compact. Since z D ker ad  g? it follows that z \ g0 D 0:
Hence, g D z g0 as linear spaces. Since obviously z; g0 D 0; the mentioned direct sum is a
direct sum of Lie algebras.
Let G 0 be a compact Lie group with algebra isomorphic to g0 : Let n D dimz: Then z ' Rn
as abelian Lie algebras. Hence, the compact torus T WD Rn =Zn has Lie algebra isomorphic z:
It follows that the compact group G WD T  G 0 has Lie algebra isomorphic to z g0 D g: (a)
follows.
Finally, let g0 be as in (c). Then by the above reasoning, g D z g0 as Lie algebras. It follows
that Dg  g0 ; g0  g0 : Now apply (b) to conclude that Dg D g0 :

120

The Lie algebra g is called simple if it is not abelian and has no ideals besides 0 and g: It is
called semisimple if it is a direct sum of simple ideals.
Lemma 35.7 Let g be semisimple, then z D 0 and Dg D g:
Proof: Let g D g1   gn be a decomposition into simple ideals. We observe that each ideal gj
is non-abelian, hence gj ; gj is a non-trivial ideal in gj : Since the latter is simple, we conclude
that Dgj D gj : Since the gj mutually commute, it follows that Dg D Dg1 C    C Dgn D g: If
X belongs to the center of g; write X D X1 C    C Xn ; according to the decomposition (52).
Then X commutes with gi and each Xj ; for j i; commutes with gi : Hence, Xi commutes
with gi as well. Hence X belongs to the center zi of gi : This center is an ideal different from gi ;
since gi is not abelian. Since gi is simple, zi D 0: We conclude that X D 0: Hence, z D 0:

Proposition 35.8 The following assertions are equivalent.
(a) The algebra g is compact and has trivial center;
(b) The Killing form of g is negative definite;
(c) The algebra g is compact and semisimple.
Proof: Assume (a). Then by Theorem 35.5, g D Dg and (b) follows. Since the implication (c)
) (a) follows from Lemma 35.7, it remains to establish the implication (b) ) (c). Assume (b).
If a; b are ad.g/-invariant subspaces of g with a  b; then a1 WD a? \ b is an ad.g/-invariant
subspace of g and b D a a1 is a direct sum decomposition of b: Applying this observation
repeatedly, we obtain a direct sum decomposition
g D g1    gn

(52)

of non-trivial ad.g/-invariant subspaces, such that gj has no ad.g/-invariant subspaces besides 0


and gj ; for each j 2 g: The assertion that gj is ad.g/-invariant is equivalent to the assertion that
gj is an ideal in g: Hence, (52) is a direct sum of ideals. It follows that gi ; gj D 0 for i j:
Hence every ideal a of gi is also an ideal of g: We see that each algebra gi has no ideals besides
0 and itself. If gi were abelian, it would centralize gi and gj for all j i; hence g: This would
imply that gi  g? D 0; contradiction. Thus, each ideal gi is simple, and it follows that g is
semisimple.

Lemma 35.9 Let g be a finite dimensional Lie algebra, and let a be a simple ideal of g: If
g D g1 : : : gn is a direct sum of ideals, then there exists a unique 1  j  n such that
gj  a:
Proof: We note that g; a  a since a is an ideal, and g; a  a; a D a since a is simple.
Hence g; a D a: From the direct sum decomposition it now follows that
a D g1 ; a C    C gn ; a:
Hence there exists a j such that gj ; a 0: Since a is simple and gj ; a is an ideal in a we must
have gj ; a D a: This implies that a D a; gj  gj : Of course j is uniquely determined by the
latter property.

121

Lemma 35.10 Let g be semisimple, and let S be the collection of simple ideals in g: Every ideal
a  g is the direct sum of the ideals from S that are contained in a: In particular, g is the direct
sum of the ideals from S:
Proof: We may express g as a direct sum of simple ideals of the form (52). If a 2 S then, by
the previous lemma, a  gj for some j: Since gj is simple, it follows that a D gj : We conclude
that S D fg1 ; : : : ; gn g:
Let now b  g be any ideal. We will show that b is the direct sum of the simple ideals from
S.b/ WD fa 2 S j a  bg by induction on #S: First, assume that #S D 1: Then g is simple,
hence b D 0 or b D g and the result follows. Now assume that #S > 1 and that the result has
been established for g with S of strictly smaller cardinality. If b D 0 there is nothing to prove. If
b 0; then g; b 0 since z D 0: It follows that gj ; b 0 for some j: But gj ; b is an ideal
in the simple algebra gj ; hence gj D gj ; b  b: Let g0 be the direct sum of the ideals from the
non-empty set S.b/; and let g00 be the direct sum of the ideals from S n S.b/: Then g0  b and
g D g0 g00 ; hence b D g0 .b \ g00 /: Now b \ g00 is an ideal in the semisimple algebra g00 : By
the induction hypothesis, b \ g00 is the direct sum of the ideals from S contained in both b and
g00 : This set is empty, hence b \ g00 D 0 and the result follows.


36 Root systems for compact algebras


In this section we assume that g is a compact Lie algebra with trivial center. Then B; the Killing
form of g; is symmetric and negative definite, see Lemma 34.1 and Proposition 35.8. The extension of B to a complex bilinear form on gC equals the Killing form of gC and is also denoted by
B: We fix a maximal torus t in g: Let R be the associated root system and let RC be a choice of
positive roots.
If 2 R; then 2 i t : Therefore, ker is a hyperplane in tC ; which is the complexification
of the hyperplane ker a \ i t in i t: The Killing form B is negative definite on t; hence positive
definite on i t: It follows that there exists a unique element H 2 i t with the properties
H ? ker

and

.H / D 2:

(53)

Lemma 36.1 Let ;  2 tC :


(a) If  C  0; then B D 0 on gC  gC :
(b) If 2 R and X 2 gC ; Y 2 gC

then X; Y 2 tC and

B.X; Y ; H / D B.X; Y / .H /
(c) gC ; gC

 CH :

122

.H 2 tC /:

(54)

Proof: Let X 2 gC and Y 2 gC : Then by invariance of the Killing form we have, for all
H 2 tC ;
.H / C .H / B.X; Y / D B.H; X; Y / C B.X; H; Y / D 0:
From this, (a) follows.
Let now ; X; Y be as in (b). Then X; Y 2 gC0 D tC ; by Corollary 31.11 and Lemma 31.4.
Moreover, for all H 2 tC we have
B.X; Y ; H / D

B.Y; X; H / D B.Y; H; X/ D B.Y; .H /X/ D B.X; Y /.H /:

Hence, (b).
Finally, for (c) we note that (b) implies that X; Y ? ker ; relative to B; for X 2 gC and
Y 2 gC : In view of (53) this implies that gC ; gC  CH :

Let  W gC ! gC be the conjugation with respect to the real form g of gC : Thus,  D I on
g and  D I on i g: We recall that .gC / D gC ; for all 2 R; see proof of Lemma 31.14.
We denote the positive definite inner product B.  ;  /jg by h  ;  i and extend it to a Hermitian
positive definite inner product on gC : Then
hX ; Y i D

.X; Y 2 gC /:

B.X; Y /;

The following result is now immediate.


Corollary 36.2 The root space decomposition
gC D tC

gC

2R

is orthogonal with respect to the inner product h  ;  i:


Lemma 36.3 Let 2 R: Then there exists a X 2 gC such that H ; X and Y WD
a standard sl.2; C/-triple.

X form

Proof: We observe that h  ;  i is positive definite on gC gC : Let X be a non-zero element


in gC : Then X C X 0; hence 0 < hX C X ; X C Xi D B.X C X; X C X/ D
2B.X; X/; since X belongs to the space gC which is perpendicular to gC with respect
to the Killing form. Put Y D X; then
B.X; Y / > 0:
Moreover, X; X D X; X D
follows from Lemma 36.1 (c) that

X; X; hence X; Y D
X; Y D cH ;

X; X 2 i g \ tC D i t: It now

for some c 2 R: Substituting this in (54) with H D H ; we obtain B.cH ; H / D 2B.X; Y /:


p 1
Since B is positive definite on i t; we have B.H ; H / > 0: Hence c > 0: Take X D c X:
Then X ; X D c 1 X; X D H : Hence, X satisfies the requirements.

123

Example 36.4 Let g D su.2/: Then gC D sl.2; C/ and the conjugation  is given by A D
A ; where the star indicates that the Hermitian adjoint is taken. Let H; X; Y be the usual
standard triple in sl.2; C/: Thus, H is diagonal with entries C1; 1 and X is upper triangular
with 1 in the upper right corner, Y is lower triangular with 1 in the lower left corner. Then
t D i RH is a maximal torus in g: Moreover, R D f; g; where 2 i t is determined by
.H / D 2: Finally, X D Y; and we see that the above lemma with X D X gives us the
usual standard triple.
If l is a Lie subalgebra of g; then via the adjoint representation, gC may be viewed as a
l-module.
Lemma 36.5 Let l be a Lie subalgebra of g and let V  gC be a ad.l/-invariant subspace.
Then V decomposes as a direct sum of irreducible l-modules.
Proof: We observe that, for every X 2 l; the endomorphism ad.X/ of gC is anti-symmetric with
respect to h  ;  i: Indeed, this follows from invariance of the Killing form. Thus, if W  V is a
ad.l/-invariant subspace, then so is W ? \ V: The lemma follows by repeated application of this
observation.

