Notes Group Theory in Physics
Notes Group Theory in Physics
Physics
Lecture Notes
MA PH 464 - Group Theory in
Physics
Lecture Notes
Vincent Bouchard
University of Alberta
These are lecture notes for MA PH 464 offered at the University of Al-
berta.
Motivation
vi
References
These lecture notes are not meant to be original. They are heavily influenced
by already existing lecture notes and textbooks on group theory in physics.
The main references that I have been using are:
vii
Contents
Motivation vi
References vii
2 Representation theory 29
2.1 Representations . . . . . . . . . . . . . . . . . . 30
2.2 Properties of representations . . . . . . . . . . . . . . 33
2.3 Unitary representations . . . . . . . . . . . . . . . 37
2.4 Semisimplicity . . . . . . . . . . . . . . . . . . . 41
2.5 Schur’s lemmas . . . . . . . . . . . . . . . . . . 43
2.6 The great orthogonality theorem . . . . . . . . . . . . 45
2.7 Characters . . . . . . . . . . . . . . . . . . . . 48
2.8 Orthogonality for characters . . . . . . . . . . . . . . 49
2.9 Reducibility and decomposition. . . . . . . . . . . . . 52
2.10 An example: S3 . . . . . . . . . . . . . . . . . . 55
2.11 An example: the regular representation . . . . . . . . . . 58
2.12 Real, pseudoreal and complex representations . . . . . . . 59
3 Applications 65
3.1 Crystallography . . . . . . . . . . . . . . . . . . 65
3.2 Quantum Mechanics . . . . . . . . . . . . . . . . . 69
3.3 Coupled harmonic oscillators. . . . . . . . . . . . . . 74
viii
CONTENTS ix
1.1 Groups
Objectives
You should be able to:
• Recall the definition of a group.
• Recognize various groups that occur frequently in physics and mathe-
matics.
1.1.1 Definition
In physics, we think of symmetries in terms of transformations of a system
(think of rotations), that we can compose with each other. What we want to
do here is capture the abstract properties that are shared by all such symmetry
transformations. This is what the abstract concept of a group does. Basically,
a group is a set with a binary operation (composition) that satisfies a bunch
of axioms.
Definition 1.1.1 Group. A group is a set G with a binary operation ·,
which we call composition or multiplication, that satisfies the following axioms:
• Closure: For all a, b ∈ G, a · b ∈ G.
• Associativity: For all a, b, c ∈ G, (a · b) · c = a · (b · c).
1
CHAPTER 1. BASIC CONCEPTS OF GROUP THEORY 2
1.1.2 Examples
You already know many examples of groups. Let us look at a few of them:
Example 1.1.6 Rotations. Rotations in two-dimensional space form a
group, denoted by SO(2), which is an abelian continuous group. Rotations in
three-dimensional space also form a group, which is denoted by SO(3), which
happens to be non-abelian. Similarly, rotations in n dimensions form the non-
abelian group SO(n).
Example 1.1.7 R and Z under addition. The set R with the operation
of addition forms an abelian continuous group. If we restrict to the subset of
integers Z, we still get an infinite abelian group, but it is now discrete.
Example 1.1.8 The trivial group. The group with only one element {1}
forms a group under multiplication, but it is a rather boring one. It is called
the trivial group.
Example 1.1.9 The cyclic group ZN . The two square roots of 1, {1, −1},
form the group Z2 under multiplication, which is an abelian discrete group of
order 2. More generally, the N ’th roots of unity, {e2πik/N }k=0,...,N −1 form the
group ZN under multiplication, which is an abelian discrete group of order N .
Example 1.1.10 The unitary group U (1). Complex numbers of norm 1,
namely eiθ with θ ∈ [0, 2π), form an abelian continuous group under multipli-
cation called U (1).
Example 1.1.11 Addition of integers mod N . Addition of integers mod
N generates an abelian discrete group of order N . The elements of the set are
{0, 1, . . . , N − 1}, with the operation of addition mod N . This has in fact the
same abstract group structure as ZN defined above.
Example 1.1.12 The symmetric group Sn . The set of invertible map-
pings f : S → S of a finite set S with n elements forms a group, which is
denoted Sn and called symmetric group (or permutation group). The ele-
ments of the group Sn are permutations of n objects. There are n! such distinct
permutations, so Sn has order n!, and is non-abelian for n ≥ 3.
CHAPTER 1. BASIC CONCEPTS OF GROUP THEORY 3
Example 1.1.13 The general linear groups GL(n, R) and GL(n, C). The
set of invertible linear maps f : V → V of a vector space V forms a continuous
group, known as the general linear group of V and denoted by GL(V ). If
V = Rn , we write GL(n, R); if V = Cn , we write GL(n, C). We can also think
of those as the sets of invertible n × n matrices with the operation of matrix
multiplication (with either real or complex entries).
Example 1.1.14 The special linear groups SL(n, R) and SL(n, C). The
set of n × n matrices with unit determinant under matrix multiplication also
forms a continuous group, since det(AB) = det(A) det(B). This is called the
special linear group SL(n, R) for matrices with real entries, and SL(n, C)
for matrices with complex entries.
Example 1.1.15 The Lorentz and Poincare groups. Lorentz trans-
formations in Minkowski space form a continuous group. If we also include
translations, we obtain the Poincare group.
1.2 Subgroups
Objectives
You should be able to:
• Determine whether a subset of a group is a subgroup.
• Construct the orthogonal and unitary groups as subgroups of the general
linear groups.
• Determine the centre of a group.
1.2.1 Definition
The notion of a subgroup is rather obvious. But these are important in physics,
for instance in the context of symmetry breaking. Suppose that a physical
system is symmetric under a certain group of transformation. Then one could
think of adding a perturbation to the system, for instance an extra term in the
Lagrangian or Hamiltonian, that lowers the symmetry group to a subgroup of
symmetries. This is captured by the concept of a subgroup.
The idea of subgroups is just like subspaces for vector spaces, which consist
of subspaces that are vector spaces in their own right.
Definition 1.2.1 Subgroups. A subset H of a group G is a subgroup if it
is a group in its own right. We write H ⊂ G. ♦
Since H is a group in its own right, it follows that any subgroup of G
contains the identity element e ∈ G. According to the definition, the group
G is a subgroup of itself. Also, for any G, the subset {e} containing only
the identity element is a subgroup, known as the trivial subgroup. Those two
subgroups are rather boring; we call them improper. We are interested in
studying proper subgroups.
1.2.2 Examples
Some examples of subgroups:
CHAPTER 1. BASIC CONCEPTS OF GROUP THEORY 4
group of special relativity! Indeed, one can check that elements of O(1, 3)
are Lorentz transformations. So one can think of Lorentz transformations as
"rotations in Minkowski space".
(More precisely, O(1, 3) is obtained by composing Lorentz transformations
with parity and time reversal transformations, just as O(n) is obtained by
composing reflections with rotations. Lorentz transformations correspond to
the component of O(1, 3) connected to the identity.)
In general, if we start with Rn with a scalar product of the form
the subset of linear transformations that preserve this inner product is denoted
by O(p, n − p), and is a subgroup of GL(n, R).
Example 1.2.14 The symplectic group Sp(2n, R). Another interesting
subgroup that plays an important role in physics is the symplectic
group. Let
0 I
x, y ∈ R2n , and J be the 2n × 2n symplectic matrix J = , where I
−I 0
is the n × n identity matrix. The subset of GL(2n, R) that leave the “anti-
symmetric bilinear form” xT Jy invariant is a subgroup of GL(2n, R), which is
called the sympletric group and is denoted by Sp(2n, R). It is for instance
fundamental in the more geometrical treatment of Hamiltonian mechanics.
Many of those groups will play a fundamental in physics. In particular, we
will see that SO(3) and SU (2) are intimately related, and so are SO(1, 3) and
SL(2, C).
Remark 1.2.18 The definitions that we are making look very similar to the
analogous definitions for vector places, but with 0 replaced by the identity
element e. Indeed, the definitions that we are proposing reduce to the previous
propositions for vector spaces when the group operation is considered to be
vector space addition. In this case, the identity element for addition in a
vector space is the zero vector, so e would be replaced by 0, as needed.
We can also define the “commutator” of two elements of a group G. But we
need to keep in mind that we only have one operation to work with, namely
group multiplication. The commutator should somehow measure how non-
abelian a group G is.
Definition 1.2.19 The commutator subgroup of a group. Let a, b ∈ G.
The commutator [a, b] of a and b is defined by
for all g, g 0 ∈ G and h, h0 ∈ H. The resulting group is the direct product of the
subgroups (G, e) and (e, H), regarded as subgroups of G × H.
Checkpoint 1.3.3 Prove this proposition.
Example 1.3.4 The Klein four-group, aka Z2 × Z2 . An example of a
direct product is the group Z2 × Z2 (also called the “Klein four-group”). Using
the construction above, one can think of this group as consisting of the four
elements (1, 1), (1, −1), (−1, 1), (−1, −1), with the operation being component-
wise multiplication. Note that this group is different from the cyclic group of
order 4 Z4 ; for instance, in Z4 = {1, −i, −1, i} only two elements (1 and −1)
square to the identity, while in Z2 × Z2 all four elements square to the identity.
This notion of direct product is what is needed to make sense, for instance,
of the gauge group of the Standard Model in particle physics, which is given
by SU (3) × SU (2) × U (1).
gk G = {gk g1 , . . . , gk gn }
higher order, but if you like this kind of thing feel free to do it yourself as an
exercise. Try for instance the symmetric group S3 , which has order 3! = 6.
Lots of fun!
1.5 Presentation
Objectives
You should be able to:
• Describe a group in terms of its presentation.
• Recall the general properties of the modular group.
If you have constructed the multiplication table for S3 as an exercise, you
have probably come to the realization that writing down multiplication tables
becomes quite annoying rather quickly for higher order groups. There has to be
a better way of encoding the abstract group structure of finite groups without
having to write down the multiplication table each time. One such way is to
write down the “presentation” of a group.
The idea is to specify a set of generators S of the group, which is a subset
of elements of the group from which all other elements can be obtained by group
multiplication, and the essential relations R that the generators satisfy. We
write the presentation as hS|Ri. Without getting into too much detail, let
me simply give a few examples of presentations.
Example 1.5.1 Z2 , Zn and Z2 × Z2 . As a basic example, consider the group
Z2 of two elements. Its presentation would be given as:
ha|a2 = ei.
ha|an = ei.
Example 1.5.2 The modular group. As a more interesting example, let
us look at the so-called modular group, which is important in various areas
theory and conformal field theory. Consider 2 × 2
of physics, suchas string
a b
matrices M = with integer entries and such that det M = 1. This
c d
group is denoted by SL(2, Z), the special linear group of 2 × 2 matrices with
integer entries. If we also impose the identification M = −M in SL(2, Z) we
end up with the group P SL(2, Z), the projective special linear group of 2 × 2
matrices with integer entries, which is also known as the modular group. This
group is important because it is the group of linear fractional transformations
of the upper half of the complex plane, where it acts as:
az + b
z 7→ , (1.5.1)
cz + d
for z ∈ C with =(z) > 0.
This group looks rather complicated, but in fact one can show that all such
CHAPTER 1. BASIC CONCEPTS OF GROUP THEORY 11
1.6 Cosets
Objectives
You should be able to:
• Determine the cosets of a subgroup of a finite group.
• State and prove Lagrange’s theorem.
1.6.1 Definition
Now that we understand subgroups, we can talk about cosets. In fact we have
already seen the idea of a coset when we stated the rearrangement theorem, but
now instead of taking the whole group G, we will consider subgroups H ⊂ G.
Definition 1.6.1 Cosets. Let H = {e, h1 , . . . , hr } be a subgroup of G, and
pick an element g ∈ G. Then the set:
Hg = {eg, h1 g, . . . , hr g}
Since the group is abelian, the left cosets are the same as the right cosets. In
general however left cosets and right cosets are not the same. We also notice
that all elements of G appear exactly once in the distinct cosets; as we will see,
this is a general result.
Cosets will turn out to be very useful in physics. Often we need to deal with
the fact that many different mathematical quantities give the same physical
observable. For instance, the electromagnetic potential is not uniquely fixed;
add to it the gradient of any function, and you get the same electric and
magnetic fields. Similarly, in quantum mechanics, any two states in a Hilbert
space that only differ by a phase give the same physical system. The way
to deal with this kind of redundancy mathematically is to define equivalence
classes of objects that give the same physics. It turns out that cosets are
perfectly suited for that; under suitable conditions, we can see the cosets as
being equivalence classes, hence from a physics point of view what we are often
really interested in are the cosets of a given group, rather than the group itself.
Thus what we would like to see is whether we can upgrade the cosets to
groups. It is clear from the definition that cosets of a subgroup H ⊂ G are
generally not subgroups of G themselves. However, what we can do is look at
the “set of cosets” all together; these in fact form a group, which will be called
the quotient group. This will be the group of equivalence classes formed by
the cosets, which is often the relevant group of objects that we are interested
in physically. But we are jumping ahead of ourselves; we will come back to
quotient groups shortly.
G must appear in at least one coset of H. But since two cosets are either
identical as sets or have no common elements, it follows that all elements of G
must appear in exactly one distinct coset.
Furthermore, one can prove that all cosets have the same number of ele-
ments, which is equal to the order of the subgroup |H|. Consider a coset gH for
some g ∈ G. Suppose that two elements of the coset are identical, ghi = ghj
for some hi 6= hj . Multiplying by g −1 on the left, we get hi = hj , which is a
contradiction. Therefore all elements of gH must be distinct, and hence the
number of elements in gH is the same as the number of elements in H.
Putting this together, we conclude that the distinct cosets are partitioning
the group G into non-intersecting bins of size equal to |H|. It follows that the
number of distinct cosets times |H| must be equal to |G|, that is, |H| divides
|G|.
Definition 1.6.5 The index of a subgroup. The number of distinct cosets
of a subgroup H ⊂ G is called the index of the subgroup. ♦
Lagrange’s theorem is in fact quite powerful to rule out when a subset
cannot be a subgroup of a finite group. For instance, Lagrange’s theorem
implies the following corollaries:
Corollary 1.6.6 More than half implies abelian. If more than half of the
elements of a finite group commute with each other, then the group is abelian.
Proof. We know that the centre of a group is a subgroup. Thus, by Lagrange’s
theorem, we know that if the order of the centre of the group is larger than
half of the order of the group, then the centre must be the whole group, i.e.
the group is abelian.
Corollary 1.6.7 Groups of prime order. All finite groups whose order
are prime numbers are cyclic, and have no proper subgroups.
Proof. By Lagrange’s theorem, if a group G has order given by a prime number,
then its only subgroups must have either order one (the trivial subgroup) or the
same order as the group itself (the group itself). Thus it cannot have proper
subgroups. Moreover, if you pick any element g ∈ G, then it generates a cyclic
subgroup hgi ⊆ G. But since G has no proper subgroups, then hgi is either
the trivial subgroup if g is the identity element, or it is the whole group G for
any other element g ∈ G. That is, G is a cyclic group, as it is equal to its
non-trivial cyclic subgroups.
1.7.1 Definition
We have already discussed briefly the idea of equivalence classes. Given a group
G, the idea is to define an equivalence relation, which we generally denote by
≡, between elements of the group, which is reflexive (a ≡ a), symmetric (if
a ≡ b then b ≡ a) and transitive (if a ≡ b and b ≡ c then a ≡ c). Once we have
an equivalence relation, we can define equivalence classes, which are subsets of
elements of the group that are equivalent according to ≡.
CHAPTER 1. BASIC CONCEPTS OF GROUP THEORY 14
1.8.2 Notation
Now let us study the symmetric group Sn in more detail. First we introduce
standard notation. An element of Sn corresponds to a permutation of n ele-
ments. It is customary to denote it by its action on the n elements {1, . . . , n}.
For instance, the cyclic rotation of the n elements, which takes 1 to 2, 2 to 3,
etc., would be denoted by
1 2 ··· n − 1 n
π= .
2 3 ··· n 1
1.8.3 Cycles
Since we are dealing with permutations of set of n elements, it is clear that
after applying a given permutation a certain number of times, we will come
back to the initial elements. In other words, consider for instance the element
1. A given permutation π will take 1 → π(1). Applying it again, we will get
1 → π(1) → π(π(1)). After a number of application, say r, we will necessarily
get back π r (1) = 1. The numbers 1, π(1), π 2 (1), . . . , π r−1 (1), that are reached
from 1 by π form what we call a “cycle”. More precisely:
Definition 1.8.3 Cycle. Let π ∈ Sn , i ∈ {1, . . . , n} and let r be the
smallest positive integer such that π r (i) = i. Then the set of r distinct elements
{π k (i)}r−1k=0 is called a cycle of π of length r, or an r-cycle generated by i.
We denote a given cycle by (i π(i) . . . π r−1 (i)).
It is clear that a given permutation π breaks up the set of n elements
{1, . . . , n} into disjoint cycles. We can then denote the permutation π by the
cycle decomposition of {1, . . . , n} that it implies. ♦
CHAPTER 1. BASIC CONCEPTS OF GROUP THEORY 17
Note however that the decomposition is not unique; in particular, since the
square of any transposition is the identity, we could insert the square of any
transposition in the product on the RHS without changing the end result.
