Liegroup Math
Liegroup Math
Liegroup Math
Dietmar Salamon
ETH-Zürich
17 October 2022
Contents
1 Lie Groups 2
3 Closed Subgroups 7
6 The Center 18
7 Inner Automorphisms 24
9 Isotropy Subgroups 28
10 Centralizers 29
11 Simple Groups 30
12 Examples 33
References 40
1
1 Lie Groups
A Lie Group is a smooth manifold with a group structure such that the
multiplication and the inverse map are smooth (C ∞ ). A Lie subgroup of
a Lie group is a subgroup that is also a submanifold. Assume throughout
this section that G is a Lie group. The tangent space of G at the identity
element 1l ∈ G is called the Lie algebra of G and will be denoted by
g := Lie(G) := T1l G.
For every g ∈ G the right and left multiplication maps Rg , Lg : G → G are
defined by
Rg (h) := hg, Lg (h) := gh
for h ∈ G. We shall denote the derivatives of these maps by
vg := dRg (h)v ∈ Thg G, gv := dLg (h)v ∈ Tgh G
for v ∈ Th G. In particular, for h = 1l and ξ ∈ T1l G = g, we have ξg, gξ ∈ Tg G
and hence ξ determines two vector fields g 7→ gξ and g 7→ ξg on G. These
are called the left-invariant respectively right-invariant vector fields gen-
erated by ξ. We shall prove in Lemma 1.2 below that the integral curves of
both vector fields through g0 = 1l agree.
Exercise 1.1. (i) Prove that
(v0 g1 )g2 = v0 (g1 g2 )
for v0 ∈ Tg0 G and g1 , g2 ∈ G. Similarly
(g0 v1 )g2 = g0 (v1 g2 ), (g0 g1 )v2 = g0 (g1 v2 ).
(ii) Prove that with the above notation the Leibniz rule holds, i.e. if α, β :
R → G are smooth curves, then
d
α(t)β(t) = α̇(t)β(t) + α(t)β̇(t).
dt
Hint: Differentiate the map R2 → G : (s, t) 7→ α(s)β(t).
(iii) Deduce that
d
γ(t)−1 = −γ(t)−1 γ̇(t)γ(t)−1
dt
for every curve γ : R → G.
(iv) Prove that the vector fields g 7→ gξ and g 7→ ξg are complete for every
ξ ∈ g. Hint: Prove that the length of the existence interval is independent
of the initial condition.
2
Lemma 1.2. Let ξ ∈ g and let γ : R → G be a smooth curve. Then the
following conditions are equivalent.
(i) For all s, t ∈ R
γ(t + s) = γ(s)γ(t), γ(0) = 1l, γ̇(0) = ξ. (1.1)
3
In other words the linear map Ad(g) : g → g is the derivative of the smooth
map G → G : h 7→ ghg −1 at h = 1l. The map G → GL(g) : g 7→ Ad(g) is a
group homomorphism, i.e.
Ad(gh) = Ad(g)Ad(h), Ad(1l) = id,
for all g, h ∈ G, and is called the adjoint action of G on its Lie algebra. The
derivative of this map at g = 1l in the direction ξ ∈ g is denoted by ad(ξ).
The Lie bracket of two elements ξ, η ∈ g is defined by
d
[ξ, η] := ad(ξ)η = dt t=0
exp(tξ)η exp(−tξ).
Lemma 1.3. For all ξ, η, ζ ∈ g we have
[ξ, η] = −[η, ξ], (1.6)
[[ξ, η], ζ] + [[η, ζ], ξ] + [[ζ, ξ], η] = 0. (1.7)
Proof. We prove that the map g → Vect(G) : ξ 7→ Xξ defined by Xξ (g) := ξg
for ξ ∈ g and g ∈ G is a Lie algebra homomorphism. To see this, denote
by ψt ∈ Diff(G) the flow generated by Xξ , i.e. ψt (g) := exp(tξ)g for t ∈ R
and g ∈ G. Then, by definition of the Lie bracket of vector fields,
d
[Xξ , Xη ](g) = dt t=0
dψt (ψ−t (g))Xη (ψ−t (g))
d
= dt t=0
exp(tξ)η exp(−tξ)g
= [ξ, η]g
Here we have used Exercise 1.1 (i). Now the assertions follow from the
properties of the Lie bracket for vector fields.
Definition 1.4. A Lie algebra is a real vector space g equipped with a skew-
symmetric bilinear map g × g → g : (ξ, η) 7→ [ξ, η] that satisfies the Jacobi
identity (1.7).
Lemma 1.5. Let ξ, η ∈ g = Lie(G) and define γ : R → G by
γ(t) := exp(tξ) exp(tη) exp(−tξ) exp(−tη).
√
Then γ̇(0) = 0 and dtd t=0 γ( t) = [ξ, η].
Proof. As in the proof of Lemma 1.3, the flow of the vector field Xξ (g) = ξg
on G is given by t 7→ Lexp(tξ) and [Xξ , Xη ] = X[ξ,η] for ξ, η ∈ g. Hence the
result follows from the corresponding formula for general vector fields.
4
Lemma 1.6. Let R2 → G : (s, t) 7→ g(s, t) be a smooth map. Then
∂s g −1 ∂t g − ∂t g −1 ∂s g + g −1 ∂s g, g −1 ∂t g = 0.
(1.8)
Proof. If M is a smooth manifold, R2 → M : (s, t) 7→ γ(s, t) is a smooth
map, and X(s, t), Y (s, t) ∈ Vect(M ) are smooth families of vector fields such
that
∂s γ = X ◦ γ, ∂t γ = Y ◦ γ,
then
(∂s Y − ∂t X − [X, Y ]) ◦ γ = 0.
To obtain the formula (1.8), apply this identity to the manifold M := G, the
map γ := g : R2 → G, and the vector fields X := Xξ and Y := Xη (as in the
proof of Lemma 1.3), where ξ := (∂s g)g −1 and η := (∂t g)g −1 .