Proposition 36.6 Let 2 R: Then dimC gC D 1: Moreover, R \ R D f ; g: The algebra
s WD gC

CH gC

is isomorphic with sl.2; C/: Its intersection with g is isomorphic with su.2/: Finally, t WD i RH
is a maximal torus in s \ g and the associated root system is equal to fjt ; jt g:
Proof: We fix X 2 gC as in Lemma 36.3 and put s D CH CX CY : Then s is
isomorphic with sl.2; C/: Moreover, s is invariant under  and g \ s D ker. I / \ s D
i RH C R.X Y / C i R.X C Y / ' su.2/: The two last assertions of the proposition hold
with s in place of s :
We consider the subspace
X
V D V ./ WD
gC CH :
2R\R

and leave it to the reader to verify that V is invariant under the adjoint action of s: It follows
that V splits as a direct sum of irreducible s-modules. The decomposition of each irreducible
s-submodule in CH -weight spaces is compatible with the given weight space decomposition of
V: All weights of the irreducible representations of s belong to 12 Z0 ; with 0 D jCH : Thus,
if 2 R \ R; then jCH 2 21 0 ; from which we conclude that 2 R \ 12 Z: It follows that
R \ R D R \ 21 Z: Put
X
Vev D
gC CH :
2R\Z

124

and let Vod d be the sum of the remaining root spaces in V: Then both Vev and Vod d are s invariant. The first of these spaces splits as a direct sum of irreducible s-modules all of whose
weights belong to Z0 : By the classification of irreducible sl.2; C/-modules it follows that each
of the irreducible summands has a zero weight space, which must be contained in CH : It follows
that Vev has only one irreducible summand, hence is irreducible. Since s  Vev is an invariant
subspace, it follows that s D Vev : This implies that R \ Z D f; g and s D s :
It remains to be shown that Vod d is the zero space. Assume not. Then V has a CH weight of the form .2n C 1/=20 ; with n 2 Z: This weight occurs in an irreducible summand
of the s -module Vod d : From the classification of the irreducible sl.2; C/-modules, we see that
1
then also occurs as a weight in the irreducible summand, hence in V: Put 0 D 12 : Then
2 0
it follows that gC0 0; hence 0 2 R: Define s0 as above, with 0 in place of : Then
V ./ D V . 0 / D V . 0 /eve n : By the first part of the proof, applied with 0 in place of ; it
follows that V ./ D V . 0 /ev D s0 : This contradicts the fact that gC  V ./:
We conclude that Vod d D 0: Hence, V D Vev D s D s and all assertions follow.

Let 2 R: Then by s we denote the B-orthogonal reflection in the hyperplane ker in i t:
Thus, s .H / D H and s D I on ker ; from which one readily deduces that
s .H / D H

.H 2 i t/:

.H /H

The complex linear extension of s to tC ; also denoted s ; is given by the same formula, for
H 2 tC :
If V is a finite dimensional real linear space, equipped with a positive definite inner product
h  ;  i; then the map v 7! hv ;  i defines a linear isomorphism j W V ! V  : We equip V 
with the so called dual inner product. This is defined to be the unique inner product that makes
j orthogonal. Thus, if ;  2 V  then
h ; i D hj

; j

i D .j

/ D .j

/:

If A W V ! V is orthogonal, then so is j A j 1 W V  ! V  : Using the definitions one


readily verifies that j A j 1 D A 1 : In this case we agree to write A for the orthogonal map
A 1 W V  ! V  : Thus, for  2 V  we write A D  A 1 :
Following the above convention for V D i t equipped with the positive definite inner product
B; we obtain an orthogonal map s W i t ! i t defined by s  D  s 1 D  s ; for  2
i t : Let H 2 i tI then it follows by application of the above formula for the reflection s that
s ./.H / D .H .H /H / D .H / .H /.H /: From this we see that
s  D 

.H /;

. 2 i t /:

Thus, s maps to and is the identity on the hyperplane H0 WD f j .H / D 0g: Since s
is orthogonal it follows that H0 D ? and that s is the orthogonal reflection in the hyperplane
? : The reflection s 2 End.i t / is therefore also given by the formula
s  D 

h ; i
;
h ; i

. 2 i t /:

Comparing this formula with the previous one we see that j.H / D 2=h ; i:
125

Lemma 36.7 Let 2 R: There exists an automorphism ' of gC ; which leaves tC invariant
and has restriction s to this space. The induced endomorphism s of tC leaves R invariant.
Moreover, if 2 R then s ./ 2 Z:
Proof: We fix X ; Y 2 s such that H ; X ; Y is a standard triple. Moreover, we put
U D


.X
2

Y /

and ' WD e adU : Then ' is an automorphism of gC : Since adU annihilates every element of the
subset ker it follows that ' D I on ker : On the other hand, we claim that '.H / D H :
To establish the claim, we observe that the identity is entirely formulated in terms of the
structure of the Lie algebra s : By isomorphism it suffices to show the similar identity in case
H ; X ; Y is the usual standard triple in sl.2; C/: The advantage in that situation is that we can
use computations in the group SL.2; C/: In fact, we have



0 1
;
U D
1 0
2
hence
exp U D

cos 2 sin 2
sin 2 cos 2

0 1
1 0

from which we see that e adU H D Ad.exp.U //H D exp.U /H exp.U / 1 D H : This
establishes the claim.
We conclude that ' 2 Aut.gC /; '.tC / D tC and 'jtC D s : It follows from Lemma 32.3 that
the induced map s 2 GL.tC / maps R to itself.
Finally, let 2 R: Then
X
V D
gC. Ck/
k2Z

is readily seen to be an ad.s /-invariant subspace. Since U 2 s it follows that V is ad.U /invariant, and since V is closed, it follows that ' leaves V invariant. Therefore,
gCs . / D gC'. / D '.gC /  V
and we conclude that s ./ 2 C Z:

Definition 36.8 The subgroup W D W .g; t/ of GL.i t / generated by the reflections s ; for
2 R; is called the Weyl group of .g; t/:
Lemma 36.9 The Weyl group W is finite.
Proof: By Lemma 36.7, each reflection s leaves R invariant. Hence w.R/  R for each
w 2 W: Since w is injective and R finite, it follows that wjR belongs to the group Sym.R/ of
bijections from R onto itself. Clearly the restriction map r W w 7! wjR ; W ! Sym.R/; is
a group homomorphism. Moreover, since R spans t ; by Lemma 31.9, it follows that ker r is
trivial. Hence #W  #Sym.R/ < 1:

126

Let E be a finite dimensional linear space. If 2 E n f0g then by a reflection in we mean


a linear map s W E ! E with s./ D and
E D R ker.s

I /:

Note that any reflection satisfies s 2 D I: Hence s 2 GL.E/ and s

D s:

Lemma 36.10 Let E be a finite dimensional real linear space, and R  E a finite subset that
spans E: Then for every 2 R there is at most one reflection s in such that s.R/ D R:
Proof: Let K be the group of A 2 GL.E/ with A.R/ D R: The restriction map r W A 7! AjR is
a group homomorphism from K to the group of bijections of R: Moreover, r has trivial kernel,
since R spans E: It follows that K is a finite group. Hence, there exists an inner product on E
for which K acts by orthogonal transformations (use averaging). If s is any reflection in a nonzero element of E which preserves R; then it must be an orthogonal transformation, hence the
orthogonal reflection in the hyperplane ? : In particular, there exists at most one such reflection.

Definition 36.11 A (general) root system is a pair .E; R/ consisting of a finite dimensional real
linear space E and a finite subset R  E n f0g such that the following conditions are fulfilled.
(a) R spans E:
(b) If 2 R; then R \ R D f ; g:
(c) If 2 R then there exists a (necessarily unique) reflection s in that maps R to itself.
(d) If ; 2 R then s ./ 2 C Z:
According to the results of this section, the pair consisting of E D i t and R D R.g; t/ is a
root system in the sense of the above definition.
For a general root system, the subgroup W of GL.E/; generated by the reflections s ; for
2 R; is called the Weyl group of the root system .E; R/: By the same proof as that of Lemma
36.9, it follows that W is finite. By averaging we see that E may be equipped with a positive
definite inner product h  ;  i that is W -invariant. It follows that each reflection s ; for 2 R is
orthogonal relative to h  ;  i: Hence, it is given by the formula
s ./ D 

h ; i
;
h ; i

. 2 E/:

(55)

In terms of the inner product the condition (d) in the definition of root system may therefore be
rephrased as
h ; i
2
2Z
.; 2 R/:
h ; i

Two root systems .E; R/ and .E 0 ; R0 / are called isomorphic if there exists a linear isomorphism T W E ! E 0 with T .R/ D R0 : If g is a compact semisimple Lie algebra, then it follows
127

from Lemmas 32.2 and 32.3 that the isomorphism class of the root system R.g; t/ is independent
of the choice of the maximal torus t:
We now have the following result, which we state without proof. It reduces the classification
of all compact semisimple Lie algebras to the classification of all root systems.
Theorem 36.12 The map g ! R.gC ; tC / induces a map from (a) the isomorphism classes of real
Lie algebras with negative definite Killing form to (b) the isomorphism classes of root systems.
This map is a bijection.