Proof. The proof simply involves multiplying the transpositions on the RHS.
Start with i1 . Under the product of transpositions, we get i1 → i2 → . . . → i2 .
Consider i2 : we get i2 → i1 → i3 → . . . → i3 . Iterating, we get that ik → ik+1
for all k = 1, . . . , r − 1, and ir → i1 . Thus the resulting permutation is the
cyclic permutation of length r that has cycle structure (i1 i2 . . . ir ).
Definition 1.8.11 Parity. We define the parity of the cycle as being even
if the number of transpositions in the decomposition is even, and odd if the
number if transpositions is odd. Note that even though the decomposition is
not unique, the parity of the decomposition is uniquely defined. ♦
We can then apply these results to permutations themselves, not just to
individual cycles. We obtain:
Proposition 1.8.12 Decomposition of permutations into products of
transpositions. Any permutation can be decomposed as a product of transpo-
sitions (and length 1 cycles). The parity (even or odd, according to whether the
decomposition has an even or odd number of transpositions) of the permutation
is unique.
In fact, the parity of a permutation can be determined directly from its
cycle structure, without having to work out its decomposition into transposi-
tions, since each r-cycle can always be decomposed into the product of r − 1
transpositions.
CHAPTER 1. BASIC CONCEPTS OF GROUP THEORY 19
π = (15)(14)(23)(6).
since σ −1 takes the element σ(i) and sends it to i. Thus we can see the
permutation σ ◦π ◦σ −1 as applying σ to the symbols in the cycle decomposition
of π. The result will generally be different from π, but its cycle structure will
necessarily be the same.
The argument may become clearer with a simple example. Consider π =
(123)(4) ∈ S4 , and σ = (12)(34) ∈ S4 . The corresponding permutations are
1 2 3 4 1 2 3 4
π= , σ= .
2 3 1 4 2 1 4 3
CHAPTER 1. BASIC CONCEPTS OF GROUP THEORY 20
λ1 + λ2 + λ3 + . . . + λn = n.
Let us check that this formula is consistent with what we wrote above in S4 .
Consider for instance the cycle structure given by one cycle of length 3 and
one cycle of length 1. The number of permutations with that cycle structure
4!
is 3·1!·1·1! = 8, as written above. Similarly, the number of permutations with
4!
one cycle of length 2 and two cycles of length 1 is 2·1!·1 2 ·2! = 6, and so on and
so forth.
Remark 1.8.17 Note that partitions of an integer n can be encoded diagram-
matically into Young diagrams. A Young diagram is a finite collection of
boxes arranged in left-justified rows, with the rows “weakly decreasing” (i.e.,
such that each row has the same or shorter length than the row right above
it). It is then clear that Young diagrams with n boxes contain precisely the
same information as integer partitions of n, with the integers in the partition
given by the length of the rows of the Young diagram. Young diagrams are
useful for studying representations of the symmetric group, in which case they
are upgraded to Young tableaux.
S1 S2 = {s1 s2 | s1 ∈ S1 and s2 ∈ S2 }.
By definition, for normal subgroups left and right cosets agree, so we can
talk about cosets without specifying left or right. Given a normal subgroup
H ⊂ G, we can then define a group structure on the set of cosets. This is
called the quotient group:
Theorem 1.10.1 Quotient groups. If H is a normal subgroup of G, then
the collection of all distinct cosets of H, denoted by G/H, is a group (with
the operation being product of subsets defined in Definition 1.9.1), called the
quotient group of G by H. The order of the quotient group is the number
of distinct cosets, which is given by |G/H| = |G|/|H|, also called the index of
the subgroup H ⊂ G.
Proof. Consider the collection of all distinct cosets of H in G, which we denote
by G/H. We want to give it a group structure. We define group multiplication
as being the multiplication of sets defined in Definition 1.9.1. Let us first
show that the multiplication of two cosets yield another coset. Consider the
left cosets aH and bH for a, b ∈ G. Elements of (aH)(bH) have the form
ah1 bh2 = a(h1 b)h2 for h1 , h2 ∈ H. Since H is normal, we know that there
exists a h3 ∈ N such that h1 b = bh3 . Thus we can write ah1 bh2 = (ab)h3 h2 ,
which is an element of the coset (ab)H. Hence multiplication of sets defines a
CHAPTER 1. BASIC CONCEPTS OF GROUP THEORY 24
Remark 1.10.2 The reason for this construction to be called a quotient group
comes from division of integers. Suppose you consider 12/4 = 3. You can
understand this calculation of the quotient as taking 12 objects, and dividing
them into disjoint classes of 4 objects; the result is 3 disjoint classes. Here
we are doing the same thing, but instead of only talking about the number of
things in a set, we have a group structure, so we start with a group, partition
the set into subsets, and equip the resulting partition with the structure of a
group.
Example 1.10.3 Trivial examples. Note that the quotient group G/G is
clearly isomorphic to the trivial group (since it has only one element, which is
the identity); and G/hei is isomorphic to G, since the cosets of the subgroup
hei are just the elements of G themselves.
Example 1.10.4 The parity quotient group for permutations. Con-
sider the normal subgroup A3 ⊂ S3 . Using the notation of Definition 1.8.14,
there are two cosets, consisting of the even permutations E = {π1 , π5 , π6 } and
the odd permutations O = {π2 , π3 , π4 }, so the order of the quotient group is 2.
We know that there is only one abstract group of order 2, with multiplication
table given in Table 1.4.3, and using the group operation on cosets we can
check that indeed, E · E = O · O = E, E · O = O · E = O, since composing
two even permutations or two odd permutations gives and even permutation,
while composing and odd and an even permutation gives an odd permutation.
Example 1.10.5 Finite cyclic groups as quotients. A fundamental
example of quotient groups consists in the construction of finite cyclic groups
as quotients. Start with the integers Z under addition. Let H = mZ be the
subgroup of multiples of the positive integer m. Since Z is abelian, mZ is
necessarily normal. We then construct the quotient group Z/mZ. What do
elements of Z/mZ look like? Consider any integer k ∈ Z. Recalling that the
group operation here is addition, the coset kZ generated by the integer k is
k + mZ, that is k plus mutiples of m. Since any two k1 and k2 that differ
by a mutiple of m generate the same coset, we see that there are exactly m
disjoint cosets k + mZ, generated by k ∈ {0, 1, . . . , m − 1}. It follows that the
quotient Z/mZ is a finite group of order m. Product of cosets (in the sense of
Definition 1.9.1)
Zm as we used for the group of roots of unity; those are the same abstract
groups.
Example 1.10.6 SO(2) as a quotient group. Another interesting example
is constructed similarly. Start with the real numbers R, and consider the
normal subgroup consisting of integers Z. The cosets have the form a + Z for
a ∈ R. It is clear that two a1 and a2 that are related by the addition of an
integer generate the same cosets, hence the cosets are indexed by a ∈ [0, 1).
The quotient group R/Z is the group of these cosets, with operation given by
adding the cosets a1 + a2 + Z, with a1 + a2 being understood as addition such
that if the result is greater than one, than we substract one so that the result
is always between 0 and 1.
Let us now argue that R/Z is isomorphic to the group of complex numbers
of absolute value 1 under multiplication, also called U (1). Consider the map-
ping f : R/Z → U (1) given by f (a+Z) = e2πia . Under this mapping, the group
operation on cosets (addition of numbers) is mapped to multiplication of com-
plex numbers. We have not defined group homomorphisms and isomorphisms
yet (we will revisit this example in Example 1.11.5), but this mapping is an
isomorphism, as it is a bijection, and it preserves the group structure. Thus we
can think of R/Z and U (1) as being the same abstract group. Further, from
the point of view of complex numbers of absolute value 1, we can understand
2πa as an angle, and we see that U (1) is isomorphic to the group of rotations
in the complex plane (SO(2)). The group operation on U (1), multiplication
of exponentials, is then mapped to composition of rotations in SO(2). All in
all, we obtain that the quotient group R/Z is the same abstract group as the
group of two-dimensional rotations SO(2)!
determinant, that is SL(n, C). The first isomorphism theorem then implies
that SL(n, C) is a normal subgroup of GL(n, C).
We have seen another example of group homomorphism when we discussed
parity of permutations in Sn and the definition of the alternating group An .
In this case the first isomorphism theorem can be used to show directly that
An is always a normal subgroup of Sn .
Example 1.11.11 The parity homomorphism for Sn . Consider the
group H = {1, −1} under multiplication. Consider the mapping : Sn → H
that assigns 1 to even permutations and −1 to odd permutations. Let us first
show that is a group homomorphism.
For any permutations π1 , π2 ∈ Sn , we need to show that (π1 ◦ π2 ) =
(π1 )(π2 ). What this means is that we must show that the composition of two
even or odd permutations is even, while the composition of an even and an odd
permutation is odd. But this is clear by definition of parity. If r1 (resp. r2 ) is
the number of transpositions in the decomposition of π1 (resp. π2 ), then the
number of transpositions in the decomposition of π1 ◦ π2 is r1 + r2 , and thus if
r1 and r2 are both even or odd, then r1 + r2 is even, while if r1 is even and r2
odd, or vice-versa, then r1 + r2 is odd.
Then is a group homomorphism, and its kernel is the alternating group
An consisting of even permutations. It thus follows from the first isomorphism
theorem that An is a normal subgroup of Sn , and that the quotient group
Sn /An is isomorphic to H.
Representation theory
29
CHAPTER 2. REPRESENTATION THEORY 30
invariant, the fields themselves will not be. Thus they must transform in a
certain way. This is where representation theory comes in; they will transform
in some representation of the group of symmetry. This is why group theory
and representation theory is so fundamental in modern physics!
Let me give a slightly more concrete example of this to end this introduc-
tion. In quantum mechanics, we may consider a Hamiltonian that is invariant
under a group of transformations. A very important question then is to find
how the solutions of the Schrodinger equation corresponding to this particular
Hamiltonian behave under these symmetry transformations. This is answered
by representation theory; we can think of the group of symmetries as linear
transformations on the vector space of solutions, which is precisely the no-
tion of a representation of the group. This can then be used to classify the
eigenfunctions of the Hamiltonian according to how they transform under the
symmetry group. Very powerful! So let us study representation theory!
2.1 Representations
Objectives
You should be able to:
• Recall the definition of a representation.
2.1.1 Definition
Let us start by defining a group representation, and then study a bunch of
examples. We will focus here on representations on finite-dimensional complex
vector spaces (but we could easily upgrade vector spaces to Hilbert spaces, and
we could also consider vector spaces over other fields than C). After choosing
a basis, we can think of those as matrix representations.
Definition 2.1.1 Representation of a group. Let G be a group. A
representation of G is a group homomorphism T : G → GL(V ) (see Defini-
tion 1.11.3 for the definition of a group homomorphism), where V is a finite-
dimensional complex vector space of dimension n. We call n the dimension
of the representation.
After choosing a basis on V , we can think of the representation as being
given by n × n complex invertible matrices T (g) for each group element g ∈ G.
♦
In other words, a representation is a mapping that takes group elements into
n×n complex invertible matrices, and this mapping is such that it preserves the
group structure. Just as when we discussed homomorphisms vs isomorphisms,
we want to distinguish between representations that keep the whole group
structure intact, and those that lose some information:
Definition 2.1.2 Faithful representations. If T is one-to-one (that is,
if it is a group isomorphism on its image), then we say that it is a faithful
representation. We say that it is unfaithful otherwise. ♦
Now we can ask many questions about groups. Do every group have a repre-
sentation (aside from the trivial representation)? If so, how many does it have?
What are their dimensions? How do we characterize representations, and how
CHAPTER 2. REPRESENTATION THEORY 31
2.1.2 Examples
But before we study these questions, let us look at a few examples of repre-
sentations.
Example 2.1.3 Identity representation. We start with the most boring
example. For any group G, consider the mapping G → GL(V ), with V = C,
given by sending g 7→ 1 for all g ∈ G. This is certainly a group homomorphism,
but it is a rather boring one. It exists for any group G. It is called the
identity representation (or trivial representation). For any non-trivial G, it
is a dimension one, unfaithful, representation.
Example 2.1.4 The permutation representation of the symmetric
group Sn . Let us look at a more interesting example. Consider the sym-
metric group Sn , which can be understood as the group of permutations of
n objects. Let us focus on S3 forsimplicity.
Wecan think of the three ob-
1 0 0
jects as vectors v1 = 0, v2 = 1 and v3 = 0. Then a permutation
0 0 1
can be represented as a 3 × 3 matrix acting on the vector space spanned by
these vectors. That is, we can define a representation T : S3 → GL(V ) with
V the three-dimensional vector space spanned by the vectors {v1 , v2 , v3 }. For
instance, the cyclic permutation π = (123) can be represented by the matrix:
0 0 1
T (π) = 1 0 0 ,
0 1 0
You can check that these matrices satisfy the multiplication rules in Z3 .
CHAPTER 2. REPRESENTATION THEORY 32
Example 2.1.7 Two-dimensional representation of the quaternion
group. The quaternion group Q is defined by the set Q = {1, −1, i, −i, j, −j, k, −k}
with multiplication rules i2 = j 2 = k 2 = −1, ij = −ji = k, jk = −kj = i, and
ki = −ik = j. One can show that the following 2 × 2 complex-valued matrices
(i here is the imaginary number) form a faithful representation of Q:
1 0 −1 0 i 0
T (1) = , T (−1) = , T (i) = ,
0 1 0 −1 0 −i
−i 0 0 1 0 −1
T (−i) = , T (j) = , T (−j) = ,
0 i −1 0 1 0
0 i 0 −i
T (k) = , T (−k) = ,
i 0 −i 0
Example 2.1.8 One-dimensional representations of Zn . In Exam-
ple 1.1.9, we defined the cyclic group Zn as the set of n’th roots of unity
under multiplication. This is in fact a one-dimensional representation of the
abstract group Zn . Indeed, writing the group elements as ak = e2πik/n for
k ∈ {0, 1, . . . , n − 1} is in fact a faithful one-dimensional representation of
Zn . But there are other one-dimensional representations. One could write
ak = e2πik`/n for some fixed ` ∈ {0, 1, . . . , n − 1}; this gives in total n different
one-dimensional representations for Zn . We recover the original one for ` = 1.
Are these all faithful? Certainly not. Setting ` = 0 gives the trivial rep-
resentation ak = 1 for all k ∈ {0, 1, . . . , n − 1}, which is clearly not faithful.
In general, the representation will be faithful if ` and n are coprime. Can you
check that?
For instance, for Z3 , if we denote the third root of unity ω = e2πi/3 ,
we get three one-dimensional representations, namely {1, 1, 1}, {1, ω, ω 2 } and
{1, ω 2 , ω}. The last two are faithful.
For Z4 , if we denote the fourth root of unity α = e2πi/4 , we get four one-
dimensional representations, namely {1, 1, 1, 1}, {1, α, α2 , α3 } = {1, i, −1, −i}, {1, α2 , 1, α2 } =
{1, −1, 1, −1} and {1, α3 , α2 , α} = {1, −i, −1, i}. The first and third ones are
not faithful, while the second and fourth ones are.
Example 2.1.9 Representations of (R, +). Consider now the group of real
numbers R under addition. To build a representation for this group, we need
CHAPTER 2. REPRESENTATION THEORY 33
So, in this new basis with b0 = Bb and a0 = Ba, we get the equation
b0 = A0 a0 ,
♦
Representations that are constructed as direct sums are certainly not min-
imal. More precisely, with our definition of minimality above, we can see that
all direct sum representations are reducible. Indeed, consider a direct sum
representation T ⊕ S : G → GL(V ⊕ W ). Then the subspaces V ⊂ V ⊕ W and
W ⊂ V ⊕ W are certainly T -invariant, as all matrices T (g) are block diagonal,
and hence do not mix elements of V with elements of W . Thus the restrictions
of T ⊕ S to the subspaces V and W are subrepresentations, and hence T ⊕ S
is reducible.
So we have seen that all direct sum representations are reducible. But
this raises an interesting question: is the converse true? Can all reducible
representations be written as direct sums of irreducible representations? In
other words, can all reducible representations be written in block diagonal
form? We will see shortly that the answer is yes for finite groups, but no in
general. Thus it makes sense to distinguish between representations that can
or cannot be written as direct sums of irreducible representations.
Definition 2.2.6 Semisimple representations. A representation is semisim-
ple (also called completely reducible) if it is a direct sum of irreducible repre-
sentations. ♦
With this definition, the question becomes: are all representations of a
given group semisimple? (i.e. either irreducible or direct sums of irreducible
representations.) We will come back to this shortly.
Remark 2.2.7 A consequence of the direct sum construction is that for any
group, there exists an infinite number of reducible representations, which we
can construct as direct sums. This is why, from a classification viewpoint, it is
much more interesting to focus on irreducible representations.