Exercise 1.7. Prove that for every g ∈ G and every ξ ∈ g
g exp(ξ)g −1 = exp(Ad(g)ξ) (1.9)
Hint: Consider the curve γ(t) = g exp(tξ)g −1 and use Exercise 1.1.
Exercise 1.8. Prove that for every ξ ∈ g
Ad(exp(ξ)) = exp(ad(ξ)). (1.10)
Hint: See Lemma 2.1 below.
Exercise 1.9. Prove that any two elements ξ, η ∈ g satisfy [ξ, η] = 0 if and
only if exp(sξ) and exp(tη) commute for all s, t ∈ R.
5
Lemma 2.1. Let ϕ : G → H be a Lie group homomorphism. Then its
derivative Φ := dϕ(1l) : g → h at the identity is a Lie algebra homomorphism.
Proof. We show first that Φ and ϕ intertwine the exponential maps, i.e.
exp(Φ(ξ)) = ϕ(exp(ξ)) (2.1)
for all ξ ∈ g. To see this, consider the curve γ(t) := ϕ(exp(tξ)) ∈ H. This
curve satisfies γ(s + t) = γ(s)γ(t) for all s, t ∈ R and γ̇(0) = Φ(ξ). Hence,
by Lemma 1.2, γ(t) = exp(tΦ(ξ)). With t = 1 this proves (2.1).
Next we prove that
Φ(gξg −1 ) = ϕ(g)Φ(ξ)ϕ(g)−1 (2.2)
for ξ ∈ g and g ∈ G. Consider the curve γ(t) := g exp(tξ)g −1 . By (2.1), we
have ϕ(γ(t)) = ϕ(g) exp(tΦ(ξ))ϕ(g)−1 . Differentiate this curve at t = 0 to
obtain (2.2). By (2.1) and (2.2), we have
d
Φ([ξ, η]) = dt t=0
Φ(exp(tξ)η exp(−tξ))
d
= dt t=0
exp(tΦ(ξ))Φ(η) exp(−tΦ(ξ))
= [Φ(ξ), Φ(η)]
for all ξ, η ∈ g. This proves Lemma 2.1.
A representation of G is a Lie group homomorphism ρ : G → GL(V )
where V is a real or complex vector space. Differentiating such a map at g = 1l
gives a Lie algebra homomorphism ρ̇ : g → End(V ) defined by
d
ρ̇(ξ) := dt t=0
ρ(exp(tξ)) for ξ ∈ g.
Examples are the obvious action of U(n) on Cn and the induced actions on
spaces of symmetric polynomials or exterior forms.
Definition 2.2. Let g be a finite-dimensional Lie algebra. A Lie algebra
automorphism of g is a bijective Lie algebra homomorphism Φ : g → g, so
its inverse is also a Lie algebra homomorphism. The group of Lie algebra
automorphisms of g is denoted by Aut(g) ⊂ GL(g). A derivation of g is
a linear map δ : g → g that satisfies δ[ξ, η] = [δξ, η] + [ξ, δη] for all ξ, η ∈ g.
The space of all derivations of g is denoted by Der(g) ⊂ End(g).
Exercise 2.3. Prove that Aut(g) is a Lie subgroup of GL(g) with the Lie
algebra Lie(Aut(g)) = Der(g) ⊂ End(g). Hint: Use Theorem 3.1 below.
Exercise 2.4. Let G be a Lie group with the Lie algebra g := Lie(G). Prove
that Ad : G → Aut(g) is a Lie group homomorphism and ad : g → Der(g) is
the corresponding Lie algebra homomorphism. Hint: See Lemma 1.3.
6
3 Closed Subgroups
Assume throughout that G is a Lie group (not necessarily compact) and
denote by g := Lie(G) its Lie algebra. Whenever necessary, we assume that g
is equipped with an inner product and define a Riemannian metric on G
by |v| := |g −1 v| for g ∈ G and v ∈ Tg G. The following theorem was first
proved in 1929 by John von Neumann [5] for the special case G = GL(n, R)
and then in 1930 by Élie Cartan [1] in full generality.
for every t ∈ R.
Proof. Assume for simplicity that G is a Lie subgroup of GL(n, R). Fix a
nonzero real number t and for k ∈ N define
ξk := k γ(t/k) − 1l ∈ Rn×n .
Then
γ(t/k) − γ(0)
lim ξk = t lim = tγ̇(0) = tξ
k→∞ k→∞ t/k
and hence k
ξk
exp(tξ) = lim 1l + = lim γ(t/k)k .
k→∞ k k→∞
7
Lemma 3.3. Let H ⊂ G be a closed subgroup. Then the set
h := η ∈ g exp(tη) ∈ H for all t ∈ R
in (3.1) is a Lie subalgebra of g
Proof. Let ξ, η ∈ h and define the curve γ : R → H by
γ(t) := exp(tξ) exp(tη)
for t ∈ R. This curve is smooth and satisfies γ(0) = 1l and γ̇(0) = ξ + η.
Since H is closed, it follows from Lemma 3.2 that
exp(t(ξ + η)) = lim γ(t/k)k ∈ H
k→∞
Then ξ ∈ h.
Proof. Fix a real number t. Then, for each i ∈ N, there exists a unique
integer mi ∈ Z such that mi |ξi | ≤ t < (mi + 1)|ξi |. The sequence mi satisfies
ξi
lim mi |ξi | = t, lim mi ξi = lim mi |ξi | = tξ.
i→∞ i→∞ i→∞ |ξi |
Hence
exp(tξ) = lim exp(mi ξi ) = lim exp(ξi )mi ∈ H
i→∞ i→∞
8
Proof of Theorem 3.1. We prove that (i) implies (ii) and (3.1). Thus assume
that H is a Lie subgroup of G and let h ⊂ g be defined by (3.1). Then h
is the Lie algebra of H. Namely, if η ∈ h, then the curve γ : R → G de-
fine by γ(t) := exp(tη) takes values in H and so η = γ̇(0) ∈ T1l H = Lie(H).