37 Weyls formulas
We retain the notation of the previous section. In this section we will describe the classification of
of all irreducible representations of the compact semisimple Lie algebra g: Moreover, in terms of
this classification we will state the beautiful character and dimension formulas due to Hermann
Weyl.
The weight lattice D .g; t/ of the pair .g; t/ is defined as the set
D f 2 i t j 8 2 R W s  2  C Zg:
Equip i t with any W -invariant positive definite inner product h  ;  i: Then from (55) we see
that, alternatively, .g; t/ may be defined as the set of elements  2 i t such that
2

h ; i
2Z
h ; i

for all

2 R:

It follows from the definition of root system that the Z-lattice spanned by R is contained in :
The collection of dominant weights (relative to RC ) is defined by
C D f 2 i t j 8 2 RC W s  2  C N. /g:
Thus, C consists of the collection of weights in that are contained in the convex cone
C C D f 2 i t j h ; i  0

for all

2 RC g:

The following results amount to the classification of all irreducible representations of g:


Theorem 37.1 For every  2 C there exists a unique (up to equivalence) irreducible representation  of g with highest weight :
From this result combined with Theorem 31.23 we obtain the following.
Corollary 37.2 The map  7!  induces a bijection from C onto the collection of equivalence classes of irreducible representations of g:

128

Let now G be a compact connected Lie group with algebra g: Let  be a finite dimensional irreducible representation of G: Then the associated representation of the Lie algebra is equivalent
to  for a unique  2 C : It turns out that in terms of this parametrization, there exist beautiful
formulas for the character and dimension of : The character  of  is conjugation invariant.
In view of the following result, which we state without proof, it is completely determined by its
restriction to T WD exp.t/:
Proposition 37.3 The group T D exp.t/ is a compact torus in G: Moreover, each element of G
is conjugate to an element of T:
If w 2 W we write .w/ D det.w/ for the determinant of wI since w is orthogonal with
respect to a suitable inner product, we have .w/ D 1: We define the element 2 i t by
1 X
:
D
2
C
2R

Theorem 37.4 (Weyls formulas). Let  be an irreducible representation of G of highest weight


: Then the character  is given by
P
w.C/.X/
w2W .w/e
 .exp X/ D P
;
w.X/
w2W .w/e

for all X 2 t for which the denominator is non-zero; these X form an open dense subset (Weyls
character formula). Moreover, the dimension of  is given by
Y h C ; i
dim D
h ; i
C
2R

(Weyls dimension formula).


Example 37.5 We consider the example of g D su.2/; with the associated standard triple
H; X; Y in gC D sl.2; C/: We recall that t D i RH is a maximal torus in g: Moreover, R D
f ; g; where 2 i t is determined by .H / D 2: Also, RC D fg is a choice of positive
roots. The associated Weyl group consists of two elements, 1 and s : Moreover, D 12 : Hence,
C D fn j n 2 Ng:
The representation with highest weight n was earlier denoted by n : We note that .nC/.H / D
n C 1 and s .n C /.H / D .n C 1/: According to the above formula the character of n is
therefore given by the formula
e i.nC1/t e i.nC1/t
n .exp i tH / D
e it e it
which is consistent with what we computed earlier. The dimension of n is given by
dim.n / D
consistent with what we discussed before.

hn C ; i
D n C 1;
h ; i

129

38 The classification of root systems


38.1 Cartan integers
In this section we shall study some aspects of the theory of root systems. In particular we shall
describe the first step towards their classification. The starting point of the theory is the definition
of a root system as given in Definition 36.11. In the rest of this section we assume that .E; R/ is
such a root system. The dimension of E is called the rank of the root system.
By the process of averaging over the Weyl group W of the given root system, we select a
W -invariant positive definite inner product h  ;  i om E: Then, for every 2 R the reflection
s is given by the following formula, for  2 E;
s ./ D 

h ; i
:
h ; i

For two roots ; 2 R we define n to be the integer determined by


s ./ D

n ;

(56)

see Definition 36.11 (d). These integers are called the Cartan integers for the root system. It
follows from the above representation of the reflection in terms of the inner product that the
Cartan integers are alternatively given by
n D 2

h ; i
:
h ; i

(57)

Lemma 38.1 Let ' be an isomorphism from .E; R/ onto a second root system .E 0 ; R0 /: Then,
for all ; 2 R;
(a) ' s D s'./ 'I
(b) n'./ '. / D n :
Proof: It is readily seen that s WD ' s ' 1 W E 0 ! E is a reflection in './: Since s.R0 / D
's ' 1 .'R/ D R0 ; (a) follows. Assertion (b) follows by application of (56).

We shall now discuss the possible values of the Cartan integers. If ; 2 E n f0g; then by
the Cauchy-Schwarz inequality there is a unique ' 2 0;  such that
h ; i D jjjj cos ' :
The number ' is called the angle between and (with respect to the given inner product).
Assume that ; 2 R and : Then
2

h ; i
jj
cos ' D 2
D n 2 Z:
jj
h ; i
130

It follows that
n n D 4 cos2 ' 2 Z:
From this formula we see that the value of ' is independent of the particular choice of W invariant inner product on E: By Definition 36.11(b) the roots ; are not proportional, hence
j cos ' j < 1: It follows that
n n 2 f0; 1; 2; 3g:
After renaming we may assume that jj  jj: It then follows from (57) that jn j  jn j:
By integrality of the Cartan integers we find that either ? or n D 1: This leads to the
following table of possibilities for n and ' :
Lemma 38.2 Let ; 2 R be non-proportional roots with jj  jj: Then the following table
contains all possible combinations of values of n ; n and ' : The question mark indicates
that the value involved is undetermined.
cos '

'

j j2
jj2

n n


2

1
2


3

1
2

2
3

1
2


4

1
2

3
4

1
2


6

1
2

5
6

n
n

Example 38.3 Let E D R2 ; equipped with p


the standard inner product. Let be the first
1 1
standard basis vector .1; 0/; and D . 2 ; 2 3/: Then jj D jj D 1 and ' D 2=3:
p
Moreover, C D . 12 ; 21 3/ has angle =3 with both and : It is easily verified that R D
f; ; . C /g is a root system. Note that n D n D 1: This root system, called A2 ;
is depicted in the illustration following Lemma 38.16. Let r D sC s : Then r is the rotation
over angle 2=3: The reflection s D s is the reflection in the line ? D R.0; 1/: The Weyl
group W equals f1; r; r 1 ; s; s r; s r 1g:
The following lemma will be extremely useful in the further development of the theory.
Lemma 38.4 Let ; 2 R be non-proportional roots.
(a) If h ; i > 0 then

2 R:
131

(b) If h ; i < 0 then C 2 R:


Proof: It suffices to establish (a). Then (b) follows by replacing with : Since 2 R is
equivalent to 2 R we may as well assume that jj  jj: Then it follows that 0 < n 
n ; hence n D 1: In view of (56) this implies that s ./ D : Now use Definition 36.11
to conclude that 2 R:

Given non-proportional roots ; 2 R we define the -string through to be the set
L ./ WD . C Z/ \ R:
The following lemma expresses that root strings have no interruptions and have at most 4 elements.
Lemma 38.5 Let ; 2 R be non-proportional.
(a) There exist unique p; q 2 Z with p  q such that L ./ D f C k j p  k  qg:
(b) p  0  q and p C q D

n :

(c) #L ./  4:
Proof: We first establish (a). Write j WD j; for j 2 Z: Assume (a) does not hold. Then
there exist integers k < l such that k ; l 2 R but kC1 ; l 1 R: It follows by application of
Lemma 38.4 that
hk ; i  0 and hl ; i  0:
On the other hand,
hk ; i D h ; i C kjj2 < h ; i C ljj2 D hl ; i;
contradiction. We conclude that (a) holds. Since 2 L ./; the first assertion of (b) follows.
For the other assertion, we note that s maps L ./ bijectively onto itself. Hence s . C p/ D
C q; from which it follows that n p D q: This establishes (b).
For (c) we note that D C p is a root. Clearly, L ./ D L . / D f C j j 0 
j  q pg; so that #L . / D q p C 1: It now follows from (b) applied to the pair ; that
q p D n ; hence n  0 and q p 2 f0; 1; 2; 3g:


38.2 Fundamental and positive systems


If F is a finite subset of E we write
NF D f

f 2F

nf f j nf 2 Ng:

Here N D f0; 1; 2; : : :g:


132

Definition 38.6 A fundamental system or basis for .E; R/ is a subset S  R such that
(a) S is a basis for EI
(b) R  NS [ N. S/:
Conditions (a) and (b) of the above definition may be restated as follows. Every root 2 R
admits a unique expression of the form
X
D
k ;
2S

with k 2 R: Moreover, either k 2 N for all 2 S or k 2


relative to S is defined by
X
ht./ D
k

N for all 2 S: The height of

2S

Lemma 38.7 Let S be a fundamental system for the root system R: Then for all roots ; 2 S
with one has h ; i  0 (or, equivalently, '  =2).
Proof: Since is a linear combination of the elements of S with both plus and minus signs,
it cannot be a root. It follows from Lemma 38.4 that h ; i  0:

To ensure the existence of fundamental systems, we introduce the notion of a positive system.
By an open half space of E we mean a set of the form E C ./ D fx 2 E j .x/ > 0g; where
 is a non-zero element of the dual space E  WD HomR .E; R/: Via the given inner product, we
sometimes identify E with E  : Accordingly, if 2 E n f0g; we write
E C . / D fx 2 E j hx ; i > 0g:
Definition 38.8 A positive system or choice of positive roots for R is a subset P  R with the
following properties.
(a) There exists an open half space containing P:
(b) R  P [ . P /:
Let P be a positive system for R: An indecomposable or simple root in P is defined to be
a root that cannot be written as the sum of two roots from P: The set of these simple roots is
denoted by S.P /:
Lemma 38.9 Let P be a positive system for R. Then S.P / is a fundamental system for R and
P D NS.P / \ R: The map P 7! S.P / is a bijective map from the collection P of positive
systems for R onto the collection S of fundamental systems for R:

133

Proof: Put S D S.P /: Let 2 P: Then either 2 S; or can be written as a sum C with
; 2 P: Proceeding in this way we see that P  NS; hence condition (b) of Definition 38.6
holds. In particular, it follows that S spans E: It remains to be shown that the elements of S are
linearly independent. Let ; be distinct roots in S: By definition of S; neither nor
does belong to P: Hence, R: It follows by application of Lemma 38.4 that h ; i  0:
From the lemma below it now follows that the elements of S are linearly independent. Hence S
is a fundamental system. In the above we established P  NS; whence P  N. S/ and since
R D P [ . P / it follows that P D NS \ R:
We have shown that the map P 7! S.P / is injective P ! S and will finish the proof
by establishing its surjectivity. For S a fundamental system of R we define RC D RC .S/ D
NS \ R: Since S is a basis for E; the linear functionals h ;  i; for 2 S; form a basis for
the dual space E  : It follows that there exists a 2 E such that h ; i > 0 for all 2 S: We
conclude that S; hence RC ; is contained in a half space. From R  NS [ . NS/ it follows that
R  RC [ . RC /: Hence RC .S/ is a positive system. From S  RC .S/  NS it follows that
S.RC .S//  S: Both sets of of this inclusion are bases for EI hence, they must be equal. We
conclude that the map R 7! S.P /; P ! S is bijective with inverse S 7! RC .S/:

Lemma 38.10 Let E be a finite dimensional real linear space, equipped with a positive definite
inner product. Let S  E be a finite subset contained in a fixed open half space and such that
h ; i  0 for all distinct ; 2 S: Then the elements of S are linearly independent.
Proof:
P There exists a  2 E such that h ; i > 0 for all 2 S: Let  2 R; for 2 S; be such
that 2S 
PD 0: We define S WD f 2 S j  > 0g: Then SC and S are disjoint. We
define  WD 2S j j: If one of the sets of summation is empty, the sum is understood to
be zero. The linear relation between the elements of S may now be expressed as C  D 0 or
C D  :
From the fact that h ; i  0 for all 2 SC and all 2 S it now follows that hC ; C i D
hC ;  i  0: Hence,
P hC ; C i D 0 and we conclude that C D  D 0: We now observe that
0 D h ; C i D 2SC j jh ; i: Since each of the inner products h ; i is strictly positive,
it follows that SC D ;: Similarly, S D ; and we conclude that  D 0 for all 2 S: This
establishes the linear independence.

For each 2 R; the set P WD ker.I s / is called the root hyperplane associated with :
Relative to the given W -invariant inner product, P D ? : We define the set of regular points in
E by
E reg WD E n [2R P :
(58)

Since R is a finite set, it is easy to show that E reg is an open dense subset of EI in particular, the
set of regular points is non-empty.
We can now establish the existence of positive systems, hence also of fundamental systems.
For 2 E reg we define
RC . / D R \ E C . / D f 2 R j h ; i > 0g:
134

Lemma 38.11 For every 2 E reg the set RC . / is a positive system for R: Moreover, every
positive system arises in this way.
Proof: That RC . / is a positive system is immediate from the definitions. Conversely, let P be
a positive system for R; and let S D S.P / be the associated fundamental system. The linear
functionals h  ; i; for 2 S; form a basis for the dual space E  : Hence there exists a 2 E
such that h ; i > 0 for all 2 S: From P  NS it now follows that P  RC . /; hence also
P \ RC . / D ;: Since P is a positive system, RC . /  P [ . P /; so we must have that
P D RC . /:

Definition 38.12 Let S be a fundamental system for R: The integers n ; for ; 2 S; are
called the Cartan integers associated with S: The square matrix n W S  S ! Z; .; / 7! n
is called the Cartan matrix for S:
We will end this section with a result that asserts that every root system is completely determined by the Cartan matrix of a fundamental system. It depends crucially on the following
lemma and Lemma 38.5.
Lemma 38.13 Let S be a fundamental system for R and RC D R \ NS the associated positive
system. If 2 RC n S; then there exists an 2 S such that 2 RC :
P
Proof: Since is not simple, it is of the form 2S k ; k 2 N; with at least two coefficients
non-zero. Thus, if 2 S and 2 R; then at least one of the coefficients of is still
positive, and it follows that 2 RC :
Assume that RC for all 2 S: Then it follows that R for all 2 S:
By Lemma 38.4 this implies that h ; i  0 for all 2 S; hence h ; i  0; hence D 0;
contradiction.

Given any finite set S we write ES for the real linear space with basis S: As a concrete
model we may take the space RS of functions S ! RI here S is embedded in RS by identifying
an element
2 S with the function W S ! R given by 7! : If v 2 ES ; we put
P
v D 2S v : With the above identification, as an element of RS ; the vector v is given by
7! v :
Let E be a real linear space and f W S ! E a map, then f has a unique extension to a linear
map ES ! E; again denoted by f: Moreover, if f W S ! S 0 is a map, then f may be viewed
as a map S ! ES 0 which in turn has a unique linear extension to a map f W ES ! ES 0 :
Theorem 38.14 There exists a map R assigning to every pair consisting of a finite set S and a
function n W S  S ! Z a finite subset R.S; n/  ES with the following properties.

(a) If ' W S 0 ! S is a bijection of finite sets, and n W S  S ! Z a function, then the induced
map ' W ES 0 ! ES maps R.S 0 ; '  n/ bijectively onto R.S; n/:

(b) If .E; R/ is a root system with fundamental system S and Cartan matrix n W S  S ! Z;
then the natural map ES ! E maps R.S; n/ bijectively onto R: In particular, .RS ; R.S; n//
is a root system isomorphic to .E; R/:
135

Remark 38.15 The above result guarantees that the isomorphism class of a root system can be
retrieved from the Cartan matrix of a fundamental system. Later we will see that all fundamental
systems are conjugate under the Weyl group, so that all Cartan matrices of a given root system
are essentially equal, cf. Lemma 38.1.
In the proof of the above result the set R will be defined by means of a recursive algorithm
with input data S; n: This algorithm will provide us with a finite procedure for finding all root
systems of a given rank. Let such a rank r be fixed. Let S be a given set with r elements. Each
root system R of rank r can be realized in the linear space ES ; having the standard basis as
fundamental system.
The possible Cartan matrices run over the finite set of maps S  S ! f0; 1; 2; 3g:
For each such map n it can be checked whether or not .ES ; R.S; n// is a root system with
fundamental system S: Condition (b) guarantees that all root systems of rank r are obtained in
this way.
Proof: We shall describe the map R and then show that it satisfies the requirements. Requirement (b) is motivational for the definition.
For each 2 S we define the map n W S ! Z by n ./ D n.; /: As said above this map
induces a linear map n W ES ! R: If the linear maps n ; for 2 S; are linearly dependent, we
define R.S; n/ D ; (we need not proceed, since n can impossibly be a Cartan matrix of a root
system). Thus, assume that the n are linearly independent linear functionals.
We consider the semi-lattice D NS  ES : Then for each 2 S the map n has integral
values on : We define a height function on in an obvious manner,
X
ht./ D
 :
2S

Let k be the finite set of  2 with ht./ D k: We put P1 D S and more generally will define
sets Pk  k by induction on k:
Let P1 ; : : : ; Pk be given, then PkC1 is defined as the subset of kC1 consisting of elements
that can be expressed in the form C with .; / 2 S  Pk satisfying the following conditions.
(i) and are not proportional.
(ii) jn . C /j  3:
(iii) Let p be the smallest integer such that C p 2 P1 [    [ Pk I then p

n ./ > 0:

We define P.S; n/ to be the union of the sets Pk ; for k  1 and put R.S; n/ D P.S; n/ [
. P.S; n//:
The set F of 2 ES with n ./ 2 f0; 1; 2; 3g for all 2 S is finite, because the n are
linearly independent functionals. In fact, #F  .#S/7 : From the above construction it follows
that R.S; n/  F; hence is finite. In particular, we see that the above inductive definition starts
producing empty sets at some level. In fact, let N be an upper bound for the height function on
F; then P.S; n/ D P1 [    [ PN :
136

