Download as pdf or txt
Download as pdf or txt
You are on page 1of 80

Groups and Algebras

for
Theoretical Physics

Masters course in theoretical physics at


The University of Bern
Spring Term 2016

S USANNE R EFFERT
Contents

Contents

1 Complex semi-simple Lie Algebras 2


1.1 Basic notions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 The CartanWeyl basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 The Killing form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Weights . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.5 Simple roots and the Cartan matrix . . . . . . . . . . . . . . . . . . . . . . . 7
1.6 The Chevalley basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.7 Dynkin diagrams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.8 The Cartan classification for finite-dimensional simple Lie algebras . . . . . 10
1.9 Fundamental weights and Dynkin labels . . . . . . . . . . . . . . . . . . . . 12
1.10 The Weyl group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.11 Normalization convention . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.12 Examples: rank 2 root systems and their symmetries . . . . . . . . . . . . . 17
1.13 Visualizing the root system of higher rank simple Lie algebras . . . . . . . . 19
1.14 Lattices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
1.15 Highest weight representations . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.16 Conjugate representations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
1.17 Remark about real Lie algebras . . . . . . . . . . . . . . . . . . . . . . . . . 27
1.18 Characteristic numbers of simple Lie algebras . . . . . . . . . . . . . . . . . 27
1.19 Relevance for theoretical physics . . . . . . . . . . . . . . . . . . . . . . . . 27

2 Generalizations and extensions: Affine Lie algebras 30


2.1 From simple to affine Lie algebras . . . . . . . . . . . . . . . . . . . . . . . . 30
2.2 The Killing form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.3 Simple roots, the Cartan matrix and Dynkin diagrams . . . . . . . . . . . . . 34
2.4 Classification of the affine Lie algebras . . . . . . . . . . . . . . . . . . . . . 35
2.5 A remark on twisted affine Lie algebras . . . . . . . . . . . . . . . . . . . . . 38
2.6 The Chevalley basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.7 Fundamental weights . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.8 The affine Weyl group . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.9 Outer automorphisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
2.10 Visualizing the root systems of affine Lie algebras . . . . . . . . . . . . . . . 47
2.11 Highest weight representations . . . . . . . . . . . . . . . . . . . . . . . . . 50

3 Advanced topics: Beyond affine Lie algebras 55


3.1 The Virasoro algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
3.2 Lie superalgebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.3 Quantum groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

1 FS2016
Part 1

Complex semi-simple Lie Algebras

Symmetries, and with them, groups and algebras are of paramount importance in theo-
retical physics. The basic concepts have already been introduced in the course Advanced
Concepts in Theoretical Physics. This course will build on the material treated there, with
a special emphasis on techniques that prove to be useful to the theorist. In order to be
self-contained, a certain amount of repetition is however inevitable.
This course consists of three parts. The first part is dedicated to simple Lie algebras,
which are basically a theorists daily bread. The second part treats affine Lie algebras and
the third generalizations beyond the affine case which keep appearing in various contexts
in theoretical physics.1
While the topic is certainly mathematical, treating the structure theory of Lie algebras,
this course is aimed at physicists. Proofs are generally not given and I do not work at the
highest possible level of generality (I do not e.g. work over a general field F).

1.1 Basic notions


In the following, we will be working over the field C, which simplifies many things as it is
algebraically closed.
A Lie algebra g is a vector space equipped with an antisymmetric binary operation

[ , ] : gg g (1.1)

which satisfies the Jacobi identity

[ X, [Y, Z ]] + [ Z, [ X, Y ]] + [Y, [ Z, X ]] = 0, X, Y, Z g. (1.2)

This binary operation is usually referred to as either a commutator or the Lie bracket.
As you know from ACTP, a Lie algebra g describes the Lie group G in the vicinity of the
identity via the exponential map

eiaX G for X g, (1.3)

where a is a parameter.
A representation associates to every element of g a linear operator on a vector space V
which respects the commutation relations of the algebra. The maximal number of linearly
independent states that generate V is the dimension of the representation. Relative to a
1 In retrospect, I would have chosen a different title for the course, since it ended up being centered around

simple Lie algebras and their generalizations while groups have not made an important appearance.