Conversely, if η ∈ Lie(H), then exp(tη) ∈ H for all t ∈ R by Lemma 1.2 and
so η ∈ h. Thus h = Lie(H).
To prove that H is closed, define the map ϕ : H × h⊥ → G by
ϕ(h, ξ) := h exp(ξ), h ∈ H, ξ ∈ h⊥ .
Its derivative at (1l, 0) is bijective. Hence ϕ restricts to a diffeomorphism
from the product of two open neighborhoods V ⊂ H of 1l and W ⊂ h⊥ of
the origin onto the open neighborhood U := ϕ(V × W ) ⊂ G of 1l. Shrinking
these neighborhoods, if necessary, we may assume that
U ∩ H = V. (3.5)
(Otherwise there exists a sequence (hi , ξi ) ∈ V × W converging to (1l, 0) such
that ϕ(hi , ξi ) ∈ H \ V for all i, contradicting the fact that V ⊂ H is a neigh-
borhood of 1l.) Also, there is an open neighborhood U0 ⊂ G of 1l such that
g, g ′ ∈ U0 =⇒ g −1 g ′ ∈ U. (3.6)
Now let hi ∈ H be a sequence that converges to an element g ∈ G. Then
the sequence h−1 −1
i g converges to 1l. Choose i0 ∈ N such that hi g ∈ U0 for
′ −1 −1
all i ≥ i0 , and define (hi , ξi ) := ϕ (hi g) ∈ V × W for i ≥ i0 . Then
h−1 ′
i g = hi exp(ξi ), h′i ∈ V ⊂ H, ξi ∈ W ⊂ h⊥ (3.7)
for i ≥ i0 and
lim h′i = 1l, lim ξi = 0. (3.8)
i→∞ i→∞
For i, j ≥ i0 this implies h′i exp(ξi ) exp(−ξj )(h′j )−1 = h−1
i hj . and hence
9
We prove that (ii) implies (i). Let H ⊂ G be a closed subgroup of G
and let h ⊂ g be the Lie subalgebra defined in equation (3.1) in Lemma 3.3.
Define k := dim(h) and ℓ := dim(g) ≥ k, and choose a basis e1 , . . . , eℓ of g
such that the vectors e1 . . . . , ek form a basis of h and ei ∈ h⊥ for i > k.
Let h0 ∈ H and define the map ψ : Rℓ → G by
Assume, by contradiction, that such an open set Ω0 does not exist. Then
there exists a sequence xi = (x1i , . . . , xℓi ) ∈ Rℓ such that
of the torus is not closed. A similar example can be constructed in any Lie
group that contains a torus of dimension at least two.
10
4 The Haar Measure
Let G be a compact Lie group and denote by C(G) the space of continuous
functions f : G → R with the norm
∥f ∥ := sup |f (g)| .
g∈G
Theorem 4.1. Let G be a compact Lie group. Then there exists a bounded
linear functional M : C(G) → R that satisfies the following conditions.
(i) M (1) = 1.
M is uniquely determined by (i) and either (ii) or (iii). It is called the Haar
measure on G.
P Pℓ
where αi ∈ Q and i αi = 1. If B = j=1 βj δbj is another such measure,
denote
k X
X ℓ
A · B := αi βj δai bj .
i=1 j=1
11
This defines a group structure on A. For A ∈ A we define two linear operators
LA , RA : C(G) → C(G) by
m
X m
X
(LA f )(g) := αi f (ai g), (RA f )(g) := αi f (gai )
i=1 i=1
12
Observation 2: For every f ∈ C(G) the set
L(f ) := {LA f | A ∈ A}
(ag)−1 (ah) = g −1 h ∈ U
13
Observation 3: For every f ∈ C(G), inf A∈A Osc(LA f ) = 0.
Choose a sequence Aν ∈ A such that
The penultimate equality follows from (4.2) and the last inequality from (4.7).
By Observation 1, f0 is constant. Hence Osc(f0 ) = 0 and so Observation 3
follows from (4.7).
Observation 3 shows that there is a sequence Aν ∈ A such that LAν f
converges uniformly to a constant p ∈ R (called a left mean of f ). Similarly,
there exists a sequence Bν ∈ A such that RBν f converges uniformly to a
constant q ∈ R (called a right mean of f ). Since
for f ∈ C(G) and A, B ∈ A it follows that the right and left means agree and
hence are independent of the choices of the sequences Aν and Bν . Namely,
min f ≤ M (f ) ≤ max f.
14
Now let f, f ′ ∈ C(M ) and choose sequences Aν , Bν ∈ A such that
M (f + f ′ ) = M (f ) + M (f ′ )
as claimed.
Next we prove that M is uniquely determined by conditions (i) and (ii).
To see this, let M ′ be another bounded linear functional on C(G) that satis-
fies (i) and (ii). Then M ′ (c) = c for every constant c and
M ′ (LA f ) = M ′ (f )
M (f ) = M ′ (M (f )) = lim M ′ (LAν f ) = M ′ (f ).
ν→∞
This proves uniqueness. That M satisfies condition (v) follows from unique-
ness and the fact that the map C(G) → R : f 7→ M (f ◦ ϕ) is a bounded
linear functional that satisfies (i) and (ii). This proves Theorem 4.1.
15
5 Invariant Inner Products
An inner product ⟨., .⟩ on g is called invariant iff it is invariant under the
adjoint action of G, i.e.
⟨g −1 ξg, g −1 ηg⟩ = ⟨ξ, η⟩
for ξ, η ∈ g and g ∈ G. A Riemannian metric on G is called bi-invariant
iff the left and right translations Lh and Rh are isometries for every h ∈ G.
