Part III Symmetries, Fields and Particles
Based on lectures by N. Dorey
Notes taken by Dexter Chua
Michaelmas 2016
These notes are not endorsed by the lecturers, and I have modified them (often
significantly) after lectures. They are nowhere near accurate representations of what
was actually lectured, and in particular, all errors are almost surely mine.
This course introduces the theory of Lie groups and Lie algebras and their applications
to high energy physics. The course begins with a brief overview of the role of symmetry
in physics. After reviewing basic notions of group theory we define a Lie group as a
manifold with a compatible group structure. We give the abstract definition of a Lie
algebra and show that every Lie group has an associated Lie algebra corresponding to
the tangent space at the identity element. Examples arising from groups of orthogonal
and unitary matrices are discussed. The case of
SU
(2), the group of rotations in three
dimensions is studied in detail. We then study the representations of Lie groups and
Lie algebras. We discuss reducibility and classify the finite dimensional, irreducible
representations of
SU
(2) and introduce the tensor product of representations. The
next part of the course develops the theory of complex simple Lie algebras. We
define the Killing form on a Lie algebra. We introduce the Cartan-Weyl basis and
discuss the properties of roots and weights of a Lie algebra. We cover the Cartan
classification of simple Lie algebras in detail. We describe the finite dimensional,
irreducible representations of simple Lie algebras, illustrating the general theory for the
Lie algebra of
SU
(3). The last part of the course discusses some physical applications.
After a general discussion of symmetry in quantum mechanical systems, we review the
approximate
SU
(3) global symmetry of the strong interactions and its consequences
for the observed spectrum of hadrons. We introduce gauge symmetry and construct a
gauge-invariant Lagrangian for Yang-Mills theory coupled to matter. The course ends
with a brief introduction to the Standard Model of particle physics.
Pre-requisites
Basic finite group theory, including subgroups and orbits. Special relativity and
quantum theory, including orbital angular momentum theory and Pauli spin matrices.
Basic ideas about manifolds, including coordinates, dimension, tangent spaces.
Contents
1 Introduction
2 Lie groups
2.1 Definitions
2.2 Matrix Lie groups
2.3 Properties of Lie groups
3 Lie algebras
3.1 Lie algebras
3.2 Differentiation
3.3 Lie algebras from Lie groups
3.4 The exponential map
4 Representations of Lie algebras
4.1 Representations of Lie groups and algebras
4.2 Complexification and correspondence of representations
4.3 Representations of su(2)
4.4 New representations from old
4.5 Decomposition of tensor product of su(2) representations
5 Cartan classification
5.1 The Killing form
5.2 The Cartan basis
5.3 Things are real
5.4 A real subalgebra
5.5 Simple roots
5.6 The classification
5.7 Reconstruction
6 Representation of Lie algebras
6.1 Weights
6.2 Root and weight lattices
6.3 Classification of representations
6.4 Decomposition of tensor products
7 Gauge theories
7.1 Electromagnetism and U(1) gauge symmetry
7.2 General case
8 Lie groups in nature
8.1 Spacetime symmetry
8.2 Possible extensions
8.3 Internal symmetries and the eightfold way
1 Introduction
In this course, we are, unsurprisingly, going to talk about symmetries. Unlike
what the course name suggests, there will be relatively little discussion of fields
or particles.
So what is a symmetry? There are many possible definitions, but we are
going to pick one that is relevant to physics.
Definition
(Symmetry)
.
A symmetry of a physical system is a transformation
of the dynamical variables which leaves the physical laws invariant.
Symmetries are very important. As Noether’s theorem tells us, every sym-
metry gives rise to a conserved current. But there is more to symmetry than
this. It seems that the whole physical universe is governed by symmetries. We
believe that the forces in the universe are given by gauge fields, and a gauge
field is uniquely specified by the gauge symmetry group it works with.
In general, the collection of all symmetries will form a group.
Definition
(Group)
.
A group is a set
G
of elements with a multiplication rule,
obeying the axioms
(i) For all g
1
, g
2
G, we have g
1
g
2
G. (closure)
(ii)
There is a (necessarily unique) element
e G
such that for all
g G
, we
have eg = ge = g. (identity)
(iii)
For every
g G
, there exists some (necessarily unique)
g
1
G
such that
gg
1
= g
1
g = e. (inverse)
(iv) For every g
1
, g
2
, g
3
G, we have g
1
(g
2
g
3
) = (g
1
g
2
)g
3
. (associativity)
Physically, these mean
(i) The composition of two symmetries is also a symmetry.
(ii) “Doing nothing” is a symmetry.
(iii) A symmetry can be “undone”.
(iv) Composing functions is always associative.
Note that the set of elements G may be finite or infinite.
Definition
(Commutative/abelian group)
.
A group is abelian or commutative
if g
1
g
2
= g
2
g
1
for all g
1
, g
2
G. A group is non-abelian if it is not abelian.
In this course, we are going to focus on smooth symmetries. These are
symmetries that “vary smoothly” with some parameters. One familiar example
is rotation, which can be described by the rotation axis and the angle of rotation.
These are the symmetries that can go into Noether’s theorem. These smooth
symmetries form a special kind of groups known as Lie groups.
One nice thing about these smooth symmetries is that they can be studied by
looking at the “infinitesimal” symmetries. They form a vector space, known as
a Lie algebra. This is a much simpler mathematical structure, and by reducing
the study of a Lie group to its Lie algebra, we make our lives much easier. In
particular, one thing we can do is to classify all (simple) Lie algebras. It turns
out this isn’t too hard, as the notion of (simple) Lie algebra is very restrictive.
After understanding Lie algebras very well, we will move on to study gauge
theories. These are theories obtained when we require that our theories obey
some sort of symmetry condition, and then magically all the interactions come
up automatically.
2 Lie groups
2.1 Definitions
A Lie group is a group and a manifold, where the group operations define smooth
maps. We already know what a group is, so let’s go into a bit more details about
what a manifold is.
The canonical example of a manifold one might want to keep in mind is the
sphere
S
2
. Having lived on Earth, we all know that near any point on Earth,
S
2
looks like
R
2
. But we also know that
S
2
is not
R
2
. We cannot specify a point
on the surface of the Earth uniquely by two numbers. Longitude/latitude might
be a good start, but adding 2
π
to the longitude would give you the same point,
and things also break down at the poles.
Of course, this will not stop us from producing a map of Cambridge. One
can reasonably come up with consistent coordinates just for Cambridge itself.
So what is true about
S
2
is that near each point, we can come up with some
coordinate system, but we cannot do so for the whole space itself.
Definition
(Smooth map)
.
We say a map
f
:
R
n
R
m
is smooth if all partial
derivatives of all orders exist.
Definition
(Manifold)
.
A manifold (of dimension
n
) is a set
M
together with
the following data:
(i) A collection U
α
of subsets of M whose union is M;
(ii)
A collection of bijections
ϕ
α
:
U
α
V
α
, where
V
α
is an open subset of
R
n
.
These are known as charts.
The charts have to satisfy the following compatibility condition:
For all
α, β
, we have
ϕ
α
(
U
α
U
β
) is open in
R
n
, and the transition function
ϕ
α
ϕ
1
β
: ϕ
β
(U
α
U
β
) ϕ
α
(U
α
U
β
)
is smooth.
We write (M, ϕ
α
) for the manifold if we need to specify the charts explicitly.
U
β
U
α
ϕ
β
ϕ
α
ϕ
α
ϕ
1
β
Definition
(Smooth map)
.
Let
M
be a manifold. Then a map
f
:
M R
is smooth if it is smooth in each coordinate chart. Explicitly, for each chart
(U
α
, ϕ
α
), the composition f ϕ
1
α
: ϕ
α
(U
α
) R is smooth (in the usual sense)
p
U
ϕ(p)
f ϕ
1
R
f
ϕ
More generally, if
M, N
are manifolds, we say
f
:
M N
is smooth if for
any chart (
U, ϕ
) of
M
and any chart (
V, ξ
) of
N
, the composition
ξ f ϕ
1
:
ϕ(U) ξ(V ) is smooth.
ϕ
ξ
f
ξ f ϕ
1
Finally, we note that if
M, N
are manifolds of dimensions
m
and
n
respectively,
then
M × N
is a manifold of dimension
m
+
n
, where charts (
U, ϕ
:
U
R
m
)
,
(
V, ξ
:
V R
n
) of
M
and
N
respectively give us a chart
ϕ × ξ
:
U × V
R
m
× R
n
= R
m+n
of M × N .
All those definitions were made so that we can define what a Lie group is:
Definition
(Lie group)
.
A Lie group is a group
G
whose underlying set is given
a manifold structure, and such that the multiplication map
m
:
G ×G G
and
inverse map
i
:
G G
are smooth maps. We sometimes write
M
(
G
) for the
underlying manifold of G.
Example. The unit 2-sphere
S
2
= {(x, y, z) R
3
: x
2
+ y
2
+ z
2
= 1}
is a manifold. Indeed, we can construct a coordinate patch near
N
= (0
,
0
,
1).
Near this point, we have
z
=
p
1 x
2
y
2
. This works, since near the north pole,
the
z
-coordinate is always positive. In this case,
x
and
y
are good coordinates
near the north pole.
However, it is a fact that the 2-sphere S
2
has no Lie group structure.
In general, most of our manifolds will be given by subsets of Euclidean space
specified by certain equations, but note that not all subsets given by equations
are manifolds! It is possible that they have some singularities.
Definition
(Dimension of Lie group)
.
The dimension of a Lie group
G
is the
dimension of the underlying manifold.
Definition
(Subgroup)
.
A subgroup
H
of
G
is a subset of
G
that is also a group
under the same operations. We write H G if H is a subgroup of G.
The really interesting thing is when the subgroup is also a manifold!
Definition
(Lie subgroup)
.
A subgroup is a Lie subgroup if it is also a manifold
(under the induced smooth structure).
We now look at some examples.
Example.
Let
G
= (
R
D
,
+) be the
D
-dimensional Euclidean space with addition
as the group operation. The inverse of a vector
x
is
x
, and the identity is
0
.
This is obviously locally homeomorphic to
R
D
, since it is
R
D
, and addition and
negation are obviously smooth.
This is a rather boring example, since
R
D
is a rather trivial manifold, and
the operation is commutative. The largest source of examples we have will be
matrix Lie groups.
2.2 Matrix Lie groups
We write
Mat
n
(
F
) for the set of
n×n
matrices with entries in a field
F
(usually
R
or
C
). Matrix multiplication is certainly associative, and has an identity, namely
the identity I. However, it doesn’t always have inverses not all matrices are
invertible! So this is not a group (instead, we call it a monoid). Thus, we are
led to consider the general linear group:
Definition (General linear group). The general linear group is
GL(n, F) = {M Mat
n
(F) : det M 6= 0}.
This is closed under multiplication since the determinant is multiplicative,
and matrices with non-zero determinant are invertible.
Definition (Special linear group). The special linear group is
SL(n, F) = {M Mat
n
(F) : det M = 1} GL(n, F).
While these are obviously groups, less obviously, these are in fact Lie groups!
In the remainder of the section, we are going to casually claim that all our favorite
matrix groups are Lie groups. Proving these take some work and machinery,
and we will not bother ourselves with that too much. However, we can do it
explicitly for certain special cases:
Example. Explicitly, we can write
SL(2, R) =

a b
c d
: a, b, c, d R, ad bc = 1
.
The identity is the matrix with a = d = 1 and b = c = 0. For a 6= 0, we have
d =
1 + bc
a
.
This gives us a coordinate patch for all points where
a 6
= 0 in terms of
b, c, a
,
which, in particular, contains the identity
I
. By considering the case where
b 6
= 0, we obtain a separate coordinate chart, and these together cover all of
SL(2, R), as a matrix in SL(2, R) cannot have a = b = 0.
Thus, we see that SL(2, R) has dimension 3.
In general, by a similar counting argument, we have
dim(SL(n, R)) = n
2
1 dim(SL(n, C)) = 2n
2
2
dim(GL(n, R)) = n
2
, dim(GL(n, C)) = 2n
2
.
Most of the time, we are interested in matrix Lie groups, which will be subgroups
of GL(n, R).
Subgroups of GL(n, R)
Lemma. The general linear group:
GL(n, R) = {M Mat
n
(R) : det M 6= 0}
and orthogonal group:
O(n) = {M GL(n, R) : M
T
M = I}
are Lie groups.
Note that we write O(
n
) instead of O(
n, R
) since orthogonal matrices make
sense only when talking about real matrices.
The orthogonal matrices are those that preserve the lengths of vectors. Indeed,
for v R
n
, we have
|Mv|
2
= v
T
M
T
Mv = v
T
v = |v|
2
.
We notice something interesting. If M O(n), we have
1 = det(I) = det(M
T
M) = det(M)
2
.
So
det
(
M
) =
±
1. Now
det
is a continuous function, and it is easy to see that
det
takes both
±
1. So O(
n
) has (at least) two connected components. Only one of
these pieces contains the identity, namely the piece
det M
= 1. We might expect
this to be a group on its own right, and indeed it is, because
det
is multiplicative.
Lemma. The special orthogonal group SO(n):
SO(n) = {M O(n) : det M = 1}
is a Lie group.
Given a frame
{v
1
, ··· , v
n
}
in
R
n
(i.e. an ordered basis), any orthogonal
matrix
M
O(
n
) acts on it to give another frame
v
a
R
n
7→ v
0
a
7→ M v
a
R
n
.
Definition
(Volume element)
.
Given a frame
{v
1
, ··· , v
n
}
in
R
n
, the volume
element is
Ω = ε
i
1
...i
n
v
i
1
1
v
i
2
2
···v
i
n
n
.
By direct computation, we see that an orthogonal matrix preserves the sign
of the volume element iff its determinant is +1, i.e. M SO(n).
We now want to find a more explicit description of these Lie groups, at
least in low dimensions. Often, it is helpful to classify these matrices by their
eigenvalues:
Definition
(Eigenvalue)
.
A complex number
λ
is an eigenvalue of
M M
(
n
)
if there is some (possibly complex) vector v
λ
6= 0 such that
Mv
λ
= λv
λ
.
Theorem.
Let
M
be a real matrix. Then
λ
is an eigenvalue iff
λ
is an eigenvalue.
Moreover, if M is orthogonal, then |λ|
2
= 1.
Proof.
Suppose
Mv
λ
=
λv
λ
. Then applying the complex conjugate gives
Mv
λ
= λ
v
λ
.
Now suppose
M
is orthogonal. Then
Mv
λ
=
λv
λ
for some non-zero
v
λ
. We
take the norm to obtain
|Mv
λ
|
=
|λ||v
λ
|
. Using the fact that
|v
λ
|
=
|Mv
λ
|
, we
have |λ| = 1. So done.
Example.
Let
M SO
(2). Since
det M
= 1, the eigenvalues must be of the
form λ = e
, e
. In this case, we have
M = M (θ) =
cos θ sin θ
sin θ cos θ
,
where θ is the rotation angle in S
1
. Here we have
M(θ
1
)M(θ
2
) = M(θ
2
)M(θ
1
) = M(θ
1
+ θ
2
).
So we have M(SO(2)) = S
1
.
Example.
Consider
G
=
SO
(3). Suppose
M SO
(3). Since
det M
= +1, and
the eigenvalues have to come in complex conjugate pairs, we know one of them
must be 1. Then the other two must be of the form e
, e
, where θ S
1
.
We pick a normalized eigenvector
n
for
λ
= 1. Then
Mn
=
n
, and
n · n
= 1.
This is known as the axis of rotation. Similarly,
θ
is the angle of rotation. We
write M (n, θ) for this matrix, and it turns out this is
M(n, θ)
ij
= cos θδ
ij
+ (1 cos θ)n
i
n
j
sin θε
ijk
n
k
.
Note that this does not uniquely specify a matrix. We have
M(n, 2π θ) = M(n, θ).
Thus, to uniquely specify a matrix, we need to restrict the range of
θ
to 0
θ π
,
with the further identification that
(n, π) (n, π).
Also note that M(n, 0) = I for any n.
Given such a matrix, we can produce a vector
w
=
θn
. Then
w
lies in the
region
B
3
= {w R
3
: kwk π} R
3
.
This has a boundary
B
3
= {w R
3
: kwk = π}
=
S
2
.
Now we identify antipodal points on
B
3
. Then each vector in the resulting
space corresponds to exactly one element of SO(3).
Subgroups of GL(n, C)
We can similarly consider subgroups of GL(n, C). Common examples include
Definition (Unitary group). The unitary group is defined by
U(n) = {U GL(n, C) : U
U = I}.
These are important in physics, because unitary matrices are exactly those
that preserve the norms of vectors, namely kvk = kUvk for all v.
Again, if
U
U
= 1, then
|det
(
U
)
|
2
= 1. So
det U
=
e
for some
δ R
.
Unlike the real case, the determinant can now take a continuous range of values,
and this no longer disconnects the group. In fact, U(n) is indeed connected.
Definition (Special unitary group). The special unitary group is defined by
SU(n) = {U U(n) : det U = 1}.
It is an easy exercise to show that
dim[U(n)] = 2n
2
n
2
= n
2
.
For
SU
(
n
), the determinant condition imposes an additional single constraint,
so we have
dim[SU(n)] = n
2
1.
Example. Consider the group G = U(1). This is given by
U(1) = {z C : |z| = 1}.
Therefore we have
M[U(1)] = S
1
.
However, we also know another Lie group with underlying manifold
S
1
, namely
SO(2). So are they “the same”?
2.3 Properties of Lie groups
The first thing we want to consider is when two Lie groups are “the same”. We
take the obvious definition of isomorphism.
Definition
(Homomorphism of Lie groups)
.
Let
G, H
be Lie groups. A map
J : G H is a homomorphism if it is smooth and for all g
1
, g
2
G, we have
J(g
1
g
2
) = J(g
1
)J(g
2
).
(the second condition says it is a homomorphism of groups)
Definition
(Isomorphic Lie groups)
.
An isomorphism of Lie groups is a bijective
homomorphism whose inverse is also a homomorphism. Two Lie groups are
isomorphic if there is an isomorphism between them.
Example. We define the map J : U(1) SO(2) by
J(e
) 7→
cos θ sin θ
sin θ cos θ
SO(2).
This is easily seen to be a homomorphism, and we can construct an inverse
similarly.
Exercise. Show that M(SU(2))
=
S
3
.
We now look at some words that describe manifolds. Usually, manifolds that
satisfy these properties are considered nice.
The first notion is the idea of compactness. The actual definition is a bit
weird and takes time to get used to, but there is an equivalent characterization
if the manifold is a subset of R
n
.
Definition
(Compact)
.
A manifold (or topological space)
X
is compact if every
open cover of X has a finite subcover.
If the manifold is a subspace of some
R
n
, then it is compact iff it is closed
and bounded.
Example.
The sphere
S
2
is compact, but the hyperboloid given by
x
2
y
2
z
2
=
1 (as a subset of R
3
) is not.
Example.