However, in physics one often encounters representations that are reducible
(in fact semisimple). So it is also interesting to develop methods to find the
decomposition of a semisimple representation as a direct sum of irreducible
representations. For instance, in the case of a quantum mechanical Hamiltonian
H with a symmetry group G, we know that solutions will transform according
to some semisimple representation T of G. In other words, we can decompose
CHAPTER 2. REPRESENTATION THEORY 36
One can show that this representation is irreducible. Let us now consider the
order two subgroup H = {e, c} ⊂ S3 . The restriction of T to the subgroup
H ' Z2 gives the two-dimensional representation:
1 0 −1 0
T (e) = , T (c) = ,
0 1 0 1
Then the 4 × 4 matrices T ⊗ S(g) are constructed by taking the product of all
matrix entries (this is not a matrix product of course):
T11 (g)S(g) T12 (g)S(g)
T ⊗ S(g) =
T21 (g)S(g) T22 (g)S(g)
T11 (g)S11 (g) T11 (g)S12 (g) T12 (g)S11 (g) T12 (g)S12 (g)
T11 (g)S21 (g) T11 (g)S22 (g) T12 (g)S21 (g) T12 (g)S22 (g)
= T21 (g)S11 (g) T21 (g)S12 (g) T22 (g)S11 (g) T22 (g)S12 (g) .
T21 (g)S21 (g) T21 (g)S22 (g) T22 (g)S21 (g) T22 (g)S22 (g)
where of course one could expand out by writing down the matrices S(g)
explicitly as above.
We call the corresponding representation the tensor product represen-
tation. Such tensor product representations occur frequently in physics, for
instance when taking the product of wave-functions in quantum mechanics.
Checkpoint 2.2.9 Check that the tensor product representation constructed
above is indeed a representation, i.e. that T ⊗ S : G → GL(mn, C) is a group
homomorphism.
One important point is that even if T and S are irreducible representations
of G, the tensor product T ⊗ S may be reducible. In fact often in particle
physics one needs to determine how the tensor product of two irreducible rep-
resentations decomposes into a sum of irreducible representations.
you want to know more about this, but we will come back to this later on in
this course.
Mathematically, unitary representations are also generally easier to handle
than general representations: they satisfy nice properties not shared by general
representations. It turns out that for finite groups, all representations are
equivalent to unitary representations. This is not so simple for infinite groups
though.
This is why unitary operators are important in quantum mechanics: they can
be used to do changes of bases while preserving orthogonality between basis
vectors according to the inner product. They behave just like rotations but in
a complex vector space.
Another nice property of unitary matrices concerns their eigenvalues. While
eigenvalues of Hermitian matrices are real, eigenvalues of unitary matrices are
generally complex, but of a very particular form: they must lie on the unit
circle. In other words, they must have modulus 1. This means that they can
be written as eiθ for some angle θ ∈ [0, 2π).
It also turns out that unitary and Hermitian matrices are closely related.
Indeed, the theorem about eigenvalues of Hermitian matrices above relies on
the fundamental statement that any Hermitian matrix can be diagonalized by
CHAPTER 2. REPRESENTATION THEORY 39
=H. (2.3.1)
The last line follows because of the Rearrangement theorem 1.4.2, since mul-
tiplying the group elements g ∈ G by a fixed element g 0 only rearranges the
terms in the sum. This invariance property will be useful later on.
CHAPTER 2. REPRESENTATION THEORY 40
D = U † HU
X
= U † T † (g)T (g)U
g∈G
X
= A† (g)A(g),
g∈G
where we defined A(g) = T (g)U . Now consider the j’th diagonal entry Djj . It
is given by summing over g ∈ G the contributions given by A†j (g)Aj (g), where
Aj (g) denotes the j’th column vector in A(g). Since for each g ∈ G and each
j, Aj (g) is a non-zero vector, then A†j (g)Aj (g) > 0, and hence Djj > 0.
We then define the diagonal matrix D1/2 whose entries are the square roots
of the entries of D, and D−1/2 as the inverse of D1/2 . We now form the matrices
B(g) = D1/2 U † T (g)U D−1/2 , and their adjoints B † (g) = D−1/2 U † T † (g)U D1/2 .
Why are we constructing these matrices? Note that, since U is unitary, and
hence U † = U −1 , the transformation B(g) = D1/2 U † T (g)U D−1/2 is a similar-
ity transformation. Thus the representations furnished by the matrices B(g)
and the T (g) are equivalent. Our goal is to show that the new representation
B(g) is explicitly unitary, which would prove the theorem, namely that any
representation of a finite group is equivalent to a unitary representation.
So let us show that the matrices B(g) are unitary. We have:
where we used the invariance property (2.3.1). Thus for any finite-dimensional
representation T of a finite group G, we have constructed a new, equiva-
lent unitary representation B, given by the set of unitary matrices B(g) =
D1/2 U † T (g)U D−1/2 . We have thus proved that all finite-dimensional repre-
sentations of finite groups are equivalent to unitary representations.
Remark 2.3.9 Note here that the requirement of having a finite group G was
crucial in the proof. Otherwise, the expression
X
H= T † (g)T (g)
g∈G
doesn’t even make sense, since the sum would be over an infinite-dimensional
set (or a continuous space if the group is continuous).
Remark 2.3.10 In view of Remark 2.3.9, we may ask: is the unitarity theorem
still true for infinite groups, either discrete or continuous? Consider for example
CHAPTER 2. REPRESENTATION THEORY 41
the infinite continuous group (R, +), and the two-dimensional representation:
1 0
T (u) = , u∈R
u 1
which is not the identity for non-zero u ∈ R. So this is not a unitary transforma-
tion, and one can show that it cannot be brought into a unitary transformation
by a similarity transformation. In fact, the same representation restricted to
u ∈ Z is also not unitary for the infinite but discrete group (Z, +).
So for what kind of groups, beyond finite groups, is the unitarity theorem
true? The general result is that the theorem holds if the group is compact.
To define compact groups, we first need to define the concept of topological
groups. Topological groups are groups G that are given the extra structure of
a topology on G, such that the group’s binary operation and function mapping
elements to their inverses are continuous with respect to this topology. Then,
a compact group is a topological group such that its topology is compact.
In the end, they key point is that because of the existence of this compact
topology on G, for compact groups we can “replace” the sum over elements of
the group by an appropriate integral over the continuous group with respect to
some measure (the Haar measure), and the resulting integral then converges.
Using this the proof above goes through with minor modifications, and the
unitarity theorem holds for compact groups. Examples of compact groups
include SO(n) and SU (n), which appear frequently in physics. Note however
that the Lorentz group is not compact.
Note that not only the unitarity theorem holds for compact groups, but
Theorem 2.2.4 also applies to compact groups: all irreducible representations
of compact groups are finite-dimensional.
2.4 Semisimplicity
Objectives
You should be able to:
• Recall that every representation of a finite group is semisimple.
In Definition 2.2.3 and Definition 2.2.6 we introduced reducible and semisim-
ple representations. We then asked whether all representations of a given group
are semisimple. That is, whether all representations are either irreducible or
can be written as direct sums of irreducible representations. With our de-
tour study of unitary representations, we are now in a position to answer this
question for finite groups.
We said earlier that unitary representations are quite nice, and satisfy prop-
erties that are not shared by all representations. An example of this is the
following important theorem.
Theorem 2.4.1 Unitary representations are semisimple. All finite-
dimensional unitary representations are semisimple. That is, all finite-dimensional
CHAPTER 2. REPRESENTATION THEORY 42
P T (g)P = T (g)P
for all g ∈ G. This is because the projection operator P acts as the identity
operator on the subspace U ⊂ V , and P v ∈ U for all v ∈ V , so the condition
that T (g)u ∈ U for all u ∈ U will be satisfied if and only if P T (g)P v = T (g)P v
for all v ∈ V .
Take the Hermitian conjugate on both sides. We get the condition P T † (g)P =
P T † (g). We know that the matrices T (g) are unitary, and hence T † (g) =
T −1 (g) = T (g −1 ). Thus the condition becomes P T (g −1 )P = P T (g −1 ), but
since this must be true for all g ∈ G, we can write it simply as
P T (g)P = P T (g)
that is,
(1 − P )T (g)(1 − P ) = T (g)(1 − P ).
But the projection matrix 1 − P projects on the orthogonal complement W of
the subspace U ⊂ V . Since V = U ⊕ W , this means that T must decompose
in block diagonal form, i.e. it is a direct sum of representations.
We continue this process inductively. Since the representation is finite-
dimensional, the process must stop at some point, and we end up with T being
a direct sum of irreducible representations, that is, a semisimple representation.
Now we can use this key result to study whether representations of finite
groups can be written as direct sums of irreducible representations. Indeed, in
Theorem 2.3.8 we showed that all finite-dimensional representations of finite
groups are equivalent to unitary representations. Therefore Theorem 2.4.1
implies the following corollary:
CHAPTER 2. REPRESENTATION THEORY 43
AT (g)P = S(g)AP = 0.
But then, this implies that T (g)P ∈ W , or, in other words, T (g)w ∈ W for any
w ∈ W . This means that W is a T -invariant subspace. Since W is irreducible,
the only possible invariant subspaces are the trivial ones, namely either W = V
or W = {0}. In the first case, this implies that A = 0. In the second case, this
means that Av is never zero for all non-zero v ∈ V .
We can redo the same argument by starting with a subspace W 0 ∈ U such
that uA = 0 if and only if u ∈ W 0 . Following the same steps as above, we
end up with two possibilities; either A = 0 or uA is never zero for all non-zero
u ∈ U.
So we are left with two cases: either A = 0, or Av and uA are never zero
for non-zero v ∈ V and u ∈ U . We now show that the second case implies that
A is an invertible square matrix. First, if the number of rows is less than the
number of columns, than there must exist a v ∈ V such that Av = 0, which is
a contradiction. Similarly, if the number of columns is less than the number of
rows, than there must exist a u ∈ U such that uA = 0, which is a contradiction.
Therefore A must be square, that is, m = n. Finally, we know that a square
matrix has a non-trivial kernel if and only its determinant is zero. Therefore
A must have non-zero determinant, or equivalently, it must be invertible.
Therefore, we conclude that either A = 0 or A is an invertible square matrix.
In the latter case, if then follows that
AT (g)A−1 = S(g),
AT (g) = T (g)A ∀g ∈ G,
From Schur’s first lemma Lemma 2.5.1, it then follows that A − λI = 0, since
A − λI is not invertible. Therefore A = λI for some λ ∈ C.
For the matrices of a single irreducible unitary representation (or two equiv-
alent ones), the relation is:
X |G|
T † (g)ij T (g)kl = δil δjk ,
d
g∈G
Proof. This theorem follows from Schur’s lemmas. Let T and S be inequiv-
alent irreducible unitary representations of dimensions d and d0 respectively.
Consider an arbitrary d × d0 matrix X, and let
X X
A= T (g)XS(g −1 ) = T (g)XS † (g),
g∈G g∈G
where the second equality follows since S is unitary. Now multiply by T (h) on
the left for any h ∈ G:
X
T (h)A = T (h)T (g)XS(g −1 )
g∈G
X
= T (h)T (g)XS(g −1 )S(h−1 )S(h)
g∈G
X
= T (hg)XS † (g)S † (h)S(h)
g∈G
X
= T (hg)X(S(h)S(g))† S(h)
g∈G
X
= T (hg)XS((hg)−1 ) S(h).
g∈G
T (h)A = AT (h).
Schur’s second lemma Lemma 2.5.2 now implies that A = λX I for some com-
plex number λX ∈ C. We wrote λX to remind ourselves that the constant
depends on the choice of matrix X in the definition of A. Thus we have
X
T (g)XT (g −1 ) = λX I.
g∈G
CHAPTER 2. REPRESENTATION THEORY 47
Next, we need to fix the constants λX . We take the trace on both sides of
the equation. On the left-hand-side, we get:
X X
Tr T (g)XT (g −1 ) = Tr XT (g −1 )T (g)
g∈G g∈G
X
= Tr X
g∈G
=|G| Tr X.
λX Tr I = dλX ,
|G|
λX = Tr X,
d
and the relation becomes
X |G|
T (g)XT (g −1 ) = I Tr X.
d
g∈G
To conclude the proof, we choose the matrices X as above, with zero entries
everywhere except in position (j, k), wherePit is taken to be 1. Using index
notation, if we write the left-hand-side as g∈G T (g)ij T † (g)kl , we can write
I = δil , and Tr X = δjk , since the trace is non-zero only if the entry 1 in X is
on the diagonal. Therefore, we obtain the second orthogonality relation:
X |G|
T (g)ij T † (g)kl = δil δjk .
d
g∈G
The great orthogonality theorem gives orthogonality relations between the
matrices of the irreducible representations of any finite group G. Those are very
useful, as will become even clearer when we introduce characters. But for the
moment we can already deduce an immediate consequence of the orthogonality
theorem:
Theorem 2.6.2 The number of irreducible representations of a finite
group. A finite group can only have a finite number of inequivalent irreducible
representations.
Proof. This is clear from the orthogonality theorem. We can think of irre-
ducible representations as giving “vectors of matrices” (T (g)ij )g∈G in a vector
space of dimension |G|. The theorem tells us that those vectors must be or-
thogonal. But there are at most |G| orthogonal vectors in a vector space of
dimension |G|, and so the number of irreducible representations must be finite.
In fact we will calculate the number of irreducible representations for any finite
group explicitly when we introduce characters.
CHAPTER 2. REPRESENTATION THEORY 48
2.7 Characters
Objectives
You should be able to:
• Compute the character of a representation.
• Recognize that the character is a function of class.
2.7.1 Definition
Definition 2.7.1 The character of a representation. Let T : G → GL(V )
be a representation of G. The character of this representation is given by
the map χ : G → C, which assigns to every group element the trace of the
corresponding matrix in the representation T :
What this tells us is that if two representations have different characters,
then they must be inequivalent. We will see that the converse statement (that
two inequivalent representations must have different characters) follows from
the orthogonality theorem.
Another simple property of characters is how they behave for direct sums
and tensor products of representations. We will leave the proof of the following
lemma as an exercise:
Lemma 2.7.4 Character of a direct sum and a tensor product. Let
T : G → GL(V ) and S : G → GL(V ) be two representations of G. Then
χ(T ⊕S) (g) = χ(T ) (g) + χ(S) (g), χ(T ⊗S) (g) = χ(T ) (g)χ(S) (g).
In particular, the character of a semisimple representation T can be written
as
n
X
(T )
χ (g) = ai χ(Ti ) (g),
i=1
where the Ti are the irreducible representations appearing in the direct sum
with multiplicity ai .
where ni is the number of elements in the i’th conjugacy class, and we use
(T )
χi to denote the character of these elements. What this means is that we
can think of the characters as vectors in a c-dimensional complex vector space.
Note that we have reduced the dimension of the vector space from |G| to c,
the number of conjugacy classes in G.
What we will now see is that the simple characters (characters of irreducible
representations) are in fact orthogonal vectors. In particular, inequivalent
irreducible representations must have different (in fact orthogonal) characters.
Theorem 2.8.1 Orthogonality of characters. The simple characters of
a finite group G (the characters of its irreducible unitary representations) are
orthogonal:
X ∗ c ∗
(T ) (S)
X
χ(T ) (g)χ(S) (g) = ni χi χi = |G|δT S ,
g∈G i=1
where δT S is zero if T and S are inequivalent, and one if they are equivalent.
Here ni denotes the number of elements in the i’th conjugacy class.
Proof. This is a direct consequence of the great orthogonality theorem Theo-
rem 2.6.1. Using the notation δT S , we can state Theorem 2.6.1 as
X |G|
T † (g)ij S(g)kl = δil δjk δT S .
d
g∈G
(T ) ∗
(g)χ(S) (g). For the right-hand-side,
P
The left-hand-side is simply g∈G χ
we notice that
X X Xd
δik δik = δik = 1 = d.
i,k i,k i=1
Therefore X ∗
χ(T ) (g)χ(S) (g) = |G|δT S .
g∈G
We have shown that the characters are orthogonal vectors in the vector
space of class functions, which has dimensions c, the number of conjugacy
classes. It thus follows immediately that:
Corollary 2.8.2 Upper bound on the number of irreps. The number ρ
of inequivalent irreducible representations of a finite group is less or equal than
the number c of conjugacy classes.
Using this result, we can prove the converse of Theorem 2.5.3, which stated
that all irreducible representations of finite abelian groups are one-dimensional.
Theorem 2.8.7 Irreducible representations of abelian groups. A
finite group is abelian if and only if all its irreducible representations are one-
dimensional.
CHAPTER 2. REPRESENTATION THEORY 52
where mα is an integer that denotes the number of times that T (α) appears in
the decomposition. Taking the trace, we get a similar relation for the charac-
ters:
Xc
(T )
χ (g) = mα χ(α) (g).
α=1
Therefore, the character of a semisimple representation is a linear combination
of simple characters with non-negative coefficients. (Any class function can be
written as a linear combination of simple characters, but what is special here
is that the coefficients are non-negative.)
A semisimple representation will then be irreducible if and only if only
one of the coefficients mα is non-zero and equal to one. Our goal is to find
a simple criterion in terms of the characters of a representation to determine
when this is the case, without first having to calculate the explicit direct sum
decomposition of the representation.