FS2016 2
Part 1. Complex semi-simple Lie Algebras

given basis, each element of g can be represented as a square matrix, and the basis vectors
are represented as column matrices.
A representation is irreducible if the matrices representing the elements of g cannot
all be brought into a block-diagonal form by a change of basis.
A Lie algebra can be specified by a set of generators { J a } and their commutation
relations
[ J a , J b ] = i f cab J c . (1.4)
c

The numbers f cab


are the so-called structure constants of the Lie algebra. The number of
generators is the dimension of the Lie algebra.
A simple Lie algebra is a Lie algebra that contains no proper ideal (no proper subset of
generators { L a } such that [ L a , J b ] { L a } J b ). A semi-simple Lie algebra is a direct sum
of simple Lie algebras. In the following, we will focus only on the simple and semi-simple
cases.

1.2 The CartanWeyl basis


In many cases, we want to work independently from a specific basis. In the case of Lie
algebras, however, choosing a particular basis is most convenient. In the following, we
want to construct the generators of g in the standard CartanWeyl basis. While this choice
is canonical in the case of finite-dimensional semi-simple Lie algebras, for more general
cases, no fully canonical form exists. First we need to find the maximal set of commuting
Hermitian generators H i , i = 1, . . . , r, where r is the rank of the algebra g:

[ H i , H j ] = 0. (1.5)

This set of generators H i forms the Cartan subalgebra h. The generators in h can be
simultaneously diagonalized.
The remaining generators E of g are chosen such that they satisfy

[ H i , E ] = i E . (1.6)

The vector = (1 , . . . , r ) is called a root2 . The E are ladder operators. A basis


satisfying both (1.5) and (1.6) is called a CartanWeyl basis (also standard or canonical
basis).
As h is the maximal Abelian subalgebra of g, the roots are non-degenerate. The root
naturally maps an element H i h to the number i :

( H i ) = i . (1.7)

The roots are therefore elements of the dual of the Cartan subalgebra,

h . (1.8)

With ( E ) = E , we see from hermitian conjugation of Eq. (1.6) that whenever is a


root, so is . We use the notation

= {set of all roots}, (1.9)

also called the root system. The root components i can be regarded as the non-zero
eigenvalues of the H i in the adjoint representation, in which the Lie algebra g itself is
2 The name is chosen as the i are roots of the characteristic equation for H i .

3 FS2016
Part 1. Complex semi-simple Lie Algebras

the vector space on which the generators act. A matrix representation of the adjoint
representation in the basis { J a } is given by
( J a )bc = i f abc . (1.10)
In the adjoint representation, we have
E 7 | E i | i, (1.11)
H i 7 | H i i, (1.12)
identifying the generators and the states of the representation. The action of a generator X
in the adjoint representation is
ad( X )Y = [ X, Y ] (1.13)
so that
ad( H i ) E = i E 7 H i |i = i |i. (1.14)
The one-to-one correspondence between the |i and the reflects the fact that the roots
E
are non-degenerate. In the adjoint representation, the zero eigenvalue has degeneracy r
(associated to the | H i i).
dim(adj. rep.) = dim(algebra) = r + # roots. (1.15)
We now need to specify the remaining commutation relation of the algebra g in the
CartanWeyl basis { H i , E }. From the Jacobi identity, we find
[ H i , [ E , E ]] = (i + i )[ E , E ]. (1.16)
Therefore,
[ E , E ] E if + ,

+

[ E , E ] = 0 / ,
if + (1.17)

[ E , E ] is a linear comb. of H i if = .

We will use in the following the expressions


r r
H = i H i , | |2 = i i . (1.18)
i =1 i =1

The full set of commutation relations of g in the CartanWeyl basis is given by


[ H i , H j ] = 0, (1.19)
[ H i , E ] = i E (1.20)
+ if +

N, E

[ E , E ] = |22 | H if = (1.21)

0 otherwise,

where N, = const.

Example: The CartanWeyl basis for sl (2, C). The basis


     
0 1 1 0 0 0
E= , H= , F= (1.22)
0 0 0 1 1 0
of sl (2, C) has commutation relations
[ H, E] = 2E, [ H, F ] = 2F, [ E, F ] = H. (1.23)
This fulfills the commutation relations of a CartanWeyl basis.