From the definition it is readily seen that the map R defined above satisfies condition (a)
of the theorem. We will finish the proof by showing that condition (b) holds. Sssume that
S is a fundamental system for a root system .E; R/: Let RC D R \ NS be the associated
positive system and n W S  S ! Z the associated Cartan matrix. The inclusion map S  E
induces a linear isomorphism ES ! E via which we shall identify. Then it suffices to show that
R.S; n/ D R: Since n is a genuine Cartan matrix, the functionals n ; for 2 S are linearly
independent. Thus it suffices to show that Pk D R \ k ; for every k 2 N: We will do this by
induction on k: For k D 1 we have R \ k D S D P1 ; and the statement holds. Let k  1 and
assume that Pj D R \ j for all j  k: We will show that PkC1 D R \ kC1 :
First, consider an element of PkC1 : It may be written as C with .; / 2 S Pk satisfying
the conditions (i)-(iii). By the inductive hypothesis, 2 RC : Moreover, there exists a smallest
integer p 0  0 such that C p 0 2 RC : By the inductive hypothesis it follows that p 0 D p:
The -string through now takes the form L ./ D f C k j p  k  qg with q the nonnegative integer determined by p C q D n ./: From condition (iii) it follows that q > 0;
hence C 2 RC : It follows that PkC1  R \ kC1 : For the converse inclusion, consider an
element 1 2 RC of weight k C 1: Since k C 1  2; the root 1 does not belong to S: By Lemma
38.13 there exists a 2 S such that WD 1 2 RC : Clearly, ht./ D k; so 2 Pk by the
inductive hypothesis. We will proceed to show that the pair .; / satisfies conditions (i)-(iii).
This will imply that 1 2 PkC1 ; completing the proof.
Since 1 is a root, ; hence (i). Since n .1 / D n1 ; condition (ii) holds by Lemma
38.2. The -root string through has the form L ./ D f C k j p  k  qg; with p the
smallest integer such that C p is a root and with q the largest integer such that C q is a
root. We note that p  0 and q  1: By the inductive hypothesis, p is the smallest integer such
that 2 P1 [    [ Pk : Moreover, by Lemma 38.5, n ./ D n D .p C q/ and (iii) follows.


38.3 The rank two root systems


We can use the method of the proof of Theorem 38.14 to classify the (isomorphism classes
of) rank two root systems. Let .E; R/ be a rank two root system. Then R has a fundamental
system S consisting of two elements, and : Without loss of generality we may assume that
jj  jj: Moreover, changing the inner product on E by a positive scalar we may as well
assume that jj D 1: From Lemma 38.7 it follows that there are 4 possible values for n ; namely
0; 1; 2; 3; with corresponding angles ' equal to =2; 2=3; 3=4; 5=6: If np
D 0 then
p
the length of is undetermined. In the remaining cases, the length of equals 1; 2 and 3;
respectively. It follows from Theorem 38.14 that for each of these cases there exists at most one
isomorphism class of root spaces. We shall discuss these cases separately.
Case n D 0: In the notation of the proof of Theorem 38.14, P1 D f; g It follows that
P2 can only contain the element C : In the notation of condition (iii) of the mentioned proof,
we have p D 0 and n ./ D 0; hence C P2 : It follows that Pj D ; for j  2: Therefore,
R D f; g is the only possible root system with the given Cartan matrix. We leave it to the
reader to check that this is indeed a root system. It is called A1  A1 :
137

Case n D 1. In this case P1 D f; g: There is only one possible element in P2 ; namely


C : Here p D 0 and n ./ D 1 whence p n ./ > 0 and it follows that C 2 P2 :
The possible elements in P3 are . C / C or . C / C : For the first element, Now p D 1
and n . C / D 1; whence 2 C P3 : Similarly, . C / C P3 : It follows that Pj D ;
for j  3: Hence, R D f; ; . C /g is the only possible root system. We leave it to the
reader to check that it is indeed a root system. It is called A2 :
Case n D 2: We have P1 D f; g and P2 D f Cg: The only possible elements in P3
are . C/C and . C/C: For the first of these we have From p D 1 and n . C/ D 0;
so that p n .C g/ > 0 and C 2 2 P3 : For the second element we have From p D 1
and n . C / D 1; whence p n .C / D 0; from which we infer that 2 C P3 : Thus,
P3 D f C 2g:
The possible elements of P4 are . C2/C and . C2/C: For the first element, p D 2
and n . C 2/ D 2; hence C 3 P4 : For the second element, p D 0 and n . C 2/ D 0;
hence 2 C 2 P4 : We conclude that Pj D ; for j  4:
Thus, in the present case the only possible root system is R D f; ; . C /; . C
2/g: Again we leave it to the reader to check that this is a root system. It is called B2 :
Case n D 3: We have P1 D f; g and P2 D f C g: The possible elements of P3 are
C 2 and 2 C : For the first element we have p;C D 1 and n . C / D 1; hence
C 2 2 P3 : For the second we have p;C D 1 and n . C / D 1; hence 2 C P3 :
Thus, P3 D f C 2g:
The possible elements of P4 are C3 and 2C2: For the first element we have p;2C D
2 and n .2 C/ D 1; hence C3 2 P3 : For the second, p;2C D 0 and n .2 C/ D 0;
hence 2 C 2 P4 : Thus, P4 D f C 3g:
The possible elements of P5 are C 3 C and C 3 C : For the first element we have
p D 3 and n . C 3/ D 3; whence C 4 P5 : For the second element we have p D 1
and n . C 3/ D 1; whence 2 C 3 2 P5 ; and we conclude that P5 D f2 C 3g:
The possible elements of P6 are 2 C 3 C and 2 C 3 C : For the first element we
have p D 0 and n .2 C 3/ D 0; and for the second p D 1 and n .2 C 3/ D 1: Hence
Pj D ; for j  6:
We conclude that the only possible root system is R D f; ; C ; 2 C ; 3 C ; 3 C
2g: We leave it to the reader to check that this is indeed a root system, called G2 :
Lemma 38.16 Up to isomorphism, the rank two root systems are completely classified by the
integer n n ; for f; g a fundamental system. The integer takes the values f0; 1; 2; 3g; giving
the root systems A1  A1 ; A2 ; B2 and G2 ; respectively.


Proof: This has been established above.


The rank 2 root systems are depicted below.

138

A1  A1

A2

2 C 3

C 2

C 2

C 3

B2

G2

38.4 Weyl chambers


We proceed to investigate the collection of fundamental systems of the root system .E; R/: An
important role is played by the connected components of E reg ; see (58), called the Weyl chambers
of R:
For every 2 R; the complement E n P is the disjoint union of the open half spaces E C ./
and E C . /: Since E reg is the intersection of the complements E n P ; each Weyl chamber can
be written in the form \2F E C ./; with 2 F; F  R: It follows that each Weyl chamber
139

is an open polyhedral cone. We denote the set of Weyl chambers by C: If C 2 C then for
every 2 R the functional h ;  i is nowhere zero on C; hence either everywhere positive or
everywhere negative. We define
RC .C / D f 2 R j h ;  i > 0 on C g:
Note that for every 2 C we have RC .C / D RC . /: Thus, by Lemma 38.11 the set RC .C / is
a positive system for R and every positive system arises in this way.
If C is a Weyl chamber, then by S.C / we denote the collection of simple roots in the positive
system RC .C /: According to Lemma 38.9 this is a fundamental system for R:
Proposition 38.17
(a) The map C 7! RC .C / defines a bijection between the collection of Weyl chambers and
the collection of positive systems for R:
(b) The map C 7! S.C / defines a bijection between the collection of Weyl chambers and the
collection of fundamental systems for R:
(c) Is C is a Weyl chamber, then
C D fx 2 E j 8 2 RC .C / W hx ; i > 0g D fx 2 E j 8 2 S.C / W hx ; i > 0g:
Proof: Recall that we denote the collections of Weyl chambers, positive systems and fundamental systems by C; P and S; respectively.
If P 2 P we define C.P / WD fx 2 E j 8 2 P W hx ; i > 0g; and if S 2 S we
put C.S/ WD fx 2 E j 8 2 S W hx ; i > 0g: With this notation, assertion (c) becomes
C D C.RC .C // D C.S.C // for every C 2 C:
Let S 2 S: Then the set C.S/ is non-empty and convex, hence connected. Since R 
NS [ NS; it follows that C.S/  E reg : We conclude that there exists a connected component
C 2 C such that C.S/  C: Every root from R has the same sign on C as on C.S/I hence,
C  C.S/: We conclude that C.S/ D C: In particular, S 7! C.S/ maps S into C:
Let P 2 P and let S be the collection of simple roots in P: From S  P  NS it readily
follows that C.S/ D C.P /: In particular, C.P / 2 C:
From Lemma 38.11 it follows that the map C 7! RC .C / is surjective. If C 2 C then from
the definitions it is obvious that C  C.RC .C //  C.S.C //: The extreme members in this
chain of inclusions are Weyl chambers, i.e., connected components of E reg ; hence equal. Thus
(c) follows. Moreover, C.RC .C // D C; from which it follows that C 7! RC .C / is injective,
whence (a). Finally, (b) follows from (a) and (c) combined with Lemma 38.9.