Every invariant inner product on g determines a bi-invariant metric on G via
⟨v, w⟩ := ⟨g −1 v, g −1 w⟩ = ⟨vg −1 , wg −1 ⟩ (5.1)
for v, w ∈ Tg G. In turn, such a metric determines a volume form and hence
a bi-invariant measure on G. By Theorem 4.1 this agrees with the Haar
measure up to a constant factor. Conversely, if G is compact, one can use
the existence of a translation invariant measure to prove the existence of an
invariant inner product.
Proposition 5.1. Let G be a compact Lie group. Then g carries an invariant
inner product.
Proof. Let M : C(G) → R denote the Haar measure and Q : g×g → R be any
inner product. For ξ, η ∈ g define fξ,η : G → R by fξ,η (g) := Q(gξg −1 , gηg −1 ).
Then the formula ⟨ξ, η⟩ := M (fξ,η ) defines an inner product on g. That it is
invariant follows from the formula fhξh−1 ,hηh−1 = fξ,η ◦ Rh .
Remark 5.2. (i) The proof of Proposition 5.1 shows that the existence of a
right invariant measure on G suffices to establish the existence of an invariant
inner product on g, and hence the existence of a bi-invariant measure on G.
(ii) On any Lie group the existence of a right invariant measure is easy to
prove. Choose any inner product on g and extend it to a Riemannian metric
on G by left translation. Then the right translations are isometries and hence
the volume form defines a right invariant measure on G.
(iii) Combining (i) and (ii) gives rise to a simpler proof of the existence of a
Haar measure for compact Lie groups.
(iv) Uniqueness in Theorem 4.1 implies that every left invariant measure is
right invariant. Here is a direct proof for compact Lie Groups: If ω is a left
invariant volume form on G, then so is Rg ∗ ω. Hence there exists a group
homomorphism λ : G → R such that Rg ∗ ω = eλ(g) ω. Since G is compact, the
only group homomorphism from G to R is λ = 0.
16
Lemma 5.3. Let G be a compact Lie group with a bi-invariant Riemannian
metric. Then the geodesics have the form
γ(t) = exp(tξ)g
for g ∈ G and ξ ∈ g.
Proof. Let I = [a, b] ⊂ R be a closed interval and γ0 : I → G be a geodesic.
Let ξ : I → g be a smooth curve such that ξ(a) = ξ(b) = 0 and consider the
map
γ(s, t) := γ0 (t) exp(sξ(t)).
Then γ −1 ∂s γ = ξ and hence
1 d b −1
Z Z b
−1
⟨γ ∂t γ, γ ∂t γ⟩ dt = ⟨∂s (γ −1 ∂t γ), γ −1 ∂t γ⟩ dt
2 ds a a
Z b
= ⟨∂t (γ −1 ∂s γ), γ −1 ∂t γ⟩ dt
a
Z b
=− ⟨ξ, ∂t (γ −1 ∂t γ)⟩ dt.
a
Here the penultimate equality follows from Lemma 1.6 and Exercise 5.5.
Now γ0 is a geodesic if and only if the left hand side vanishes at s = 0 for
every ξ, and the right hand side vanishes at s = 0 for every ξ if and only
if ∂t (γ0 −1 ∂t γ0 ) ≡ 0. This proves Lemma 5.3.
Exercise 5.4. (i) Prove that the group GL+ (n, R) of real n × n-matrices
with positive determinant is connected.
(ii) Prove that the exponential map exp : Rn×n → GL+ (n, R) is not sur-
jective for n > 1. Hint: Every negative eigenvalue of an exponential ma-
trix Φ = exp(A) must have even multiplicity.
(iii) Prove that Φ2 is an exponential matrix for every Φ ∈ GL(n, R).
(iv) Prove that for every compact connected Lie group G the exponential
map exp : g → G is surjective. Hint: Use Proposition 5.1 (existence of an
invariant inner product), Lemma 5.3 (geodesics and exponential map), and
the Hopf-Rinow theorem (the existance of minimal geodesics).
Exercise 5.5. Let G be a compact connected Lie group. Prove that an inner
product ⟨·, ·⟩ on g := Lie(G) is invariant if and only if
⟨[ξ, η], ζ⟩ = ⟨ξ, [η, ζ]⟩
for ξ, η, ζ ∈ g.
17
6 The Center
Let G be a connected Lie group. The subgroup
Z(G) := {g ∈ G | gh = hg ∀ h ∈ G}
Z(g) := {ξ ∈ g | [ξ, η] = 0 ∀ η ∈ g} .
Note that Z(G) is a normal subgroup and the center of the quotient G/Z(G)
is trivial. The following theorem is due to Herman Weyl.
Theorem 6.1. Let G be a compact connected Lie group. Then the first Betti
number of G is given by dim H 1 (G; R) = dim Z(g).
Proof. The proof consists of three steps.
Step 1: Suppose G is equipped with a bi-invariant Riemannian metric. Then
1
∇v X(g) = dξ(g)v + [ξ(g), η] g,
2
where ξ : G → g, η ∈ g, v = ηg ∈ Tg G, and X(g) = ξ(g)g.
Suppose first that ξ(g) ≡ ξ is constant. Then, by Lemma 5.3, the integral
curves of Xξ are geodesics. Hence
∇Xξ Xξ = 0
∇Xη Xξ + ∇Xξ Xη = 0
it follows that
1
∇Xη Xξ = X[ξ,η] .
2
This proves Step 1 in the case where ξ : G → g is constant. The general case
is an immediate consequence.
18
Step 2: The Riemann curvature tensor of G is given by
1
R(ξg, ηg)ζg = − [[ξ, η], ζ]g
4
for g ∈ G and ξ, η, ζ ∈ g.
Consider the right invariant vector fields Xξ (g) = ξg for ξ ∈ g. By Step 1,
1
∇Xη Xξ = X[ξ,η] .