Theorem 2.9.1 Criterion for reducibility. A representation T of a finite
group G is irreducible if and only if
c
(T )
X
ni |χi |2 = |G|,
i=1
are given by
c
1 X (T ) (α)
mα = ni χi (χi )∗ ,
|G| i=1
(α)
where ni is the number of elements in the i’th conjugacy class, and χi are
the characters of the irreducible representation T (α) .
Proof. We know that
c
(T ) (β)
X
χi = m β χi .
β=1
(α)
We multiply by ni (χi )∗ and sum over i. We get:
c c c
!
(T ) (α) (β) (α)
X X X
ni χi (χi )∗ = mβ ni χi (χi )∗
i=1 β=1 i=1
=|G|mα ,
The decomposition theorem is useful to determine the direct sum decom-
position of reducible representations. If one knows the characters of all irre-
ducible representations of a given group, then calculating the coefficients mα
CHAPTER 2. REPRESENTATION THEORY 55
2.10 An example: S3
Objectives
You should be able to:
• Apply the orthogonality theorems to calculate the character table of a
finite group.
The only possibility is that two of the representations are one-dimensional (we
know them already), and the third one is two-dimensional.
Even if we do not know the third two-dimensional irreducible representation
explicitly, we can still construct the character table of S3 using orthogonality
of characters. The character table should look like:
Table 2.10.1 Constructing the character table for S3
C1 C2 C3
(1)
T 1 1 1
T (2) 1 −1 1
T (3) 2 a b
CHAPTER 2. REPRESENTATION THEORY 56
We have already filled many entries, and left two unknowns (a and b),
which we will determine shortly. The first column is given by the trace of
the matrices corresponding to the identity element of the group, which are just
identity matrices. So these entries are just the dimension of the representations.
This is true in general: the first column is always given by the dimensions of
the representations.
The first row corresponds to the identity representation, which assigns 1 to
all group elements. So the characters are always 1.
The second row corresponds to the parity representation, which assigns 1
to even permutations and -1 to odd permutations. The permutations in the
conjugacy classes C1 and C3 are even, while those in C2 are odd, so we can fill
the second row appropriately.
There are two entries left, a and b. To find them, we use orthogonality of
characters. We will use orthogonality of columns, which is just the statement
of the second orthogonality relation Theorem 2.8.3. This relation says that
3
X (α) (α) |G|
(χi )∗ χj = δij .
α=1
ni
Applying this to the first and second conjugacy classes (the first and second
columns), that is, setting i = 1 and j = 2, the right-hand-side becomes zero,
and we get:
(1)(1) + (1)(−1) + (2)(a) = 0.
Therefore a = 0. Similarly, using the first and third columns (i = 1 and j = 3),
we get:
(1)(1) + (1)(1) + (2)(b) = 0,
that is b = −1. We have thus completed the character table of S3 :
Table 2.10.2 The character table for S3
C1 C2 C3
T (1) 1 1 1
T (2) 1 −1 1
T (3) 2 0 −1
For fun we can check that the other orthogonality relations are also satisfied.
For instance, sticking with the second orthogonality relation Theorem 2.8.3,
but applying it to the third column (i = j = 3), we get:
|G| 6
(1)(1) + (1)(1) + (−1)(−1) = = = 3,
2 2
which is indeed true.
The first orthogonality relation Theorem 2.8.1 correspond to orthogonality
between rows. Recall that it says that:
3
(α) (β)
X
ni (χi )∗ χi = |G|δαβ .
i=1
For instance, using it for the second and third irreducible representations (α =
2 and β = 3), we get:
1(1)(2) + 3(1)(0) + 2(1)(−1) = 0,
which is indeed correct. Applying to the third row (α = β = 3), we get:
1(2)(2) + 3(0)(0) + 2(−1)(−1) = |G| = 6,
CHAPTER 2. REPRESENTATION THEORY 57
Let us first show that this representation is reducible (we already know that
since S3 has only three irreducible representations of dimensions 1, 1 and 2.)
We use the reducibility criterion Theorem 2.9.1. We calculate:
3
(T )
X
ni |χi |2 = 1(3)2 + 3(1)2 + 2(0)2 = 12,
i=1
Let us calculate these coefficients using the character table Table 2.10.2. We
get:
1
m1 = (1(3)(1) + 3(1)(1) + 2(0)(1)) = 1,
6
1
m2 = (1(3)(1) + 3(1)(−1) + 2(0)(1)) = 0,
6
1
m3 = (1(3)(2) + 3(1)(0) + 2(0)(−1)) = 1.
6
CHAPTER 2. REPRESENTATION THEORY 58
T = T (1) ⊕ T (3) .
In the case of the regular representation of a group of order n, since the only
(S)
non-vanishing character is χ1 = n, for any α we calculate:
1 (α)
(α)
mα = n(χ1 )∗ = (χ1 )∗ .
n
But for any irreducible representation, the character of the identity element is
the trace of the identity matrix, which is just the dimension of the representa-
tion:
(α)
χ1 = dα .
Therefore mα = dα , for all irreducible representations!
Therefore for any finite group the regular representation decomposes as the
direct sum of irreducible representations:
c
M
(S)
T = dα T (α) ,
α=1
Objectives
You should be able to:
• Determine whether a given representation is real, pseudoreal or complex.
In this course we focus on group representations on complex vector spaces.
However, sometimes such a representation may be real, just like complex num-
bers may be real. So a natural question is: given a representation on a complex
vector space, can we determine easily when a representation is real or not?
First, how do we know whether a complex number z is actually real? Well,
it is easy, you look at it, and see whether its imaginary part is zero! More
precisely, the statement is that z is real if and only if z = z ∗ , that is, z is equal
to is complex conjugate, which means that its imaginary part vanishes.
We can do the same thing with matrices. We say that a matrix M is real
if and only if M ∗ . Note that this is really the complex conjugate, not the
complex conjugate transpose. So this means that all the matrix entries are
real numbers.
Now what about representations? When is a representation real? This is of
great interest in physics; for instance, once one understands particles in terms
of representations, then anti-particles transform in the complex conjugate rep-
resentations. So understanding the relation between a representation and is
complex conjugate is crucial.
T ∗ (g) = ST (g)S −1 .
♦
(T )
Remark 2.12.3 In terms of characters, it follows that if the characters χi
are complex, i.e. not equal to their complex conjugate, then T is complex.
The converse is not true however; if the characters are real, then we cannot
CHAPTER 2. REPRESENTATION THEORY 61
T ∗ (g) = ST (g)S −1 ,
T † (g) = (S −1 )T T (g)T S T .
T (g) = (S −1 )T T (g −1 )T S T .
What this means is that S −1 S T commutes with all matrices T (g) of an irre-
ducible representation. By Schur’s lemma Lemma 2.5.2, we know that S −1 S T =
λI for some λ ∈ C. Thus S T = λS. This means that
S = (S T )T = (λS)T = λS T = λ2 S,
T ∗ (g) = ST (g)S −1 ,
W −1 T ∗ (g)W = W T (g)W −1 .
It thus follows that W T (g)W −1 has only real entries, thus T is equivalent to
a representation with only real entries.
note here that if g and h are in the same conjugacy class, than g 2 and h2 also
are; thus the Frobenius-Schur indicator is also a class function, and could be
written as a sum over conjugacy classes.
Remark 2.12.9 The classification between real, pseudoreal and complex irre-
ducible representations also hold for compact groups. Moreover, those can be
distinguished using the Frobenius-Schur indicator as above, but with the sum
replaced by an integral over the compact group.
Example 2.12.10 Irreducible representations of S3 . As an example, let
us look at the irreducible representations of S3 , as studied in Section 2.10. To
evaluate the Frobenius-Schur indicator we will need the square of the group
elements. We note that the square of the identity is still the identity. As for the
transpositions (conjugacy class C2 ), their square is the identity. For the cyclic
permutations (conjugacy class C3 ), we calculate. Take for instance (123). The
square is (123)(123) = (132), and hence it is still a cyclic permutation of length
three. Thus the square of elements in the conjugacy class C3 are still in C3 .
Let us now determine whether the irreducible representations are real, pseu-
doreal or complex. First, the two one-dimensional representations are mani-
festly real, since they map all group elements to real numbers. Let us, for fun,
evaluate the Frobenius-Schur indicator for these two representations.
For the trivial representation, we get:
1 X (1) 2 1
χ (g ) = ((1)(1) + (3)(1) + (2)(1)) = 1,
|S3 | 6
g∈S3
and hence the representation is real, as expected (this is obviously always the
case for the trivial representation).
For the non-trivial one-dimensional representation T (2) (using the notation
in Section 2.10). The sum is:
1 X (2) 2 1
χ (g ) = ((1)(1) + (3)(1) + (2)(1)) = 1,
|S3 | 6
g∈S3
and thus this representation is also real! This is interesting, because depending
on how you think about these 2 × 2 matrices, they may be written as matrices
with complex entries. But in fact in Example 2.1.6 we had already seen one
way to write these matrices with real entries, and thus the representation must
be real indeed.
Example 2.12.11 The two-dimensional representation of the quater-
nion group. Let us now look at the two-dimensional representation of the
quaternion group given in Example 2.1.7. First, we should check that it is ir-
reducible. Using the criterion for irreducibility Theorem 2.9.1, and calculating
the traces of the matrices, we get:
Applications
In this section we study a few simple physical applications of group and rep-
resentation theory, focusing on finite groups.
3.1 Crystallography
Objectives
You should be able to:
• Define what point and space groups are.
• Prove the crystallographic restriction theorem.
Group theory is really all about symmetries. In this section we will see
how group theory can be applied to study objects with symmetries, such as
molecules and crystals. We will see that symmetries are highly constraining,
and that studying symmetries of physical objects such as crystals gives rise to
highly non-trivial, and perhaps unexpected, results.
One should note that there is extensive literature on this subject, in fact
there are whole books dedicated to the study of symmetries of crystals. Physi-
cists and chemists have invented their own notation to denote the symmetry
groups, and reading through this literature can quickly become confusing and
even overwhelming. So what we will do here is just quickly go through some of
the basic results that highlight how group theory is useful to understand the
physics and chemistry of crystals.
65
CHAPTER 3. APPLICATIONS 66
of a lattice of points that are invariant under these translations. For particular
such lattices, there are sometimes extra symmetries (such as rotations, reflec-
tions, etc.), which, together with translations, form the structure of the space
group.
Just as for point groups, space groups can be classified.
• In one dimension, there are two space groups (also known as “line groups”):
the group of translations, which is isomorphic to the group of integers Z,
and the infinite dihedral group, which includes reflections.
• In two dimensions, there are 17 space groups, also known as wallpaper
groups (have a look at this page, it’s super fun!). Those are obtained by
combining the 5 types of two-dimensional lattices (known as “Bravais lat-
tices”) specifying the translational symmetries with the two-dimensional
point groups that are consistent with translational symmetries.
• In three dimensions, there is a total of 230 different space groups. Those
are obtained by combining the 14 types of three-dimensional Bravais
lattices with the 32 point groups that are consistent with translational
symmetries.
3.1.3 Crystals
We are now ready to discuss symmetry groups of crystals, and the famous
crystallographic restriction theorem, which shows how powerful the study
of symmetries can be in physics and chemistry.
To start with: what is a crystal? In physics and chemistry, a crystal is a
solid material which has a highly ordered microscopic structure: it consists in
a lattice of atoms that extends in all directions and is invariant under transla-
tions. Examples include snowflakes, diamonds, table salt, etc.
Let us focus on crystals in three dimensions. Mathematically, the lattice
of atoms in a crystal is a Bravais lattice: this is an infinite array of discrete
points that are generated by a translation:
where n1 , n2 , n3 ∈ Z and v~1 , v~2 , v~3 are primitive vectors in three dimensions
which lie in different directions (not necessarily perpendicular) and span the
lattice. Thus a crystal is invariant under translations generated by T~ .
Furthermore, a given crystal may be invariant under symmetry operations
that leave a point fixed, such as rotations and reflections. The group of such
symmetry operations is a point group; in this context, it is known as a crys-
tallographic point group. Together with translations, it forms the space group
of the crystal.
We have seen that finite point groups in three dimensions can be classified:
there are 7 infinite familes, and 7 additional point groups. But can all of those
be point groups of crystals? In other words, can all of these point groups arise
as symmetry groups of lattices? The answer to this question is the essence
of the crystallographic restriction theorem, and may appear surprising: only
a very small, finite, number of point groups are symmetry groups of crystals!
For instance, there is no crystal that is invariant under rotations by an angle
2π/5! Isn’t that surprising?
Theorem 3.1.1 Crystallographic restriction theorem. Consider a crys-
tal in three dimensions that is invariant under rotations by an angle 2π/n
around an axis. Then the only allowed possibilities are n = 1, 2, 3, 4, 6, with
the case n = 1 being the trivial case with no rotational symmetry.
CHAPTER 3. APPLICATIONS 68
Note that the same result holds for a crystal in two dimensions invariant
under rotations.
What? Really? Those are the only possibilities? Imposing translational
symmetry is that constraining? Let’s prove this! There are a number of differ-
ent proofs of this theorem. We will first present a geometric proof, and then
present a representation theoretic, or matrix based, proof, for fun.
Geometric proof. First, we forget about the case n = 1 since it is trivial, as it
involves no rotational symmetry.
Now suppose that a lattice is invariant under rotations by an angle 2π/n
about an axis. Consider two lattice points A and B in the plane perpendicular
to the axis of rotation, and separated by a translation vector ~v of minimal
length (by which I mean that there is no lattice point on the line between A
and B). Let us rotate point A about point B by an angle 2π/n and call the
resulting point A0 . Similarly, let us rotate point B about A by the same angle
2π/n but in the opposite direction, and call the resulting point B 0 .
Let us call v~0 the vector joining A0 and B 0 . We notice that v~0 is parallel to
~v . Thus, if rotation by an angle 2π/n is a symmetry of the lattice, it follows
that the new translation vector v~0 must be an integer multiple of ~v , that is,
Since we must have v~0 = k~v for some k ∈ Z, and hence |v~0 | = |k||vecv|, we get
the constraint:
2π
|k| = 1 − 2 cos ,
n
that is, 1 − 2 cos 2π
n must be a non-negative integer. Since | cos α| ≥ 1 for any
angle α, the only possible values of |k| are |k| = 0, 1, 2, 3. Solving for n for each
cases, we get, respectively, n = 6, 4, 3, 2. Those are the only possible rotational
symmetries of a crystal, which concludes the proof.
Representation theoretic proof. Let us now give a representation theoretic, or
matrix based, proof. A rotation by an angle 2π/n about an axis can be seen
as a linear operator acting on the two-dimensional vector space perpendicular
to the axis of rotation. After choosing a basis for this vector space, such a
rotation can be written in terms of a 2 × 2 rotation matrix. In the language of
group theory, this gives a two-dimensional representation of the cyclic rotation
group Zn . On general grounds, we know how to write such a rotation matrix
with respect to an orthonormal basis for R2 :
cos 2π sin 2π
n n .
− sin 2π
n cos 2π
n
This matrix expression for the linear operator is only valid with respect to an
orthonormal basis, but its trace (that is, the character of the representation),
which is equal to 2 cos(2π/n), is invariant under changes of basis, as we know
(that is, the characters are the same for equivalent representations related by
similarity transformations). So in a different choice of basis, the 2 × 2 rotation
matrix may look different, but its trace will always be equal to 2 cos(2π/n).
CHAPTER 3. APPLICATIONS 69
In this section we give a very brief introduction to the power of group and
representation theory in quantum mechanics.
H 0 = R−1 HR.
CHAPTER 3. APPLICATIONS 71
H = R−1 HR,
or, equivalently,
RH = HR.
We can look at all such operators R that commute with H: those form a group
G, which we call the symmetry group of the Hamiltonian.
More precisely, we could start with the symmetry group G, realized in terms
of properties of the Hamiltonian (for instance some explicit symmetry of the
system, such as periodicity). Then the operators R would form a representation
of the group G, acting on the space of states of the system.
So what this means is that we can group the eigenfunctions of the systems
into subspaces that are invariant under the action of the symmetry group; all
eigenfunctions in each of these subspaces share the same eigenvalue.
This statement follows neatly from representation theory. If R is an irre-
ducible presentation, then, since HR = RH, by Schur’s lemma we must have
H = EI for some number E. What this means is that all states have the
same energy eivenvalue. However, in general R is not irreducible. But one can
L wuch that R decomposes as direct sum
choose a basis for the space of states
of irreducible representations R = α R(α) , and by Schur’s lemma, it follows
that with respect to this choice of basis
L the Hamiltonian also decomposes as a
direct sum of diagonal matrices H = α E (α) I, where I is the identity matrix
of the size of the corresponding irreducible representation R(α) .
To summarize, what this means is that we can choose a basis for the space of
states, composed of eigenfunctions of the Hamiltonian, in which the state space
decomposes as a direct sum of subspaces that transform according to irreducible
representations of the symmetry group. All states in each such subspace are
eigenfunctions of the Hamiltonian with the same energy eigenvalue.