The orthogonal groups are compact. Recall that the definition of
an orthogonal matrix requires
M
T
M
=
I
. Since this is given by a polynomial
equation in the entries, it is closed. It is also bounded since each entry of
M
has
magnitude at most 1.
Similarly, the special orthogonal groups are compact.
Example.
Sometimes we want to study more exciting spaces such as Minkowski
spaces. Let
n
=
p
+
q
, and consider the matrices that preserve the metric on
R
n
of signature (p, q), namely
O(p, q) = {M GL(n, R) : M
T
ηM = η},
where
η =
I
p
0
0 I
q
.
For
p, q
both non-zero, this group is non-compact. For example, if we take
SO(1, 1), then the matrices are all of the form
M =
cosh θ sinh θ
sinh θ cosh θ
,
where θ R. So this space is homeomorphic to R, which is not compact.
Another common example of a non-compact group is the Lorentz group
O(3, 1).
Another important property is simply-connectedness.
Definition
(Simply connected)
.
A manifold
M
is simply connected if it is
connected (there is a path between any two points), and every loop
l
:
S
1
M
can be contracted to a point. Equivalently, any two paths between any two
points can be continuously deformed into each other.
Example.
The circle
S
1
is not simply connected, as the loop that goes around
the circle once cannot be continuously deformed to the loop that does nothing
(this is non-trivial to prove).
Example.
The 2-sphere
S
2
is simply connected, but the torus is not.
SO
(3) is
also not simply connected. We can define the map by
l(θ) =
(
θn θ [0, π)
(2π θ)n θ [π, 2π)
This is indeed a valid path because we identify antipodal points.
The failure of simply-connectedness is measured by the fundamental group.
Definition
(Fundamental group/First homotopy group)
.
Let
M
be a manifold,
and
x
0
M
be a preferred point. We define
π
1
(
M
) to be the equivalence classes
of loops starting and ending at
x
0
, where two loops are considered equivalent if
they can be continuously deformed into each other.
This has a group structure, with the identity given by the “loop” that stays
at x
0
all the time, and composition given by doing one after the other.
Example. π
1
(
S
2
) =
{
0
}
and
π
1
(
T
2
) =
Z ×Z
. We also have
π
1
(
SO
(3)) =
Z/
2
Z
.
We will not prove these results.
3 Lie algebras
It turns out that in general, the study of a Lie group can be greatly simplified by
studying its associated Lie algebra, which can be thought of as an infinitesimal
neighbourhood of the identity in the Lie group.
To get to that stage, we need to develop some theory of Lie algebra, and also
of differentiation.
3.1 Lie algebras
We begin with a rather formal and weird definition of a Lie algebra.
Definition
(Lie algebra)
.
A Lie algebra
g
is a vector space (over
R
or
C
) with
a bracket
[ ·, ·] : g × g g
satisfying
(i) [X, Y ] = [Y, X] for all X, Y g (antisymmetry)
(ii)
[
αX
+
βY, Z
] =
α
[
X, Z
] +
β
[
Y, Z
] for all
X, Y, Z g
and
α, β F
((bi)linearity)
(iii)
[
X,
[
Y, Z
]] + [
Y,
[
Z, X
]] + [
Z,
[
X, Y
]] = 0 for all
X, Y, Z g
.(Jacobi identity)
Note that linearity in the second argument follows from linearity in the first
argument and antisymmetry.
Some (annoying) pure mathematicians will complain that we should state
anti-symmetry as [
X, X
] = 0 instead, which is a stronger condition if we are
working over a field of characteristic 2, but I do not care about such fields.
There isn’t much one can say to motivate the Jacobi identity. It is a property
that our naturally-occurring Lie algebras have, and turns out to be useful when
we want to prove things about Lie algebras.
Example.
Suppose we have a vector space
V
with an associative product (e.g.
a space of matrices with matrix multiplication). We can then turn
V
into a Lie
algebra by defining
[X, Y ] = XY Y X.
We can then prove the axioms by writing out the expressions.
Definition
(Dimension of Lie algebra)
.
The dimension of a Lie algebra is the
dimension of the underlying vector space.
Given a finite-dimensional Lie algebra, we can pick a basis B for g.
B = {T
a
: a = 1, ··· , dim g}.
Then any X g can be written as
X = X
a
T
a
=
n
X
a=1
X
a
T
a
,
where X
a
F and n = dim g.
By linearity, the bracket of elements X, Y g can be computed via
[X, Y ] = X
a
Y
b
[T
a
, T
b
].
In other words, the whole structure of the Lie algebra can be given by the bracket
of basis vectors. We know that [
T
a
, T
b
] is again an element of
g
. So we can write
[T
a
, T
b
] = f
ab
c
T
c
,
where f
ab
c
F are the structure constants.
Definition
(Structure constants)
.
Given a Lie algebra
g
with a basis
B
=
{T
a
}
,
the structure constants are f
ab
c
given by
[T
a
, T
b
] = f
ab
c
T
c
,
By the antisymmetry of the bracket, we know
Proposition.
f
ba
c
= f
ab
c
.
By writing out the Jacobi identity, we obtain
Proposition.
f
ab
c
f
cd
e
+ f
da
c
f
cb
e
+ f
bd
c
f
ca
e
= 0.
As before, we would like to know when two Lie algebras are the same.
Definition
(Homomorphism of Lie algebras)
.
A homomorphism of Lie algebras
g, h is a linear map f : g h such that
[f(X), f(Y )] = f([X, Y ]).
Definition
(Isomorphism of Lie algebras)
.
An isomorphism of Lie algebras is a
homomorphism with an inverse that is also a homomorphism. Two Lie algebras
are isomorphic if there is an isomorphism between them.
Similar to how we can have a subgroup, we can also have a subalgebra
h
of
g
.
Definition
(Subalgebra)
.
A subalgebra of a Lie algebra
g
is a vector subspace
that is also a Lie algebra under the bracket.
Recall that in group theory, we have a stronger notion of a normal subgroup,
which are subgroups invariant under conjugation. There is an analogous notion
for subalgebras.
Definition
(Ideal)
.
An ideal of a Lie algebra
g
is a subalgebra
h
such that
[X, Y ] h for all X g and Y h.
Example. Every Lie algebra g has two trivial ideals h = {0} and h = g.
Definition (Derived algebra). The derived algebra of a Lie algebra g is
i = [g, g] = span
F
{[X, Y ] : X, Y g},
where F = R or C depending on the underlying field.
It is clear that this is an ideal. Note that this may or may not be trivial.
Definition (Center of Lie algebra). The center of a Lie algebra g is given by
ξ(g) = {X g : [X, Y ] = 0 for all Y g}.
This is an ideal, by the Jacobi identity.
Definition
(Abelian Lie algebra)
.
A Lie algebra
g
is abelian if [
X, Y
] = 0 for
all X, Y g. Equivalently, if ξ(g) = g.
Definition
(Simple Lie algebra)
.
A simple Lie algebra is a Lie algebra
g
that is
non-abelian and possesses no non-trivial ideals.
If
g
is simple, then since the center is always an ideal, and it is not
g
since
g
is not abelian, we must have
ξ
(
g
) =
{
0
}
. On the other hand, the derived algebra
is also an ideal, and is non-zero since it is not abelian. So we must have
i
(
g
) =
g
.
We will later see that these are the Lie algebras on which we can define a non-
degenerate invariant inner product. In fact, there is a more general class, known
as the semi-simple Lie algebras, that are exactly those for which non-degenerate
invariant inner products can exist.
These are important in physics because, as we will later see, to define the
Lagrangian of a gauge theory, we need to have a non-degenerate invariant inner
product on the Lie algebra. In other words, we need a (semi-)simple Lie algebra.
3.2 Differentiation
We are eventually going to get a Lie algebra from a Lie group. This is obtained
by looking at the tangent vectors at the identity. When we have homomorphisms
f
:
G H
of Lie groups, they are in particular smooth, and taking the derivative
will give us a map from tangent vectors in
G
to tangent vectors in
H
, which
in turn restricts to a map of their Lie algebras. So we need to understand how
differentiation works.
Before that, we need to understand how tangent vectors work. This is
completely general and can be done for manifolds which are not necessarily Lie
groups. Let
M
be a smooth manifold of dimension
D
and
p M
a point. We
want to formulate a notion of a “tangent vector” at the point
p
. We know how
we can do this if the space is
R
n
a tangent vector is just any vector in
R
n
.
By definition of a manifold, near a point p, the manifold looks just like R
n
. So
we can just pretend it is R
n
, and use tangent vectors in R
n
.
However, this definition of a tangent vector requires us to pick a particular
coordinate chart. It would be nice to have a more “intrinsic” notion of vectors.
Recall that in
R
n
, if we have a function
f
:
R
n
R
and a tangent vector
v
at
p
, then we can ask for the directional derivative of
f
along
v
. We have a
correspondence
v
v
.
This directional derivative takes in a function and returns its derivative at a
point, and is sort-of an “intrinsic” notion. Thus, instead of talking about
v
, we
will talk about the associated directional derivative
v
.
It turns out the characterizing property of this directional derivative is the
product rule:
v
(fg) = f (p)
v
g + g(p)
v
f.
So a “directional derivative” is a linear map from the space of smooth functions
M R to R that satisfies the Leibnitz rule.
Definition
(Tangent vector)
.
Let
M
be a manifold and write
C
(
M
) for the
vector space of smooth functions on
M
. For
p M
, a tangent vector is a linear
map v : C
(M) R such that for any f, g C
(M), we have
v(fg) = f(p)v(g) + v(f)g(p).
It is clear that this forms a vector space, and we write
T
p
M
for the vector space
of tangent vectors at p.
Now of course one would be worried that this definition is too inclusive, in
that we might have included things that are not genuinely directional derivatives.
Fortunately, this is not the case, as the following proposition tells us.
In the case where
M
is a submanifold of
R
n
, we can identify the tangent
space with an actual linear subspace of
R
n
. This is easily visualized when
M
is a
surface in
R
3
, where the tangent vectors consists of the vectors in
R
3
“parallel to”
the surface at the point, and in general, a “direction” in
M
is also a “direction”
in
R
n
, and tangent vectors of
R
n
can be easily identified with elements of
R
n
in
the usual way.
This will be useful when we study matrix Lie groups, because this means the
tangent space will consist of matrices again.
Proposition.
Let
M
be a manifold with local coordinates
{x
i
}
i=1,··· ,D
for some
region U M containing p. Then T
p
M has basis
x
j
j=1,··· ,D
.
In particular, dim T
p
M = dim M .
This result on the dimension is extremely useful. Usually, we can manage to
find a bunch of things that we know lie in the tangent space, and to show that
we have found all of them, we simply count the dimensions.
One way we can obtain tangent vectors is by differentiating a curve.
Definition
(Smooth curve)
.
A smooth curve is a smooth map
γ
:
R M
.
More generally, a curve is a C
1
function R M.
Since we only want the first derivative, being C
1
is good enough.
There are two ways we can try to define the derivative of the curve at time
t
= 0
R
. Using the definition of a tangent vector, to specify
˙γ
(0) is to tell how
we can differentiate a function
f
:
M R
at
p
=
γ
(0) in the direction of
˙γ
(0).
This is easy. We define
˙γ(0)(f) =
d
dt
f(γ(t)) R.
If this seems too abstract, we can also do it in local coordinates.
We introduce some coordinates
{x
i
}
near
p M
. We then refer to
γ
by
coordinates (at least near p), by
γ : t R 7→ {x
i
(t) R : i = 1, ··· , D}.
By the smoothness condition, we know
x
i
(
t
) is differentiable, with
x
i
(0) = 0.
Then the tangent vector of the curve γ at p is
v
γ
= ˙x
i
(0)
x
i
T
p
(M), ˙x
i
(t) =
dx
i
dt
.
It follows from the chain rule that this exactly the same thing as what we
described before.
More generally, we can define the derivative of a map between manifolds.
Definition
(Derivative)
.
Let
f
:
M N
be a map between manifolds. The
derivative of f at p M is the linear map
Df
p
: T
p
M T
f(p)
N
given by the formula
(Df
p
)(v)(g) = v(g f )
for v T
p
M and g C
(N).
This will be useful later when we want to get a map of Lie algebras from a
map of Lie groups.
3.3 Lie algebras from Lie groups
We now try to get a Lie algebra from a Lie group G, by considering T
e
(G).
Theorem.
The tangent space of a Lie group
G
at the identity naturally admits
a Lie bracket
[ ·, ·] : T
e
G × T
e
G T
e
G
such that
L(G) = (T
e
G, [ ·, ·])
is a Lie algebra.
Definition
(Lie algebra of a Lie group)
.
Let
G
be a Lie group. The Lie algebra
of
G
, written
L
(
G
) or
g
, is the tangent space
T
e
G
under the natural Lie bracket.
The general convention is that if the name of a Lie group is in upper case
letters, then the corresponding Lie algebra is the same name with lower case
letters in fraktur font. For example, the Lie group of SO(n) is so(n).
Proof.
We will only prove it for the case of a matrix Lie group
G Mat
n
(
F
).
Then
T
I
G
can be naturally identified as a subspace of
Mat
n
(
F
). There is then
an obvious candidate for the Lie bracket the actual commutator:
[X, Y ] = XY Y X.
The basic axioms of a Lie algebra can be easily (but painfully) checked.
However, we are not yet done. We have to check that if we take the bracket
of two elements in
T
I
(
G
), then it still stays within
T
I
(
G
). This will be done by
producing a curve in G whose derivative at 0 is the commutator [X, Y ].
In general, let
γ
be a smooth curve in
G
with
γ
(0) =
I
. Then we can Taylor
expand
γ(t) = I + ˙γ(0)t + ¨γ(0)t
2
+ O(t
3
),
Now given
X, Y T
e
G
, we take curves
γ
1
, γ
2
such that
˙γ
1
(0) =
X
and
˙γ
2
(0) =
Y
.
Consider the curve given by
γ(t) = γ
1
1
(t)γ
1
2
(t)γ
1
(t)γ
2
(t) G.
We can Taylor expand this to find that
γ(t) = I + [X, Y ]t
2
+ O(t
3
).
This isn’t too helpful, as [
X, Y
] is not the coefficient of
t
. We now do the slightly
dodgy step, where we consider the curve
˜γ(t) = γ(
t) = I + [X, Y ]t + O(t
3/2
).
Now this is only defined for
t
0, but it is good enough, and we see that its
derivative at
t
= 0 is [
X, Y
]. So the commutator is in
T
I
(
G
). So we know that
L(G) is a Lie algebra.
Example.
Let
G
=
GL
(
n, F
), where
F
=
R
or
C
. Then
L
(
GL
(
n, F
)) =
gl
(
n, F
) =
Mat
n
(
F
) because we know it must be an
n × n
-dimensional sub-
space of Mat
n
(F).
More generally, for a vector space
V
, say of dimension
n
, we can consider
the group of invertible linear maps
V V
, written
GL
(
V
). By picking a basis
of
V
, we can construct an isomorphism
GL
(
V
)
=
GL
(
n, F
), and this gives us a
smooth structure on
GL
(
V
) (this does not depend on the basis chosen). Then
the Lie algebra gl(V ) is the collection of all linear maps V V .
Example. If G = SO(2), then the curves are of the form
g(t) = M(θ(t)) =
cos θ(t) sin θ(t)
sin θ(t) cos θ(t)
SO(2).
So we have
˙g(0) =
0 1
1 0
˙
θ(0).
Since the Lie algebra has dimension 1, these are all the matrices in the Lie
algebra. So the Lie algebra is given by
so(2) =

0 c
c 0
, c R
.
Example.
More generally, suppose
G
=
SO
(
n
), and we have a path
R
(
t
)
SO(n).
By definition, we have
R
T
(t)R(t) = I.
Differentiating gives
˙
R
T
(t)R(t) + R
T
(t)
˙
R(t) = 0.
for all t R. Evaluating at t = 0, and noting that R(0) = I, we have
X
T
+ X = 0,
where
X
=
˙
R
(0) is a tangent vector. There are no further constraints from
demanding that
det R
= +1, since this is obeyed anyway for any matrix in O(
n
)
near I.
By dimension counting, we know the antisymmetric matrices are exactly the
matrices in L(O(n)) or L(SO(n)). So we have
o(n) = so(n) = {X Mat
n
(R) : X
T
= X}.
Example.
Consider
G
=
SU
(
n
). Suppose we have a path
U
(
t
)
SU
(
n
), with
U(0) = I. Then we have
U
(t)U(t) = I.
Then again by differentiation, we obtain
Z
+ Z = 0,
where Z =
˙
U(0) su(n). So we must have
su(n) {Z Mat
n
(C) : Z
= Z}.
What does the condition
det U
(
t
) = 1 give us? We can do a Taylor expansion by
det U(t) = 1 + tr Z · t + O(t
2
).
So requiring that det U(t) = 1 gives the condition
tr Z = 0.
By dimension counting, we know traceless anti-Hermitian matrices are all the
elements in the Lie algebra. So we have
su(n) = {Z Mat
n
(C), Z
= Z, tr Z = 0}.
Example.
We look at
SU
(2) in detail. We know that
su
(2) is the 2
×
2 traceless
anti-Hermitian matrices.
These are given by multiples of the Pauli matrices
σ
j
, for
j
= 1
,
2
,
3, satisfying
σ
i
σ
j
= δ
ij
I +
ijk
σ
k
.
They can be written explicitly as
σ
1
=
0 1
1 0
, σ
2
=
0 i
i 0
, σ
3
=
1 0
0 1
One can check manually that the generators for the Lie algebra are given by
T
a
=
1
2
a
.
Indeed, each
T
a
is in
su
(2), and they are independent. Since we know
dim su
(2) =
3, these must generate everything.
We have
[T
a
, T
b
] =
1
4
[σ
a
, σ
b
] =
1
2
abc
σ
c
= f
ab
c
T
c
,
where the structure constants are
f
ab
c
= ε
abc
.
Example.
Take
G
=
SO
(3). Then
so
(3) is the space of 3
×
3 real anti-symmetric
matrices, which one can manually check are generated by
˜
T
1
=
0 0 0
0 0 1
0 1 0
,
˜
T
2
=
0 0 1
0 0 0
1 0 0
,
˜
T
3
=
0 1 0
1 0 0
0 0 0
We then have
(
˜
T
a
)
bc
= ε
abc
.
Then the structure constants are
[
˜
T
a
,
˜
T
b
] = f
ab
c
˜
T
c
,
where
f
ab
c
= ε
abc
.
Note that the structure constants are the same! Since the structure constants
completely determine the brackets of the Lie algebra, if the structure constants
are the same, then the Lie algebras are isomorphic. Of course, the structure
constants depend on which basis we choose. So the real statement is that if
there are some bases in which the structure constants are equal, then the Lie
algebras are isomorphic.
So we get that
so
(3)
=
su
(2), but
SO
(3) is not isomorphic to
SU
(2). Indeed,
the underlying manifold is
SU
(2) is the 3-sphere, but the underlying manifold of
SO
(3) has a fancy construction. They are not even topologically homeomorphic,
since SU(2) is simply connected, but SO(3) is not. More precisely, we have
π
1
(SO(3)) = Z/2Z
π
1
(SU(2)) = {0}.