2
Hence Step 2 follows by straight forward calculation from the identity
Hence Ric(ξg, ξg) ≥ 0 with equality if and only if ξ ∈ Z(g). Now let α ∈
Ω1 (G) and choose ξ : G → g such that
19
Theorem 6.2. Let G be a compact Lie group. Then the following holds.
(i) The fundamental group of G is abelian.
(ii) If Z(G) is finite, then so is π1 (G).
Proof. Assertion (i) holds for every topological group. To see this, choose
two curves α, β : [0, 1] → G with the endpoints
α(0) = α(1) = β(0) = β(1) = 1l.
Denote
α(2t), if 0 ≤ t ≤ 1/2,
α#β(t) :=
α(1)β(2t − 1), if 1/2 ≤ t ≤ 1.
(Here the term α(1) can be dropped, but the more general form will be
needed below.) Define
α(2t − s), if s/2 ≤ t ≤ (s + 1)/2,
αs (t) :=
1l, otherwise,
and
β(2t + s − 1), if (1 − s)/2 ≤ t ≤ 1 − s/2,
βs (t) :=
1l, otherwise,
for 0 ≤ s, t ≤ 1. Then γs (t) = αs (t)βs (t) is a homotopy from γ0 = α#β
to γ1 = β#α. This proves (i).
To prove (ii), note that by (i) the fundamental group
Γ := π1 (G)
is abelian, and by Theorem 6.1
Hom(Γ, R) ∼
= H 1 (G; R) = 0.
This implies that Γ is finite. To see this, note first that Γ is finitely generated.
Let γ1 , . . . , γn ∈ Γ be generators. Since Γ is abelian, the set R ⊂ Zn of all
integer vectors m = (m1 , . . . , mn ) that satisfy
γ1 m1 · · · γn mn = 1
form a subgroup of Zn and there is a natural isomorphism
Γ∼
= Zn /R.
Since Hom(Γ, R) = R⊥ = {0}, it follows that R spans Rn . Hence the quotient
Rn /R is compact, so Γ ∼
= Zn /R is a finite set. This proves Theorem 6.2.
20
Now let us denote by G
e the universal cover of G. In explicit terms,
Define
α((1 + s)t), if 0 ≤ t ≤ 1/(s + 1),
αs (t) :=
α(1), otherwise,
and
β((2t − s)/(2 − s)), if s/2 ≤ t ≤ 1,
βs (t) :=
1l, otherwise,
for 0 ≤ s, t ≤ 1. Since α(1) ∈ Z(G), we have
α1 β1 = β1 α1 = α#β.
[G, G] ⊂ G
21
Proposition 6.4. Let G be a compact connected Lie group. Then Z(G) is
finite if and only if [G, G] = G.
Proof. Choose an invariant inner product on the Lie algebra g = Lie(G)
(Proposition 5.1) and consider the subbundle
By Exercise 5.5, the Lie bracket of any two right invariant vector fields
Xξ (g) = ξg and Xη (g) = ηg is contained in E. Hence, by Frobenius’ theorem,
E is integrable. Let H be the leaf of E through 1l, i.e.
If α, β : [0, 1] → G are paths that are tangent to E, then so are αβ and α−1 .
Hence H is a subgroup of G. Next we prove that
[G, G] ⊂ H.
[G, G] ⊂ H.
22
This map is a continuously differentiable embedding near t = 0 and it is
everywhere tangent to E. Hence the image U0 of a sufficiently small neigh-
bourhood of 0 ∈ Rk under ϕ is a neighbourhood of 1l in H with respect to
the intrinsic topology of H and it is contained in [G, G]. More generally,
for every h ∈ H the set U = U0 h ⊂ H is a neighbourhood of h with re-
spect to the intrinsic topology and U h−1 ⊂ [G, G]. Hence the sets H ∩ [G, G]
and H \ [G, G] are both open with respect to the intrinsic topology of H.
Since H ∩ [G, G] ̸= ∅ it follows that H ⊂ [G, G], as claimed.
Thus we have proved that
[G, G] = H
α1 , . . . , αg , β1 , . . . , βg ∈ G
e
that satisfy
g
Y
[αj , βj ] = γ.
j=1
Exercise 6.6. The nontrivial SO(3)-bundle over the 2-sphere does not carry
a flat connection.
23
7 Inner Automorphisms
Let g be a finite-dimensional Lie algebra. An automorphism Φ ∈ Aut(g) is
called an inner automorphism iff there exists a smooth path η : [0, 1] → g
such that Φ(ξ(0)) = ξ(1) for every solution ξ : [0, 1] → g of the ordinary
differential equation ξ˙ + [ξ, η] = 0. The inner automorphisms of g form a
group which will be denoted by Inn(g) ⊂ Aut(g). If g = Lie(G) is the Lie
algebra of a connected Lie group G, then Inn(g) is the image of the adjoint
representation Ad : G → Aut(g). The following example shows that Inn(g)
need not be a closed subset of Aut(g).
Example 7.1. Let A ∈ GL(V ) be an automorphism of a real or complex
vector space V . Then G = R × V is a Lie group with the product
(s, x) · (t, y) := (s + t, x + exp(sA)y) (7.1)
for s, t ∈ R and x, y ∈ V . The unit is the origin in R × V , the inverse
of (s, x) ∈ G is the pair (s, x)−1 = (−s, − exp(−sA)x), the adjoint action
of G on itself is given by (s, x) · (t, y) · (s, x)−1 = (t, x + exp(sA)y − exp(tA)x),
so the adjoint action of (s, x) ∈ G on the Lie algebra g = R × V is given by
Ad(s, x)(τ, η) = (τ, exp(sA)η − τ Ax) (7.2)
for (τ, η) ∈ g, and hence the Lie bracket on g is given by
[(σ, ξ), (τ, η)] = (0, σAη − τ Aξ) (7.3)
for (σ, ξ), (τ, η) ∈ g. Since A is nonsingular, the center of g is trivial.