This leads to the very important physical concept of degeneracy.
Definition 3.2.2 Degeneracy of states. We say that two eigenfunctions
of the Hamiltonian are degenerate if they share the same energy eigenvalue.
♦
At first, in the history of physics, when eigenstates that shared the same
energy eigenvalues were discovered, it was very puzzling to physicists. What
should states have the same energy eigenvalue? When you diagonalize a generic
matrix, you expect distinct eigenvalues.
But from the viewpoint of the representation theory, degeneracy is ex-
pected, and in fact can be constrained using representation theory. Indeed,
we know that the “degree of degeneracy” of an eigenfunction (that is, the
number of degenerate states) is at least as large as the dimension of the irre-
ducible representation that it belongs to. It can be larger if two eigenspaces
corresponding to different irreducible representations have the same energy
eigenvalue; we call such degeneracy “accidental”, since it is not determined by
symmetry. However, most degeneracies are determined by symmetry.
CHAPTER 3. APPLICATIONS 72
3.2.4 Examples
We will look at two simple examples of the construction. But most relevant
examples to quantum mechanics involve continuous groups and their represen-
tations, thus for this we will need to wait until later.
Example 3.2.3 Parity. Let us start with a very simple one-dimensional
example. The Hamiltonian of a one-dimensional system takes the form H =
d2
− 12 dx 2 + V (x) for some potential function V (x). Now suppose that V (−x) =
that is, they are even. States that transform according to the parity represen-
tation transform as:
ψ(−x) = S(a)ψ(x) = −ψ(x),
that is, they are odd. The upshot of the representation theory is that we can
label the states according to whether they are even or odd.
We also know that there is no degeneracy here (barring accidental degen-
eracy), since all irreducible representations of Z2 are one-dimensional.
This example is of course almost trivial: the goal was just to get familiar
with how to use the language of representation theory to understand states of
a quantum mechanical system.
CHAPTER 3. APPLICATIONS 73
for some real number k. Since eika = eika+i2π , the representations are indexed
by a real number k such that
π π
− ≤k< .
a a
So the states of the system can be labeled by such a real number. This is an
example of a Brillouin zone.
We can look at the special case of a one-dimensional crystal (note that it
can be easily generalized to more than one dimension) with a finite number of
sites, say N . We can impose a periodic boundary condition, meaning that if
we translate from the N th site, we go back to the first one. Thus the group of
symmetry is generated by the translation T , but with the condition T n = e.
This group is the cyclic group ZN .
The abelian group ZN has N one-dimensional irreducible representations
Sk . Those are given by representing the generator T by the roots of unity
Sk (T ) = wk , for k = 0, 1, . . . , N − 1, where w = e2πi/N . So we can label
the states according to how they transform. For instance, a state ψk that
transforms in the kth irreducible representation would transform as
2πk
A= .
Na
Thus φk (x) = 2πk
N a x + B. Subsituting back in (3.2.3), and redefining uk (x) to
include in it the arbitrary constant B, we get:
2πikx
ψk (x) = e Na uk (x).
CHAPTER 3. APPLICATIONS 74
for some N D × N D matrix of coefficients H. The key point here is that the
equation is still linear. But given a particular spring-mass system, the precise
form of the matrix H may be mess, and in fact not so easy to write down. The
goal of this section will be to extract a lot of information about the motion of
the system without ever having to write down H explicitly. As you could guess,
what we will use is simply the symmetries of the system, and representation
theory.
for some positive constant ω and a vector of coefficients ~a (for the case ω = 0,
we would take the ansatz ~η (t) = ~at + ~b). Plugging back in (3.3.1), we obtain
the equation
H~a = ω 2~a,
which is nothing but an eigenvalue problem for the N D × N D matrix of coef-
ficients H!
In other words, what we have found is that the normal modes are given
by eigenvectors of H, with oscillating frequency given by the square root of
their eigenvalues. Because the equations of motion are linear, we can conclude
that the general motion of the system will be given by a linear combination
(superposition) of these N D normal modes.
A few things to note: for the motion to be oscillatory, all eigenvalues of H
must be non-negative (otherwise we would have an unstable equilibrium point).
For a normal mode with positive ω, all masses are oscillating with the same
frequency ω. For a normal mode with a zero eigenvalue, the corresponding
motion is uniform translation of all masses.
The upshot of this discussion is that to understand the motion of the sys-
tem, we need to find its normal modes, and to find its normal modes, we need to
find the eigenvectors of the N D × N D matrix H. This is where representation
theory comes into play!
3.3.3 Examples
In this section we consider two examples. The first one is very simple, and is
studied simply to set the stage; the second one is more involved.
Example 3.3.2 Two masses connected by a spring in one dimension.
We start with the simple example of two masses connected by a spring and
restricted to move in one dimension. We want to understand the normal modes
of the system. In this case we can do everything explicitly, so let us start by
doing that. But the goal of this example is to show how representation theory
can be used to gain information on the system without solving explicitly.
Let η1 (t) and η2 (t) be the displacements of the two masses. The equations
of motion are (setting the constants m1 = m2 = k = 1 for simplicity):
d2 d2
η1 = −(η1 − η2 ), η2 = −(η2 − η1 ).
dt2 dt2
In matrix form, those reads:
d2
1 −1
~
η = − ~η .
dt2 −1 1
We could of course solve this equation explicitly here. But let us use represen-
tation theory instead. To get the normal modes we want to find eigenvectors
for the matrix
1 −1
H= . (3.3.2)
−1 1
We want to do so using symmetries of the system.
The symmetry group G here is rather simple. It has two elements: the
identity, and the exchange of the two masses. Thus G ∼ = Z2 . It acts on the
two-dimensional space of states in a two-dimensional representation T . We
can write down the two-dimensional representation explicitly (we will do that
below for completeness), but to find its decomposition in terms of irreducible
representations we only need its characters. The identity element has char-
acter χ(e) = 2, while the non-trivial element does not leave any mass fixed,
hence it must have character χ(a) = 0. For completeness, the two-dimensional
CHAPTER 3. APPLICATIONS 77
So there is one normal mode with zero frequency, and one normal mode with a
non-zero frequency. It is easy to identify these normal modes physically. The
first mode corresponds to linear translation of the whole system (the spring is
not stretched at all). The second mode corresponds to the masses oscillating
with the same frequency in opposite directions (the “breathing mode”).
Of course this example was rather trivial. We could have solved the equa-
tions of motion explicitly. But in more complicated systems, H may be difficult
to write down, and studying symmetries provides valuable information on the
normal modes of the system without having to solve any equation of motion.
Example 3.3.3 The triangular molecule. We now consider the more
interesting example of 3 masses in two dimensions connected by three springs in
the form of an equilateral triangle. Each mass has two displacement coordinates
(horizontal and vertical directions), and hence the displacement vector is 3×2 =
6-dimensional.
Writing down the matrix H for this system would be pretty annoying. So
let us not do that. We will use symmetry instead to gain information on the
normal modes of the system.
The symmetry group here is the dihedral group D3 which is isomorphic
to the symmetric group S3 . It acts on the space of states in the form of a
6-dimensional representation T . We will not write down the representation
explicitly, but simply deduce its characters on the three conjugacy classes of
S3 . We will use the notation in Section 2.10.
In the conjugacy class C1 containing the identity element, the character of
CHAPTER 3. APPLICATIONS 78
Hamilton’s principle then states that the motion of the system from time
t1 to t2 is such that the action has a stationary value (is extremized). This is
a very profound statement. Out of all possible paths in configuration space,
Nature chooses the path which extremizes the action. The physics is entirely
encoded in its Lagrangian; the equations of motion can be obtained by ex-
tremizing the action. Isn’t that beautiful? All of classical mechanics can be
summarized in this neat variational principle.
Using variational calculus, one finds that the action S has a stationary
value if and only if the Euler-Lagrange equations
d ∂L ∂L
− =0
dt ∂ q̇i ∂qi
are satisfied, for i = 1, . . . , s, where the qi are coordinates for the system, and
q̇i denotes time derivative of the coordinates.
Thus the physics of a system in classical mechanics is encoded in a single
object: its Lagrangian. This makes understanding symmetries of the theory
much easier. In terms of the equations of motion, one would say that a system
has a symmetry if both sides of the equations of motion transform in the same
way, so that the equations are invariant under the symmetry operation. But
here the statement is much simpler. A system if symmetric if its action is
invariant under the symmetry operation. Or, in the language of representation
theory, the action must transform in the trivial representation of the symmetry
group. In fact, in most cases, invariance of the action comes from invariance
of the Lagrangian itself.
In fact, in modern physics one often starts with the desired symmetries
of a given physical system, and use this to determine the form of the La-
grangian, by writing down the most general action which is invariant under
these symmetries. Let me give a very simple example. Consider a free particle
in three-dimensional space. Homogeneity of space and time implies that L can
only depend on |~v |2 : it cannot depend explicitly on ~x or t, nor can it depend
on the direction of ~v . Then, one can show that L must be proportional to |~v |2
by requiring that it is invariant under Galilean transformations of the frame
of reference. This constant of proportionality defines the mass of the particle,
which can be shown to be positive, since otherwise the action would have no
extremum. Thus, simply by requiring invariance under given symmetries, we
are able to recover the Lagrangian of the system! The same can be done for a
system of N particles, see “Mechanics” by Landau and Lifschitz, Sections 1.3
and 1.4.
This approach is what is done for instance in particle physics. Hamilton’s
principle (and its quantum version) is postulated. We then study the symme-
tries of the system, and figure out the most general Lagrangian that is com-
patible with these symmetries. This Lagrangian will involve a certain number
of constants; we then use experiments to fix the value of these constants.
CHAPTER 3. APPLICATIONS 80
where S is the action of the system. This infinite sum over all paths is what is
called in quantum physics a path integral. When ~ → 0, the classical limit
is recovered, since the path integral “localizes” on the classical configuration
given by the extremum of S. This gives a beautiful conceptual explanation of
the difference between quantum and classical physics.
We have just seen that quantum physics can be understood neatly in terms
of the action of a system. In fact, all of the fundamental laws of physics can
be written in terms of an action principle. This includes electromagnetism,
general relativity, the Standard Model of particle physics, and string theory.
For instance, almost everything we know about Nature can be captured in the
Lagrangian
√ 1
L = g(R + Fµν F µν + ψ̄6Dψ),
2
where the first term is the Einstein term, the second the Yang-Mills (or Maxwell)
term, and the last the Dirac term. Those describe gravity, the forces of Nature
(like electromagnetism and the nuclear forces), and the dynamics of particles
like electrons and quarks. You are welcome to try to understand what these
terms really mean! :-)
constant during motion. These conserved quantities can often be found rela-
tively easily and then used to reduce the differential system to a simpler one
which is easier to solve. In fact, in many cases these conserved quantities are
even more interesting than a full solution to the equations of motion. It is
therefore of interest to learn how to find conserved quantities.
The key result here is that we can associate conserved quantities to sym-
metries of the system! This is the fundamental statement of Noether’s the-
orem: For any continuous symmetry of the action of a system, there is a
corresponding conserved quantity. And this is not just an abstract statement;
we can construct this conserved quantity explicitly. For instance, invariance
under time translations gives rise to conservation of energy. Invariance under
space translations gives rise to conservation of momentum. Invariance under
space rotations gives rise to conservation of angular momentum. Once again,
group theory is fundamental!
Chapter 4
82
CHAPTER 4. LIE GROUPS AND LIE ALGEBRAS 83
with a ∈ R>0 and b ∈ R. This set forms a group under matrix multiplication.
Its underlying manifold is given by R>0 ×R, which is a smooth manifold. Using
matrix multiplication, we see that
a1 b1 a2 b2 a1 a2 a1 b2 + b1
T1 T2 = = ,
0 1 0 1 0 1
its inverse is
− ab
1
T = a .
0 1
CHAPTER 4. LIE GROUPS AND LIE ALGEBRAS 84
is 2(n2 − 1).
Note that SL(n, R) and SL(n, C) are non-compact.
Example 4.1.9 O(n) and SO(n). The group O(n) of real orthogonal ma-
trices is a closed subgroup of GL(n, R), and hence a Lie group. The number of
independent real parameters in an n × n orthogonal matrix can be calculated
to be 21 n(n − 1), which is the dimension of the Lie group.
The group SO(n) of special orthogonal matrices (with determinant one)
is also a closed subgroup of GL(n, R), and hence a Lie group. In this case,
imposing that the determinant of an orthogonal matrix A is equal to one is a
discrete condition, as any orthogonal matrix must satisfy det A = ±1 to start
with. Thus it does not reduce the number of degrees of freedom, and hence
the dimension of SO(n) is also 12 n(n − 1).
The orthogonal groups O(n) and special orthogonal groups SO(n) are com-
pact Lie groups.
Example 4.1.10 U (n) and SU (n). The group U (n) of unitary matrices is
a closed subgroup of GL(n, C), and hence a Lie group. The number of real
independent parameters in an n × n unitary matrix can be calculated to be n2 ,
which is the dimension of U (n).
The special unitary group SU (n) is also a closed subgroup of GL(n, C), and
hence a Lie group. Imposing the condition that the determinant is equal to
one reduces the number of real degrees of freedom by one. Indeed, any unitary
matrix A has | det A|2 = 1, which means that its determinant can be written
as det A = eiθ for some real number θ. Imposing that det A = 1 fixes this real
parameter θ. As a result, the real dimension of SU (n) is n2 − 1.
Just as for orthogonal groups, both U (n) and SU (n) are compact Lie
groups.
In this course we will focus on matrix Lie groups, such as the orthogonal
and unitary groups, which are realized as closed subgroups of the general linear
groups. This is why we will rarely need to go back to the formal abstract
definition of Lie groups.
x0 = a1 x + b1 ,
This is nothing else than the group operation on the two-dimensional Lie group
constructed in Example 4.1.4.
We also introduced matrix Lie groups above as groups of matrices. But
it is also customary to think of them as groups of transformations of target
space.
CHAPTER 4. LIE GROUPS AND LIE ALGEBRAS 87
α −β ∗
A=
β α∗
a2 + b2 + c2 + d2 = 1,
two and three dimensions, and then define the formal idea that applies to all
Lie groups.
in terms of an angle of rotation θ ∈ [0, 2π). The deep idea of Lie is that we
can recover all rotations by doing a rotation by a very small angle many many
times. Doesn’t sound very deep, but it is. :-) More precisely, we think of
infinitesimal rotations. If we take the angle θ to be small, we can approximate
cos θ ∼
= 1 + O(θ)2 and sin θ ∼ = θ + O(θ)3 . Keeping only terms of first order in
θ, we can thus approximate the rotation matrix as
cos θ sin θ ∼ 1 θ 0 1
= =I +θ .
− sin θ cos θ −θ 1 −1 0
(I + M T )(I + M ) ∼
= I + M + M T = I,
A∼
= I + θX
for some real parameter θ as before. So we have recovered the same first order
expansion as for rotations. We call X the generator of the Lie group (more
precisely, it is a matrix representation of the generator of the Lie group).
Checkpoint 4.2.1 Starting from orthogonal matrices, we recovered the first
order expansion of rotations. But this should only work for special orthogonal
matrices. It looks like we never imposed the condition that det A = 1. Why?
The key insight of Lie is that we can now recover finite rotations from the
knowledge of the generator X alone. Let R(θ) be a rotation by a finite angle
θ. Pick a large integer N . We can recover R(θ) by doing R(θ/N ) N times. In
the limit as N → ∞, R(θ/N ) becomes an infinitesimal rotation. Thus we get:
N
N θ
R(θ) = lim R(θ/N ) = lim I+ X .
N →∞ N →∞ N
x N
If we naively use the relation ex = limN →∞ (1 + N) for matrices, we would
conclude that
R(θ) = eθX .
In other words, we can recover finite rotations by exponentiating the infinites-
imal generator! That is cool. This is due to the group structure, which says
CHAPTER 4. LIE GROUPS AND LIE ALGEBRAS 90
Then any two-dimensional rotation R(θ) can be written in terms of the gener-
ator X via exponentiation:
R(θ) = eθX .
Proof. Let us start with a two-dimensional rotation R(θ). Since it is a Lie
group, we can Taylor expand near the identity:
dR 1 d2 R
R(θ) = I + θ+ θ2 + . . . (4.2.1)
dθ θ=0 2 dθ2 θ=0
(By derivative of a matrix here we mean derivative of its entries.) Let us now
identity the derivatives of the rotation matrix. Rotations form a Lie group.
The group structure is given by composition of rotations, which can be written
as the requirement that
Taking the derivative on both sides with respect to θ1 , and then setting θ1 = 0,
we get
dR(θ1 + θ2 ) dR(θ1 )
= R(θ2 ) . (4.2.2)
dθ1 θ1 =0 dθ1 θ1 =0
We can calculate the left-hand-side of (4.2.2) via the chain rule:
dR(θ1 + θ2 ) d(θ1 + θ2 ) dR(θ2 )
= .
d(θ1 + θ2 ) dθ1 θ1 =0 dθ2
dn R(θ)
= X n.
dθn θ=0
Thus all derivatives of the rotation matrix at the origin are determined by the
generator X! As is clear from the proof, this follows because of the group
structure of rotations.