So we see that we don’t have a perfect correspondence between Lie algebras and
Lie groups. However, usually, two Lie groups with the same Lie algebra have
some covering relation. For example, in this case we have
SO(3) =
SU(2)
Z/2Z
,
where
Z/2Z = {I, I} SU(2)
is the center of SU(2).
We can explicitly construct this bijection as follows. We define the map
d : SU(2) SO(3) by
d(A)
ij
=
1
2
tr(σ
i
j
A
) SO(3).
This is globally a 2-to-1 map. It is easy to see that
d(A) = d(A),
and conversely if
d
(
A
) =
d
(
B
), then
A
=
B
. By the first isomorphism theorem,
this gives an isomorphism
SO(3) =
SU(2)
Z/2Z
,
where Z/2Z = {I, I} is the center of SU(2).
Geometrically, we know
M
(
SU
(2))
=
S
3
. Then the manifold of
SO
(3) is
obtained by identifying antipodal points of the manifold.
3.4 The exponential map
So far, we have been talking about vectors in the tangent space of the identity
e
. It turns out that the group structure means this tells us about the tangent
space of all points. To see this, we make the following definition:
Definition
(Left and right translation)
.
For each
h G
, we define the left and
right translation maps
L
h
: G G
g 7→ hg,
R
h
: G G
g 7→ gh.
These maps are bijections, and in fact diffeomorphisms (i.e. smooth maps with
smooth inverses), because they have smooth inverses
L
h
1
and
R
h
1
respectively.
In general, there is no reason to prefer left translation over right, or vice
versa. By convention, we will mostly talk about left translation, but the results
work equally well for right translation.
Then since
L
h
(
e
) =
h
, the derivative D
e
L
h
gives us a linear isomorphism
between
T
e
G
and
T
h
G
, with inverse given by D
h
L
h
1
. Then in particular, if we
are given a single tangent vector
X L
(
G
), we obtain a tangent vector at all
points in G, i.e. a vector field.
Definition (Vector field). A vector field V of G specifies a tangent vector
V (g) T
g
G
at each point
g G
. Suppose we can pick coordinates
{x
i
}
on some subset of
G
,
and write
v(g) = v
i
(g)
x
i
T
g
G.
The vector field is smooth if
v
i
(
g
)
R
are all differentiable for any coordinate
chart.
As promised, given any
X T
e
G
, we can define a vector field by using
L
g
:= D
e
L
g
to move this to all places in the world. More precisely, we can define
V (g) = L
g
(X).
This has the interesting property that if
X
is non-zero, then
V
is non-zero
everywhere, because L
g
is a linear isomorphism. So we found that
Proposition.
Let
G
be a Lie group of dimension
>
0. Then
G
has a nowhere-
vanishing vector field.
This might seem like a useless thing to know, but it tells us that certain
manifolds cannot be made into Lie groups.
Theorem
(Poincare-Hopf theorem)
.
Let
M
be a compact manifold. If
M
has
non-zero Euler characteristic, then any vector field on M has a zero.
The Poincare-Hopf theorem actually tells us how we can count the actual
number of zeroes, but we will not go into that. We will neither prove this, nor
use it for anything useful. But in particular, it has the following immediate
corollary:
Theorem
(Hairy ball theorem)
.
Any smooth vector field on
S
2
has a zero.
More generally, any smooth vector field on S
2n
has a zero.
Thus, it follows that
S
2n
can never be a Lie group. In fact, the full statement
of the Poincare-Hopf theorem implies that if we have a compact Lie group of
dimension 2, then it must be the torus! (we have long classified the list of all
possible compact 2-dimensional manifolds, and we can check that only the torus
works)
That was all just for our own amusement. We go back to serious work.
What does this left-translation map look like when we have a matrix Lie group
G Mat
n
(
F
)? For all
h G
and
X L
(
G
), we can represent
X
as a matrix in
Mat
n
(F). We then have a concrete representation of the left translation by
L
h
X = hX T
h
G.
Indeed,
h
acts on
G
by left multiplication, which is a linear map if we view
G
as
a subset of the vector space
Mat
n
(
F
). Then we just note that the derivative of
any linear map is “itself”, and the result follows.
Now we may ask ourselves the question given a tangent vector
X T
e
G
,
can we find a path that “always” points in the direction
X
? More concretely,
we want to find a path γ : R T
e
G such that
˙γ(t) = L
γ(t)
X.
Here we are using
L
γ(t)
to identify
T
γ(t)
G
with
T
e
G
=
L
(
G
). In the case of a
matrix Lie group, this just says that
˙γ(t) = γ(t)X.
We also specify the boundary condition γ(0) = e.
Now this is just an ODE, and by general theory of ODE’s, we know a solution
always exists and is unique. Even better, in the case of a matrix Lie group, there
is a concrete construction of this curve.
Definition
(Exponential)
.
Let
M Mat
n
(
F
) be a matrix. The exponential is
defined by
exp(M) =
X
`=0
1
!
M
`
Mat
n
(F).
The convergence properties of these series are very good, just like our usual
exponential.
Theorem.
For any matrix Lie group
G
, the map
exp
restricts to a map
L
(
G
)
G.
Proof.
We will not prove this, but on the first example sheet, we will prove this
manually for G = SU(n).
We now let
g(t) = exp(tX).
We claim that this is the curve we were looking for.
To check that it satisfies the desired properties, we simply have to compute
g(0) = exp(0) = I,
and also
dg(t)
dt
=
X
`=1
1
( 1)!
t
`1
X
`
= exp(tX)X = g(t)X.
So we are done.
We now consider the set
S
X
= {exp(tX) : t R}.
This is an abelian Lie subgroup of G with multiplication given by
exp(tX) exp(sX) = exp((t + s)X)
by the usual proof. These are known as one-parameter subgroups.
Unfortunately, it is not true in general that
exp(X) exp(Y ) = exp(X + Y ),
since the usual proof assumes that
X
and
Y
commute. Instead, what we’ve got
is the Baker–Campbell–Hausdorff formula.
Theorem (Baker–Campbell–Hausdorff formula). We have
exp(X) exp(Y ) = exp
X + Y +
1
2
[X, Y ] +
1
12
([X, [X, Y ]] [Y, [X, Y ]]) + ···
.
It is possible to find the general formula for all the terms, but it is messy.
We will not prove this.
By the inverse function theorem, we know the map
exp
is locally bijective. So
we know
L
(
G
) completely determines
G
in some neighbourhood of
e
. However,
exp
is not globally bijective. Indeed, we already know that the Lie algebra
doesn’t completely determine the Lie group, as
SO
(3) and
SU
(2) have the same
Lie algebra but are different Lie groups.
In general,
exp
can fail to be bijective in two ways. If
G
is not connected,
then
exp
cannot be surjective, since by continuity, the image of
exp
must be
connected.
Example.
Consider the groups O(
n
) and
SO
(
n
). Then the Lie algebra of O(
n
)
is
o(n) = {X Mat
n
(R) : X + X
T
= 0}.
So if X o(n), then tr X = 0. Then we have
det(exp(X)) = exp(tr X) = exp(0) = +1.
So any matrix in the image of
exp
has determinant +1, and hence can only lie
inside SO(n). It turns out that the image of exp is indeed SO(n).
More generally, we have
Proposition.
Let
G
be a Lie group, and
g
be its Lie algebra. Then the image
of g under exp is the connected component of e.
On the other hand,
exp
can also fail to be injective. This happens when
G
has a U(1) subgroup.
Example. Let G = U(1). Then
u(1) = {ix, x R}
We then have
exp(ix) = e
ix
.
This is certainly not injective. In particular, we have
exp(ix) = exp(i(x + 2π))
for any x.
4 Representations of Lie algebras
4.1 Representations of Lie groups and algebras
So far, we have just talked about Lie groups and Lie algebras abstractly. But we
know these groups don’t just sit there doing nothing. They act on things. For
example, the group
GL
(
n, R
) acts on the vector space
R
n
in the obvious way. In
general, the action of a group on a vector space is known as a representation.
Definition
(Representation of group)
.
Let
G
be a group and
V
be a (finite-
dimensional) vector space over a field
F
. A representation of
G
on
V
is given
by specifying invertible linear maps
D
(
g
) :
V V
(i.e.
D
(
g
)
GL
(
V
)) for each
g G such that
D(gh) = D(g)D(h)
for all
g, h G
. In the case where
G
is a Lie group and
F
=
R
or
C
, we require
that the map D : G GL(V ) is smooth.
The space
V
is known as the representation space, and we often write the
representation as the pair (V, D).
Here if we pick a basis
{e
1
, ··· , e
n
}
for
V
, then we can identify
GL
(
V
) with
GL
(
n, F
), and this obtains a canonical smooth structure when
F
=
R
or
C
. This
smooth structure does not depend on the basis chosen.
In general, the map D need not be injective or surjective.
Proposition.
Let
D
be a representation of a group
G
. Then
D
(
e
) =
I
and
D(g
1
) = D(g)
1
.
Proof. We have
D(e) = D(ee) = D(e)D(e).
Since D(e) is invertible, multiplying by the inverse gives
D(e) = I.
Similarly, we have
D(g)D(g
1
) = D(gg
1
) = D(e) = I.
So it follows that D(g)
1
= D(g
1
).
Now why do we care about representations? Mathematically, we can learn
a lot about a group in terms of its possible representations. However, from a
physical point of view, knowing about representations of a group is also very
important. When studying field theory, our fields take value in a fixed vector
space
V
, and when we change coordinates, our field will “transform accordingly”
according to some rules. For example, we say scalar fields “don’t transform”,
but, say, the electromagnetic field tensor transforms “as a 2-tensor”.
We can describe the spacetime symmetries by a group
G
, so that specifying
a “change of coordinates” is equivalent to giving an element of
G
. For example,
in special relativity, changing coordinates corresponds to giving an element of
the Lorentz group O(3, 1).
Now if we want to say how our objects in
V
transform when we change
coordinates, this is exactly the same as specifying a representation of
G
on
V
!
So understanding what representations are available lets us know what kinds of
fields we can have.
The problem, however, is that representations of Lie groups are very hard.
Lie groups are very big geometric structures with a lot of internal complexity.
So instead, we might try to find representations of their Lie algebras instead.
Definition
(Representation of Lie algebra)
.
Let
g
be a Lie algebra. A represen-
tation ρ of g on a vector space V is a collection of linear maps
ρ(X) gl(V ),
for each
X g
, i.e.
ρ
(
X
) :
V V
is a linear map, not necessarily invertible.
These are required to satisfy the conditions
[ρ(X
1
), ρ(X
2
)] = ρ([X
1
, X
2
])
and
ρ(αX
1
+ βX
2
) = αρ(X
1
) + βρ(X
2
).
The vector space
V
is known as the representation space. Similarly, we often
write the representation as (V, ρ).
Note that it is possible to talk about a complex representation of a real Lie
algebra, because any complex Lie algebra (namely
gl
(
V
) for a complex vector
space
V
) can be thought of as a real Lie algebra by “forgetting” that we can
multiply by complex numbers, and indeed this is often what we care about.
Definition
(Dimension of representation)
.
The dimension of a representation
is the dimension of the representation space.
We will later see that a representation of a Lie group gives rise to a represen-
tation of the Lie algebra. The representation is not too hard to obtain if we
have a representation
D
:
G GL
(
V
) of the Lie group, taking the derivative of
this map gives us the confusingly-denoted D
e
D
:
T
e
G T
e
(
GL
(
V
)), which is
a map
g gl
(
V
). To check that this is indeed a representation, we will have
to see that it respects the Lie bracket. We will do this later when we study the
relation between representations of Lie groups and Lie algebras.
Before we do that, we look at some important examples of representations of
Lie algebras.
Definition
(Trivial representation)
.
Let
g
be a Lie algebra of dimension
D
.
The trivial representation is the representation
d
0
:
g F
given by
d
0
(
X
) = 0
for all X g. This has dimension 1.
Definition
(Fundamental representation)
.
Let
g
=
L
(
G
) for
G Mat
n
(
F
). The
fundamental representation is given by d
f
: g Mat
n
(F) given by
d
f
(X) = X
This has dim(d
f
) = n.
Definition
(Adjoint representation)
.
All Lie algebras come with an adjoint
representation
d
Adj
of dimension
dim
(
g
) =
D
. This is given by mapping
X g
to the linear map
ad
X
: g g
Y 7→ [X, Y ]
By linearity of the bracket, this is indeed a linear map g gl(g).
There is a better way of thinking about this. Suppose our Lie algebra
g
comes from a Lie group
G
. Writing
Aut
(
G
) for all the isomorphisms
G G
, we
know there is a homomorphism
Φ : G Aut(G)
g 7→ Φ
g
given by conjugation:
Φ
g
(x) = gxg
1
.
Now by taking the derivative, we can turn each Φ
g
into a linear isomorphism
g g, i.e. an element of GL(g). So we found ourselves a homomorphism
Ad : G GL(g),
which is a representation of the Lie group
G
! It is an exercise to show that
the corresponding representation of the Lie algebra
g
is indeed the adjoint
representation.
Thus, if we view conjugation as a natural action of a group on itself, then
the adjoint representation is the natural representation of g over itself.
Proposition. The adjoint representation is a representation.
Proof.
Since the bracket is linear in both components, we know the adjoint
representation is a linear map g gl(g). It remains to show that
[ad
X
, ad
Y
] = ad
[X,Y ]
.
But the Jacobi identity says
[ad
X
, ad
Y
](Z) = [X, [Y, Z]] [Y, [X, Z]] = [[X, Y ], Z] = ad
[X,Y ]
(Z).
We will eventually want to find all representations of a Lie algebra. To do
so, we need the notion of when two representations are “the same”.
Again, we start with the definition of a homomorphism.
Definition
(Homomorphism of representations)
.
Let (
V
1
, ρ
1
)
,
(
V
2
, ρ
2
) be rep-
resentations of
g
. A homomorphism
f
: (
V
1
, ρ
1
)
(
V
2
, ρ
2
) is a linear map
f : V
1
V
2
such that for all X g, we have
f(ρ
1
(X)(v)) = ρ
2
(X)(f(v))
for all v V
1
. Alternatively, we can write this as
f ρ
1
= ρ
2
f.
In other words, the following diagram commutes for all X g:
V
1
V
2
V
1
V
2
f
ρ
1
(X) ρ
2
(X)
f
Then we can define
Definition
(Isomorphism of representations)
.
Two
g
-vector spaces
V
1
, V
2
are
isomorphic if there is an invertible homomorphism f : V
1
V
2
.
In particular, isomorphic representations have the same dimension.
If we pick a basis for
V
1
and
V
2
, and write the matrices for the representations
as
R
1
(
X
) and
R
2
(
X
), then they are isomorphic if there exists a non-singular
matrix S such that
R
2
(X) = SR
1
(X)S
1
for all X g.
We are going to look at special representations that are “indecomposable”.
Definition
(Invariant subspace)
.
Let
ρ
be a representation of a Lie algebra
g
with representation space
V
. An invariant subspace is a subspace
U V
such
that
ρ(X)u U
for all X g and u U.
The trivial subspaces are U = {0} and V .
Definition
(Irreducible representation)
.
An irreducible representation is a rep-
resentation with no non-trivial invariant subspaces. They are referred to as
irreps.
4.2 Complexification and correspondence of representa-
tions
So far, we have two things Lie algebras and Lie groups. Ultimately, the thing
we are interested in is the Lie group, but we hope to simplify the study of a
Lie group by looking at the Lie algebra instead. So we want to understand
how representations of Lie groups correspond to the representations of their Lie
algebras.
If we have a representation
D
:
G GL
(
V
) of a Lie group
G
, then taking
the derivative at the identity gives us a linear map
ρ
:
T
e
G T
I
GL
(
V
), i.e. a
map
ρ
:
g gl
(
V
). To show this is a representation, we need to show that it
preserves the Lie bracket.
Lemma.
Given a representation
D
:
G GL
(
V
), the induced representation
ρ : g gl(V ) is a Lie algebra representation.
Proof.
We will again only prove this in the case of a matrix Lie group, so that
we can use the construction we had for the Lie bracket.
We have to check that the bracket is preserved. We take curves
γ
1
, γ
2
:
R G
passing through I at 0 such that ˙γ
i
(0) = X
i
for i = 1, 2. We write
γ(t) = γ
1
1
(t)γ
1
2
(t)γ
1
(t)γ
2
(t) G.
We can again Taylor expand this to obtain
γ(t) = I + t
2
[X
1
, X
2
] + O(t
3
).
Essentially by the definition of the derivative, applying D to this gives
D(γ(t)) = I + t
2
ρ([X
1
, X
2
]) + O(t
3
).
On the other hand, we can apply D to () before Taylor expanding. We get
D(γ) = D(γ
1
1
)D(γ
2
2
)D(γ
1
)D(γ
2
).
So as before, since
D(γ
i
) = I + (X
i
) + O(t
2
),
it follows that
D(γ)(t) = I + t
2
[ρ(X
1
), ρ(X
2
)] + O(t
3
).
So we must have
ρ([X
1
, X
2
]) = [ρ(X
1
), ρ(X
2
)].
How about the other way round? We know that if
ρ
:
g gl
(
V
) is induced
by D : G GL(V ), then have
D(exp(X)) = I + (X) + O(t
2
),
while we also have
exp(ρ(X)) = I + rρ(X) + O(t
2
).
So we might expect that we indeed have
D(exp(X)) = exp(ρ(X))
for all X g.
So we can try to use this formula to construct a representation of
G
. Given
an element
g G
, we try to write it as
g
=
exp
(
X
) for some
X g
. We then
define
D(g) = exp(ρ(X)).
For this to be well-defined, we need two things to happen:
(i) Every element in G can be written as exp(X) for some X
(ii) The value of exp(ρ(X)) does not depend on which X we choose.
We first show that if this is well-defined, then it indeed gives us a representation
of G.
To see this, we use the Baker-Campbell-Hausdorff formula to say
D(exp(X) exp(Y )) = exp(ρ(log(exp(X) exp(Y ))))
= exp
ρ
log
exp
X + Y +
1
2
[X, Y ] + ···

= exp
ρ
X + Y +
1
2
[X, Y ] + ···

= exp
ρ(X) + ρ(Y ) +
1
2
[ρ(X), ρ(Y )] + ···
= exp(ρ(X)) exp(ρ(Y ))
= D(exp(X))D(exp(Y )),
where
log
is the inverse to
exp
. By well-defined-ness, it doesn’t matter which
log we pick. Here we need to use the fact that
ρ
preserves the Lie bracket, and
all terms in the Baker-Campbell-Hausdorff formula are made up of Lie brackets.
So when are the two conditions satisfied? For the first condition, we know
that
g
is connected, and the continuous image of a connected space is connected.
So a necessary condition is that
G
must be a connected Lie group. This rules out
groups such as O(
n
). It turns out this is also sufficient. The second condition is
harder, and we will take note of the following result without proof:
Theorem.