The following result gives a useful characterization of the simple roots in terms of the associated Weyl chamber.
Lemma 38.18 Let C be an open Weyl chamber. A root 2 R belongs to the associated
fundamental system S.C / if and only if the following two conditions are fulfilled.
(a) h ;  i > 0 on C I
140

(b) C \ ? has non-empty interior in ? :


Proof: Put S D S.C / and assume that 2 S: Then (a) follows by definition. From Proposition
38.17 we know that C consists of the points x 2 E with hx ; i > 0 for all 2 S: Since
S is a basis of the linear space E; it is readily seen that C consists of the points x 2 E with
hx ; i  0 for all 2 S: The functionals h ;  ij? ; for 2 S n fg; form a basis of ? ; hence
the set CN \ ? contains the non-empty open subset of ? consisting of the points x 2 ? with
hx ; i > 0 for all 2 S n fg: This implies (b).
Conversely, assume that is a root and that (a) and (b) are fulfilled. From (a) it follows that
2 RC .C /: It remains to be shown that is indecomposable. Assume the latter were not true.
Then D C ; for ; 2 RC .C /: From (b) it follows that h ;  i  0 and h ;  i  0 on an
open subset U of ? On the other hand, h C ;  i D 0 on U: It follows that h ;  i and h ;  i
are zero on U; hence on ? by linearity. From this it follows in turn that ? D ? D ? : Hence
and are proportional to ; contradiction.

The Weyl group leaves R; hence E reg ; invariant. It follows that W acts on the set of connected
components on E reg ; i.e., on the set C of Weyl chambers. Clearly, W acts on the set of positive
systems and on the set of fundamental systems, and the actions are compatible with the maps
of Proposition 38.17. More precisely, if w 2 W and C 2 C; then RC .wC / D wRC .C / and
S.wC / D wS.C /:
Lemma 38.19 Let RC be a positive system for R and let be an associated simple root. Then
s maps RC n fg onto itself.
P
Proof: Let S be the set of simple roots in RC and let 2 RC ; : Then

D
2S k ;
P
with k 2 N and k 0 > 0 for at least one 0 different from : Now s ./ D 2S nfg k C l
for some l 2 Z: Since s is a root, it either belongs to NS or to NS: The latter possibility is
excluded by k 0 > 0: Hence s 2 NS \ R D RC :


If RC is a positive system for R; we define .RC / D to be half the sum of the positive
roots, i.e.,
1 X
D
:
2
C
2R

Corollary 38.20 If is simple in RC ; then s D :


P
Proof: Write D 12 2RC nfg C 21 : The sum in the first term is fixed by s ; whereas the

term 12 is mapped onto 21 :

Two Weyl chambers C; C 0 are said to be separated by the root hyperplane ? if the linear
functional h ;  i has different signs on C and C 0 : We will write d.C; C 0 / for the number of root
hyperplanes separating C and C 0 : If P is any positive system for R then d.C; C 0 / is the number
of 2 P such that h ;  i has different signs on C and C 0 (use that R is the disjoint union of
P and P and that roots define the same hyperplane if and only if they are proportional). In
particular,
d.C; C 0 / D # RC .C / n RC .C 0 / :
141

Definition 38.21 Two Weyl chambers C and C 0 are called adjacent if d.C; C 0 / D 1; i.e., the
chambers are separated by precisely one root hyperplane.
Lemma 38.22 Let C; C 0 be Weyl chambers. Then C; C 0 are adjacent if and only if C 0 D s .C /
for some 2 S.C /: If the latter holds, then 2 S.C 0 /:
Proof: Let C and C 0 be adjacent. Then RC .C / n RC .C 0 / D fg for a unique root : From
S.C / n RC .C 0 / D ; it would follow that S.C /  RC .C 0 /; whence RC .C /  RC .C 0 /: Since
both members of this inclusion have half the cardinality of R; they must be equal, contradiction.
Hence S.C / n RC .C 0 / contains a root, which must be : Similarly, S.C 0 / contains the root
: Since RC .C 0 / and RC .C / have the same cardinality, we infer that RC .C 0 / D RC .C / n
fg [ f g D s .RC .C //; by Lemma 38.19. It follows that RC .C 0 / D RC .s .C //; hence
C 0 D s .C /:
Conversely, assume that 2 S.C / and s .C / D C 0 : Then RC .C 0 / D s RC .C / D RC .C /n
fg [ f g from which one sees that #RC .C / n RC .C 0 / D 1: Hence, C and C 0 are adjacent.

Lemma 38.23 Let C; C 0 be distinct Weyl chambers. Then there exists a chamber C 00 that is
adjacent to C 0 and such that d.C; C 00 / D d.C; C 0 / 1:
Proof: There must be a root 2 S.C 0 / n RC .C /; for otherwise S.C 0 /  RC .C /; hence
RC .C 0 /  RC .C /; contradiction. Let C 00 D s .C 0 /: Then C 0 and C 00 are adjacent by the
previous lemma. Also, by Lemma 38.19, RC .C 00 / D s RC .C 0 / D RC .C 0 / n fg [ f g:
From this we see that RC .C 0 / n RC .C / is the disjoint union of RC .C 00 / n RC .C / and fg: It
follows that d.C; C 00 / D d.C; C 0 / 1:

Lemma 38.24 Let C be a Weyl chamber and S D S.C / the associated fundamental system.
Then for every Weyl chamber C 0 C there exists a sequence s1 ; : : : sn of reflections in roots
from S such that C 0 D s1    sn .C /:
Proof: We give the proof by induction on d D d.C; C 0 /: If d D 1; then the result follows
from Lemma 38.22. Thus, let d > 1 and assume the result has been established for C 0 with
d.C; C 0 / < d: By the previous lemma, there exists a chamber C 00 ; adjacent to C 0 and such that
d.C; C 00 / D d.C; C 0 / 1: By Lemma 38.22, C 00 D s .C 0 / for a simple root 2 S.C 0 /:
By the induction hypothesis there exists a w 2 W that can be expressed as a product of
reflections in roots from S.C / such that w.C / D C 00 : Thus, s w.C / D s .C 00 / D C 0 : Moreover,
s w D wsw 1 D ws w 1 ; and since 2 S.C 00 /; it follows that WD w 1 belongs to
S.C / D w 1 S.C 00 /: We conclude that C 0 D ws.C / with w a product of reflections from roots
in S.C / and with s D s ; reflection in a root from S.C /:

Lemma 38.25 Let S be a fundamental system for R: Then every root from R is conjugate to
a root from S by an element of W that can be written as a product of simple reflections, i.e.,
reflections in roots from S:
142

Proof: Let 2 R: There exists a Weyl chamber C such that ? \ C has non-empty interior in
? : By Lemma 38.18 it follows that either or belongs to S.C /: Replacing C by s .C / if
necessary, we may assume that 2 S.C /: Let C C be the unique Weyl chamber with S.C C / D
S: Then there exists a Weyl group element of the form stated such that w 1 .C / D C C : It follows
that w 2 S.C C / D S:

Corollary 38.26 Let S be a fundamental system for R: Then W is already generated by the
associated collection of simple reflections.
Proof: Let W0 be the subgroup of W generated by reflections in roots from S: Let 2 R: Then
by the previous lemma there exists a w 2 W0 such that D w; with 2 S: It follows that
s D ws w 1 2 W0 : Since W is generated by the s ; for 2 R; it follows that W D W0 :

Definition 38.27 Let S be a fundamental system for R: If w 2 W then an expression w D
s1    sn of w in terms of simple reflections is called a reduced expression if it is not possible to
extract a non-empty collection of factors without changing the product.
Lemma 38.28 Let 1 ; : : : ; n 2 S be simple roots (possibly with repetitions), and let sj D sj
be the associated simple reflections. Assume that s1    sn .n / is positive relative to S: Then
s1    sn is not a reduced expression. More precisely, there exists a 1  k < n such that
s1    sn D s1    sk 1 skC1    sn 1 :
Proof: Write j D sj C1 : : : sn 1 .n /; for 0  j < n: Let P be the positive system determined
by S: Then 0 2 P and n 1 D n 2 P; hence there exists a smallest index 1  k  n 1
such that k 2 P: We have that sk .k / D k 1 2 P; hence, by Lemma 38.19, k D k : We
now observe that for every w 2 W we have wsn D swn w: Applying this with w D skC1 : : : sn 1
we obtain skC1    sn 1 sn D sk skC1    sn 1 D sk    sn 1 : This implies that
s1    sn D s1    sk sk    sn

D s1    sk 1 skC1    sn 1 :


Lemma 38.29 The Weyl group acts simply transitively on the set of Weyl chambers.
Proof: Let C denote the collection of Weyl chambers. The transitivity of the action of W on
C follows from Lemma 38.24. To establish that the action is simple, we must show that for all
C 2 C and w 2 W; wC D C ) w D 1:
Fix C 2 C and let S D S.C / be the associated fundamental system for R: Let w 2 W n f1g:
Then w 1 has a reduced expression of the form w 1 D s1    sn ; with n  1; sj D sj ; j 2
S.C /: From Lemma 38.28 it follows that w 1 n < 0 on C; hence n < 0 on w.C /: It follows
that w.C / C:

143

Remark 38.30 It follows from the above result, combined with Proposition 38.17, that the Weyl
group acts simply transitively on the collection of fundamental systems for R as well as on the
collection of positive systems.
Let S; S 0 be two fundamental systems, and let w be the unique Weyl group element such that
w.S/ D S 0 : Let n W S  S ! Z and n0 W S 0  S 0 ! Z be the associated Cartan matrices.
Then it follows from Lemma 38.1 that n0 .w; w/ D n.; / for all ; 2 S; or more briefly,
w  n0 D n: Thus, the Cartan matrices are essentially equal.
Let S be a fixed fundamental system for R: From now on we denote the associated positive
system by RC : The elements of S are called the simple roots, those of RC are called the positive
roots. The associated Weyl chamber
E C D fx 2 E j 8 2 RC W h ; xi > 0g
is called the associated positive chamber. Given a root ; we will use the notation > 0 to
indicate that 2 RC I this is equivalent to h ;  i > 0 on E C :
We define numbers lS .w/ D l.w/ and nS .w/ D n.w/ for a Weyl group element w 2 W:
Firstly, l.w/; the length of w; is by definition the shortest length of a reduced expression for
w: Secondly, n.W / is the number of positive roots 2 RC such that w is negative, i.e.,
w 2 RC :
Remark 38.31 In general, the numbers lS .w/ and nS .w/ do depend on the particular choice of
fundamental system. This can already be verified for the root system A2 :
Lemma 38.32 For every w 2 W;

n.w/ D l.w/ D d.E C ; w

.E C // D d.E C ; w.E C //:

Moreover, any reduced expression for w; relative to S; has length l.w/:


Proof: d.E C ; w 1 .E C // equals the number of positive roots 2 RC such that < 0 on
w 1 .E C /: The latter condition is equivalent with w < 0 on E C or w 2 RC : Thus, n.w/ D
d.E C ; w 1 .E C //: On the other hand, clearly
d.E C ; w 1 .E C // D d.wE C ; ww

E C / D d.E C ; wE C /:

It follows from the proof of Lemma 38.24 that any reduced expression has length at most
d.E; wE C /: In particular, l.w/  d.E C ; wE C /:
We will finish the proof by showing that n.w/  l.w/; by induction on l.w/: If l.w/ D 1;
then w is a simple reflection, and the inequality is obvious. Thus, let n > 1 and assume the
estimate has been established for all w with l.w/ < n: Let w 2 W with l.w/ D n: Then w
has a reduced expression of the form w D s1    sn 1 s ; with 2 S.C /: Put v D s1 : : : sn 1 I
this expression must be reduced, hence l.v/ < n and it follows that n.v/  n 1 by the
inductive hypothesis. On the other hand, from Lemma 38.28 it follows that w 2 RC ; hence
WD v > 0: The root belongs to S.vE C /; hence RC .wE C / D RC .s vE C / D RC .vE C / n
fg [ f g: It follows that RC n RC .wE C / is the disjoint union of RC n RC .vE C / and fg:
Hence n.w/ D d.E C ; wE C / D d.E C ; vE C/ C 1 D n.v/ C 1  l.v/ C 1  l.w/:

144

38.5 Dynkin diagrams


Let .E; R/ be a root system, S a fundamental system for R: The Coxeter graph attached to S is
defined as follows. The vertices of the graph are in bijective correspondence with the roots of SI
two vertices ; are connected by n  n edges. Thus, every pair is connected by 0; 1; 2 or 3
edges, see the table in Lemma 38.2.
The Dynkin diagram of S consists of the Coxeter graph together with the symbol > or <
attached to each multiple edge, pointing towards the shorter root. From Lemma 38.16 it follows that (up to isomorphism) the Dynkin diagrams of the rank-2 root systems are given by the
following list:

A1  A1

A2

B2

G2

It follows from Remark 38.30 that the Dynkin diagrams for two different choices of fundamental
systems for R are isomorphic (in an obvious sense). We may thus speak of the Dynkin diagram
of a root system. The following result expresses that the classification of root systems amounts
to describing the list of all possible Dynkin diagrams.
Theorem 38.33 Let R1 ; R2 be two root systems. If the Dynkin diagrams associated with R1
and R2 are isomorphic, then R1 and R2 are isomorphic as well.
Proof: Let S1 and S2 be fundamental systems for R1 and R2 ; respectively. It follows from
Lemma 38.2 that the Cartan matrices n1 and n2 of S1 and S2 are completely determined by their
Dynkin diagrams. An isomorphism between these Dynkin diagrams gives rise to a bijection
' W S1 ! S2 such that, n1 D '  n2 : By Theorem 38.14 it follows that R1 and R2 are isomorphic.

Remark 38.34 It follows from the above result combined with Theorem 36.12 that the (isomorphism classes of) Dynkin diagrams are in bijective correspondence with the isomorphism classes
of semisimple compact Lie algebras.
Let S be a fundamental system. The decomposition of its Dynkin diagram D into connected
components Dj ; .1  j  p/; determines a decomposition of S into a disjoint union of subsets
Sj ; .1  j  p/: Here Sj consists of the roots labelling the vertices in Dj : The decomposition
of S is uniquely determined by the conditions that Si ? Sj if i j; and that every Sj cannot be
145

written as a disjoint union of proper subsets Sj1 ; Sj 2 with Sj1 ? Sj 2 : We will investigate what
this means for the root system R:
If .Ej ; Rj /; with j D 1; 2; are two root systems, we define their direct sum .E; R/ as
follows. First, E WD E1 E2 : Via the natural embeddings Ej ! E; the sets R1 and R2 may be
viewed as subsets of EI accordingly we define R to be their union. If 2 R1 ; the map s I
is a reflection in .; 0/ preserving R: By a similar remark for R2 ; we see that R is a root system.
Moreover, for all 2 R1 and 2 R2 ; n D 0: From this we see that E1 ? E2 for every
W -invariant inner product on E: Every reflection preserves both R1 and R2 ; hence E1 and E2
are invariant subspaces for the Weyl group. Moreover, the maps v 7! v I and w 7! I w
define embeddings W1 ,! W and W2 ,! W via which we shall identify. Accordingly we have
W D W1  W2 : Similar remarks hold for the direct sum of finitely many root systems.
Definition 38.35 A root system .E; R/ is called reducible if R is the union of two non-empty
subsets R1 and R2 such that E D span.R1 / span.R2 /: It is called irreducible if it is not
reducible.
The following result expresses that every root system allows a decomposition as a direct sum
of irreducibles, which is essentially unique.
Proposition 38.36 Let .E; R/ be a root system. Then there exist finitely many linear subspaces
Ej ; 1  j  n; such that Rj WD Ej \ R is an irreducible root system for every j; and such that
R D [j Rj : The Ej are uniquely determined up to order.
If Sj is a fundamental system of Rj ; for j D 1    n; then S D S1 [    [ Sn is a fundamental
system for R: Every fundamental system for R arises in this way.
If Pj is a positive system of Pj ; for j D 1    n; then P D P1 [    [ Pn is a positive system
for R: Every positive system of R arises in this way.
Proof: From the definition of irreducibility, it follows that .E; R/ has a decomposition as stated.
We will establish its uniqueness at the end of the proof.
If the Sj are fundamental systems as stated, then it is readily checked from the definition that
their union S is a fundamental system for R: Let Pj be positive systems as stated, then again
from the definition it is readily verified that their union P is a positive system for R:
Conversely, let P be a positive system for R: Then it is readily verified that every set Pj WD
P \ Rj is a positive system for Rj : Moreover, let S be a fundamental system for R: Since R is
the disjoint union of the sets Rj ; it follows that S is the disjoint union of the sets Sj WD S \ Rj :
Each Sj is linearly independent, hence for dimensional reasons a basis of Ej : Now Rj  .NS [
. NS// and Rj  RSj : By linear independence this implies that Rj  NSj [. NSj / for every
j: Hence every Sj is a fundamental system.
We now turn to uniqueness of the decomposition as stated. Let E D 1j m Ej0 be a decomposition with similar properties. Fix a fundamental system Sj0 for Rj0 D R \ Ej0 ; for every
j: The union S 0 is a fundamental system for R hence of the form S D S1 [    [ Sn ; with Sj a
fundamental system for Rj ; for each j: It follows that S10 is the disjoint union of the sets S10 \ Sj ;
1  j  n: Hence E10 is the direct sum of the spaces E10 \ Ej and R10 is the union of the sets
R10 \ Rj D R10 \ Ej : From the irreducibility of E10 it follows that there exists a unique j such
that E10 D Ej : The other components may be treated similarly.