Plugging this back into (4.2.1), with X 0 := I, we get
∞
X 1 n n
R(θ) = θ X = eθX ,
n=0
n!
AT A = (I + M T )(I + M ) ∼
= I + M T + M = I.
RR0 R−1 ∼
= (I + M )(I + M 0 )(I − M ) = I + M 0 + (M M 0 − M 0 M ),
Then:
3
X
[Li , Lj ] = i ijk Lk . (4.2.5)
k=1
• Reconstruct the group elements from the Lie algebra generators by ex-
ponentiation.
In this section we generalize the discussion about two-dimensional and
three-dimensional rotations to more general Lie groups. We will try to come
up with an abstract definition for the algebra of generators of a Lie group.
We will proceed in three steps. First, we will study how to extract the Lie
algebra corresponding to a given matrix Lie group. Second, we will define the
abstract notion of a Lie algebra. Third, we will show how we can recover the
group structure corresponding to a Lie algebra through exponentiation.
Geometrically, the Lie algebra is the linearization of the Lie group at the
origin. In other words, we replace the group manifold locally near the origin
by its tangent space. Since we are working with matrix Lie groups, we can
calculate this linearization explicitly by expanding the matrices of the funda-
mental representation near the identity, and keeping only terms of first order.
Let us be a little more precise.
We start with a matrix Lie group. That is, we start with its defining
representation, so we can think of the group elements as matrices A that depend
on n real parameters. We want to construct the associated Lie algebra. We
proceed in three steps.
1. We expand the group elements near the identity, to first order:
A = I + iL.
3. For any two group elements A, A0 with first order expansions A = I + iL,
A0 = I + iL0 , the commutator of matrices [L, L0 ] encodes whether the
group elements commute. Since L and L0 are linear combinations of the
generators Li , to know all such commutators we only need to calculate
the commutators of the generators Li . Because of the group structure
we know that the commutator closes, that is, the commutator of two
generators will be itself a linear combination of the generators. So we
can write:
d
X
[Li , Lj ] = cijk Lk (4.3.1)
k=1
for all x, y, z ∈ g.
♦
This is the abstract definition of a Lie algebra. Concretely, in this course the
Lie algebras that we will study will be obtained as linearizations of matrix Lie
groups. That is, they will be constructed as in Subsection 4.3.1, starting from
a matrix representation of a Lie group. Then the vector space g is constructed
as real linear combinations of the infinitesimal generators of the group, and the
Lie bracket is realized explicitly as a commutator of the matrix representation
for these generators. We know that the Lie bracket closes, that is, for any two
x, y ∈ g, [x, y] ∈ g, as required for a Lie algebra.
In this context, the three axioms in the definition of Lie algebras are clear.
The commutator of matrices is certainly bilinear and antisymmetric, by defi-
nition. As for the Jacobi identity, one can check that given any three matri-
ces A, B, C, the commutator [A, B] = AB − BA satisfies the Jacobi identity
[A, [B, C]] + [B, [C, A]] + [C, [A, B]] = 0.
Checkpoint 4.3.3 Check that for any three matrices A, B, C, the commuta-
tor [A, B] = AB − BA satisfies the Jacobi identity [A, [B, C]] + [B, [C, A]] +
[C, [A, B]] = 0.
Thus the procedure of Subsection 4.3.1 always produces a Lie algebra. The
structure of the algebra is encoded in the structure constants as in (4.3.1).
CHAPTER 4. LIE GROUPS AND LIE ALGEBRAS 96
g(~0) = e0 = I,
This follows because for any matrix A, we have eA e−A = I, which can be
checked by expanding the Taylor series for the exponential.
The tricky statement is whether the groupPmultiplication P closes. In other
words, we needP to show that given any two i ti Li and i si Li in g, then
there exists a i ui Li ∈ g such that
g(~t)g(~s) = g(~u).
eA eB = eC .
eA eB = eA+B .
4.4 SU (2)
Objectives
You should be able to:
• Determine the Lie algebra su(2).
• Explain the statement that SU (2) is a double cover of SO(3).
In this course we will mostly focus on the two most important Lie groups
in physics, namely SO(3) and SU (2). We have already calculated the Lie
algebra so(3) associated to the Lie group SO(3) in Section 4.2. We obtained
that so(3) is the three-dimensional vector spanned by elements L1 , L2 , L3 with
the bracket
3
X
[Li , Lj ] = i ijk Lk .
k=1
U † U = I, det U = 1.
U †U ∼
= (I + iL)† (I + iL) ∼
= I − iL† + iL = I.
Since we keep only terms of first order in L, we can ignore any terms that are
order two in the complex numbers a, b, c, d. Thus, to first order, we have
det U ∼
= 1 + a + d.
For this to be equal to one we must have a = −d, that is, the matrix L is
traceless (has vanishing trace).
The conclusion is that L must be a 2 × 2 traceless Hermitian matrix. Any
such matrix can be written as
t1 t2 + it3
L=
t2 − it3 −t1
What? This is the same abstract algebra as so(3)! Indeed, it turns out
that the Lie algebras so(3) and su(2) are isomorphic. Really, they are the
same. This is absolutely fundamental, and, as we will see, is the reason for
the appearance of fermions (particles with half-integer spin) in non-relativistic
physics.
This does not mean however that the Lie groups SO(3) and SU (2) are
isomorphic: in fact they are not. But they are certainly closely related. More
precisely, there exists a two-to-one group homomorphism from SU (2) to SO(3).
We say that SU (2) is a “double covering” of SO(3).
and SO(3), which share the same Lie algebras su(2) ∼ = so(3), are only locally
isomorphic.
We provide a formal proof of this statement using the adjoint representation
of the Lie group SU (2) below. But before we do that, let us write down an
informal argument, which may be more enlightening.
Both SU (2) and SO(3) share the same Lie algebra. For SU (2), we obtained
the Lie algebra by constructing the generators of SU (2), which is given by the
vector space spanned by the Pauli matrices (4.4.1), which we reproduce here
for convenience:
1 0 1 1 0 −i 1 1 0
T1 = , T2 = , T3 = .
2 1 0 2 i 0 2 0 −1
For SO(3), we obtained the algebra by looking at the generators of rotations
in three dimensions (4.2.4), which we also reproduce here:
0 0 0 0 0 −1 0 1 0
L1 = −i 0 0 1 , L2 = −i 0 0 0 , L3 = −i −1 0 0 .
0 −1 0 1 0 0 0 0 0
One can show that all special unitary matrices in SU (2) can be obtained by
exponentiation, and similarly for special orthogonal matrices in SO(3) (this is
because SU (2) and SO(3) are compact and connected). Thus we can write an
arbitrary U ∈ SU (2) as
3
!
X
U (θ1 , θ2 , θ3 ) = exp i θk Tk ,
k=1
This means that the vectors in R3 corresponding to X and AdU (X) have the
same length. In other words, AdU is a linear operator on R3 that preserves
the length of vectors: it is an orthogonal transformation in O(3) ⊂ GL(3, R).
Thus Ad : SU (2) → O(3) ⊂ GL(3, R). But in fact, we know even more. From
Example 4.1.10, we know that the underlying manifold structure of SU (2) is S 3 .
In particular, it is connected (in fact, it is also simply connected). Therefore,
the image of the continuous mapping Ad : SU (2) → O(3) must be connected to
the identity in O(3), which means that it must be a subgroup of SO(3) ⊂ O(3).
Thus Ad : SU (2) → SO(3).
What remains to be shown is that the group homomorphism Ad : SU (2) →
SO(3) is surjective, and that it is two-to-one.
We will leave the proof that Ad : SU (2) → SO(3) is surjective as an
exercise. (Note that there are many ways that this can be done.)
To show that Ad : SU (2) → SO(3) is two-to-one, pick any unitary matrix
U1 , and set U2 := −U1 . Then
since in both cases AdU (X) = X, i.e. AdU is the identity in SO(3). Thus
ker(f ) ∼
= Z2 , where Z2 is realized as ±1 times the identity matrix. Since the
kernel has two elements, this means that Ad : SU (2) → SO(3) is two-to-one.
Remark 4.4.2 We have shown that there is a two-to-one group homomorphism
Ad : SU (2) → SO(3). The kernel is Z2 , realized as ±1 times the identity matrix
in SU (2). Thus, by the first isomorphism theorem, we can write
∼ SU (2)/Z2 .
SO(3) =
The groups however are said to be “locally isomorphic”, since they have iso-
morphic Lie algebras. Equivalently, this can be seen since if U ∈ SU (2) is near
the identity, then −U ∈ SU (2) is not. And hence in a neighborhoud of the
identity the map is a local isomorphism.
103
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 104
Turing this around, one could think of vectors as being objects that transform
according to the three-dimensional defining representation of SO(3).
Remark 5.1.1 This is a good time to introduce a notation that is standard in
physics, which is to denote the representation of a Lie group by its dimension
in bold face. For instance, we would write 3 for the defining representation of
SO(3). The notation is slightly ambiguous though, because there may be more
than one representation of the same dimension. For instance, if a representa-
tion is complex, then its complex conjugate has the same dimension (we would
then write, say, 5 and 5 for the complex conjugate representations). Never-
theless, this is common notation, and it is usually clear from context what
representation is being discussed.
Now you probably see a pattern. How can we get more representations of
SO(3)? A natural way to think about representations is to think about how
they act. Namely, we think of objects that transform in certain ways under
rotations. We have already studied how scalars and vectors transform, which
gave us the one-dimensional and three-dimensional representation of SO(3).
What next?
The next objects to look at are matrices. How do matrices tranform under
rotations? Think of a matrix M as being a linear operator on R3 . Then under
a rotation R (which is a change of basis for R3 ) the matrix M will change
by a similarity transformation M 0 = RM RT , since R is orthogonal. In index
notation, this becomes
3
X 3
X
0 T
Mab = Rai Mij Rjb = Rai Rbj Mij .
i,j=1 i,j=1
This defines a new representation of SO(3). Indeed, we could place the nine
components of the matrix M in a vector, and then the transformation rule
above would give us a nine-dimensional representation of SO(3), where we
associate to every group element of SO(3) the 9 × 9 matrix with components
given by Rai Rbj , where R is the corresponding 3 × 3 rotation matrix. In other
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 105
9 = 3 ⊗ 3.
3
X
= Rai Rbj Aij .
i,j=1
A similar argument holds for Sij . Thus the subspaces of symmetric and anti-
symmetric rank 2 tensors are invariant subspaces, and thus provide subrepre-
sentations. The number of independent components in a rank 2 anti-symmetric
tensor in three dimensions is 3, while the number of independent components
in a rank 2 symmetric tensor in three dimensions is 6. We thus obtain the
decomposition 9 = 6 ⊕ 3 for the nine-dimensional tensor representation.
Now, are the 6 and 3 representations irreducible? It turns out that the
three-dimensional one is, and is in fact dual to the vector representation (since
in three dimensions rank 2 anti-symmetric tensors are dual to vectors, see Zee’s
book, section IV.1, for more on this). However the six-dimensional represen-
tation is still reducible. P3
Indeed, consider the subspace consisting of the trace S = i=1 Sii . Then
one sees that
3
X
S0 = 0
Saa
a=1
d X
X 3
= Rai Raj Sij
a=1 i,j=1
3 d
!
X X
T
= Ria Raj Sij
i,j=1 a=1
3
X
= δij Sij
i,j=1
= S,
Lemma 5.1.2 Let Sk1 ···kj be a rank j tensor that is fully symmetric under
permutations of indices, and such that it is traceless:
3
X
δk1 k2 Sk1 k2 ...kj = 0.
k1 ,k2 =1
(Note that it does not matter what pair of indices we choose here, since S is
fully symmetric.) Then S has 2j + 1 independent components.
Proof. We need to count the number of independent components of S. First,
let us count the number of components of a symmetric rank j tensor. Suppose
first that each index can only take values 1 and 2. There the possibilities
are 2 · · · 2, 2 · · · 21, 2 · · · 211, and so on. This gives j + 1 possibilities. Then
we add a 3. If we have a 3 for the first index, then we have j possibilities
for 2s and 1s in the remaining j − 1 indices. If we have two 3s, then we
have j − 1 possibilities
P for the remaining indices. And so on. Overall we get:
Pj j 1 1
k=0 (k + 1) = k=0 k + (j + 1) = 2 j(j + 1) + (j + 1) = 2 (j + 1)(j + 2)
possibilities. So a symmetric rank j tensor has 21 (j + 1)(j + 2) independent
components. P3
The tracelessness condition k1 ,k2 =1 δk1 k2 Sk1 k2 ...kj = 0 consists in a num-
ber of independent conditions. The number of conditions here is the number
of values that the indices k3 · · · kn can take. From the previous paragraph, this
is the number of independent components of a rank j − 2 symmetric tensor,
which is 12 j(j − 1). Therefore, the total number of independent components of
a symmetric traceless rank j tensor is
1 1
(j + 1)(j + 2) − j(j − 1) = 2j + 1.
2 2
The end result is that there is an infinite number of irreducible representa-
tions for SO(3), indexed by a non-negative integer j, with dimensions 2j + 1.
The objects that transform according to these representations are symmetric
traceless rank j tensors.
This is the key result of this section. And, if you have done some physics,
you may have encountered the formula 2j + 1 before. It is rather famous in
the history of quantum mechanics and atomic physics. It corresponds to the
degeneracy of states of the hydrogen atom (or spherical harmonics). It also
corresponds to the multiplicity, or quantum states, of particles with integer
spin. This is certainly not a coincidence, as we will see!
algebra, one may obtain the corresponding representation of the Lie group by
exponentiation. Since representations of Lie algebras are easier to construct
than for the corresponding Lie groups, we now move on to the construction of
irreducible representations of Lie algebras.
In this section we will define more precisely what representations of Lie
algebras are, and then look at one important representation that exists for all
Lie algebras: the adjoint representation.
5.2.1 Definition
Recall that a Lie algebra g is a vector space with a bracket [·, ·] that satisfies a
bunch of axioms. Suppose that g is n-dimensional, and pick a basis L1 , . . . , Ln
for g. Then we can write
n
X
[Li , Lj ] = cijk Lk , (5.2.1)
k=1
Γ : g → gl(V )
that is compatible with the brackets (this is called a Lie algebra homomor-
phism):
Γ([x, y]) = [Γ(x), Γ(y)] := Γ(x)Γ(y) − Γ(y)Γ(x),
for all x, y ∈ g. The dimension of the representation is the dimension of V .
♦
As for groups, we can talk about irreducible representations, which
are representations such that V has no proper invariant subspace, that is, no
proper subrepresentation.
Remark 5.2.2 When we define representations of groups, we map group ele-
ments to linear operators in GL(V ), that is, invertible matrices. Invertibility is
required because of the group properties, since we map the group operation to
matrix multiplication. For Lie algebras, we map elements of the vector space
to linear operators in gl(V ), i.e. matrices that are not necessarily invertible.
This is an important distinction.
We have already seen examples of Lie algebra representations.
Example 5.2.3 Trivial representation. As for groups, the trivial repre-
sentation Γ always exists for Lie algebra. It is defined by sending all elements
of g to the zero endomorphism 0 of a one-dimensional vector space V , which
maps all elements of V to the origin. Indeed, if Γ(Li ) = 0 for i = 1, . . . , n,
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 109
which defines the three-dimensional Lie algebra su(2). More precisely, the Pauli
matrices (5.2.2) furnish a two-dimensional representation of the Lie algebra
su(2), since they map the generators to 2×2 matrices that obey the appropriate
commutation relations.
In the case of SO(3), we started with the 3-dimensional representation of
SO(3) consisting of 3 × 3 special orthogonal matrices. We found in (4.2.4) that
the infinitesimal generators are:
0 0 0 0 0 −1 0 1 0
L1 = −i 0 0 1 , L2 = −i 0 0 0 , L3 = −i −1 0 0 .
0 −1 0 1 0 0 0 0 0
(5.2.3)
Those also satisfy the commutation relations:
3
X
[Li , Lj ] = i ijk Lk .
k=1
Thus (5.2.3) defines another representation of the same Lie algebra su(2) ∼
=
so(3), this time a three-dimensional one.
What this equation means is that the jk’th component of the matrix Γi is given
by the structure constant −cijk . Then these Γ1 , . . . , Γn form a n-dimensional
representation of g.
Proof. To show that it is a representation, we need to check that the mapping
respects the bracket, that is, that the matrices Γi satisfy the commutation
relations
n
X
[Γi , Γj ] = cijk Γk .
k=1
Well, it turns out that this is a direct consequence of the Jacobi identity. Recall
that for a Lie algebra g, the bracket [·, ·] satisfies the Jacobi identity
Since the Lb are linearly independent, this means that the coefficients of the lin-
ear combination must be identically zero, so we obtain the following condition
on the structure constants:
n
X
(cjka ciab + ckia cjab + cija ckab ) = 0.
a=1
This is an equation for the kb’th component of a matrix, where the sum over
a comes from matrix multiplication. Thus it can be rewritten in matrix form
as the equation
n
X
Γi Γj − Γj Γi = cija Γa ,
a=1
We can define the adjoint representation a little more abstractly, for those
who like this stuff. The adjoint representation is a representation that maps
elements of the vector space g to linear operators on the vector space g itself.