Let
G
be a simply connected Lie group with Lie algebra
g
, and let
ρ
:
g gl
(
V
) be a representation of
g
. Then there is a unique representation
D : G GL(V ) of G that induces ρ.
So if we only care about simply connected Lie groups, then studying the
representations of its Lie algebra is exactly the same as studying the representa-
tions of the group itself. But even if we care about other groups, we know that
all representations of the group give representations of the algebra. So to find a
representation of the group, we can look at all representations of the algebra,
and see which lift to a representation of the group.
It turns out Lie algebras aren’t simple enough. Real numbers are terrible,
and complex numbers are nice. So what we want to do is to look at the
complexification of the Lie algebra, and study the complex representations of
the complexified Lie algebra.
We will start with the definition of a complexification of a real vector space.
We will provide three different definitions of the definition, from concrete to
abstract, to suit different people’s tastes.
Definition
(Complexification I)
.
Let
V
be a real vector space. We pick a basis
{T
a
}
of
V
. We define the complexification of
V
, written
V
C
as the complex
linear span of {T
a
}, i.e.
V
C
=
n
X
λ
a
T
a
: λ
a
C
o
.
There is a canonical inclusion
V V
C
given by sending
P
λ
a
T
a
to
P
λ
a
T
a
for
λ
a
R.
This is a rather concrete definition, but the pure mathematicians will not
be happy with such a definition. We can try another definition that does not
involve picking a basis.
Definition
(Complexification II)
.
Let
V
be a real vector space. The complex-
ification of
V
has underlying vector space
V
C
=
V V
. Then the action of a
complex number λ = a + bi on (u
1
, u
2
) is given by
λ(u
1
, u
2
) = (au
1
bu
2
, au
2
+ bu
1
).
This gives
V
C
the structure of a complex vector space. We have an inclusion
V V
C
by inclusion into the first factor.
Finally, we have a definition that uses some notions from commutative algebra,
which you may know about if you are taking the (non-existent) Commutative
Algebra course. Otherwise, do not bother reading.
Definition
(Complexification III)
.
Let
V
be a real vector space. The com-
plexification of
V
is the tensor product
V
R
C
, where
C
is viewed as an
(R, C)-bimodule.
To define the Lie algebra structure on the complexification, we simply declare
that
[X + iY, X
0
+ iY
0
] = [X, X
0
] + i([X, Y
0
] + [Y, X
0
]) [Y, Y
0
]
for X, Y V V
C
.
Whichever definition we decide to choose, we have now a definition, and we
want to look at the representations of the complexification.
Theorem.
Let
g
be a real Lie algebra. Then the complex representations of
g
are exactly the (complex) representations of g
C
.
Explicitly, if
ρ
:
g gl
(
V
) is a complex representation, then we can extend
it to g
C
by declaring
ρ(X + iY ) = ρ(X) + (Y ).
Conversely, if
ρ
C
:
g
C
gl
(
V
), restricting it to
g g
C
gives a representation of
g.
Proof. Just stare at it and see that the formula works.
So if we only care about complex representations, which is the case most of
the time, we can study the representations of the complexification instead. This
is much easier. In fact, in the next chapter, we are going to classify all simple
complex Lie algebras and their representations.
Before we end, we note the following definition:
Definition
(Real form)
.
Let
g
be a complex Lie algebra. A real form of
g
is a
real Lie algebra h such that h
C
= g.
Note that a complex Lie algebra can have multiple non-isomorphic real forms
in general.
4.3 Representations of su(2)
We are now going to study the complex representations of
su
(2), which are
equivalently the representations of the complexification
su
C
(2). In this section,
we will write
su
(2) when we actually mean
su
C
(2) for brevity. There are a
number of reasons why we care about this.
(i)
We’ve already done this before, in the guise of “spin”/“angular momentum”
in quantum mechanics.
(ii)
The representations are pretty easy to classify, and form a motivation for
our later general classification of representations of complex simple Lie
algebras.
(iii)
Our later work on the study of simple Lie algebras will be based on our
knowledge of representations of su(2).
Recall that the uncomplexified su(2) has a basis
su(2) = span
R
T
a
=
1
2
a
: a = 1, 2, 3
,
where
σ
a
are the Pauli matrices. Since the Pauli matrices are independent over
C, they give us a basis of the complexification of su(2):
su
C
(2) = span
C
{σ
a
: a = 1, 2, 3}.
However, it is more convenient to use the following basis:
H = σ
3
=
1 0
0 1
, E
±
=
1
2
(σ
1
±
2
) =
0 1
0 0
,
0 0
1 0
.
We then have
[H, E
±
] = ±2E
±
, [E
+
, E
] = H.
We can rewrite the first relation as saying
ad
H
(E
±
) = ±2E
±
.
Together with the trivial result that
ad
H
(
H
) = 0, we know
ad
H
has eigenvalues
±2 and 0, with eigenvectors E
±
, H. These are known as roots of su(2).
Now let’s look at irreducible representations of
su
(2). Suppose (
V, ρ
) is a
representation. By linear algebra, we know ρ(H) has an eigenvector v
λ
, say
ρ(H)v
λ
= λv
λ
.
The eigenvalues of
ρ
(
H
) are known as the weights of the representation
ρ
. The
operators E
±
are known as step operators. We have
ρ(H)ρ(E
±
)v
λ
= (ρ(E
±
)ρ(H) + [ρ(H), ρ(E
±
)])v
λ
= (ρ(E
±
)ρ(H) + ρ([H, E
±
])v
λ
= (λ ± 2)ρ(E
±
)v
λ
.
So
ρ
(
E
±
)
v
λ
are also eigenvectors of
ρ
(
H
) with eigenvalues
λ ±
2, provided
ρ
(
E
±
)
v
λ
are non-zero. This constrains what can happen in a finite-dimensional
representation. By finite-dimensionality, we cannot have infinitely many eigen-
values. So at some point, we have to stop. In other words, a finite-dimensional
representation must have a highest weight Λ C, with
ρ(H)v
Λ
= Λv
Λ
, ρ(E
+
)v
Λ
= 0.
Now if we start off with
v
Λ
we can keep applying
E
to get the lower weights.
Indeed, we get
v
Λ2n
= (ρ(E
))
n
v
Λ
for each
n
. Again, this sequence must terminate somewhere, as we only have
finitely many eigenvectors. We might think that the irrep would consist of the
basis vectors
{v
Λ
, v
Λ2
, v
Λ4
, ··· , v
Λ2n
}.
However, we need to make sure we don’t create something new when we act on
these by
ρ
(
E
+
) again. We would certainly get that
ρ
(
E
+
)
v
Λ2n
is an eigenvector
of eigenvalue Λ
2
n
+ 2, but we might get some other eigenvector that is not
v
Λ2n+2
. So we have to check.
We can compute
ρ(E
+
)v
Λ2n
= ρ(E
+
)ρ(E
)v
Λ2n+2
= (ρ(E
)ρ(E
+
) + [ρ(E
+
), ρ(E
)])v
Λ2n+2
= ρ(E
)ρ(E
+
)v
Λ2n+2
+ 2n + 2)v
Λ2n+2
.
using the fact that [
E
+
, E
] =
H
. We now get a recursive relation. We can
analyze this by looking at small cases. When n = 1, the first term is
ρ(E
)ρ(E
+
)v
Λ2n+2
= ρ(E
)ρ(E
+
)v
Λ
= 0,
by definition of v
Λ
. So we have
ρ(E
+
)v
Λ2
= Λv
Λ
.
When n = 2, we have
ρ(E
+
)v
Λ4
= ρ(E
)ρ(E
+
)v
Λ2
+ 2)v
Λ2
= (2Λ 2)v
Λ2
.
In general, ρ(E
+
)v
Λ2n
is some multiple of v
Λ2n+2
, say
ρ(E
+
)v
Λ2n
= r
n
V
Λ2n+2
.
Plugging this into our equation gives
r
n
= r
n1
+ Λ 2n + 2,
with the boundary condition
ρ
(
E
+
)
v
Λ
= 0, i.e.
r
1
= Λ. If we solve this recurrence
relation, we determine an explicit relation for r
n
given by
r
n
= (Λ + 1 n)n.
Now returning to the problem that the sequence must terminate, we figure that
we also need to have a lowest weight of the form Λ
2
n
. Hence we must have
some non-zero vector v
Λ2n
6= 0 such that
ρ(E
)v
Λ2n
= 0.
For this to be true, we need
r
N+1
= 0 for some non-negative integer
N
. So we
have
+ 1 (N + 1))(N + 1) = 0.
This is equivalent to saying
N)(N + 1) = 0.
Since
N
+ 1 is a non-negative integer, we must have Λ
N
= 0, i.e. Λ =
N
. So
in fact the highest weight is an integer!
In summary, we’ve got
Proposition.
The finite-dimensional irreducible representations of
su
(2) are
labelled by Λ Z
0
, which we call ρ
Λ
, with weights given by
{−Λ, Λ + 2, ··· , Λ 2, Λ}.
The weights are all non-degenerate, i.e. each only has one eigenvector. We have
dim(ρ
Λ
) = Λ + 1.
Proof.
We’ve done most of the work. Given any irrep, we can pick any eigenvector
of
ρ
(
H
) and keep applying
E
+
to get a highest weight vector
v
Λ
, then the above
computations show that
{v
Λ
, v
Λ2
, ··· , v
Λ
}
is a subspace of the irrep closed under the action of
ρ
. By irreducibility, this
must be the whole of the representation space.
The representation
ρ
0
is the trivial representation;
ρ
1
is the fundamental one,
and ρ
2
is the adjoint representation.
Now what do these tell us about representations of the Lie group
SU
(2)? We
know
SU
(2) is simply connected, so by our discussion previously, we know the
complex representations of
SU
(2) are exactly the representations of
su
(2). This
is not too exciting.
We know there are other Lie groups with Lie algebra
su
(2), namely
SO
(3).
Now that
SO
(3) is not simply connected, when does a representation of
su
(2)
give us a representation of SO(3)? Recall that we had
SO(3)
=
SU(2)
Z/2Z
.
So an element in
SO
(3) is a pair
{A, A}
of elements in
SU
(2). Given a
representation
ρ
Λ
of
su
(2), we obtain a corresponding representation
D
Λ
of
SU
(2). Now we get a representation of
SO
(3) if and only if
D
Λ
respects the
identification between A and A. In particular, we need
D
Λ
(I) = D
Λ
(I). ()
On the other hand, if this were true, then multiplying both sides by
D
Λ
(
A
) we
get
D
Λ
(A) = D
Λ
(A).
So (
) is a necessary and sufficient condition for
D
Λ
to descend to a representation
of SO(3).
We know that
I = exp(H).
So we have
D
Λ
(I) = exp(ρ
Λ
(H)).
We know that ρ
Λ
has eigenvalues Λ S
Λ
. So D
Λ
(I) has eigenvalues
exp(λ) = (1)
λ
= (1)
Λ
,
since
λ
and Λ differ by an even integer. So we know
D
Λ
(
I
) =
D
Λ
(
I
) if and
only if Λ is even. In other words, we get a representation of
SO
(3) iff Λ is even.
In physics, we have already seen this as integer and half-integer spin. We
have integer spin exactly when we have a representation of
SO
(3). The fact that
half-integer spin particles exist means that “spin” is really about
SU
(2), rather
than SO(3).
4.4 New representations from old
We are now going to look at different ways of obtaining new representations
from old ones. We start with a rather boring one.
Definition
(Conjugate representation)
.
Let
ρ
be a representation of a real Lie
algebra g on C
n
. We define the conjugate representation by
¯ρ(X) = ρ(X)
for all X g.
Note that this need not be an actual new representation.
To obtain more interesting new representations, we recall the following
definitions from linear algebra:
Definition
(Direct sum)
.
Let
V, W
be vector spaces. The direct sum
V W
is
given by
V W = {v w : v V, w W }
with operations defined by
(v
1
w
1
) + (v
2
w
2
) = (v
1
+ v
2
) (w
1
+ w
2
)
λ(v w) = (λv) (λw).
We often suggestively write v w as v + w. This has dimension
dim(V W ) = dim V + dim W.
Definition
(Sum representation)
.
Suppose
ρ
1
and
ρ
2
are representations of
g
with representation spaces
V
1
and
V
2
of dimensions
d
1
and
d
2
. Then
V
1
V
2
is
a representation space with representation ρ
1
ρ
2
given by
(ρ
1
ρ
2
)(X) · (v
1
v
2
) = (ρ
1
(X)(v
1
)) (ρ
2
(X)(v
2
)).
In coordinates, if
R
i
is the matrix for
ρ
i
, then the matrix of
ρ
1
ρ
2
is given by
(R
1
R
2
)(X) =
R
1
(X) 0
0 R
2
(X)
The dimension of this representation is d
1
+ d
2
.
Of course, sum representations are in general not irreducible! However, they
are still rather easy to understand, since they decompose into smaller, nicer bits.
In contrast, tensor products do not behave as nicely.
Definition
(Tensor product)
.
Let
V, W
be vector spaces. The tensor product
V W
is spanned by elements
v w
, where
v V
and
w W
, where we identify
v (λ
1
w
1
+ λ
2
w
2
) = λ
1
(v w
1
) + λ
2
(v w
2
)
(λ
1
v
1
+ λ
2
v
2
) w = λ
1
(v
1
w) + λ
2
(v
2
w)
This has dimension
dim(V W ) = (dim V )(dim W ).
More explicitly, if
e
1
, ··· , e
n
is a basis of
V
and
f
1
, ··· , f
m
is a basis for
W
,
then {e
i
f
j
: 1 i n, 1 j m} is a basis for V W .
Given any two maps
F
:
V V
0
and
G
:
W W
0
, we define
F G
:
V W V
0
W
0
by
(F G)(v w) = (F (v)) (G(w)),
and then extending linearly.
The operation of tensor products should be familiar in quantum mechanics,
where we combine two state spaces to get a third. From a mathematical point
of view, the tensor product is characterized by the fact that a bilinear map
V ×W U is the same as a linear map V W U .
Definition
(Tensor product representation)
.
Let
g
be a Lie algebra, and
ρ
1
, ρ
2
be representations of
g
with representation spaces
V
1
, V
2
. We define the tensor
product representation ρ
1
ρ
2
with representation space V
1
V
2
given by
(ρ
1
ρ
2
)(X) = ρ
1
(X) I
2
+ I
1
ρ
2
(X) : V
1
V
2
V
1
V
2
,
where I
1
and I
2
are the identity maps on V
1
and V
2
.
Note that the tensor product representation is not given by the “obvious”
formula ρ
1
(X) ρ
2
(X). This does not work.
Now suppose (
ρ, V
) is a reducible representation. Then it has a non-trivial
invariant subspace
U V
. Now we pick a basis of
U
and extend it to
V
. Then
ρ sends any member of U to itself, so the matrix representation of ρ looks like
R(X) =
A(X) B(X)
0 C(X)
.
However, there is no a priori reason why
B
(
X
) should vanish as well. When
this happens, we say the representation is completely reducible.
Definition
(Completely reducible representation)
.
If (
ρ, V
) is a representation
such that there is a basis of V in which ρ looks like
ρ(X) =
ρ
1
(X) 0 ··· 0
0 ρ
2
(X) ··· 0
.
.
.
.
.
.
.
.
.
.
.
.
0 0 ··· ρ
n
(X)
,
then it is called completely reducible. In this case, we have
ρ = ρ
1
ρ
2
··· ρ
n
.
Here is an important fact:
Theorem.
If
ρ
i
for
i
= 1
, ··· , m
are finite-dimensional irreps of a simple Lie
algebra
g
, then
ρ
1
··· ρ
m
is completely reducible to irreps, i.e. we can find
˜ρ
1
, ··· , ˜ρ
k
such that
ρ
1
··· ρ
m
= ˜ρ
1
˜ρ
2
··· ˜ρ
k
.
We will not prove this.
4.5 Decomposition of tensor product of su
(2)
representa-
tions
We now try to find an explicit description of the decomposition of tensor products
of irreps of su(2).
We let
ρ
Λ
and
ρ
Λ
0
be irreps of
su
(2), where Λ
,
Λ
0
N
. We call the represen-
tation spaces V
Λ
and V
Λ
0
.
We can form the tensor product ρ
Λ
ρ
Λ
0
with representation space
V
Λ
V
Λ
0
= span
R
{v v
0
: v V
Λ
, v
0
V
Λ
0
}.
By definition, for X su(2), the representation is given by
(ρ
Λ
ρ
Λ
0
)(X)(v v
0
) = (ρ
Λ
(X)v) v
0
+ v (ρ
Λ
0
(X)v
0
).
This gives us a completely reducible representation of su(2) of dimension
dim(ρ
Λ
ρ
Λ
0
) = (Λ + 1)(Λ
0
+ 1).
We can then write
ρ
Λ
ρ
Λ
0
=
M
Λ
00
Z,Λ
00
0
L
Λ
00
Λ,Λ
0
ρ
Λ
00
,
where
L
Λ
00
Λ,Λ
0
are some non-negative integers we want to find out. These coefficients
are usually known as Littlewood-Richardson coefficients in general.
Recall that V
Λ
has a basis {v
λ
}, where
λ S
Λ
= {−Λ, Λ 2, ··· , }.
Similarly, V
Λ
0
has a basis {v
0
λ
0
}.
Then we know that the tensor product space has basis
B = {v
λ
v
0
λ
0
: λ S
Λ
, λ
0
S
Λ
0
}.
We now see what H does to our basis vectors. We have
(ρ
Λ
ρ
Λ
0
)(H)(v
λ
v
0
λ
0
) = (ρ
Λ
(H)v
λ
) v
0
λ
0
+ v
λ
(ρ
Λ
0
(H)v
0
λ
0
)
= (λ + λ
0
)(v
λ
v
0
λ
0
).
We thus see that the weights of the tensor product are just the sum of the
weights of the individual components. In other words, we have
S
Λ,Λ
0
= {λ + λ
0
: λ S
Λ
, λ
0
S
Λ
0
}
Note that here we count the weights with multiplicity, so that each weight can
appear multiple times.
We see that the highest weight is just the sum of the largest weights of the
irreps, and this appears with multiplicity 1. Thus we know
L
Λ+Λ
0
Λ,Λ
0
= 1,
i.e. we have one copy of
ρ
Λ+Λ
0
in the decomposition of the tensor product. We
write
ρ
Λ
ρ
Λ
0
= ρ
Λ+Λ
0
˜ρ
Λ,Λ
0
,
where ˜ρ
Λ,Λ
0
has weight set
˜
S
Λ,Λ
0
satisfying
S
Λ,Λ
0
= S
Λ+Λ
0
˜
S
Λ,Λ
0
.
We now notice that there is only one Λ + Λ
0
2 term in
˜
S
Λ,Λ
0
. So there must be
a copy of ρ
Λ+Λ
0
2
as well. We keep on going.
Example. Take Λ = Λ
0
= 1. Then we have
S
1
= {−1, +1}.