146

In view of the above result we may now call the uniquely determined .Ej ; Rj / the irreducible
components of the root system .E; R/:
Lemma 38.37
(a) Let R be a root system. Then the Dynkin diagram of R is the disjoint union of the Dynkin
diagrams of the irreducible components of R:
(b) A root system is irreducible if and only if the associated Dynkin diagram is connected.
Proof: Let .E; R/ be an root system, with irreducible components .Ej ; Rj /: Select a fundamental system Sj for each Rj and let S be their union. The inclusion Sj  S induces an inclusion of
Dj ,! D via which we may identify. For distinct indices i; j we have n D 0 for all 2 Si ;
2 Sj : Hence no vertex of Di is connected with any vertex of Dj : It follows that D is the
disjoint union of the Dj ; and (a) follows.
We turn to (b). If R is reducible, then by (a), the associated Dynkin diagram is not connected.
Conversely, assume that the Dynkin diagram of R is not connected. Then it may be written as the
disjoint union of two non-empty diagrams D1 and D2 : Fix a fundamental system S of R: Then
S decomposes into a disjoint union of two non-empty subsets S1 and S2 such that the elements
of Sj label the vertices of Dj : It follows that for all 2 S1 and all 2 S2 ; n D 0: Put
Ej D span.Sj /; then it follows that for each 2 S the reflection s leaves the decomposition
E D E1 E2 invariant. Hence, the Weyl group W of R leaves the decomposition invariant. Let
2 R; then there exists a w 2 W such that w 2 S D S1 [ S2 : It follows that lies either in
E1 or in E2 : Hence R D R1 [ R2 with Rj D Ej \ R; and we see that R is reducible.

The following result relates the notion of irreducibiliy of a root system with decomposability
of a semisimple Lie algebra.
Proposition 38.38 Let g be a compact semisimple Lie algebra with Dynkin diagram D: Let
D D D1 [ : : : [ Dn be the decomposition of D into its connected components. Then every Dj
is the Dynkin diagram of a compact simple Lie algebra gj : Moreover,
g ' g1    gn :
In particular, g is simple if and only if D is connected.
Remark 38.39 Note that in view of Lemma 35.10 the above result implies that the connected
components of D are in bijective correspondence with the simple ideals of g:
Proof: Let g D j hj be the decomposition of g into its simple ideals. For each j we fix a
maximal torus tj  hj : Then t WD t1    tn is a maximal torus in g (use that hi commutes
with hj for every i j ). Via the direct sum decomposition of t; we view tj as the linear
subspace of elements of t that vanish on tk for every k j: Accordingly, t D t1    tn ;
and a similar decomposition of the complexification. Let Rj be the root system of tj in hj : Since
gC is the direct sum of tC and the root spaces gC ; for 2 R1 [    [ Rn ; it follows that the
root system R of t in g equals the disjoint union of the Rj : Hence, R is the direct sum of the Rj :
147

The Dynkin diagram of R is the disjoint union of the Dynkin diagrams of the Rj : The proof will
be finished if we can show that the Dynkin diagram of Rj ; is connected, for each j: By Lemma
38.37 this is equivalent to the assertion that each Rj is irreducible.
Thus, we may assume g is simple, t a maximal torus in g; and then we must show that
R D R.g; t/ is irreducible. Assume not. Then we may decompose R as the disjoint union of
two non-empty subsets R1 and R2 whose spans have zero intersection. Put E D i t ; and for
j D 1; 2; define Ej D span.Rj /: Then E D E1 E2 : Let
t1 WD \2R2 ker

and

t2 WD \ 2R1 ker :

Then t D t1 t2 and, accordingly, Ej ' i tj : For j D 1; 2; let


gj D tj . g \

gC /:

2Rj

Then g D g1 g2 as a vector space. Moreover, adt normalizes this decomposition, t1 centralizes


g2 and t2 centralizes g1 : If ; 2 R and C 2 R; then we must have that f; g is a subset
of either R1 or R2 : From this we readily see that g1 and g2 are subalgebras of g: Moreover,
if 2 R1 and 2 R2 ; then C R; hence gC.C / D 0: It follows that g1 ; g2 D 0: We
conclude that g D g1 g2 as a direct sum of ideals, contradicting the assumption that g is simple.

In view of the above the following result amounts to the classification of all simple compact
Lie algebras.
Theorem 38.40 The following is a list of all connected Dynkin diagrams of root systems. These
diagrams are in bijective correspondence with the (isomorphism classes of) the simple compact
Lie algebras.

148

An W

:::

n1

SU.n C 1/

Bn W

:::

n2

SO.2n C 1/

Cn W

:::

n3

Sp.n/

Dn W

:::

n4

SO.2n/

G2 W
F4 W

E6 W

E7 W

E8 W

Acknowledgement: I warmly thank Lotte Hollands for providing me with LATEX files for these
and all other pictures in the lecture notes.
149

Index
abelian group, 4
action of a group, 43
adjacent Weyl chambers, 142
adjoint representation, of G, 18
angle, between roots, 130
anti-symmetry, of Lie bracket, 20
arcwise connected, 8
associativity, 4
automorphism, of a Lie group, 7
Banach space, 69
Banach-Steinhaus theorem, 71
barrelled space, 71
base space, of principal bundle, 48
basis, of root system, 133
Cartan integer, 130
Cartan integers, 135
Cartan matrix, 135
center of a group, 5
center, of a Lie algebra, 118
character of a representation, 80
character, multiplicative, 94
choice of positive roots, 133
class function, 93
closed subgroup, 12
commutative group, 4
commutative Lie algebra, 22
commuting elements, of the Lie algebra, 22
compact Lie algebra, 118
complex Hilbert space, 72
complexification, of a Lie algebra, 101
component of the identity, 23
conjugation, 5
connected, 8
continuous action, 43
continuous representation, 69
contragredient representation, 82
coset space, 6
coset, left, 6

Coxeter graph, 145


cyclic vector, 111
densities, bundle of, 63
density, invariant, 64
density, on a linear space, 62
density, on a manifold, 63
derivation, 117
direct sum of representations, 83
dominant, 128
dual of a representation, 82
Dynkin diagram, 145
equivalence class, 5
equivalence relation, 5
equivalent representations, 74
equivariant map, 43, 74
exponential map, 16
fiber, of a map, 5
finite dimensional representation, 69
frame bundle, 48
free action, 50
fundamental system, 133
G-space, 43
group, 4
group of automorphisms, 116
group of interior automorphisms, 117
Haar measure, 66
Haar measure, normalized, 66
half space, 133
height of a root, 133
Hermitian inner product, 72
highest weight, 113
highest weight vector, 111
Hilbert space, 69
homogeneous, 55
homomorphism, of groups, 4
homomorphism, of Lie groups, 7
150

normal subgroup, 6
normalized Haar measure, 66

ideal, 60
image, of a homomorphism, 4
indecomposable root, 133
induced infinitesimal representation, 107
integral curve, 16
integral operator, 89
integral, of a density, 64
intertwining map, 43, 74
invariance, of Killing form, 118
invariant density, 64
invariant subspace, 72
inverse function theorem, 17
irreducible representation, 72
isomorphic, 4
isomorphism, 4
isomorphism of Lie groups, 7
isomorphism of root systems, 127

one parameter subgroup, 18


open half space, 133
open subgoup, 24
orbit space, 44
orbits, for group action, 44
orthogonal group, 13
partition, 5
Peter-Weyl theorem, 86
positive density, 63
positive density, on a manifold, 63
positive root, 110
positive system, 133
primitive vector of an sl(2)-module, 104
principal fiber bundle, 48
product density, 88
proper action, 49
proper map between topological spaces, 49

Jacobi identity, 21
kernel, of a group homomorphism, 4
kernel, of an integral operator, 89
killing form, 117
Lebesgue measure, 63
left action, 43
left invariant vector field, 15
left regular representation, 70
left translation, 5
Lie algebra, 21
Lie algebra homomorphism, 21
Lie subgroup, 28
local trivialization, of principal bundle, 48
locally convex space, 69
Lorentz group, 13
matrix coefficient, of representation, 73
maximal torus, 107
module, for a Lie algebra, 70
module, of a Lie group, 71
monomorphism, 4
multiplicative character, 94
multiplicity, of an irreducible representation, 84
neutral element, 4

Radon measure, 65
rank, of a root system, 130
real symplectic group, 13
reducible root system, 146
reflection, 127
regular element, 115
relation, 5
representation, of a Lie algebra, 70
representative functions, 85
right action, 43
right regular representation, 70
right translation, 5
root space, 108
root space decompostion, 108
root system, general, 127
roots, 108
Schur orthogonality, 79
Schur orthogonality relations, 79
Schurs lemma, 75
semisimple Lie algebra, 121
sesquilinear form, 72

151

simple ideal, 121


simple Lie algebra, 121
simple root, 133
slice, 50
smooth action, 45
special linear group, 10
special orthogonal group, 13
special unitary group, 13
spectral theorem, 87
standard sl(2)-triple, 103
structure group, of principal bundle, 48
subalgebra of a Lie algebra, 30
subgroup, 4
submersion theorem, 11
substitution of variables, for density, 64
symplectic form, 13
symplectic group, compact form, 15
symplectic group, complex form, 15
system of positive roots, 110
tensor product, of representations, 83
topological group, 43, 66
torus, 107
total space, of principal bundle, 48
uniform boundedness theorem, 71
unimodular group, 67
unitarizable representation, 72
unitary group, 13
unitary representation, 72
vector field, 15
weight, 106
weight lattice, 128
weight space, 106
Weyl chamber, 110, 139
Weyl group, of a compact algebra, 126
Weyl group, of root system, 127
Weyls character formula, 129
Weyls dimension formula, 129

152

S-ar putea să vă placă și