FS2016 4
Part 1. Complex semi-simple Lie Algebras

1.3 The Killing form


In order to define a scalar product on g, we will in the following define the Killing form.
This will allow us, among other things, to fix normalizations. It is defined as follows:
e( X, Y ) = Tr(adX adY ),
K X, Y g. (1.24)

The Killing form obeys X, Y g

Ke( X, Y ) = Ke(Y, X ) symmetry, (1.25)


e([ X, Y ], Z ) = K
K e( X, [Y, Z ]) invariance. (1.26)

For semi-simple Lie algebras, the Killing form is non-degenerate:


e( X, Y ) = 0
K Y X = 0. (1.27)

This is an alternative way of defining semi-simplicity.

Example: The Killing form of sl (2, C). Reusing the basis elements E, H, F in this order-
ing and their commutation relation from the last example, we first express them in the
adjoint representation:

0 2 0

2 0 0 0 0 0
adE = 0 0 1 , adH = 0 0 0 , adF = 1 0 0 . (1.28)
0 0 0 0 0 2 0 2 0

Using these matrices, we can now directly calculate the Killing form on this basis:
e( E, E) = 0,
K e( E, H ) = 0,
K e( E, F ) = 4,
K (1.29)
e( H, H ) = 8,
K e( H, F ) = 0,
K e( F, F ) = 0.
K (1.30)

In matrix form,
0 0 4
e( X, Y ) = 0 8 0 .
K (1.31)
4 0 0
F

Often, a normalized version of the Killing form is used instead:


1
K ( X, Y ) = Tr(adX adY ), (1.32)
2g

where g is the dual Coxeter number. The standard basis { J a } is understood to be orthonor-
mal with respect to K,
K ( J a , J b ) = a,b . (1.33)
The same is true for the generators of the Cartan sub-algebra:

K ( H i , H j ) = i,j . (1.34)

As the Killing form acts as a scalar product, it can be used to raise and lower indices:

f abc = f adc [K( J d , J b )]1 . (1.35)


d

5 FS2016
Part 1. Complex semi-simple Lie Algebras

f abc is anti-symmetric in all three indices. In the orthonormal basis { J a }, the position of the
indices is irrelevant.
From the cyclic property of the trace, it follows that

K ([ Z, X ], Y ) + K ( X, [ Z, Y ]) = 0. (1.36)

The Killing form is uniquely characterized by this property. It also follows that

[ E , E ] = K ( E , E ) H, (1.37)
2
K ( E , E ) = . (1.38)
| |2
The fundamental role of the Killing form is to establish an isomorphism between the Cartan
subalgebra h and its dual h :
K ( H i , ) : h R, (1.39)
for fixed H i . To every element h , there corresponds a H h:

( H i ) = K ( H i , H ). (1.40)

For a root , we have in particular:

H = H = i H i . (1.41)
i

With this isomorphism, we finally also have a positive definite scalar product on h :

(, ) = K ( H , H ). (1.42)

Since roots are elements of h , this defines a scalar product on root space. In particular,

||2 = (, ). (1.43)

1.4 Weights
So far, we have studied the algebra g from the point of view of the adjoint representation
which encodes the essential structure of the algebra. We have seen that in the adjoint,
the eigenvalues of the Cartan generators are called the roots and that the Killing form
induces a scalar product between them. We now study the more general context of a finite
dimensional representation.
For an arbitrary representation, we can always find a basis {|i} such that

H i | i = i | i. (1.44)

The eigenvalues i make up the vector

= ( 1 , . . . , r ) (1.45)

which is called a weight.


( H i ) = i . (1.46)
The scalar product between weights is fixed by the Killing form. In the adjoint repre-
sentation, weights are called roots. From [ H i , E ] = i E , we learn that E changes the
eigenvalue of a state by :

H i E |i = [ H i , E ]|i + E H i |i = (i + i ) E |i. (1.47)

FS2016 6
Part 1. Complex semi-simple Lie Algebras

If E is non-zero, it must be proportional to a state | + i, therefore the name ladder or


step operator for E .
We are mostly interested in finite dimensional representations. For any |i in a finite-
dimensional representation, there are p, q Z+ such that

( E ) p+1 |i E | + pi = 0, (1.48)
( E )q+1 |i E | qi = 0, (1.49)

for any root .