So we have
S
1,1
= {−2, 0, 0, 2}.
We see that the highest weight is 2, and this corresponds to a factor of
ρ
2
. In
doing so, we write
S
1,1
= {−2, 0, +2} {0} = S
2
S
0
.
So we have
ρ
1
ρ
1
= ρ
2
ρ
0
.
From the above, one can see (after some thought) that in general, we have
Proposition.
ρ
M
ρ
N
= ρ
|NM|
ρ
|NM|+2
··· ρ
N+M
.
5 Cartan classification
We now move on to the grand scheme of classifying all complex simple Lie
algebras. The starting point of everything is that we define a natural inner
product on our Lie algebra
g
. We will find a subalgebra
h
of
g
that plays
the role of the
H
we had when we studied
su
(2). The remainder of
g
will be
controlled by things known as roots, which live in
h
. We will see that the Killing
form induces a inner product on
h
, which allows us to think of these roots as
“geometric” objects that live in
R
n
. We can then find some strong conditions
that restrict what these roots and their inner products can be, and it turns out
these completely characterize our possible Lie algebras.
5.1 The Killing form
The first thing we want to figure out is an “invariant” inner product on our Lie
algebra
g
. We will do so by writing down a formula, and then checking that
for a (semi-)simple Lie algebra, it is non-degenerate. Having a non-degenerate
inner product will be very useful. Amongst many things, it provides us with a
bijection between a vector space and its dual.
We recall the following definitions:
Definition
(Inner product)
.
Given a vector space
V
over
F
, an inner product
is a symmetric bilinear map i : V ×V F.
Definition
(Non-degenerate inner product)
.
An inner product
i
is said to be
non-degenerate if for all v V non-zero, there is some w V such that
i(v, w) 6= 0.
The question we would like to ask is if there is a “natural” inner product on
g. We try the following:
Definition
(Killing form)
.
The Killing form of a Lie algebra
g
is the inner
product κ : g × g F given by
κ(X, Y ) = tr(ad
X
ad
Y
),
where
tr
is the usual trace of a linear map. Since
ad
is linear, this is bilinear in
both arguments, and the cyclicity of the trace tells us this is symmetric.
We can try to write this more explicitly. The map
ad
X
ad
Y
:
g g
is given
by
Z 7→ [X, [Y, Z]].
We pick a basis {T
a
}
a=1,...,D
for g. We write
X = X
a
T
a
, Y = Y
a
T
a
, Z = Z
a
T
a
.
We again let f
ab
c
be the structure constants satisfying
[T
a
, T
b
] = f
ab
c
T
c
.
We then have
[X, [Y, Z]] = X
a
Y
b
Z
c
[T
a
, [T
b
, T
c
]]
= X
a
Y
b
Z
c
f
ad
e
f
bc
d
T
e
= M(X, Y )
c
e
Z
c
T
e
,
where
M(X, Y )
c
e
= X
a
Y
b
f
ad
e
f
bc
d
.
So the trace of this thing is
κ(X, Y ) = tr(M (X, Y )) = κ
ab
X
a
Y
b
, κ
ab
= f
ad
c
f
bc
d
.
Why is this a natural thing to consider? This is natural because it obeys an
invariance condition:
Definition
(Invariant inner product)
.
An inner product
κ
on a Lie algebra
g
is
invariant if for any X, Y, Z g, we have
κ([Z, X], Y ) + κ(X, [Z, Y ]) = 0.
Equivalently, we have
κ(ad
Z
X, Y ) + κ(X, ad
Z
Y ) = 0.
What does this condition actually mean? If one were to arbitrarily write
down a definition of invariance, they might try
κ(ad
Z
X, Y ) = κ(X, ad
Z
Y )
instead. However, this is not the right thing to ask for.
Usually, we think of elements of the Lie algebra as some sort of “infinitesimal
transformation”, and as we have previously discussed, the adjoint representa-
tion is how an element
Z g
naturally acts on
g
. So under an infinitesimal
transformation, the elements X, Y g transform as
X 7→ X + ad
Z
X
Y 7→ Y + ad
Z
Y
What is the effect on the Killing form? It transforms infinitesimally as
κ(X, Y ) 7→ κ(X + ad
Z
X, Y + ad
Z
Y ) κ(X, Y ) + κ(ad
Z
X, Y ) + κ(X, ad
Z
Y ),
where we dropped the “higher order terms” because we want to think of the
ad
Z
terms as being infinitesimally small (justifying this properly will require going
to the more global picture involving an actual Lie group, which we shall not go
into. This is, after all, just a motivation). So invariance of the Killing form says
it doesn’t transform under this action.
So we now check that the Killing form does satisfy the invariance condition.
Proposition. The Killing form is invariant.
Proof. We have
κ([Z, X], Y ) = tr(ad
[Z,X]
ad
Y
)
= tr([ad
Z
, ad
X
] ad
Y
)
= tr(ad
Z
ad
X
ad
Y
ad
X
ad
Z
ad
Y
)
= tr(ad
Z
ad
X
ad
Y
) tr(ad
X
ad
Z
ad
Y
)
Similarly, we have
κ(X, [Z, Y ]) = tr(ad
X
ad
Z
ad
Y
) tr(ad
X
ad
Y
ad
Z
).
Adding them together, we obtain
κ([Z, X], Y ) + κ(X, [Z, Y ]) = tr(ad
Z
ad
X
ad
Y
) tr(ad
X
ad
Y
ad
Z
).
By the cyclicity of tr, this vanishes.
The next problem is to figure out when the Killing form is degenerate. This
is related to the notion of simplicity of Lie algebras.
Definition
(Semi-simple Lie algebra)
.
A Lie algebra is semi-simple if it has no
abelian non-trivial ideals.
This is weaker than the notion of simplicity simplicity requires that there
are no non-trivial ideals at all!
In fact, it is true that
g
being semi-simple is equivalent to
g
being the direct
sum of simple Lie algebras. This is on the second example sheet.
Theorem
(Cartan)
.
The Killing form of a Lie algebra
g
is non-degenerate iff
g
is semi-simple.
Proof.
We are only going to prove one direction if
κ
is non-degenerate, then
g is semi-simple.
Suppose we had an abelian ideal
a g
. We want to show that
κ
(
A, X
) = 0
for all
A a
and
X g
. Indeed, we pick a basis of
a
, and extend it to a basis of
g
. Then since [
X, A
]
a
for all
X g
and
A a
, we know the matrix of
ad
X
must look like
ad
X
=
∗ ∗
0
.
Also, if
A a
, then since
a
is an abelian ideal,
ad
A
kills everything in
a
, and
ad
A
(X) a for all X g. So the matrix must look something like
ad
A
=
0
0 0
.
So we know
ad
A
ad
X
=
0
0 0
,
and the trace vanishes. So
κ
(
A, X
) = 0 for all
X g
and
A a
. So
A
= 0. So
a
is trivial.
Now if
κ
is non-degenerate, then
κ
ab
is “invertible”. So we can find a
κ
ab
such that
κ
ab
κ
bc
= δ
c
a
.
We can then use this to raise and lower indices.
Note that this definition so far does not care if this is a real or complex
Lie algebra. From linear algebra, we know that any symmetric matrix can be
diagonalized. If we are working over
C
, then the diagonal entries can be whatever
we like if we choose the right basis. However, if we are working over
R
, then
the number of positive (or negative) diagonal entries are always fixed, while the
magnitudes can be whatever we like, by Sylvester’s law of inertia.
Thus in the case of a real Lie algebra, it is interesting to ask when the matrix
always has the same sign. It turns out it is the case with always negative sign is
the interesting case.
Definition
(Real Lie algebra of compact type)
.
We say a real Lie algebra is of
compact type if there is a basis such that
κ
ab
= κδ
ab
,
for some κ R
+
.
By general linear algebra, we can always pick a basis so that κ = 1.
The reason why it is called “compact” is because these naturally arise when
we study compact Lie groups.
We will note the following fact without proof:
Theorem.
Every complex semi-simple Lie algebra (of finite dimension) has a
real form of compact type.
We will not use it for any mathematical results, but it will be a helpful thing
to note when we develop gauge theories later on.
5.2 The Cartan basis
From now on, we restrict to the study of finite-dimensional simple complex Lie
algebras. Every time we write the symbol
g
or say “Lie algebra”, we mean a
finite-dimensional simple complex Lie algebra.
Recall that we have already met such a Lie algebra
su
C
(2) = span
C
{H, E
+
, E
}
with the brackets
[H, E
±
] = ±2E
±
, [E
+
, E
] = H.
These are known as the Cartan basis for
su
C
(2). We will try to mimic this
construction in an arbitrary Lie algebra.
Recall that when we studied
su
C
(2), we used the fact that
H
is a diagonal
matrix, and
E
±
acted as step operators. However, when we study Lie algebras
in general, we want to think of them abstractly, rather than as matrices, so it
doesn’t make sense to ask if an element is diagonal.
So to develop the corresponding notions, we look at the
ad
map associated to
them instead. Recall that the adjoint map of
H
is also diagonal, with eigenvectors
given by
ad
H
(E
±
) = ±2E
±
ad
H
(H) = 0.
This is the structure we are trying to generalize.
Definition
(
ad
-diagonalizable)
.
Let
g
be a Lie algebra. We say that an element
X g is ad-diagonalizable if the associated map
ad
X
: g g
is diagonalizable.
Example. In su
C
(2), we know H is ad-diagonalizable, but E
±
is not.
Now we might be tempted to just look at all
ad
-diagonalizable elements.
However, this doesn’t work. In the case of
su
(2), each of
σ
1
, σ
2
, σ
3
is
ad
-
diagonalizable, but we only want to pick one of them as our
H
. Instead, what
we want is the following:
Definition
(Cartan subalgebra)
.
A Cartan subalgebra
h
of
g
is a maximal
abelian subalgebra containing only ad-diagonalizable elements.
A Cartan subalgebra always exists, since the dimension of
g
is finite, and the
trivial subalgebra
{
0
} g
is certainly abelian and contains only
ad
-diagonalizable
elements. However, as we have seen, this is not necessarily unique. Fortunately,
we will later see that in fact all possible Cartan subalgebras have the same
dimension, and the dimension of h is called the rank of g.
From now on, we will just assume that we have fixed one such Cartan
subalgebra.
It turns out that Cartan subalgebras satisfy a stronger property.
Proposition.
Let
h
be a Cartan subalgebra of
g
, and let
X g
. If [
X, H
] = 0
for all H h, then X h.
Note that this does not follow immediately from
h
being maximal, because
maximality only says that ad-diagonalizable elements satisfy that property.
Proof. Omitted.
Example. In the case of su
C
(2), one possible Cartan subalgebra is
h = span
C
{H}.
However, recall our basis is given by
H = σ
3
E
±
=
1
2
(σ
1
±
2
),
where the
σ
i
are the Pauli matrices. Then, by symmetry, we know that
σ
1
=
E
+
+
E
gives an equally good Cartan subalgebra, and so does
σ
2
. So we have
many choices, but they all have the same dimension.
Now we know that everything in
h
commutes with each other, i.e. for any
H, H
0
h, we have
[H, H
0
] = 0.
Since ad is a Lie algebra representation, it follows that
ad
H
ad
H
0
ad
H
0
ad
H
= 0.
In other words, all these
ad
maps commute. By assumption, we know each
ad
H
is diagonalizable. So we know they are in fact simultaneously diagonalizable. So
g
is spanned by simultaneous eigenvectors of the
ad
H
. Can we find a basis of
eigenvectors?
We know that everything in
h
is a zero-eigenvector of
ad
H
for all
H h
,
since for H, H
0
h, we have
ad
H
(H
0
) = [H, H
0
] = 0.
We can arbitrarily pick a basis
{H
i
: i = 1, ··· , r},
where
r
is the rank of
h
. Moreover, by maximality, there are no other eigenvectors
that are killed by all H h.
We are now going to label the remaining eigenvectors by their eigenvalue.
Given any eigenvector E g and H h, we have
ad
H
(E) = [H, E] = α(H)E
for some constant
α
(
H
) depending on
H
(and
E
). We call
α
:
h C
the root of
E. We will use the following fact without proof:
Fact.
The non-zero simultaneous eigenvectors of
h
are non-degenerate, i.e. there
is a unique (up to scaling) eigenvector for each root.
Thus, we can refer to this eigenvector unambiguously by
E
α
, where
α
designates the root.
What are these roots
α
? It is certainly a function
h C
, but it is actually a
linear map! Indeed, we have
α(H + H
0
)E = [H + H
0
, E]
= [H, E] + [H
0
, E]
= α(H)E + α(H
0
)E
= (α(H) + α(H
0
))E,
by linearity of the bracket.
We write Φ for the collection of all roots. So we can write the remaining
basis eigenvectors as
{E
α
: α Φ}.
Example. In the case of su(2), the roots are ±2, and the eigenvectors are E
±
.
We can now define a Cartan-Weyl basis for g given by
B = {H
i
: i = 1, ··· , r} {E
α
: α Φ}.
Recall that we have a Killing form
κ(X, Y ) =
1
N
tr(ad
X
ad
Y
),
where
X, Y g
. Here we put in a normalization factor
N
for convenience later
on. Since g is simple, it is in particular semi-simple. So κ is non-degenerate.
We are going to evaluate κ in the Cartan-Weyl basis.
Lemma. Let H h and α Φ. Then
κ(H, E
α
) = 0.
Proof. Let H
0
h. Then
α(H
0
)κ(H, E
α
) = κ(H, α(H
0
)E
α
)
= κ(H, [H
0
, E
α
])
= κ([H
0
, H], E
α
)
= κ(0, E
α
)
= 0
But since α 6= 0, we know that there is some H
0
such that α(H
0
) 6= 0.
Lemma. For any roots α, β Φ with α + β 6= 0, we have
κ(E
α
, E
β
) = 0.
Proof. Again let H h. Then we have
(α(H) + β(H))κ(E
α
, E
β
) = κ([H, E
α
], E
β
) + κ(E
α
, [H, E
β
]),
= 0
where the final line comes from the invariance of the Killing form. Since
α
+
β
does not vanish by assumption, we must have κ(E
α
, E
β
) = 0.
Lemma. If H h, then there is some H
0
h such that κ(H, H
0
) 6= 0.
Proof.
Given an
H
, since
κ
is non-degenerate, there is some
X g
such that
κ
(
H, X
)
6
= 0. Write
X
=
H
0
+
E
, where
H
0
h
and
E
is in the span of the
E
α
.
0 6= κ(H, X) = κ(H, H
0
) + κ(H, E) = κ(H, H
0
).
What does this tell us?
κ
started life as a non-degenerate inner product
on
g
. But now we know that
κ
is a non-degenerate inner product on
h
. By
non-degeneracy, we can invert it within h.
In coordinates, we can find some κ
ij
such that
κ(e
i
H
i
, e
0
j
H
j
) = κ
ij
e
i
e
0
j
for any
e
i
,
e
0
j
. The fact that the inner product is non-degenerate means that we
can invert the matrix κ, and find some (κ
1
)
ij
such that
(κ
1
)
ij
κ
jk
= δ
i
k
Since
κ
1
is non-degenerate, this gives a non-degenerate inner product on
h
.
In particular, this gives us an inner product between the roots! So given two
roots α, β Φ h
, we write the inner product as
(α, β) = (κ
1
)
ij
α
i
β
j
,
where
α
i
:=
α
(
H
i
). We will later show that the inner products of roots will
always be real, and hence we can talk about the “geometry” of roots.
We note the following final result:
Lemma. Let α Φ. Then α Φ. Moreover,
κ(E
α
, E
α
) 6= 0
This holds for stupid reasons.
Proof. We know that
κ(E
α
, E
β
) = κ(E
α
, H
i
) = 0
for all
β 6
=
α
and all
i
. But
κ
is non-degenerate, and
{E
β
, H
i
}
span
g
. So
there must be some E
α
in the basis set, and
κ(E
α
, E
α
) 6= 0.
So far, we know that
[H
i
, H
j
] = 0
[H
i
, E
α
] = α
i
E
α
for all α Φ and i, j = 1, ··· , r. Now it remains to evaluate [E
α
, E
β
].
Recall that in the case of su
C
(2), we had
[E
+
, E
] = H.
What can we get here? For any H h and α, β Φ, we have
[H, [E
α
, E
β
]] = [E
α
, [E
β
, H]] [E
β
, [H, E
α
]]
= (α(H) + β(H))[E
α
, E
β
].
Now if α + β 6= 0, then either [E
α
, E
β
] = 0, or α + β Φ and
[E
α
, E
β
] = N
α,β
E
α+β
for some N
α,β
.
What if
α
+
β
= 0? We claim that this time, [
E
α
, E
α
]
h
. Indeed, for any
H h, we have
[H, [E
α
, E
α
]] = [[H, E
α
], E
α
] + [[E
α
, H], E
α
]
= α(H)[E
α
, E
α
] + α(H)[E
α
, E
α
]
= 0.
Since
H
was arbitrary, by the (strong) maximality property of
h
, we know that
[E
α
, E
α
] h.
Now we can compute
κ([E
α
, E
α
], H) = κ(E
α
, [E
α
, H])
= α(H)κ(E
α
, E
α
).
We can view this as an equation for [
E
α
, E
α
]. Now since we know that
[
E
α
, E
α
]
h
, and the Killing form is non-degenerate when restricted to
h
, we
know [E
α
, E
α
] is uniquely determined by this relation.
We define the normalized
H
α
=
[E
α
, E
α
]
κ(E
α
, E
α
)
.
Then our equation tells us
κ(H
α
, H) = α(H).
Writing H
α
and H in components:
H
α
= e
α
i
H
i
, H = e
i
H
i
.
the equation reads
κ
ij
e
α
i
e
j
= α
i
e
i
.
Since the e
j
are arbitrary, we know
e
α
i
= (κ
1
)
ij
α
j
.
So we know
H
α
= (κ
1
)
ij
α
j
H
i
.
We now have a complete set of relations:
Theorem.
[H
i
, H
j
] = 0
[H
i
, E
α
] = α
i
E
α
[E
α
, E
β
] =
N
α,β
E
α+β
α + β Φ
κ(E
α
, E
β
)H
α
α + β = 0
0 otherwise
.
Now we notice that there are special elements
H
α
in the Cartan subalgebra
associated to the roots. We can compute the brackets as
[H
α
, E
β
] = (κ
1
)
ij
α
i
[H
j
, E
β
]
= (κ
1
)
ij
α
i
β
j
E
β
= (α, β)E
β
,
where we used the inner product on the dual space
h
induced by the Killing
form κ.
Note that so far, we have picked
H
i
and
E
α
arbitrarily. Any scalar multiple
of them would have worked as well for what we did above. However, it is often
convenient to pick a normalization such that the numbers we get turn out to
look nice. It turns out that the following normalization is useful:
e
α
=
s
2
(α, α)κ(E
α
, E
α
)
E
α
h
α
=
2
(α, α)
H
α
.
Here it is important that (
α, α
)
6
= 0, but we will only prove it next chapter where
the proof fits in more naturally.