Consider the case V = C2 and A = diag(iα, iβ), where α, β are nonzero
real numbers such that α/β is irrational. Then
Ad(s, x)(τ, η) = (τ, esiα η1 − τ iαx1 , esiβ η2 − τ iβx2 )
for s, τ ∈ R and x1 , x2 , η1 , η2 ∈ C by (7.2). Now choose a pair λ1 , λ2 ∈ S 1 such
that (λ1 , λ2 ) ̸= (esiα , esiβ ) for all s ∈ R and define the linear map Φλ : g → g
by Φλ (τ, η) := (τ, λ1 η1 , λ2 , η2 ). Then Φλ ∈ Aut(g) \ Inn(g). Since α/β is irra-
tional, there exists a sequence si ∈ R such that limi→∞ (esi iα , esi iβ ) = (λ1 , λ2 ).
Hence limi→∞ Ad(si , 0) = Φλ and so Inn(g) is not closed in Aut(g).
For every finite-dimensional Lie algebra g the group Inn(g) is the leaf
through the identity of a foliation on Aut(g) with T1l Inn(g) = ad(g). Thus
it carries an intrinsic smooth structure which turns it into a Lie group with
the Lie algebra ad(g) ∼ = g/Z(g). This shows that every finite-dimensional Lie
algebra with a trivial center is the Lie algebra of a connected Lie group.
24
8 Maximal Toral Subgroups
Let G be a compact connected Lie group. A Lie subgroup of G is a closed
subgroup H which is a submanifold. A linear subspace h ⊂ g is called a
Lie subalgebra iff it is invariant under the Lie bracket. If H ⊂ G is a Lie
subgroup, then by definition of the Lie bracket h := T1l H is a Lie subalgebra
of g. A maximal torus in G is a connected abelian subgroup T ⊂ G which
is not properly contained in any other connected abelian subgroup. The
fundamental example is the subgroup of diagonal matrices in U(n) or SU(n).
G → R : g 7→ |gηg −1 − τ |2
25
Lemma 8.3 shows that any two maximal tori in G have the same dimen-
sion. This dimension is called the rank of G. The rank of G agrees with the
dimension of a maximal abelian Lie subalgebra of g. (Prove this!)
Lemma 8.4. Let T ⊂ G be a maximal torus. Then every element of the Lie
algebra g := Lie(G) is conjugate to an element of t := Lie(T).
Proof. Let ξ ∈ g. The set {exp(sξ) | s ∈ R} is a torus and hence is contained
in a maximal torus T′ . By Lemma 8.3, there exists a g ∈ G such that
gT′ g −1 = T. Hence exp(sgξg −1 ) ∈ T for every s ∈ R, and hence gξg −1 ∈ t.
This proves Lemma 8.4.
Lemma 8.5. Let G be a compact connected Lie group and let T ⊂ G be a
maximal torus. Then T is a maximal abelian subgroup of G.
Proof. We follow the argument of Frank Adams in Lectures on Lie groups.
Let h ∈ G be an element that commutes with T. We shall prove that h ∈ T.
To see this let S ⊂ G be a maximal torus containing h and denote by
H := cl hk | k ∈ Z
generates T.
b By Lemma 8.2, there exists a maximal torus containing b h and
hence both h and T. Since T is a maximal torus it follows that h ∈ T. This
proves Lemma 8.5.
26
Example 8.6. In general, a maximal abelian subgroup need not be a torus.
For example the n × n-matrices with diagonal entries ±1 and determinant 1
form a maximal abelian subgroup of G = SO(n).
For every maximal torus T ⊂ G denote
GT := g ∈ G | g −1 Tg = T .
27
9 Isotropy Subgroups
Let G be a compact connected Lie group and M be a compact smooth
manifold equipped with a left action of G. The action will be denoted by
G × M → M : (g, x) 7→ gx.
Gx := {h ∈ G | hx = x} .
Since Ggx = gGx g −1 the set of isotropy subgroups is invariant under conjuga-
tion. The next theorem asserts that the set of conjugacy classes of isotropy
subgroups is finite.
Theorem 9.1. There exist finitely many Lie subgroups H1 , . . . , HN of G such
that for every x ∈ M there exists a j such that Gx is conjugate to Hj .
Proof. The proof is by induction on the dimension of M . If M is zero-
dimensional, then the result is obvious. Now assume that dim M = n > 0
and that the result has been proved for all manifolds of dimensions less
than n. We prove that every point x0 ∈ M has a neighbourhood U in which
only finitely many isotropy subgroups occur up to conjugacy. To see this,
let G0 := Gx0 choose a G-invariant metric on M , denote by Lx : g → Tx M the
infinitesimal action, and define H0 := ker L∗x0 ⊂ Tx0 M . Then the exponential
map
G × H0 → M : (g, v0 ) 7→ g expx0 (v0 )
descends to a map
ϕ0 : G ×G0 H0 → M,
where (g, v0 ) ∼ (gg0 , g0−1 v0 ) for g ∈ G, v0 ∈ H0 , and g0 ∈ G0 . The restriction
of ϕ0 to a sufficiently small neighbourhood of the zero section in the vector
bundle G ×G0 H0 → G/G0 is a G-equivariant diffeomorphism onto a neigh-
bourhood of the G-orbit of x0 . It follows that the isotropy groups of points
x ∈ M belonging to this neighbourhood are all conjugate to subgroups of G0
that appear as isotropy subgroups of the action of G0 on H0 . By consider-
ing the action of G0 on the unit sphere in H0 we obtain from the induction
hypothesis that there are only finitely many such isotropy subgroups. This
proves the local statement. Cover M by finitely many such neighbourhoods
to prove the global assertion of Theorem 9.1.