Note that the triplet of generators E , E , H/||2 form an su(2) subalgebra analo-
gous to
{ J + , J , J 3 } : [ J + , J ] = 2J 3 , [ J 3 , J ] = J . (1.50)
Due to this fact, many of the properties of simple Lie algebras can be analyzed to a large
extent by making judicious use of the properties of su(2).
If |i is in a finite-dimensional representation, also its projection onto su(2) is finite-
dimensional.
Let the dimension of the subalgebra su(2) be 2j + 1. From the state |i, the state with
highest J 3 projection m = j can be reached by a finite number, p, of applications of J + . q
applications of J on the other hand lead to the state with m = j:

(, ) (, )
j= + p, j = q. (1.51)
| |2 | |2

Eliminating j leads to
(, )
2 = ( p q). (1.52)
| |2
(,)
Therefore, any weight in a finite-dimensional representation is such that | |2
is an integer.
This is true in particular for = , where is a root.

1.5 Simple roots and the Cartan matrix


As we have seen,
#roots = dim( g) rk( g)  rk( g). (1.53)
The roots are therefore in general linearly dependent.
Thanks to the Euclidean inner product on h , we can have a geometric interpretation of
the roots ("root vector"). We can in particular introduce a hyperplane in root space, which
itself must not contain any roots, and use it to split the root system into positive roots +
which are on one side of the hyperplane, and negative roots on the other side. Note
that this choice of hyperplane and therefore the splitting into positive and negative roots is
not unique and amounts to the choice of a particular basis. Since whenever is a root, also
is a root, we have
= + . (1.54)
We can thus decompose the algebra as

g = h g+ g , (1.55)

where g = spanC { E , > 0} are the subalgebras spanned by the step operators for
positive and negative roots. This is the so-called Gauss or triangle decomposition. A step

7 FS2016
Part 1. Complex semi-simple Lie Algebras

operator E associated to a positive root + is called a raising operator, while E


with + is called a lowering operator.
A simple root i is defined to be a root that cannot be written as the sum of two positive
roots. Geometrically speaking, the simple roots are those positive roots that are closest to
the hyperplane used to separate positive and negative roots, where the notion of the norm
is induced by the Killing form.
There are exactly r = rk( g) simple roots:

s = { 1 , . . . , r }, (1.56)

where the subscript i is a labeling index (not the index that refers to the root component!).
The set of simple roots provides the most convenient basis for root space. Immediate
consequences of the above definition are
/ ,
i j

any positive root is a sum of positive roots.


The basis of simple roots is however in general not orthonormal. This fact is captured in
the Cartan matrix, which is defined as follows:
2( i , j )
Aij = . (1.57)
| j |2

We will see that the Cartan matrix summarizes the structure of a semisimple complex Lie
algebra completely. Using Eq. (1.52), we see that the entries of the Cartan matrix are
necessarily integers. Moreover, Aii = 2 and in general, Aij 6= A ji . The CauchySchwarz
inequality implies that
Aij A ji < 4, i 6= j. (1.58)
Thus, for i 6= j, the entries of the Cartan matrix are negative integers and can be equal
to 0, 1, 2, 3. For Aij 6= 0, at least one of Aij , A ji must be equal to 1. In the set of
roots of a simple Lie algebra at most two different lengths of roots are possible (long and
short). The ratio between the squared lengths of long and short roots can be either two
or three. When all the roots have the same length, the algebra is called simply laced. For
convenience, we introduce the notation
2i
i = . (1.59)
| i |2
i is called the coroot associated to the simple root i . The scalar product between roots
and coroots is always an integer. We can now write the Cartan matrix very compactly as

Aij = (i , j ). (1.60)

The highest root is the unique root for which i mi is maximized in the expansion
i mi i . All elements of can be reached by repeated subtraction of simple roots from .
r r
= ai i = ai i , ai , ai N. (1.61)
i =1 i =1

The coefficients ai are called the marks or Kac labels, while the coefficients ai are called
the comarks or dual Kac labels. Marks and comarks are related via
2
ai = ai . (1.62)
| i |2

FS2016 8
Part 1. Complex semi-simple Lie Algebras

The dual Coxeter number (which we have already used in the definition of the normalized
Killing form Eq. (1.32)) is defined as
r
g= i + 1. (1.63)
i =1

1.6 The Chevalley basis


We will see that the full set of roots can be reconstructed from the simple roots, which in
turn can be straightforwardly extracted from the Cartan matrix. This fact is made manifest
in the so-called Chevalley basis. In this basis, to each simple root correspond the three
generators
2i H
e i = E i , f i = E i , h i = , (1.64)
| i |2
with commutation relations