Under this normalization, we have
[h
α
, h
β
] = 0
[h
α
, e
β
] =
2(α, β)
(α, α)
e
β
[e
α
, e
β
] =
n
αβ
e
α+β
α + β Φ
h
α
α + β = 0
0 otherwise
Note that the number of roots is
d r
, where
d
is the dimension of
g
and
r
is its
rank, and this is typically greater than
r
. So in general, there are too many of
them to be a basis, even if they spanned
h
(we don’t know that yet). However,
we are still allowed to talk about them, and the above relations are still true.
It’s just that they might not specify everything about the Lie algebra.
5.3 Things are real
We are now going to prove that many things are real. When we do so, we can
then restrict to a real form of h, and then we can talk about their geometry.
The first result we want to prove in this section is the following:
Theorem.
(α, β) R
for all α, β Φ.
To prove this, we note that any Lie algebra contains many copies of
su
(2).
If
α
Φ, then
α
Φ. For each pair
±α
Φ, we can consider the subalgebra
with basis {h
α
, e
α
, e
α
}. Then we have
[h
α
, e
±α
] = ±2e
±α
[e
α
, e
α
] = h
α
.
These are exactly the relations for the Lie algebra of
su
(2) (or rather,
su
C
(2),
but we will not write the
C
every time). So this gives a subalgebra isomorphic
to su(2). We call this su(2)
α
.
It turns out we also get a lot of representations of su(2)
α
.
Definition
(String)
.
For
α, β
Φ, we define the
α
-string passing through
β
to
be
S
α,β
= {β + ρα Φ : ρ Z}.
We will consider the case where
β
is not proportional to
α
. The remaining
case, where we may wlog take
β
= 0, is left as an exercise in the example sheet.
We then have a corresponding vector subspace
V
α,β
= span
C
{e
β+ρα
: β + ρα S
α,β
}.
Consider the action of su(2)
α
on V
α,β
. We have
[h
α
, e
β+ρα
] =
2(α, β + ρα)
(α, α)
e
β+ρα
=
2(α, β)
(α, α)
+ 2ρ
e
β+ρα
. ()
We also have
[e
±α
, e
β+ρα
]
(
e
β+(ρ±1)α
β + (ρ ± 1)α Φ
0 otherwise
V
α,β
.
So
V
α,β
is invariant under the action of
su
(2)
α
. So
V
α,β
is the representation
space for some representation of su(2)
α
.
Moreover, () tells us the weight set of this representation. We have
S =
2(α, β)
(α, α)
+ 2ρ : β + ρα Φ : ρ Z
.
This is not to be confused with the string S
α,β
itself!
Since the Lie algebra itself is finite-dimensional, this representation must
be finite-dimensional. We also see from the formula that the weights are non-
degenerate and are spaced by 2. So it must be an irreducible representation of
su(2). Then if its highest weight is Λ Z, we have
S = {−Λ, Λ + 2, ··· , }.
What does this tell us? We know that the possible values of
ρ
are bounded
above and below, so we can write
S
α,β
= {β + ρα Φ : n
ρ n
+
}.
for some n
±
Z. In particular, we know that
Proposition. For any α, β Φ, we have
2(α, β)
(α, α)
Z.
Proof. For ρ = n
±
, we have
2(α, β)
(α, α)
+ 2n
= Λ and
2(α, β)
(α, α)
+ 2n
+
= Λ.
Adding the two equations yields
2(α, β)
(α, α)
= (n
+
+ n
) Z.
We are next going to prove that the inner products themselves are in fact
real. This will follow from the following lemma:
Lemma. We have
(α, β) =
1
N
X
δΦ
(α, δ)(δ, β),
where N is the normalization factor appearing in the Killing form
κ(X, Y ) =
1
N
tr(ad
X
ad
Y
).
Proof. We pick the Cartan-Weyl basis, with
[H
i
, E
δ
] = δ
i
E
δ
for all i = 1, ··· , r and δ Φ. Then the inner product is defined by
κ
ij
= κ(H
i
, H
j
) =
1
N
tr[ad
H
i
ad
H
j
].
But we know that these matrices
ad
H
i
are diagonal in the Cartan-Weyl basis,
and the non-zero diagonal entries are exactly the δ
i
. So we can write this as
κ
ij
=
1
N
X
δΦ
δ
i
δ
j
.
Now recall that our inner product was defined by
(α, β) = α
i
β
j
(κ
1
)
ij
= κ
ij
α
i
β
j
,
where we define
β
j
= (κ
1
)
jk
β
k
.
Putting in our explicit formula for the κ
ij
, this is
(α, β) =
1
N
X
δΦ
α
i
δ
i
δ
j
β
j
=
1
N
X
δΦ
(α, δ)(δ, β).
Corollary.
(α, β) R
for all α, β Φ.
Proof. We write
R
α,β
=
2(α, β)
(α, α)
Z.
Then the previous formula tells us
2
(β, β)
R
α,β
=
1
N
X
δΦ
R
α,δ
R
β
We know that
R
α,β
are all integers, and in particular real. So (
β, β
) must be
real as well. So it follows that (α, β) is real since R
α,β
is an integer.
5.4 A real subalgebra
Let’s review what we know so far. The roots
α
Φ are elements of
h
. In
general, the roots will not be linearly independent elements of
h
, because there
are too many of them. However, it is true that the roots span h
.
Proposition. The roots Φ span h
. In particular, we know
|Φ| dim h
.
Proof.
Suppose the roots do not span
h
. Then the space spanned by the roots
would have a non-trivial orthogonal complement. So we can find
λ h
such
that (λ, α) = 0 for all α Φ. We now define
H
λ
= λ
i
H
i
h.
Then as usual we have
[H
λ
, H] = 0 for all H h.
Also, we know
[H
λ
, E
α
] = (λ, α)E
α
= 0.
for all roots
α
Φ by assumption. So
H
λ
commutes with everything in the Lie
algebra. This would make
hH
λ
i
a non-trivial ideal, which is a contradiction since
g is simple.
Since the
α
Φ span, we can find a basis
{α
(i)
Φ :
i
= 1
, ··· , r}
of roots.
Again, this choice is arbitrary. We now define a real vector subspace
h
R
h
by
h
R
= span
R
{α
(i)
: i = 1, ··· , r}.
One might be worried that this would depend on our choice of the basis
α
(i)
,
which was arbitrary. However, it does not, since any choice will give us the same
space:
Proposition. h
R
contains all roots.
So h
R
is alternatively the real span of all roots.
Proof.
We know that
h
is spanned by the
α
(i)
as a complex vector space. So
given an β h
, we can find some β
i
C such that
β =
r
X
i=1
β
i
α
(i)
.
Taking the inner product with α
(j)
, we know
(β, α
(j)
) =
r
X
i=1
β
i
(α
(i)
, α
(j)
).
We now use the fact that the inner products are all real! So
β
i
is the solution to
a set of real linear equations, and the equations are non-degenerate since the
α
(i)
form a basis and the Killing form is non-degenerate. So
β
i
must be real. So
β h
R
.
Now the inner product of any two elements of
h
R
is real, since an element in
h
R
is a real linear combination of the
α
(i)
, and the inner product of the
α
(i)
is
always real.
Proposition.
The Killing form induces a positive-definite inner product on
h
R
.
Proof.
It remains to show that (
λ, λ
)
0 for all
λ
, with equality iff
λ
= 0. We
can write
(λ, λ) =
1
N
X
δΦ
(λ, δ)
2
0.
If this vanishes, then (
λ, δ
) = 0 for all
δ
Φ. But the roots span, so this implies
that λ kills everything, and is thus 0 by non-degeneracy.
Now this
h
R
is just like any other real inner product space we know and love
from IA Vectors and Matrices, and we can talk about the lengths and angles
between vectors.
Definition (Norm of root). Let α Φ be a root. Then its length is
|α| =
p
(α, α) > 0.
Then for any α, β, there is some “angle” ϕ [0, π] between them given by
(α, β) = |α||β|cos ϕ.
Now recall that we had the quantization result
2(α, β)
(α, α)
Z.
Then in terms of the lengths, we have
2|β|
|α|
cos ϕ Z.
Since the quantization rule is not symmetric in
α, β
, we obtain a second quanti-
zation constraint
2|α|
|β|
cos ϕ Z.
Since the product of two integers is an integer, we know that
4 cos
2
ϕ Z.
So
cos ϕ = ±
n
2
, n = {0, 1, 2, 3, 4}.
So we have some boring solutions
ϕ = 0,
π
2
, π,
and non-boring ones
ϕ =
π
6
,
π
4
,
π
3
,
2π
3
,
3π
4
,
5π
6
.
These are the only possibilities!
5.5 Simple roots
Now we have a lot of roots that span
h
R
, but they obviously do not form a basis.
For example, whenever
α h
R
, then so is
α
. So it might be helpful to get rid
of half of these things. The divide is again rather arbitrary.
To do so, we pick a hyperplane in
h
R
=
R
r
, i.e. a subspace of dimension
r
1.
Since there are finitely many roots, we can pick it so that it doesn’t contain any
root. Then the plane divides
h
R
into 2 sides. We then pick a distinguished side,
and say that α is positive if it lies in that side, and negative otherwise.
Then if
α
is positive,
α
is negative. More interestingly, if
α, β
are positive
roots, then
α
+
β
is also positive. Similarly, if they are both negative, then the
sum is also negative.
It turns out restricting to positive roots is not enough. However, the following
trick does the job:
Definition
(Simple root)
.
A simple root is a positive root that cannot be written
as a sum of two positive roots. We write Φ
S
for the set of simple roots.
We can immediately deduce a few properties about these simple roots.
Proposition.
Any positive root can be written as a linear combination of simple
roots with positive integer coefficients. So every root can be written as a linear
combination of simple roots.
Proof.
Given any positive root, if it cannot be decomposed into a positive sum
of other roots, then it is simple. Otherwise, do so, and further decompose the
constituents. This will have to stop because there are only finitely many roots,
and then you are done.
Corollary. The simple roots span h
R
.
To show that they are independent, we need to do a bit of work that actually
involves Lie algebra theory.
Proposition. If α, β Φ are simple, then α β is not a root.
Proof.
Suppose
α β
were a root. By swapping
α
and
β
if necessary, we may
wlog assume that α β is a positive root. Then
α = β + (α β)
is a sum of two positive roots, which is a contradiction.
Proposition. If α, β Φ
S
, then the α-string through β, namely
S
α,β
= {β + Φ},
has length
αβ
= 1
2(α, β)
(α, α)
N.
Proof. Recall that there exists n
±
such that
S
α,β
= {β + : n
n n
+
},
We have shown before that
n
+
+ n
=
2(α, β)
(α, α)
Z.
In the case where
α, β
are simple roots, we know that
β α
is not a root. So
n
0. But we know β is a root. So we know that n
= 0 and hence
n
+
=
2(α, β)
(α, α)
N.
So there are
n
+
+ 1 = 1
2(α, β)
(α, α)
things in the string.
From this formula, we learn that
Corollary. For any distinct simple roots α, β, we have
(α, β) 0.
We are now in a position to check that we do get a basis this way.
Proposition. Simple roots are linearly independent.
Proof.
Suppose we have a non-trivial linear combination
λ
of the simple roots.
We write
λ = λ
+
λ
=
X
iI
+
c
i
α
(i)
X
jI
b
j
α
(j)
,
where
c
i
, b
j
0 and
I
+
, I
are disjoint. If
I
is empty, then the sum is a positive
root, and in particular is non-zero. Similarly, if
I
+
is empty, then the sum is
negative. So it suffices to focus on the case c
i
, b
j
> 0.
Then we have
(λ, λ) = (λ
+
, λ
+
) + (λ
, λ
) 2(λ
+
, λ
)
> 2(λ
+
, λ
)
= 2
X
iI
+
X
jI
c
i
b
j
(α
(i)
, α
(j)
)
0,
since (
α
(i)
, α
(j)
)
0 for all simple roots
α
(i)
, α
(j)
. So in particular
λ
is non-
zero.
Corollary. There are exactly r = rank g simple roots roots, i.e.
|Φ
S
| = r.
5.6 The classification
We now have a (not so) canonical choice of basis of h
R
B = {α Φ
S
} = {α
(i)
: i = 1, ··· , r}.
We want to re-express the Lie algebra in terms of this basis.
Definition
(Cartan matrix)
.
The Cartan matrix
A
ij
is defined as the
r × r
matrix
A
ij
=
2(α
(i)
, α
(j)
)
(α
(j)
, α
(j)
)
.
Note that this is not a symmetric matrix. We immediately know that
A
ij
Z
.
For each root α
(i)
, we have an su(2) subalgebra given by
{h
i
= h
α
(i)
, e
i
±
= e
±α
(i)
}.
We then have
[h
i
, e
i
±
] = ±2e
i
±
, [e
i
+
, e
i
] = h
i
.
Looking at everything in our Lie algebra, we have the relations
[h
i
, h
j
] = 0
[h
i
, e
j
±
] = ±A
ji
e
j
±
[e
i
+
, e
j
] = δ
ij
h
i
,
with no summation implied in the second row. Everything here we’ve seen,
except for the last expression [
e
i
+
, e
j
]. We are claiming that if
i 6
=
j
, then the
bracket in fact vanishes! This just follows from the fact that if
α
(i)
and
α
(j)
are
simple, then α
(i)
α
(j)
is not a root.
Note that this does not a priori tell us everything about the Lie algebra. We
have only accounted for step generators of the form
e
i
±
, and also we did not
mention how, say, [e
i
+
, e
j
+
] behaves.
We know
[e
i
+
, e
j
+
] = ad
e
i
+
(e
j
+
) e
α
(i)
+α
(j)
if
α
(i)
+
α
(j)
Φ. We know that if
α
(i)
+
α
(j)
Φ, then it belongs to a string.
In general, we have
ad
n
e
i
+
(e
j
+
) e
α
(j)
+
(i)
if α
(j)
+
(i)
Φ, and we know how long this string is. So we have
(ad
e
i
±
)
1A
ji
e
j
±
= 0.
This is the Serre relation. It turns out this is all we need to completely charac-
terize the Lie algebra.
Theorem
(Cartan)
.
Any finite-dimensional, simple, complex Lie algebra is
uniquely determined by its Cartan matrix.
We will not prove this.
To achieve the Cartan classification, we need to classify all Cartan matrices,
and then reconstruct g form the Cartan matrix.
We first do the first part. Recall that
A
ij
=
2(α
(i)
, α
(j)
)
(α
(j)
, α
(j)
)
Z.
We first read off properties of the Cartan matrix.
Proposition. We always have A
ii
= 2 for i = 1, ··· , r.
This is not a symmetric matrix in general, but we do have the following:
Proposition. A
ij
= 0 if and only if A
ji
= 0.
We also have
Proposition. A
ij
Z
0
for i 6= j.
We now get to a less trivial fact:
Proposition. We have det A > 0.
Proof. Recall that we defined our inner product as
(α, β) = α
T
κ
1
β,
and we know this is positive definite. So we know
det κ
1
>
0. We now write
the Cartan matrix as
A = κ
1
D,
where we have
D
j
k
=
2
(α
(j)
, α
(j)
)
δ
j
k
.
Then we have
det D =
Y
j
2
(α
(j)
, α
(j)
)
> 0.
So it follows that
det A = det κ
1
det D > 0.
It is an exercise to check that if the Cartan matrix is reducible, i.e. it looks
like
A =
A
(1)
0
0 A
(2)
,
then in fact the Lie algebra is not simple.
So we finally have
Proposition. The Cartan matrix A is irreducible.
So in total, we have the following five properties:
Proposition.
(i) A
ii
= 2 for all i.
(ii) A
ij
= 0 if and only if A
ji
= 0.
(iii) A
ij
Z
0
for i 6= j.
(iv) det A > 0.
(v) A is irreducible.
These are hugely restrictive.
Example.
For a rank 1 matrix, there is only one entry, and it is on the diagonal.
So we must have
A =
2
.
This corresponds to the simple Lie algebra su(2).
Example.
For a rank 2 matrix, we know the diagonals must be 2, so we must
have something of the form
A =
2 m
2
.
We know that the off-diagonals are not all zero, but if one is non-zero, then the
other also is. So they must be both non-zero, and we have
m <
4. We thus
have
(m, ) = (1, 1), (1, 2), (1, 3).
Note that we do not write out the cases where we swap the two entries, because
we will get the same Lie algebra but with a different ordering of the basis.
Now we see that we have a redundancy in the description of the Cartan
matrix given by permutation. There is a neat solution to this problem by drawing
Dynkin diagrams.
Definition
(Dynkin diagram)
.
Given a Cartan matrix, we draw a diagram as
follows:
(i) For each simple root α
(i)
Φ
S
, we draw a node
(ii) We join the nodes corresponding to α
(i)
, α
(j)
with A
ij
A
ji
many lines.
(iii)
If the roots have different lengths, we draw an arrow from the longer root
to the shorter root. This happens when A
ij
6= A
ji
.
Note that we wouldn’t have to draw too many lines. We have
A
ij
=
2|α
(i)
|
|α
(j)
|
cos ϕ
ij
, A
ji
=
2|α
(j)
|
|α
(i)
|
cos ϕ
ij
,
where ϕ
ij
is the angle between them. So we have
cos
2
ϕ
ij
=
1
4
A
ij
A
ji
.
But we know cos
2
ϕ
ij
[0, 1]. So we must have
A
ij
A
ji
{0, 1, 2, 3}.
So we have to draw at most 3 lines, which isn’t that bad. Moreover, we have the
following information:
Proposition. A simple Lie algebra has roots of at most 2 distinct lengths.
Proof. See example sheet.
It is an exercise to see that all the information about
A
ji
can be found from
the Dynkin diagram.
We can now revisit the case of a rank 2 simple Lie algebra.
Example. The Cartan matrices of rank 2 are given by
2 1
1 2
,
2 2
1 2
,
2 3
1 2
These correspond to the Dynkin diagrams
The conditions on the matrices
A
ij
now translate to conditions on the Dynkin
diagrams which we will not write out, and it turns out we can classify all the
allowed Dynkin diagrams as follows:
Theorem
(Cartan classification)
.
The possible Dynkin diagrams include the
following infinite families (where n is the number of vertices):
A
n
:
B
n
:
C
n
:
D
n
:
And there are also five exceptional cases:
E
6
:
E
7
:
E
8
:
F
4
:
G
2
:
Example.
The infinite families
A
n
, B
n
, C
n
, D
n
correspond to well-known com-
plex Lie groups by
Family Lie Group
A
n
L
C
(SU(n + 1))
B
n
L
C
(SO(2n + 1))
C
n
L
C
(Sp(2n))
D
n
L
C
(SO(2n))
where the Sp(2n) are the symplectic matrices.
Note that there is some repetition in our list. For example, we have
A
1
=
B
1
=
C
1
=
D
1
and
B
2
=
C
2
. Also,
D
2
does not give a simple Lie algebra, since
it is disconnected. We have D
2
=
A
1
A
1
. So we have
L
C
(SO(4))
=
L
C
(SU(2)) L
C
(SU(2)).