28
10 Centralizers
Let G be a compact connected Lie group. For any subset H ⊂ G the cen-
tralizer of H is defined by
C(H) := CG (H) := {g ∈ G | gh = hg ∀ h ∈ H}
By Theorem 3.1 this set is a Lie subgroup of G with Lie algebra
\
Lie(C(H)) = ξ ∈ g | hξh−1 = ξ ∀ h ∈ H =
ker(1l − Ad(h)).
h∈H
29
Theorem 10.2. Let G be a compact connected Lie group. Then there exist
finitely many centralizer subgroups H1 , . . . , Hm of G such that every central-
izer subgroup H ⊂ G is conjugate to one of the Hi .
11 Simple Groups
A Lie algebra g is called abelian iff the Lie bracket vanishes. A Lie sub-
algebra h ⊂ g is called an ideal iff [h, g] ⊂ h. The Lie algebra of a normal
Lie subgroup of G is necessarily an ideal. A Lie algebra g is called simple
iff it is not abelian and has no nontrivial ideals (that is {0} and g are the
only ideals in g). It is called semi-simple iff it is a direct sum of simple Lie
algebras. A Lie group is called simple (respectively semi-simple) iff its Lie
algebra is simple (respectively semi-simple).
Theorem 11.1. Every compact connected simply connected simple Lie group
is isomorphic to one in the following list
An := SU(n + 1), n ≥ 1,
Bn := Spin(2n + 1), n ≥ 2,
Cn := Sp(n), n ≥ 3,
Dn := Spin(2n), n ≥ 4,
30
The Killing form
Every Lie algebra carries a natural pairing
κ(ξ, η) := trace(Ad(ξ)Ad(η))
called the Killing form. On su(n) this form is negative definite. In general
the Killing form may have a kernel and/or be indefinite.
Exercise 11.3. Prove that the Killing forms on su(n) and so(2n) are given
by
κ(ξ, η) = −(2n − 1) trace(ξ ∗ η), ξ, η ∈ su(n),
κ(ξ, η) = −(n − 2) trace(ξ T η), ξ, η ∈ so(2n).
Root systems
Let G be a compact Lie group with maximal torus T. The exponential map
is onto by Exercise 5.4 (iv). It determines an isomorphism
T∼
= t/Λ
Λ := {τ ∈ t | exp(τ ) = 1l}
w:t→R
w(Λ) ⊂ Z.
31
Now consider the adjoint representation of T on g. Since the action
preserves any invariant inner product, the commuting Automorphisms ad(τ )
for τ ∈ t are simultaneously diagonalizable (over C). It follows that there
exists a decomposition M
g=t⊕ Vα
α
The weights wα are called the roots of the Lie algebra g. For each α define
τα ∈ t to be the dual element with respect to the Killing form, i.e.
κ(τα , σ) = wα (σ), σ ∈ t.
The length of the longest root is an important invariant of the Lie group G.
We denote the square of its inverse by
1
a(G) := .
supα ℓ(α)2
G dim(G) a(G)
SU(n) n2 − 1 n
1
Spin(n) 2 n(n − 1) n − 2
Sp(n) n(2n + 1) n + 1
G2 14 4
F4 52 9
E6 78 12
E7 133 18
E8 248 30
32
12 Examples
Example 12.1 (General linear group). The group GL(n, R) of invertible
real n × n-matrices is a Lie group. This space is an open set in Rn×n and
hence is obviously a manifold. Its Lie algebra is the vector space Rn×n of all
real n × n matrices with Lie bracket operation
[A, B] := AB − BA.
In this case the exponential map exp : Rn×n → GL(n, R) is the usual ex-
ponential map for matrices and the expressions gv and vg for v ∈ Th G are
given by matrix multiplication. The example GL(n, C) of invertible complex
n × n-matrices is similar. However, the group GL(n, C) is connected while
the group GL(n, R) has two components distinguished by the sign of the
determinant.
det : GL(n, C) → C∗
The formula
det(exp(A)) = exp(trace(A))
shows that the Lie algebra of SL(n, C) is given by
The Lie group SL(n, R) with Lie algebra sl(n, R) is defined analogously.
S 1 := {z ∈ C | |z| = 1}
33
Example 12.4 (Torus). Let V be an n-dimensional real vector space and
Λ ⊂ V be a lattice (a discrete additive subgroup) which spans V . Then
T := V /Λ
is a compact abelian Lie group (the group operation is the addition in V ) with
Lie algebra V . The exponential map is the projection V → V /Λ. Any such
Lie group is called a torus. Tori can be characterized as compact connected
finite-dimensional abelian Lie groups. The basic example is the standard
torus Tn := Rn /Zn and every n-dimensional torus is isomorphic to Tn .
Example 12.5 (Orthogonal group). The orthogonal n × n-matrices form
a Lie group
O(n) := Φ ∈ Rn×n | ΦT Φ = 1l .
The group SO(n) is compact and connected and the exponential map is
surjective (see Exercise 5.4).
Example 12.6 (Unitary group). The unitary n × n-matrices form a Lie
group
U(n) := U ∈ Cn×n | U ∗ U = 1l
34
Example 12.7 (Unit quaternions). Denote by
H = R4
and satisfies the rule |xy| = |x| · |y|. Hence the unit quaternions form a group
Sp(1) := {x ∈ H | |x| = 1}
with unit 1 and inverse map x 7→ x̄. Its Lie algebra consists of the imaginary
quaternions
sp(1) := {x ∈ H | x0 = 0} .
The exponential map is given by the usual formula exp(x) = ∞ n
P
n=0 x /n!.
3
The quaternion multiplication defines a group structure on S = Sp(1) and a
Lie algebra structure on R3 ≃ sp(1). This Lie algebra structure corresponds
to the vector product.
Example 12.8. The quaternion matrices Φ ∈ Hn×n with Φ∗ Φ = 1l form a
compact connected group denoted by Sp(n). Its Lie algebra sp(n) consists
of the quaternion matrices A ∈ Hn×n with A∗ + A = 0. Here A∗ denotes the
conjugate transpose as in the complex case.