That is, we think of elements of the Lie algebra as operators acting on the
algebra itself. The formal definition is the following:
Definition 5.2.6 The adjoint representation (bis). Let g be a Lie algebra,
and let gl(g) be the Lie algebra of linear endomorphisms of g. The adjoint
representation is defined as the mapping
ad :g → gl(g)
x 7→ adx := [x, ·],
adx :g → g
y 7→ [x, y].
♦
In this more abstract formulation, we send an element x ∈ g to the operator
that acts on g by evaluating the bracket of any y ∈ g with the fixed x ∈ g. It
may not be obvious at first that this is equivalent to the definition in terms of
structure constants, but it is. We will leave the comparison as an exercise.
Checkpoint 5.2.7 Show that the abstract definition of the adjoint represen-
tation is equivalent to the definition in terms of structure constants, by cal-
culating the components of the matrices corresponding to the linear operators
adLi .
Example 5.2.8 The adjoint representation of su(2). Let us now construct
the adjoint representation of su(2). We recall the commutation relations
3
X
[Ti , Tj ] = i ijk Tk .
k=1
The structure constants are thus cijk = iijk . We construct the adjoint repre-
sentation by defining the three matrices Γi , i = 1, 2, 3 with components given
by
(Γi )jk = −iijk .
Recalling the definition of the Levi-Civita symbol, we see that these matrices
take the form
0 0 0
Λ1 = − i 0 0 1 ,
0 −1 0
0 0 −1
Λ2 = − i 0 0 0 ,
1 0 0
0 1 0
Λ3 = − i −1 0 0 .
0 0 0
Those are precisely the generators of SO(3), see (5.2.3)! Therefore, the ad-
joint representation of su(2) ∼
= so(3) constructs the infinitesimal generators of
SO(3), and by exponentiation the defining representation of SO(3). Note that
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 112
At this stage, all that we know is that m is a real number, since representations
of su(2) are composed of Hermitian matrices (and hence with real eigenvalues).
Remark 5.3.1 The Dirac bra-ket notation. The Dirac bra-ket notation is
conventional in the treatment of Lie algebras in both physics and mathematics.
In this notation, we represent a complex vector v as |vi. We represent its
complex conjugate transpose (v ∗ )T as hv|. Thus the norm square of v can be
written as
n
X
hv|vi = (v ∗ )T v = |vi |2 .
i=1
hj|ji = 1.
|m − 1i := Γ− |mi.
Then, acting on the highest weight state |ji with Γ− , we produce a tower of
states:
|ji, |j − 1i, ..., |qi.
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 114
We know that the tower must end at some point, since we are constructing a
finite-dimensional representation. Thus there must be an eigenstate |qi of Γ3
such that
Γ− |qi = 0.
The next step is to figure out what q is for a given j. This will tell us the
dimension of the corresponding representation. To this end, we notice that we
have not fully determined how Γ+ acts on an eigenstate of Γ3 . All that we
know is that
Γ+ |mi = Nm |m + 1i
for some normalization constant Nm .
Lemma 5.3.2 Let |mi be eigenstates of Γ3 such that Γ− |mi = |m − 1i, with
highest weight state |ji (that is, Γ+ |ji = 0). Then
Γ+ |mi = Nm |m + 1i
with
1 1
Nm = j(j + 1) − m(m + 1).
2 2
Proof. We can compute Nm as follows:
Nm |m + 1i =Γ+ |mi
=Γ+ Γ− |m + 1i
=(Γ− Γ+ + Γ3 )|m + 1i
=(Nm+1 + (m + 1))|m + 1i.
Nm = Nm+1 + (m + 1).
We can solve this recursion as follows. First, since |ji is the highest weight
state, we know that Nj = 0. Thus
Nj−1 = j.
Then
Nj−2 = Nj−1 + (j − 1) = j + (j − 1).
By induction, we get
s−1
X 1 s
Nj−s = (j − k) = sj − s(s − 1) = (2j − s + 1),
2 2
k=0
0 =Γ+ Γ− |qi
=(Γ− Γ+ + Γ3 )|qi
=(Nq + q)|qi,
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 115
1
with Nq = 2 j(j + 1) − 21 q(q + 1). Thus we are looking for a solution to the
equation
1 1 1 1
j(j + 1) − q(q + 1) + q = j(j + 1) − q(q − 1) = 0.
2 2 2 2
There are two solutions: q = j + 1 and q = −j. The first one does not make
sense, since |ji is a highest weight state. Therefore q = −j.
What this means is that we have constructed the following tower of eigen-
states for Γ3 :
There are 2j + 1 states. Since these states span the vector space on which our
representation acts, 2j + 1 must be a positive integer. That is, j must be a
non-negative half-integer.
The outcome of this construction is the following important result:
Theorem 5.3.3 Irreducible representations of su(2). There is an infi-
nite tower of irreducible representations for su(2), indexed by a non-negative
half-integer j, with dimensions 2j + 1. We call the (2j + 1)-dimensional rep-
resentation the spin-j representation, and denote its states by |j, mi (those
are eigenstates of Γ3 with eigenvalues m = −j, −j + 1, . . . , j − 1, j ) .
We have not proved here that these representations are irreducible, and
that we have obtained all irreducible representations: this is beyond the scope
of this class.
with Nm = 12 j(j + 1) − 12 m(m + 1). One can show that these states are orthog-
onal, that is,
hj, m0 |j, mi = 0 if m0 6= m.
Checkpoint 5.3.4 Show that hj, m0 |j, mi = 0 if m0 6= m.
However, the states |j, mi are not normalized. We normalized the highest
weight state by requiring that hj, j|j, ji = 1, but hj, m|j, mi =
6 1 for m ≤
2
j − 1. Let Cm = hj, m|j, mi. We can construct a new basis of states that are
orthonormal as
^ 1
|j, mi = |j, mi.
Cm
Lemma 5.3.5 Let |j, ^ ^
mi be a tower of eigenstates Γ3 |j, ^
mi = m|j, mi, for
m = −j, −j + 1, . . . , j − 1, j that are orthonormal:
^
hj, m0 |j,
] mi = δm0 ,m .
Then
r
^ 1
Γ+ |j, mi = (j + m + 1)(j − m)|j,^
m + 1i,
2
r
^ 1
Γ− |j, mi = (j + m)(j − m + 1)|j,^
m − 1i.
2
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 116
Proof. We can construct a basis of orthonormal states from our previous basis
as
^ 1
|j, mi = |j, mi.
Cm
Clearly, this does not change how Γ3 acts, since
^
Γ3 |j, ^
mi = m|j, mi.
where for the second equality we applied Γ+ Γ− on the ket |j, mi. Thus, if we
2
write Cm = hj, m|j, mi, we get the relation
2 2 1 2
Cm−1 = Nm−1 Cm = (j + m)(j − m + 1)Cm . (5.3.3)
2
Now we know that
Γ− |j, mi = |j, m − 1i.
In terms of the normalized basis, this becomes
^ Cm−1 ^
Γ− |j, mi = |j, m − 1i
Cm
Substituting (5.3.3), we get:
r
^ 1
Γ− |j, mi = (j + m)(j − m + 1)|j,^
m − 1i.
2
As for Γ+ , we know that
j−m
Γ+ |j, mi = Nm |j, m + 1i = (j + m + 1)|j, m + 1i.
2
In terms of the normalized basis, we get:
^ Cm+1 j − m
Γ+ |j, mi = (j + m + 1)|j,^
m + 1i.
Cm 2
Substituting (5.3.3):
s
^ 2 j−m
Γ+ |j, mi = (j + m + 1)|j,^
m + 1i
(j + m + 1)(j − m) 2
r
1
= (j + m + 1)(j − m)|j,^
m + 1i.
2
Remark 5.3.6 For clarity, we will now drop the tilde symbol on the normalized
states. From now on all states will be assumed to be normalized.
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 117
Then the action of the operators Γ3 and Γ± can be rewritten in matrix form
as
1 1 0 1 0 1 1 0 0
Γ3 = , Γ+ = √ , Γ− = √ .
2 0 −1 2 0 0 2 1 0
1 i
Γ1 = √ (Γ+ + Γ− ), Γ2 = − √ (Γ+ − Γ−). (5.3.4)
2 2
Thus
1 0 1 i 0 1 1 1 0
Γ1 = , Γ2 = − , Γ3 = .
2 1 0 2 −1 0 2 0 −1
Those are none other than the Pauli matrices giving the defining representation
of SU (2)! See Example 5.2.4. Thus we have recovered the two-dimensional
representation of su(2) that is obtained from the infinitesimal generators of
SU (2).
Example 5.3.8 The spin 1 representation. Consider now the spin 1
representation with j = 1, with the three normalized states |1, 1i, |1, 0i and
|1, −1i. The action of Γ± can be calculated to be:
and
Γ+ |1, 1i = 0, Γ+ |1, 0i = |1, 1i, Γ+ |1, −1i = |1, 0i.
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 118
One can check that these matrices indeed satisfy the commutation relations of
the su(2) Lie algebra. In fact, they are equivalent to the defining representation
of SO(3) as presented in Example 5.2.4. Indeed, one can find a similarity
transformation that brings the matrices of the defining representation of SO(3)
into the matrices above (this is the similarity trasnformation that diagonalizes
Γ3 ).
T : G → GL(V )/F ∗ ,
where elements of GL(V )/F ∗ are equivalence classes of invertible linear trans-
formations of V that differ by overall rescaling. ♦
This is precisely what we want in quantum mechanics, since overall rescaling
of the wave-function by a phase factor does not change the physics. In other
words, given a group of symmetry for a quantum mechanical system, we know
that it acts on the space of states of the system as a projective representation.
So our goal to understand the system is to classify projective representations
of the group of symmetries.
But how do we construct projective representations of a group G? We have
already seen that the irreducible representations of a Lie algebra are in one-to-
one correspondence with the irreducible ordinary representations of the asso-
ciated simply-connected Lie group. As for projective representations, the key
result is Bargmann’s theorem, which tells us that for a large class of groups
(which includes SO(3), the Lorentz group and the Poincare group), every pro-
jective unitary representation of G comes from an ordinary representation of
the universal cover. Great! In fact, for finite-dimensional representation this
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 121
is always true regardless of the group. But in quantum mechanics the Hilbert
space is often infinite-dimensional, so one has to be careful, but Bargmann’s
theorem holds for the groups of interest such as rotations, translations and
Lorentz transformations.
In the context of SO(3), we know that its universal cover is SU (2). So what
this means is that all irreducible projective representations of SO(3) come from
ordinary irreducible representations of SU (2), which in turn correspond to irre-
ducible representations of the Lie algebra su(2) =∼ so(3), which we have already
constructed. So we already know all projective representations of SO(3)!
Before we look at those representations more closely, let us define a partic-
ular type of projective representations that is very important in physics.
cover of SO(3), we know that they will all be either ordinary or spin represen-
tations of SO(3). How do we distinguish between the two?
Let us first work out an example. Consider the spin-1/2 representation of
su(2), with j = 1/2. The representation is given by the Pauli matrices (4.4.1),
reproduced here for convenience:
1 0 1 1 0 −i 1 1 0
T1 = , T2 = , T3 = .
2 1 0 2 i 0 2 0 −1
Now consider two elements of the Lie algebra given by θ1 T3 and θ2 T3 for some
θ1 , θ2 ∈ R. By exponentiation, these give rise to matrices
where I is the 2×2 identity matrix. But in SO(3), rotating about an axis by 2π
should be the identity transformation, not minus the identity transformation!
Thus we conclude that this representation does not quite preserve the group
structure: it preserves it only up to sign. That is, it is a spin representation of
SO(3)!
This was somehow expected, because when we constructed representa-
tions of SO(3) in terms of how tensors transform, we only obtained (2j + 1)-
dimensional representations with j a non-negative integer: we never saw the
representations with j a half-integer.
But what about the other representations of SO(3) with half-integer j?
Are they ordinary or spin representations?
Consider the highest weight construction in the previous section. In this
construction, Γ3 was always a diagonal matrix, whose diagonal entries corre-
sponded to the values of m between −j and j. If j is a half-integer, then those
diagonal entries are all half-integers. Then if we construct matrices R(θ) by
exponentiating the matrices θΓ3 , the same argument above will go through,
and we will always end up with the statement that a rotation by 2π gives minus
the identity matrix. (Note that this does not happen if the diagonal values of
Γ3 are integers, which is the case when j is an integer.)
Therefore we obtain the very important result that finite-dimensional irre-
ducible projective representations of SO(3) come in two classes:
• An infinite class of ordinary representations, indexed by a non-negative
integer j, with dimensions 2j + 1. Those are tensor representations,
and we call objects that transform according to these representations
tensors (for j = 0, we call them scalars, and for j = 1, we call them
vectors).
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 123
ψi := (ψ ∗ )i .
This is of course just a notational convention, but it is very useful to keep track
of the transformation properties of tensors.
We then denote the components of a unitary matrix U by U i a . Its transpose
matrix U † then have components (U † )a i . With this convention the upshot is
that to construct objects with appropriate transformation properties under
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 125
unitary transformations, we should always sum only over one index that is
upper and one that is lower. This follows because of the unitary property that
U † U = I, or, in index notation,
N
X
(U † )b j U j a = δab .
j=1
PN
Therefore, the object i=1 Tiji transforms like a tensor with one single upper
index (i.e. in the N defining representation of SU (N )). So summing over one
upper and one lower index for any tensor creates a subrepresentation. We call
this operation taking the trace of a tensor.
Since taking the trace always defines a subrepresentation, to understand
irreducible representations of SU (N ) we know that we should look for traceless
tensors that have appropriate symmetry properties upstairs and downstairs.
In general, the study of irreducible representations from this point of view
involves Young tableaux, as already mentioned for SO(N ). The idea is similar,
where we now look at general tensor products N ⊗ · · · N ⊗ N ⊗ · · · ⊗ N, and
study their decompositions as direct sums of irreducible representations using
Clebsch-Gordan coefficients.
We will not discuss Young tableaux in this course, but let us at least enu-
merate the first few non-trivial tensor representations of SU (N ):
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 126
Thus,
T (g)∗ = ST (g)S −1
for some S if and only if
Tk∗ = −STk S −1
for the Pauli matrices Tk , k = 1, 2, 3.
One can check using matrix multiplication that Pauli matrices satisfy the
property that
Ti Tj = −Tj Ti for i 6= j, Ti2 = I.
Since T1 and T3 are real-valued, it thus follows that
For SU (3), the statement is that all irreducible representations are given by
traceless tensors that are fully symmetric upstairs an downstairs individually.
i ···i
That is, tensors of the form Tk11 ···kpq that are traceless and fully symmetric
under permutations of the ia and the kb separately.
Thus we conclude that irreducible representations of SU (3) are indexed by
two non-negative integers (p, q), corresponding to the number of upper and
lower indices respectively. What is the dimension of such a representation?
Lemma 5.5.3 Dimension of irreducible representations of SU (3). A
i ···i
traceless tensor Tk11 ···kpq that is fully symmetric under permutations of the ia
and the kb separately has
1
(p + 1)(q + 1)(p + q + 2)
2
independent components. Thus, the dimension of the (p, q) irreducible repre-
sentation of SU (3) is 12 (p + 1)(q + 1)(p + q + 2).
i ···i
Proof. A tensor Tk11 ···kpq that is fully symmetric under permutations of the ia
and the kb separately has
1 1
(p + 1)(p + 2) (q + 1)(q + 2) (5.5.2)
2 2
How many constraints does this give? Well, the indices i2 , . . . , ip and k2 , . . . , kp
can take any values here. So the number of constraints is the number of
independent components of a tensor with p − 1 upper indices and q − 1 lower
indices that is fully symmetric upstairs and downstairs. By (5.5.2), this is
1 1
(p)(p + 1) (q)(q + 1) . (5.5.3)
2 2
We are now ready to do some cool physics and see how particles in the
Standard Model can be labeled by representations, and how the idea of Grand
Unified Theories (GUTs) then naturally arises.
• The matter particles, i.e. those that compose matter, such as electrons,
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 131
muons, quarks, etc. all correspond to spin 1/2 particles (i.e. they are
spinors).
• The Higgs boson is the only spin 0 particle.
• However, with the recent discovery of neutrino masses and neutrino os-
cillation, it is generally believed that one more particle should be added
to the Standard Model, which would responsible for giving masses to
the neutrinos. This particle is the “right-handed neutrino”, which would
transform according to the trivial representation (1, 1)0 . The existence
of the right-handed neutrino has not been confirmed experimentally how-
ever.
This is all very cool. But, from a theoretical physics viewpoint, the represen-
tations that appear in the Standard Model appear rather random. Why are
there particles transforming in the (3, 2)1/3 representation, but not in, say, the
(3, 2)1/3 representation? Who chose this seemingly random list of representa-
tions?