Finally, we also have D
3
= A
3
, and this reflects an isomorphism
L
C
(SU(4))
=
L
C
(SO(6)).
Hence, a list without repetitions is given by
A
n
for n 1,
B
n
for n 2,
C
n
for n 3,
D
n
for n 4.
This classification is very important in modern theoretical physics, since in
many theories, we need to pick a Lie group as, say, our gauge group. So knowing
what Lie groups are around lets us know what theories we can have.
5.7 Reconstruction
Now given the Cartan matrix, we want to reconstruct the Lie algebra itself.
Recall that we had a Cartan-Weyl basis
{H
i
, E
α
: i = 1, ··· , r : α Φ}.
The first thing to do in the reconstruction is to figure out what the set of roots
is.
By definition, the Cartan matrix determines simple roots
α
(i)
for
i
= 1
, ··· , r
.
We can read off the inner products from the Cartan matrix as
A
ij
=
2(α
(i)
, α
(j)
)
(α
(j)
, α
(j)
)
=
2|α
(i)
|
|α
(j)
|
cos ϕ
ij
.
This allows us to find the simple roots.
How about the other roots? We can find them by considering the root strings,
as we know that the length of the α
(i)
-string through α
(j)
is given by
ij
= 1 A
ji
N.
So we can work out the length of the root string from each simple root. By
Cartan’s theorem, this gives us all roots.
Instead of going through a formal and general reconstruction, we do an
example.
Example. Consider g = A
2
. We have a Dynkin diagram
A
2
:
So we see that the Cartan matrix is
A =
2 1
1 2
.
So we know that g has two simple roots α, β Φ, and we know
A
12
= A
21
=
2(α, β)
(α, α)
=
2(β, α)
(β, β)
= 1.
So we find that (α, α) = (β, β), and ϕ
αβ
=
2π
3
. So we can display the roots as
α
β
2π
3
Since α, β Φ
S
, we know ±(α β) 6∈ Φ, and also we have
α,β
=
β
= 1
2(α, β)
(α, α)
= 2.
So we have roots
β
+
nα, α
+
˜
for
n, ˜n {
0
,
1
}
. So in fact our list of roots
contains
α, β, α + β Φ.
These are the positive ones. We know each of these roots has a negative
counterpart. So we have more roots
α, β, α β Φ.
Then by our theorem, we know these are all the roots. We can find the length
of α + β by
(α + β, α + β) = (α, α) + (β, β) + 2(α, β)
= (α, α) + (β, β) (β, β)
= (α, α).
So in fact all roots have the same length. So if we draw all our roots, we get
α
β
α + β
α
βα β
So the Cartan-Weyl basis consists of
B
CW
= {H
1
, H
2
, E
±α
, E
±β
, E
±(α+β)
}.
This is the complete Cartan-Weyl basis of
A
2
. So this has dimension 2 + 6 = 8.
If we want to find out the brackets of these things, we look at the ones we
have previously written down, and use the Jacobi identity many times.
6 Representation of Lie algebras
We now try to understand the possible representations of Lie algebras. It turns
out they again are pretty restricted.
6.1 Weights
Let
ρ
be a representation of
g
on
V
. Then it is completely determined by the
images
H
i
7→ ρ(H
i
) gl(V )
E
α
7→ ρ(E
α
) gl(V ).
Again, we know that
[ρ(H
i
), ρ(H
j
)] = ρ([H
i
, H
j
]) = 0.
So all
ρ
(
H
i
) commute, and by linear algebra, we know they have a common
eigenvector v
λ
. Again, for each H h, we know
ρ(H)v
λ
= λ(H)v
λ
for some λ(H), and λ lives in h
.
Definition
(Weight of representation)
.
Let
ρ
:
g gl
(
V
) be a representation
of
g
. Then if
v
λ
V
is an eigenvector of
ρ
(
H
) for all
H h
, we say
λ h
is a
weight of ρ, where
ρ(H)v
λ
= λ(H)v
λ
for all H h.
The weight set S
ρ
of ρ is the set of all weights.
For a weight
λ
, we write
V
λ
for the subspace that consists of vectors
v
such
that
ρ(H)v = λ(H)v.
Note that the weights can have multiplicity, i.e.
V
λ
need not be 1-dimensional.
We write
m
λ
= dim V
λ
1
for the multiplicity of the weight.
Example. By definition, we have
[H
i
, E
α
] = α
i
E
α
So we have
ad
H
i
E
α
= α
i
E
α
.
In terms of the adjoint representation, this says
ρ
adj
(H
i
)E
α
= α
i
E
α
.
So the roots α are the weights of the adjoint representation.
Recall that for
su
C
(2), we know that the weights are always integers. This
will be the case in general as well.
Given any representation
ρ
, we know it has at least one weight
λ
with an
eigenvector
v V
λ
. We’ll see what happens when we apply the step operators
ρ(E
α
) to it, for α Φ. We have
ρ(H
i
)ρ(E
α
)v = ρ(E
α
)ρ(H
i
)v + [ρ(H
i
), ρ(E
α
)]v.
We know that
[ρ(H
i
), ρ(E
α
)] = ρ([H
i
, E
α
]) = α
i
ρ(E
α
).
So we know that if v V
λ
, then
ρ(H
i
)ρ(E
α
)v = (λ
i
+ α
i
)ρ(E
α
)v.
So the weight of the vector has been shifted by
α
. Thus, for all vectors
v V
λ
,
we find that
ρ(E
α
)v V
λ+α
.
However, we do not know a priori if
V
λ+α
is a thing at all. If
V
λ+α
=
{
0
}
, i.e.
λ + α is not a weight, then we know that ρ(E
α
)v = 0.
So the Cartan elements
H
i
preserve the weights, and the step operators
E
α
increment the weights by α.
Consider the action of our favorite
su
(2)
α
subalgebra on the representation
space. In other words, we consider the action of
{ρ
(
h
α
)
, ρ
(
e
α
)
, ρ
(
e
α
)
}
on
V
.
Then
V
becomes the representation space for some representation
ρ
α
for
su
(2)
α
.
Now we can use what we know about the representations of
su
(2) to get
something interesting about V . Recall that we defined
h
α
=
2
(α, α)
H
α
=
2
(α, α)
(κ
1
)
ij
α
j
H
i
.
So for any v V
λ
, after some lines of algebra, we find that
ρ(h
α
)(v) =
2(α, λ)
(α, α)
v.
So we know that we must have
2(α, λ)
(α, α)
Z
for all
λ S
ρ
and
α
Φ. So in particular, we know that (
α
(i)
, λ
)
R
for all
simple roots
α
(i)
. In particular, it follows that
λ h
R
(if we write
λ
as a sum
of the
α
(i)
, then the coefficients are the unique solution to a real system of
equations, hence are real).
Again, from these calculations, we see that
L
λS
ρ
V
λ
is a subrepresentation
of
V
, so by irreducibility, everything is a linear combination of these simultaneous
eigenvectors, i.e. we must have
V =
M
λS
ρ
V
λ
.
In particular, this means all ρ(H) are diagonalizable for H h.
6.2 Root and weight lattices
Recall that the simple roots are a basis for the whole root space. In particular,
any positive root can be written as a positive sum of simple roots. So for any
β Φ, we can write
β =
r
X
i=1
β
i
α
(i)
,
where β
i
Z for i = 1, ··· , r. Hence all roots lie in a root lattice:
Definition
(Root lattice)
.
Let
g
be a Lie algebra with simple roots
α
(i)
. Then
the root lattice is defined as
L[g] = span
Z
{α
(i)
: i = 1, ··· , r}.
Note that here we are taking the integer span, not the real span.
What we want to do is to define an analogous weight lattice, where all possible
weights lie.
We start by defining some funny normalization of the roots:
Definition (Simple co-root). The simple co-roots are
ˇα
(i)
=
2α
(i)
(α
(i)
, α
(i)
)
.
Similarly, we can define the co-root lattice:
Definition
(Co-root lattice)
.
Let
g
be a Lie algebra with simple roots
α
(i)
.
Then the co-root lattice is defined as
ˇ
L[g] = span
Z
{ˇα
(i)
: i = 1, ··· , r}.
Definition
(Weight lattice)
.
The weight lattice
L
W
[
g
] is the dual to the co-root
lattice:
L
W
[g] =
ˇ
L
[g] = {λ h
R
: (λ, µ) Z for all µ
ˇ
L[g]}.
So we know λ L
W
[g] iff for all α
(i)
, we have
(λ, ˇα
(i)
) =
2(α
(i)
, λ)
(α
(i)
, α
(i)
)
Z.
So by what we have previously computed, we know that all weights
λ S
ρ
are
in L
W
[g].
Given the basis
B = {ˇα
(i)
: i = 1, ··· , r}
for
ˇ
L[g], we define the dual basis
B
= {ω
(i)
: i = 1, ··· , r}
by demanding
(ˇα
(i)
, ω
(j)
) =
2(α
(i)
, ω
(j)
)
(α
(i)
, α
(i)
)
= δ
ij
.
These
ω
(i)
are known as the fundamental weights. As the simple roots span
h
R
,
we can write
ω
(i)
=
r
X
j=1
B
ij
α
(j)
for some B
ij
R. Plugging this into the definition, we find
r
X
k=1
2(α
(i)
, α
(k)
)
(α
(i)
, α
(i)
)
B
jk
= δ
i
j
.
So we know
r
X
k=1
B
jk
A
ki
= δ
i
j
,
i.e. B is the inverse of the Cartan matrix A. Thus we can write
α
(i)
=
r
X
j=1
A
ij
ω
(j)
.
Example. Consider g = A
2
. Then we have
A =
2 1
1 2
.
Then we have
α = α
(1)
= 2ω
(1)
ω
(2)
β = α
(2)
= ω
(1)
+ 2ω
(2)
.
So we have
ω
(1)
=
1
3
(2α + β)
ω
(2)
=
1
3
(α + 2β).
We can then draw the weight lattice:
ω
(2)
ω
(1)
α
β
Note that the roots are also in the weight lattice, because they are weights of
the adjoint representation.
Given any weight λ S
ρ
L
W
[g], we can write
λ =
r
X
i=1
λ
i
ω
(i)
for some λ
i
Z. These {λ
i
} are called Dynkin labels of the weight λ.
6.3 Classification of representations
Suppose we have a representation
ρ
of
g
. We can start with any eigenvector
v V
λ
. We then keep applying things of the form
ρ
(
E
α
), where
α
Φ
+
is a
positive weight. Doing so will bring us further away from the hyperplane dividing
the positive and negative weights, and since there are only finitely many weights,
we must eventually stop. This is known as a highest weight.
Definition
(Highest weight)
.
A highest weight of a representation
ρ
is a weight
Λ S
ρ
whose associated eigenvector v
Λ
is such that
ρ(E
α
)v
Λ
= 0
for all α Φ
+
.
The Dynkin labels (or indices) of a representation are the Dynkin labels of
its highest weight.
Now if v V
λ
, then we know
ρ(E
α
)v V
λ+α
,
if
λ
+
α S
ρ
(and vanishes otherwise). So acting with the step operators
translates us in the weight lattice by the vector corresponding to the root.
In the case of
su
(2), we found out an explicit formula for when this will stop.
In this general case, we have the following result which we will not prove:
Theorem. For any finite-dimensional representation of g, if
λ =
r
X
i=1
λ
i
ω
(i)
S
ρ
,
then we know
λ m
(i)
α
(i)
S
ρ
,
for all m
(i)
Z and 0 m
(i)
λ
i
.
If we know further that
ρ
is irreducible, then we can in fact obtain all weights
by starting at the highest weight and applying this procedure.
Moreover, for any
Λ =
X
Λ
i
ω
(i)
L
W
[g],
this is the highest weight of some irreducible representation if and only if Λ
i
0
for all i.
This gives us a very concrete way of finding, at least the weights, of an
irreducible representation. In general, though, we don’t immediately know the
multiplicities of the weights.
Definition
(Dominant integral weight)
.
A dominant integral weight is a weight
Λ =
X
Λ
i
ω
(i)
L
W
[g],
such that Λ
i
0 for all i.
Example.
Take the example of
g
=
A
2
. It is a fact that the fundamental
representation f has Dynkin labels (1, 0). In other words,
Λ = ω
(1)
.
To get the remaining weights, we subtract roots:
(i) Λ = ω
(1)
S
f
. So we can subtract by α
(1)
exactly once.
(ii) So
λ = ω
(1)
(2ω
(1)
ω
(2)
)
= ω
(1)
+ ω
(2)
S
f
.
This has Dynkin labels (
1
,
1). So we can do one further subtraction to
get a new weight
λ α
(2)
= ω
(1)
+ ω
(2)
(2ω
(2)
ω
(1)
)
= ω
(2)
In the weight diagram, these are given by
ω
(2)
ω
(1)
α
β
(1, 0)
(1, 1)
(0, 1)
In general, for a dominant integral weight of the form
Λ = Λ
1
ω
(1)
+ Λ
2
ω
(2)
L
W
[A
2
]
for Λ
1
, Λ
2
Z
0
. We then get an irrep ρ
1
,Λ
2
)
of A
2
.
One can show, via characters, that the dimension of this representation is
given by
dim ρ
1
,Λ
2
)
=
1
2
1
+ 1)(Λ
2
+ 1)(Λ
1
+ Λ
2
+ 2).
This is symmetric in Λ
1
and Λ
2
. So if Λ
1
6
= Λ
2
, then we get a pair of distinct
representations of the same dimension. It is a fact that
λ S
1
,Λ
2
)
λ S
2
,Λ
1
)
,
and we have
ρ
1
,Λ
2
)
= ¯ρ
2
,Λ
1
)
.
On the other hand, if Λ
1
= Λ
2
, then this representation is self-conjugate.
The first few representations of A
2
are given by
Highest weight Dimension Name
ρ
(0,0)
1 Trivial
ρ
(1,0)
3 Fundamental
ρ
(0,1)
¯
3 Anti-fundamental
ρ
(2,0)
6
ρ
(0,2)
¯
6
ρ
(1,1)
8 Adjoint
We can figure out the weights of the adjoint representation with Λ = (1
,
1)
S
ρ
.
Since both labels are positive, we can subtract both α
(1)
and α
(2)
to get
Λ α
(1)
= (1, 2)
Λ α
(2)
= (2, 1)
ω
(2)
ω
(1)
α
β
(1, 1)(1, 2)
(2, 1)
Now we have a +2 in the Dynkin labels. So we can subtract twice. So we have
the following weights:
Λ α
(1)
α
(2)
= (0, 0)
Λ α
(1)
2α
(2)
= (1, 2)
Λ 2α
(1)
α
(2)
= (2, 1)
ω
(2)
ω
(1)
α
β
(1, 1)(1, 2)
(2, 1)
(0, 0)
(1, 2)
(2, 1)
Finally, we can subtract α
(1)
from the second or α
(2)
from the third to get
Λ 2α
(1)
2α
(2)
= (1, 1) S
ρ
.
ω
(2)
ω
(1)
α
β
(1, 1)(1, 2)
(2, 1)
(0, 0)
(2, 1)
(2, 1)
(1, 1)
Now notice that we only have seven weights, not eight. This is because our
algorithm did not tell us the multiplicity of the weights. However, in this
particular case, we know, because this is the adjoint representation, and the
Cartan generators give us two things of weight 0. So we can plot all weights
with multiplicity as
It is a general fact that the weight lattice traces out a polygon like this, and
the vertices at the perimeter are always non-degenerate. So in this case, we
could also have figured out the multiplicty by using this general fact.
6.4 Decomposition of tensor products
The last thing we need to know how to do is how we can decompose tensor
products of irreps. This is not hard. It is just what we did for
su
(2) but in a
more general context.
We let
ρ
Λ
and
ρ
Λ
0
be irreps of
g
with corresponding representation spaces
V
Λ
and V
0
Λ
0
. Then we can write
V
Λ
=
M
λS
Λ
V
λ
, V
0
Λ
0
=
M
λ
0
S
Λ
0
V
0
λ
0
.
If v
λ
V
Λ
and v
λ
0
V
0
Λ
0
, then for H h, we have
(ρ
Λ
ρ
Λ
0
)(H)(v
λ
v
λ
0
) = ρ
Λ
(H)v
λ
v
λ
0
+ v
λ
ρ
Λ
0
(H)v
λ
0
= λ(H)v
λ
v
λ
0
+ λ
0
(H)v
λ
v
λ
0
= (λ + λ
0
)(H)v
λ
v
λ
0
.
Since things of the form v
λ
v
λ
0
span V
Λ
V
Λ
0
, we know
S
ΛΛ
0
= {λ + λ
0
: λ S
Λ
, λ
0
S
Λ
0
}
with multiplicities. To figure out the decomposition, we find out what the highest
weight of the tensor product is, and then subtract off the factors corresponding
to the highest weight, and then repeat.
Example.
Consider
g
=
A
2
, and the fundamental representation
ρ
(1,0)
. We
consider
ρ
(1,0)
ρ
(1,0)
.
Recall that the weight set is given by
S
(1,0)
= {ω
(1)
, ω
(1)
+ ω
(2)
, ω
(2)
}.
We can then draw the collection of all weights in ρ
(1,0)
ρ
(1,0)
as follows:
α
β
So we see that we have a highest weight of (2
,
0). We then use the algorithm to
compute all the weights of such an irrep, remove it from the weight lattice, and
be left with
α
β
This is just the anti-fundamental representation. So we have
3 3 = 6
¯
3.
7 Gauge theories
We are now going to do something rather amazing. We are going to show that
Maxwell’s equation is “forced” upon us by requiring that our (charged) fields
possess something known as a U(1) gauge symmetry. We are going to write
down essentially the only possible theory under the assumption of this gauge
symmetry, and unwrapping it gives us Maxwell’s equation.
When we do it for Maxwell’s equation, it is not entirely clear that what we
wrote down was “the only possible thing”, but it will be when we try to do it
for general gauge theories.
Before we begin, it is important to note that everything we do here has a very
nice geometric interpretation in terms of connections on principal
G
-bundles,
and there are very good geometric reasons to pick the definitions (and names) we
use here. Unfortunately, we do not have the time to introduce all the necessary
geometry in order to do so.
7.1 Electromagnetism and U(1) gauge symmetry
In electromagnetism, we had two fields
E
and
B
. There are four Maxwell’s
equations governing how they behave. Two of them specify the evolution of the
field and how they interact with matter, while the other two just tells us we can
write the field in terms of a scalar and a vector potential Φ
, A
. Explicitly, they
are related by
E = −∇Φ +
A
t
B = × A.
The choice of potentials is not unique. We know that
E
and
B
are invariant
under the transformations
Φ 7→ Φ +
α
t
A 7→ A + α
for any gauge function α = α(x, t).
Now we believe in relativity, so we should write things as 4-vectors. It so
happens that there are four degrees of freedom in the potentials, so we can
produce a 4-vector
a
µ
=
Φ
A
i
.