Exercise 12.9. (i) Prove that the map
defined by
x0 + ix1 x2 + ix3
U (x) :=
−x2 + ix3 x0 − ix1
is a Lie group isomorphism.
35
(ii) Prove that the corresponding Lie algebra homomorphism
is given by
iξ1 ξ2 + iξ3
u(ξ) := .
−ξ2 + iξ3 −iξ1
Show that the matrices
i 0 0 1 0 i
I= , J= , K= .
0 −i −1 0 i 0
satisfy the quaternion relations. In other words, the Lie algebra su(2) is
isomorphic to the imaginary quaternions and the isomorphism is given by
i 7→ I, j 7→ J, k 7→ K. The natural orientation of SU(2) is determined by
the irdered basis I, J, K of su(2).
(iii) Prove that
for ξ, η ∈ R3 ∼
= Im(H).
Exercise 12.10 (Spin(3)). The unit quaternions act on the imaginary qua-
ternions by conjugation. This determines a homomorphism
defined by
Φ(x)ξ := xξ x̄
for x ∈ Sp(1) and ξ ∈ Im(H) ∼ = R3 . On the left the multiplication is
understood as a product of matrix and vector and on the right as a product
of quaternions.
(i) Prove that
2
x0 + x21 − x22 − x23 2(x1 x2 − x0 x3 ) 2(x0 x2 − x1 x3 )
Φ(x) = 2(x0 x3 − x1 x2 ) x20 − x21 + x22 − x23 2(x2 x3 − x0 x1 ) .
2(x1 x3 − x0 x2 ) 2(x0 x1 − x2 x3 ) x20 − x21 − x22 + x23
(ii) Verify that the map SU(2) → SO(3) : U (x) 7→ Φ(x) is a Lie group
homomorphism and a double cover. Deduce that π1 (SO(3)) = Z2 .
36
(iii) Let su(2) → so(3) : u(ξ) 7→ A(ξ) denote the corresponding Lie algebra
homomorphism. Prove that
0 −ξ3 ξ2
A(ξ) = 2 ξ3 0 −ξ1 .
−ξ2 ξ1 0
Prove that [A(ξ), A(η)] = 2A(ξ × η) and trace(A(ξ)T A(η)) = 8⟨ξ, η⟩.
Exercise 12.11 (Spin(4)). The group Sp(1) × Sp(1) acts on H by the or-
thogonal transformations x 7→ uxv̄ for (u, v) ∈ Sp(1) × Sp(1). Prove that
this action determines a double cover
and find an explicit formula for the matrix Ψ(u, v) ∈ R4×4 defined by
Lemma 12.12. (i) SO(n) is connected and in the case n ≥ 3 its fundamental
group is isomorphic to Z2 . Hence for n ≥ 3 the universal cover of SO(n) is
a compact group (with the same Lie algebra). It is denoted by Spin(n).
(ii) SU(n) is connected and simply connected and π2 (SU(n)) = 0.
(iii) The fundamental group of U(n) is isomorphic to the integers. The
determinant homomorphism det : U(n) → S 1 induces an isomorphism of
fundamental groups.
Proof. The subgroup of all matrices Φ ∈ SO(n) whose first column is the
first unit vector e1 = (1, 0, . . . , 0) ∈ Rn is isomorphic to SO(n − 1). Hence
there is a fibration
SO(n − 1) ,→ SO(n) → S n−1
where the second map sends a matrix in SO(n) to its first column. The
homotopy exact sequence of this fibration has the form
37
To prove (ii) consider the fibration
where the last map sends U ∈ SU(n) to the first column of U . The homotopy
exact sequence of this fibration has the form
For n = 1 the group SU(1) = {1} is obviously connected and simply con-
nected. For n ≥ 2 use the exact sequence inductively (over n) with k = 0, 1.
The statement about π2 is proved similarly with k = 2.
To prove (iii) consider the fibration
SU(n) ,→ U(n) → S 1 .
Lemma 12.13. For every compact oriented smooth 3-manifold Y and every
smooth map g : Y → SU(2) we have
Z
trace g −1 dg ∧ g −1 dg ∧ g −1 dg = −24π 2 deg(g).
Y
Proof. Denote
ωg := trace g −1 dg ∧ g −1 dg ∧ g −1 dg ∈ Ω3 (Y ).
ωg◦f = f ∗ ωg .
38
In particular, with ω0 := ωid ∈ Ω3 (SU(2)), we have ωg = g ∗ ω0 and hence
Z Z
ωg = deg(g) ω0 . (12.1)
Y SU(2)
39
Example 12.15 (Invertible linear operators). For invertible operators
on an infinite-dimensional Hilbert space H the relation between Lie-group
and Lie-algebra is somewhat subtle. Not every one parameter group t 7→ S(t)
of invertible linear operators is differentiable. Such groups can be generated
by unbounded operators and this leads to the theory of semigroups of linear
operators.
References
[1] Élie Joseph Cartan, La théorie des groupes finis et continus et l’Analysis
Situs. Mémorial des Sciences Mathématiques 42 (1930), 1–61.
[2] Joachim Hilgert & Karl-Hermann Neeb, Structure and Geometry of Lie
groups. Springer Monographs in Mathematics, Springer-Verlag, 2012.
[3] Anatolij Ivanovich Malcev, On the theory of the Lie groups in the large.
Matematichevskii Sbornik N. S. 16 (1945), 163–189.
Corrections to my paper ”On the theory of Lie groups in the large”. Matem-
atichevskii Sbornik N. S. 19 (1946), 523–524.
[5] John von Neumann, Über die analytischen Eigenschaften von Gruppen lin-
earer Transformationen und ihrer Darstellungen. Mathematische Zeitschrift
30 (1929), 3–42.
40