One answer to this question could be that there is no answer, and that
this is just what Nature says. That is a fine answer. But often, when there
is something that seems rather random or mysterious, this means that we are
missing something. And going a little deeper into representation theory, one
sees that indeed, these representations are not random. This is the fundamental
idea behind Grand Unified Theories (GUTs).
1 →(1, 1)0 ,
5 →(3, 1)−1/3 ⊕ (1, 2)1/2 ,
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 132
10 →5−2 ⊕ 52 ,
16 →101 ⊕ 5−3 ⊕ 15 ,
45 →240 ⊕ 10−4 ⊕ 104 ⊕ 10 .
points to is that the larger groups SU (5) and/or SO(10) may perhaps play a
role in whatever theory of physics goes beyond the Standard Model. If this
is the case, it would be a tremendous achievement of representation theory in
particle physics.
an element of SO(4), corresponding to the rotation that brings the first point
on S 3 to the second. Finally, one sees that this mapping is not one-to-one,
since the two points U, V ∈ S 3 and the opposite points −U, −V ∈ S 3 will be
mapped to the same rotation in SO(4).
From the point of view of Lie algebras, the statement that SU (2) × SU (2)
is a double cover of SO(4) implies that they must share the same Lie algebra
∼ su(2) ⊕ su(2) (the Lie algebra of a product group is the direct sum of
so(4) =
the Lie algebras of the individual factors). Let us see this explicitly.
Lemma 5.7.3 The isomorphism between so(4) and su(2) ⊕ su(2). The
Lie algebra so(4) is isomorphic to the Lie algebra su(2) ⊕ su(2).
Proof. Following along the same lines as we did for so(3) and su(2), we can
find an explicit description of the Lie algebra so(4). It is six-dimensional, and
there is a natural choice of basis, which we denote by Ji , Ki with i = 1, 2, 3,
with the bracket:
3
X
[Ji , Jj ] =i ijk Jk , (5.7.1)
k=1
3
X
[Ji , Kj ] =i ijk Kk , (5.7.2)
k=1
3
X
[Ki , Kj ] =i ijk Jk . (5.7.3)
k=1
To see the isomorphism with the Lie algebra su(2) ⊕ su(2), we do a change of
basis. We define a new basis
1
J±,i = (Ji ± Ki ), i = 1, 2, 3.
2
It is straightforward to check that the bracket becomes, in this new basis,
We recognize that the J+,i and the J−,i generate two copies of the su(2) Lie
algebra. Moreover, those commute, and hence the Lie algebra is a direct sum
su(2) ⊕ su(2). We conclude that so(4) ∼
= su(2) ⊕ su(2).
m 0 1 2 3 4 5 6 7
Type Real Real Complex Pseudo-real Pseudo-real Pseudo-real Complex Real
To end this section, let us check that the table is consistent with what we
have found so far. SO(3) corresponds to k = 0 and m = 3, which says that
the one 2-dimensional spin reprepresentation is pseudo-real, which is what we
have found. For SO(4), we have k = 0 and m = 4, and hence from the table
we conclude that the two 2-dimensional spin representations are pseudo-real,
which is indeed what we stated above. Great!
The only difference with (5.7.3) is the sign in the bracket [Ki , Kj ]. This sign
is however quite important.
To determine the representations of SO(3, 1), we want to construct its
double cover, as we did for SO(4). Let us first do it at the level of Lie algebras.
As in (5.7.3), we want to find a change of basis to construct an isomorphism
of Lie algebras.
The key is to realize that if we do a change of basis Kj 7→ iKj , the brackets
in (5.7.6) become exactly equal to those in (5.7.3). But we are only allowed to
do such a change of basis if we consider the “complexified Lie algebra”, that
is, we consider the complex vector space with basis given by the Ji , Ki . We
denote this complex Lie algebra as so(3, 1)C . Thus what we have found is that
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 138
the complexified Lie algebra so(3, 1)C and so(4)C are isomorphic. In particular,
doing the change of basis as for (5.7.3), we conclude that
so(3, 1)C ∼
= su(2)C ⊕ su(2)C .
But in the end we want to work with the real Lie algebra so(3, 1). So we want
to turn the right-hand-side into the complexification of a single Lie algebra
(instead of direct sum of complex Lie algebras).
To do so we proceed through a sequence of isomorphisms of Lie algebra.
The key is the isomorphism su(2)C = ∼ sl2 (C), which can be checked from the
commutation relations of those. Then we get:
so(3, 1)C ∼
=su(2)C ⊕ su(2)C (5.7.7)
∼
=sl2 (C) ⊕ sl2 (C) (5.7.8)
∼
=sl2 (C) ⊕ isl2 (C) (5.7.9)
∼
=sl2 (C)C , (5.7.10)
where in the last line we mean the complexification of the real Lie algebra
sl2 (C) associated to the Lie group SL(2, C).
As a result, we get so(3, 1)C ∼
= sl2 (C)C , which is an isomorphism between
the complexifications of two Lie algebras. Upon restriction to the real Lie
algebras this becomes the isomorphism
so(3, 1) ∼
= sl2 (C).
Thus the Lie groups SO(3, 1) and SL(2, C) share the same Lie algebra.
It turns out that SL(2, C) is simply connected, and is in fact the universal
cover of SO(3, 1). Moreover, the map SL(2, C) → SO(3, 1) is a double cover-
ing, and hence Spin(3, 1) ∼= SL(2, C). We have found the spin group Spin(3, 1)!
Thus all finite-dimensional irreducible projective representations of SO(3, 1)
descend from irreducible finite-dimensional representations of SL(2, C).
What are the irreducible representations of SL(2, C)? Well, it is easier to
work at the level of Lie algebras. I will be rather sketchy here for brevity. The
isomorphisms (5.7.10) are very useful. First, it turns out that the complex rep-
resentations of the complexification sl2 (C)C are in one-to-one correspondence
with the real representations of sl2 (C). Thus, the latter are in one-to-one cor-
respondence with the complex representations of su(2)C ⊕ su(2)C . But those
are equivalent to the representations we constructed earlier when we studied
SO(4). They are indexed by two half-integers j1 and j2 and have dimensions
(2j1 + 1)(2j2 + 1). Therefore, the same is true of the real irreducible represen-
tations of SL(2, C).
It turns out that as for SO(4), the representations with j1 + j2 ∈ Z descend
to ordinary irreducible representations of SO(3, 1), while the other ones are
spin. The (1/2, 0) and (0, 1/2) are the two 2-dimensional spin representations;
the (1/2, 1/2) is the four-dimensional representation (which corresponds to how
a four-vector in spacetime transforms).
As for SO(4), we can ask whether the spin representations are real, pseudo-
real, or complex. It turns out that for SO(3, 1), the two 2-dimensional spin
representations are in fact complex conjugate of each other. This is one key
difference with SO(4), in which case they were pseudo-real.
In fact, the table Table 5.7.5 can be generalized for arbitrary special or-
thogonal groups SO(p, q). The type of spin representations is again periodic
with periodic eight, but it now depends on p − q mod 8. We get the following
generalized table, which reduces to Table 5.7.5 for q = 0.
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 139
p − q mod 8 0 1 2 3 4 5 6 7
Type Real Real Complex Pseudo-real Pseudo-real Pseudo-real Complex Real
where Lµν are the components of a Lorentz transformation. Thus the Poincare
group includes both Lorentz transformations and translations.
We can extract the Lie algebra of the Poincare group as usual. It is 10-
dimensional. A convenient choice of basis uses the generators Jµν = −Jνµ for
the Lorentz transformations, and four generators Pµ for the translations. The
commutation relations can be written as
[Pµ , Pν ] =0,
[Jµν , Pρ ] = − i(ηµρ Pν − ηνρ Pµ ),
[Jµν , Jρσ ] = − i(ηµρ Mνσ − ηµσ Mνρ − ηνρ Mµσ + ηνσ Mµρ ),
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 140
where ηµν are the components of the Minkowski metric (i.e. the diagonal
matrix diag(−1, 1, 1, 1)).
Remark 5.8.1 The generators Jµν for Lorentz transformations are connected
to the generators of rotation Ji and boosts Ki introduced in (5.7.6) by
3
1 X
Ji = imn J mn , Ki = Ji0 ,
2 m,n=1
one of the label for representations of the Poincare group. Thus mass appears
naturally in this context; it is one of the label that specifies the representations
of the Poincare group.
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 141
In any case, going back to the classification, we fix an eigenvector |pi, with
eigenvalues pµ . The idea of the method of little groups is to now consider
the subgroup of the Poincare group that leaves pµ invariant: this is called the
little group (in mathematics, it is called the “stabilizer subgroup”). The idea
of the method is to induce unitary irreducible representations of the whole
group from the unitary irreducible representations of the little group. In other
words, we classify the states with a fixed momentum pµ in terms of the unitary
irreducible representations of the little group.
Concretely, we need to separate the classification into three cases, depend-
ing on the choice of starting eigenvector |pi:
1. Positive mass m > 0, in which case we choose |pi to have eigenvalues pµ =
(m, 0, 0, 0) (that is, we use Lorentz transformations to bring ourselves to
the rest frame of the particle);
2. Zero mass m = 0 but with non-zero pµ , in which case we choose |pi to
have eigenvalues pµ = (p, 0, 0, p), p > 0;
The Lie algebra se(2) associated to the special Euclidean group SE(2) is
three-dimensional. One generator J is the generator of infinitesimal rotations,
with the two other generators P1 and P2 being generators of infinitesimal trans-
lations. The commutation relations are:
Pi |ki = ki |ki, i = 1, 2.
Isn’t that great? This is why it makes sense to talk about the mass and spin
(or helicity) of particles: that’s because those index the unitary irreducible
representations of the Poincare group! Moreover, the distinction between mas-
sive and massless particles which is crucial in physics appear naturally here in
terms of representations. Beautiful!
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 143
c123 =1,
1
c147 = − c156 = c246 = c257 = c345 = −c367 = ,
√ 2
3
c458 =c678 = .
2
Looking at the Gell-Mann matrices, we see that there are two diagonal
generators with real eigenvalues: L3 and L8 . We conclude that these generate
the Cartan subalgebra of su(3). Thus, while the dimension of su(3) is 8, its
rank is 2. In general, the dimension of su(N ) is N 2 − 1 while its rank is N − 1.
We want to continue mimicking the highest weight construction for su(2).
We construct representations of g by constructing the vector space on which
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 145
they act. We take a basis for that vector space consisting of eigenvectors of the
simultaneously diagonalizable Cartan generators. To each such basis vector
|αi is associated a vector α = (α1 , . . . , αm ) ∈ Rm of eigenvalues of this basis
vector under the action of the Cartan generators H1 , . . . , Hm . That is,
Hi |αi = αi |αi.
Definition 5.9.7 Weight vector. The weight vectors of a representation
Γ of a Lie algebra g are the vectors of eigenvalues of the eigenvectors of the
Cartan generators. These vectors live in Rm , where m is the rank of g. We
call Rm the weight space of g. ♦
Remark 5.9.8 Note that the weight space is the same for all representations
of a Lie algebra g, but the weight vectors change. In fact, one can think of a
representation as being determined by a set of weights in weight space.
This is indeed the natural generalization of the highest weight construction
for su(2). In this case, there was only one Cartan generator, so the weight space
was one-dimensional, i.e. it was just R. The weight vectors of a representation
were given by the half-integers −j, −j+1, . . . , j−1, j ∈ R, for some non-negative
half-integer j.
Example 5.9.9 The weights of the fundamental representation of
su(3). Going back to su(3), with generators given by the Gell-Mann matrices
(see (5.9.3), we can extract the weights of the representation by calculating
1
the eigenvalues of the Cartan generators for the basis vectors e1 = 0,
0
0 0
e2 = 1, e3 = 0 on which the representation acts. We get
0 1
1 1
T3 e1 = e1 , T8 e1 = √ e1 ,
2 2 3
1 1
T3 e2 = − e2 , T8 e2 = √ e2 ,
2 2 3
1
T3 e3 =0, T8 e3 = − √ .
3
1
So the weight vectors of the fundamental representation are ( 12 , 2√ 3
), (− 12 , 2√
1
3
)
1 2
and (0, − 3 ), which live in the weight space R .
√
It turns out that the adjoint representation plays a very special role.
Definition 5.9.10 Root. The roots of a Lie algebra g are the non-zero
weights of its adjoint representation. ♦
The roots play an important role in the highest weight construction of
representations. First, one can show that if α is a root, then −α is also a root.
So roots come in pairs ±α. It turns out that the generators of the Lie algebra
that are not generators of the Cartan subalgebra can also be grouped in pairs
E±α , which play the role of raising and lowering operators. Indeed, one can
show that if |µi is an eigenvector of the Cartan generators Hi , i = 1, . . . , m
with eigenvalue µ = (µ1 , . . . , µm ), then
In other words, E±α |µi is also an eigenvector of the Cartan generators, but
with eigenvalues raised or lowered by the value of the corresponding root α.
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 146
In this way, we can define the notion of highest weight vector, and construct
representations using the raising and lowering operators associated to the roots
of the Lie algebra.
In fact we can also use the result above to determine the roots. If we
can rearrange the remaining generators of the Lie algebra into pairs of raising
and lowering operators E±α , when we can determine the roots by evaluating
[Hi , E±α ] = ±αi E±α . The αi give the components of the root vector associated
to this pair of lowering and raising operators. Note that to calculate these
commutators, we can use whatever representation we want. So in this way,
if we can identity the pairs of raising and lowering operators, we can obtain
the roots without ever writing down the adjoint representation, for instance
by calculating commutators using the simpler fundamental representation.
Example 5.9.11 Roots of su(3). One way to find the roots of su(3) is to
write down the 8 × 8 matrices corresponding to the Cartan generators T3 and
T8 in the adjoint representation, and extract their eigenvalues. Or, we can use
our knowledge of the Gell-Mann matrices (5.9.3) and deduce the three pairs
of raising and lowering operators. Indeed, (T1 , T2 ), (T4 , T5 ) and (T6 , T7 ) form
pair of Pauli matrices embedded in 3 × 3 matrices. So the pairs of raising and
lowering operators for su(3) should be:
1
E±α(1) = √ (T1 ± iT2 ),
2
1
E±α(2) = √ (T4 ± iT5 ),
2
1
E±α(3) = √ (T6 ± iT7 ).
2
Then we can calculate the commutators to determine the roots α(1) , α(2) , α(3) .
First, for α(1) , the commutator with T3 is just the same as for su(2), so we
get:
[T3 , E±α(1) ] = ±E±α(1) ,
(1)
which means that α1 = 1. Furthermore, T8 commutes with T1 and T2 , and
hence
[T8 , E±α(1) ] = 0,
(1)
that is α2 = 0. Thus the first two roots are ±(1, 0).
For α(2) and α(3) we can calculate the commutator explicitly using the
fundamental representation (Gell-Mann matrices). We get:
√
1 3
[T3 , E±α(2) ] = ± E±α(2) , [T8 , E±α(2) ] = ± E±α(2) ,
2 √2
1 3
[T3 , E±α(3) ] = ± E±α(3) , [T8 , E±α(3) ] = ∓ E (3) .
2 2 ±α
√ √
Thus the remaining four roots are ±( 12 , 3
2 ) and ±( 12 , − 3
2 ). The resulting
root system is shown in Figure 5.9.12.
CHAPTER 5. REPRESENTATION THEORY OF LIE GROUPS 147
Figure 5.9.12 The root system for su(3) (taken from wikipedia).
But our goal here is not to construct representations, but to classify simple
Lie algebras. It turns out that the roots are key for this as well. Indeed,
the roots tell us not only something about the adjoint representation, but in
fact they can be used to reconstruct the whole algebra itself. This is perhaps
expected, since the adjoint representation is defined in terms of the structure
constants, and hence should package the information content of the Lie algebra
itself.
To see how this goes we need a few more definitions.
Definition 5.9.13 Positive and simple root. A positive weight is a
weight of a representation such that the first non-zero component is positive.
A positive root is a positive weight for the adjoint representation. A simple
root is a positive root that cannot be written as a linear combination of other
positive roots with all coefficients being positive. ♦
One can show that knowing the simple roots is in fact sufficient to determine
all roots of a Lie algebra, because of the symmetries of a root system. The
simple roots in fact form a basis for Rm , so, in particular, there are m simple
roots, where m is the rank of the Lie algebra.
Example 5.9.14 Simple roots for su(3).√ In Example 5.9.11 √
we determined
that the roots of su(3) are ±(1, 0), ±( 21 , 23 ) and ±( 21 , − 23 ). The positive
√ √
roots are then (1, 0), ( 12 , 23 ) and ( 12 , − 23 ). Since
√ √
1 3 1 3
(1, 0) = ( , ) + ( ,− ),
2 2 2 2
√
we see that (1, 0) is not a simple root. The other two positive roots, ( 12 , ± 3
2 ),
are the simple roots of su(3).
implies that the only possible angles between root vectors are π/2, 2π/3, 3π/4
and 5π/6. We draw and edge between nodes as follows:
• No edge if the root vectors are perpendicular (π/2);
• A single edge if the angle between the root vectors is 2π/3;