We can then succinctly write the gauge transformations as
a
µ
7→ a
µ
+
µ
α.
We can recover our fields via something known as the electromagnetic field tensor,
given by
f
µν
=
µ
a
ν
ν
a
µ
.
It is now easy to see that
f
is invariant under gauge transformations. We can
expand the definitions out to find that we actually have
f
µν
=
0 E
x
E
y
E
z
E
x
0 B
z
B
y
E
y
B
z
0 B
x
E
z
B
y
B
x
0
So we do recover the electric and magnetic fields this way.
The free field theory of electromagnetism then has a Lagrangian of
L
EM
=
1
4g
2
f
µν
f
µν
=
1
2g
2
(E
2
B
2
).
For reasons we will see later on, it is convenient to rescale the a
µ
by
A
µ
= ia
µ
iR = L(U(1))
and write
F
µν
= if
µν
=
µ
A
ν
ν
A
µ
.
Now the magic happens when we want to couple this to matter.
Suppose we have a complex scalar field φ : R
3,1
C with Lagrangian
L
φ
=
µ
φ
µ
φ W (φ
φ).
The interesting thing is that this has a global U(1) symmetry by
φ 7→ gφ
φ
7→ g
1
φ
for g = e
U(1), i.e. the Lagrangian is invariant under this action.
In general, it is more convenient to talk about infinitesimal transformations
Consider an element
g = exp(X) 1 + X,
where we think of X as “small”. In our case, we have X u(1)
=
iR, and
φ 7→ φ + δ
X
φ,
φ
7→ φ
+ δ
X
φ
,
where
δ
X
φ = Xφ
δ
X
φ
= Xφ
.
Now this is a global symmetry, i.e. this is a symmetry if we do the same
transformation at all points in the space. Since we have
δ
X
L
φ
= 0,
we get a conserved charge. What if we want a local symmetry? We want to have
a different transformation at every point in space, i.e. we now have a function
g : R
3,1
U(1),
and we consider the transformations given by
φ(x) 7→ g(x)φ(x)
φ
(x) 7→ g
1
(x)φ
(x).
This is in general no longer a symmetry. Under an infinitesimal variation
X : R
3,1
u(1), we have
δ
X
φ = Xφ.
So the derivative transforms as
δ
X
(
µ
φ) =
µ
(δ
X
φ) = (
µ
X)φ + X
µ
φ.
This is bad. What we really want is for this to transform like φ, so that
µ
φ 7→ g(x)
µ
φ.
Then the term
µ
φ
µ
φ will be preserved.
It turns out the solution is to couple
φ
with
A
µ
. Recall that both of these
things had gauge transformations. We now demand that under any gauge
transformation, both should transform the same way. So from now on, a gauge
transformation X : R
3,1
u(1) transforms both φ and A
µ
by
φ 7→ φ + Xφ
A
µ
7→ A
µ
µ
X.
However, this does not fix the problem, since everything we know so far that
involves the potential
A
is invariant under gauge transformation. We now do
the funny thing. We introduce something known as the covariant derivative:
D
µ
=
µ
+ A
µ
,
Similar to the case of general relativity (if you are doing that course), the
covariant derivative is the “right” notion of derivative we should use whenever
we want to differentiate fields that are coupled with
A
. We shall now check that
this derivative transforms in the same way as φ. Indeed, we have
δ
X
(D
µ
φ) = δ
X
(
µ
φ + A
µ
φ)
=
µ
(δ
X
φ) + A
µ
δ
X
φ
µ
Xφ
= X
µ
φ + XA
µ
φ
= XD
µ
φ.
This implies that the kinetic term
(D
µ
φ)
D
µ
φ
is gauge invariant. So we can put together a gauge-invariant Lagrangian
L =
1
4g
2
F
µν
F
µν
+ (D
µ
φ)
(D
µ
φ) W (φ
φ).
So what the electromagnetic potential
A
gives us is a covariant derivative D
which then allows us to “gauge” these complex fields to give them a larger
symmetry group.
7.2 General case
We now want to do what we had above for a general Lie group
G
. In particular,
this can be non-abelian, and thus the Lie algebra has non-trivial bracket. It
turns out in the general case, it is appropriate to insert some brackets into what
we have done so far, but otherwise most things follow straight through.
In the case of electromagnetism, we had a Lie group U(1), and it acted on
C
in the obvious way. In general, we start with a Lie group
G
, and we have
to pick a representation
D
of
G
on a vector space
V
. Since we want to have a
real Lagrangian, we will assume our vector space
V
also comes with an inner
product.
Given such a representation, we consider fields that take values in
V
, i.e.
functions
φ
:
R
3,1
V
. We can, as before, try to write down a (scalar)
Lagrangian
L
φ
= (
µ
φ,
µ
φ) W ((φ, φ)).
Writing out the summation explicitly, we have
L
φ
=
3
X
µ=0
(
µ
φ,
µ
φ) W ((φ, φ)),
i.e. the sum is performed outside the inner product.
This is not always going to be invariant under an arbitrary action of the
group
G
, because the inner product is not necessarily preserved. Fortunately, by
definition of being unitary, what we require is that our representation is unitary,
i.e. each D(g) is unitary for all g G.
It is a theorem that any representation of a compact Lie group is equivalent
to one that is unitary, so we are not losing too much by assuming this (it is also
that a non-compact Lie group cannot have a non-trivial unitary representation).
Again, it will be convenient to think about this in terms of infinitesimal
transformations. Near the identity, we can write
g = exp(X)
for some X g. Then we can write
D(g) = exp(ρ(X)) gl(n, C).
We write
ρ
:
g gl
(
n, C
) for the associated representation of the Lie algebra
g
. If
D
is unitary, then
ρ
will be anti-Hermitian. Then the infinitesimal transformation
is given by
φ 7→ φ + δ
X
φ = φ + ρ(X)φ.
In general, an infinitesimal Gauge transformation is then given by specifying a
function
X : R
3,1
g.
Our transformation is now
δ
X
φ = ρ(X(x))φ.
Just as in the abelian case, we know our Lagrangian is no longer gauge invariant
in general.
We now try to copy our previous fix. Again, we suppose our universe comes
with a gauge field
A
µ
: R
3,1
L(G).
Again, we can define a covariant derivative
D
µ
φ =
µ
φ + ρ(A
µ
)φ.
In the case of a non-abelian gauge symmetry, our gauge field transformed under
X as
δ
X
A
µ
=
µ
X.
This expression still makes sense, but it turns out this isn’t what we want, if
we try to compute
δ
X
(D
µ
φ
). Since the gauge field is non-abelian, there is one
extra thing we can include, namely [
X, A
µ
]. It turns out this is the right thing
we need. We claim that the right definition is
δ
X
A
µ
=
µ
X + [X, A
µ
].
To see this, we show that
Proposition. We have
δ
X
(D
µ
φ) = ρ(X)D
µ
φ.
The proof involves writing out the terms and see that it works.
Proof. We have
δ
X
(D
µ
φ) = δ
X
(
µ
φ + ρ(A
µ
)φ)
=
µ
(δ
X
φ) + ρ(A
µ
)δ
X
φ + ρ(δ
X
A
µ
)φ
=
µ
(ρ(X)φ) + ρ(A
µ
)ρ(X)φ ρ(
µ
X)φ + ρ([X, A
µ
])φ
= ρ(
µ
X)φ + ρ(X)
µ
φ + ρ(X)ρ(A
µ
)φ
+ [ρ(A
µ
), ρ(X)]φ ρ(
µ
X)φ + ρ([X, A
µ
])φ
= ρ(X)(
µ
φ + ρ(A
µ
)φ)
= ρ(X)D
µ
φ,
as required.
Thus, we know that (D
µ
φ,
D
µ
φ
) is gauge invariant. We can then write down
a gauge invariant “matter” part of the action
L
φ
= (D
µ
φ, D
µ
φ) W [(φ, φ)].
Finally, we need to produce a gauge invariant kinetic term for
A
µ
:
R
3,1
L
(
G
).
For the case of an abelian gauge theory, we had
F
µν
=
µ
A
ν
ν
A
µ
L(G).
In this case, there is one extra term we can add, namely [
A
µ
, A
ν
]
g
, and it
turns out we need it for things to work out. Our field strength tensor is
F
µν
=
µ
A
ν
ν
A
µ
+ [A
µ
, A
ν
].
How does the field strength tensor transform? In the abelian case, we had that
F
µν
was invariant. Let’s see if that is the case here. It turns out this time the
field strength tensor transforms as the adjoint representation.
Lemma. We have
δ
X
(F
µν
) = [X, F
µν
] L(G).
Proof. We have
δ
X
(F
µν
) =
µ
(δ
X
A
ν
)
ν
(δ
X
A
µ
) + [δ
X
A
µ
, A
ν
] + [A
µ
, δ
X
A
ν
]
=
µ
ν
X +
µ
([X, A
ν
])
ν
µ
X
ν
([X, A
µ
]) [
µ
X, A
ν
]
[A
µ
,
ν
X] + [[X, A
µ
], A
ν
] + [A
µ
, [X, A
ν
]]
= [X,
µ
A
ν
] [X,
ν
A
µ
] + ([X, [A
µ
, A
n
]]
= [X, F
µν
].
where we used the Jacobi identity in the last part.
So to construct a scalar quantity out of
F
µν
, we need an inner product
invariant under an adjoint representation. We know one of these the Killing
form! So we can just pick
L
A
=
1
g
2
κ(F
µν
, F
µν
).
This is known as the Yang-Mills Lagrangian. Note that for each fixed
µ, ν
, we
have that F
µν
g. So we should read this as
L
A
=
1
g
2
X
µ,ν
κ(F
µν
, F
µν
).
Putting all these together, we get ourselves an invariant Lagrangian of the system
with a G gauge symmetry.
So the final Lagrangian looks like
L =
1
g
2
X
µ,ν
κ(F
µν
, F
µν
) + (D
µ
φ, D
µ
φ) + W ((φ, φ)).
Now if we are further told that we actually have a simple complex Lie algebra,
then it is a fact that we can find a real form of compact type. So in particular
we can find a basis B = {T
a
, a = 1, ··· , dim g} with
κ
ab
= κ(T
a
, T
b
) = κδ
ab
.
In this basis, we have
L
A
=
κ
g
2
d
X
a=1
F
µν,a
F
µν,a
,
and this does look like a copy of d many electromagnetic fields.
So to construct (a sensible) gauge theory, we need to get a (semi-)simple
Lie algebra, and then find some representations of
g
. These are things we have
already studied well before!
According to our current understanding, the gauge group of the universe is
G = U(1) × SU(2) × SU(3),
and, for example, the Higgs boson has representation
φ = (+1, ρ
1
, ρ
0
).
8 Lie groups in nature
We are now going to look around and see how the things we have done exhibit
themselves in nature. This chapter, by design, is very vague, and one shall not
expect to learn anything concrete from here.
8.1 Spacetime symmetry
In special relativity, we have a metric
ds
2
= dt
2
+ dx
2
+ dy
2
+ dz
2
.
The group of (orientation-preserving and) metric-preserving symmetries gives
us the Lorentz group
SO
(3
,
1). However, in certain cases, we can get something
more (or perhaps less) interesting. Sometimes it makes sense to substitute
τ
=
it
,
so that
ds
2
= dτ
2
+ dx
2
+ dy
2
+ dz
2
.
This technique is known as Wick rotation. If we put it this way, we now have a
symmetry group of SO(4) instead.
It happens that when we complexify this, this doesn’t really matter. The
resulting Lie algebra is
L
C
(SO(3, 1)) = L
C
(SO(4)) = D
2
.
Its Dynkin diagram is
This is not simple, and we have
D
2
= A
1
A
1
.
In other words, we have
so(4) = su(2) su(2).
At the Lie group level,
SO
(4) does not decompose as
SU
(2)
× SU
(2). Instead,
SU(2) × SU(2) is a double cover of SO(4), and we have
SO(4) =
SU(2) × SU(2)
Z
2
.
Now in general, our fields in physics transform when we change coordinates.
There is the boring case of a scalar, which never transforms, and the less boring
case of a vector, which transforms like a vector. More excitingly, we have objects
known as spinors. The weird thing is that spinors do have an
so
(3
,
1) action,
but this does not lift to an action of
SO
(3
,
1). Instead, we need to do something
funny with double covers, which we shall not go into.
In general, spinors decompose into “left-handed” and “right-handed” com-
ponents, known as Weyl fermions, and these correspond to two different rep-
resentations of
A
1
. We can summarize the things we have in the following
table:
Field so(3, 1) A
1
A
1
Scalar 1 (ρ
0
, ρ
0
)
Dirac Fermion LH RH (ρ
1
, ρ
0
) (ρ
0
, ρ
1
)
Vector 4 (ρ
1
, ρ
1
)
However, the Lorentz group
SO
(3
,
1) is not all the symmetries we can do to
Minkowski spacetime. These are just those symmetries that fix the origin. If we
add in the translations, then we obtain what is known as the Poincar´e group.
The Poincar´e algebra is not simple, as the translations form a non-trivial ideal.
8.2 Possible extensions
Now we might wonder how we can expand the symmetry group of our theory
to get something richer. In very special cases, we will get what is known as
conformal symmetry, but in general, we don’t. To expand the symmetry further,
we must consider supersymmetry.
Conformal field theory
If all our fields are massless, then our theory gains conformal invariance. Before
we look into conformal invariance, we first have a look at a weaker notion of
scale invariance.
A massless free field φ : R
3,1
V has a Lagrangian that looks like
L =
1
2
(
µ
φ,
µ
φ).
This is invariant under scaling, namely the following simultaneous transformations
parameterized by λ:
x 7→ x
0
= λ
1
x
φ(x) 7→ λ
φ(x
0
),
where ∆ = dim V is the dimension of the field.
Often in field theory, whenever we have scale invariance, we also get conformal
invariance. In general, a conformal transformation is one that preserves all angles.
So this, for example, includes some mixture and scaling, and is more general
than Lorentz invariance and scale invariance.
In the case of 4 (spacetime) dimensions, it turns out the conformal group
is
SO
(4
,
2), and has dimension 15. In general, members of this group can be
written as
0 D
D 0
P
µ
+ K
µ
P
ν
+ K
ν
M
µν
.
Here the
K
µ
are the special conformal generators, and the
P
µ
are translations.
We will not go deep into this.
Conformal symmetry gives rise to conformal field theory. Theoretically, this
is very important, since they provide “end points” of renormalization group flow.
This will be studied more in depth in the Advanced Quantum Field Theory
course.
Supersymmetry
Can we add even more symmetries? It turns out there is a no-go theorem that
says we can’t.
Theorem
(Coleman-Mondula)
.
In an interactive quantum field theory (sat-
isfying a few sensible conditions), the largest possible symmetry group is the
Poincar´e group times some internal symmetry that commutes with the Poincar´e
group.
But this is not the end of the world. The theorem assumes that we are working
with traditional Lie groups and Lie algebras. The idea of supersymmetry is to
prefix everything we talk about with “super-”, so that this theorem no longer
applies.
To being with, supersymmetry replaces Lie algebras with graded Lie algebras,
or Lie superalgebras. This is a graded vector space, which we can write as
g = g
0
g
1
,
where the elements in
g
0
as said to have grade 0, and elements in
g
1
have grade
1. We write
|X|
= 0
,
1 for
X g
0
, g
1
respectively. These will correspond to
bosonic and fermionic operators respectively.
Now the Lie bracket respects a graded anti-commutation relation
[X, Y ] = (1)
|X||Y |
[Y, X].
So for Bosonic generators, they anti-commute as usual, but with we have
two fermionic things, they commute. We can similarly formulate a super-
Jacobi identity. These can be used to develop field theories that have these
“supersymmetries”.
Unfortunately, there is not much experimental evidence for supersymmetry,
and unlike string theory, we are already reaching the scales we expect to see
supersymmetry in the LHC. . .
8.3 Internal symmetries and the eightfold way
Finally, we get to the notion of internal symmetries. Recall that for a complex
scalar field, our field was invariant under a global U(1) action given by phase
change. More generally, fields can carry some representation of a Lie group
G
.
In general, if this representation has dimension greater than 1, this means that
we have multiple different “particles” which are related under this symmetry
transformation. Then by symmetry, all these particles will have the same mass,
and we get a degeneracy in the mass spectrum.
When we have such degeneracies, we would want to distinguish the different
particles of the same mass. These can be done by looking at the weights of the
representation. It turns out the different weights correspond to the different
quantum numbers of the particles.
Often, we do not have an exact internal symmetry. Instead, we have an
approximate symmetry. This is the case when the dominant terms in the
Lagrangian are invariant under the symmetry, while some of the lesser terms are
not. In particular, the different particles usually have different (but very similar)
masses.
These internal symmetries are famously present in the study of hadrons, i.e.
things made out of quarks. The basic examples we all know are the nucleons,
namely protons and neutrons. We can list them as
Charge (Q) Mass (M)
p +1 938 MeV
n 0 940 MeV
Note that they have very similar masses. Later, we found, amongst many other
things, the pions:
Charge (Q) Mass (M)
π
+
+1 139 MeV
π
0
0 135 MeV
π
-1 139 MeV
Again, these have very similar masses. We might expect that there is some
approximate internal symmetry going on. This would imply that there is some
conserved quantity corresponding to the weights of the representation. Indeed,
we later found one, known as isospin:
Charge (Q) Isospin (J) Mass (M)
p +1 +
1
2
938 MeV
n 0
1
2
940 MeV
π
+
+1 +1 139 MeV
π
0
0 0 135 MeV
π
-1 -1 139 MeV
Isospin comes from an approximate
SU
(2)
I
symmetry, with a generator given by
H = 2J.
The nucleons then have the fundamental ρ
1
representation, and the pions have
the ρ
2
representation.
Eventually, we got smarter, and discovered an extra conserved quantum
number known as hypercharge. We can plot out the values of the isospin and
the hypercharge for our pions and some other particles we discovered, and we
found a pattern:
π
+
K
0
K
+
π
K
¯
K
0
η
π
0
At firsts, physicists were confused by the appearance of this pattern, and tried
very hard to figure out generalizations. Of course, now that we know about Lie
algebras, we know this is the weight diagram of a representation of su(3).
However, the word
SU
(3) hasn’t resolved all the mystery. In reality, we only
observed representations of dimension 1
,
8 and 10. We did not see anything
else. So physicists hypothesized that there are substructures known as quarks.
Each quark (
q
) carry a
3
representation (flavour), and antiquarks (
¯q
) carry a
¯
3
representation.
The mesons correspond to
q¯q
particles, and the representation decompose as
3
¯
3 = 1 8.
Bosons correspond to qqq particles, and these decompose as
3 3 3 = 1 8 8 10.
Of course, we now have to explain why only
q¯q
and
qqq
appear in nature,
and why we don’t see quarks appearing isolated in nature. To do so, we have
to go very deep down in QCD. This theory says that quarks have a
SU
(3)
gauge symmetry (which is a different
SU
(3)) and the quark again carries the
fundamental representation
3
. More details can be found in the Lent Standard
Model course.