[hi , h j ] = 0, [hi , e j ] = A ji e j , [hi , f j ] = A ji f j , [ei , f j ] = ij h j . (1.65)

We see that all the structure constants in the Chevalley basis are integers. The remaining
step operators are obtained by repeated commutations of the basic generators, subject to
the Serre relations,

[ad(ei )]1 A ji e j = 0, (1.66)


i 1 A ji j
[ad( f )] f = 0. (1.67)

These constraints encode the rules for reconstructing the full root system from the simple
roots. Finally,
K (hi , h j ) = (i , j ). (1.68)

The fact that the Serre relations do not mix the generators ei and f i corresponds to the fact
that the root system is separated into positive and negative roots. Since the commutation
relations and the Serre relations can be expressed in terms of the Cartan matrix A, we
see that A encodes all the information about the structure of the Lie algebra g. The
abstract formulation of Lie algebras via the Cartan matrix is in fact a good starting point
for generalizations.

1.7 Dynkin diagrams


All the information in the Cartan matrix can be encoded in a planar diagram, the so-called
Dynkin diagram. To each Cartan matrix, a diagram made of vertices and connecting lines
is associated:

Each vertex in the diagram represents a simple root (therefore #vertices = rk( g)).

Long roots are marked by , short roots by .

The vertices i and j are joined by Aij A ji lines, in particular:

Orthogonal roots correspond to disjoint vertices.


Vertices connected by one line correspond to simple roots spanning an angle of
2
3 = 120 .

9 FS2016
Part 1. Complex semi-simple Lie Algebras

Figure 1.1: Root systems of the rank 2 Lie algebras

Vertices connected by two lines correspond to simple roots spanning an angle of


3
4 = 135 .
Vertices connected by three lines correspond to simple roots spanning an angle
of 5
6 = 150 .

Dynkin diagrams in which vertices are joined only by a single line correspond to Lie algebras
in which all simple roots have the same length, hence the name simply laced. Cartan
matrices which differ only by a renumbering of roots lead to the same Dynkin diagram.

Example: Cartan matrices and Dynkin diagrams of rank 2 Lie algebras. This is the
simplest non-trivial example, where we can see the power of the Cartan matrix. For rank
two, the Cartan matrix is a 2 2 matrix. Based on its properties listed after the definition
in Eq. (1.57), we have the following inequivalent possibilities:
       
2 0 2 1 2 2 2 3
(1) : , (2) : , (3) : , (4) : . (1.69)
0 2 1 2 1 2 1 2

They correspond to the following Dynkin diagrams:

(1) : , (2) : , (3) : , (4) : (1.70)

(1) is actually not a simple Lie algebra (the Cartan matrix is block-diagonal), but semi-
simple: it corresponds to A1 A1 . (2) corresponds to A2 , (3) to B2 and (4) to the
exceptional Lie algebra G2 . The root diagrams are shown in Figure 1.1. We will meet these
simple Lie algebras again soon.

1.8 The Cartan classification for finite-dimensional simple Lie


algebras
The enumeration of all possible Cartan matrices is purely combinatorial. It is possible to
classify the finite-dimensional simple Lie algebras completely via their Cartan matrices,
respectively their Dynkin diagrams. There are four infinite series:

Ar (r 1), Br (r 3), Cr (r 2), Dr (r 4), (1.71)

plus five isolated cases:

E6 , E7 , E8 , G2 , F4 , (1.72)

FS2016 10
Part 1. Complex semi-simple Lie Algebras

where the subscript denotes the rank of the group. The algebras in the infinite series are
called the classical Lie algebras. They are isomorphic to the following matrix algebras:

Ar
= sl (r + 1), Br
= so (2r + 1), Cr
= sp(r ), Dr
= so (2r ). (1.73)

The five isolated cases are called the exceptional Lie algebras. The restrictions on the
ranks of the classical Lie algebras are imposed to avoid overcounting. Including all values
of r leads to the following isomorphisms:

A1
= B1
= C1
= D1 , B2
= C2 , D2
= A1 A1 , D3
= A3 . (1.74)

The simple Lie algebras ofthe types Ar , Dr , E6 , E7 and E8 are all simply laced. For Br , Cr
and
F4 , the long roots are 2 times longer than the short roots, while for G2 , the long root
is 3 times longer than the short root.
The Cartan matrices of the simple finite-dimensional Lie algebras are:
Ar :
2 1 0

. . . 0 0
1 2 1 . . . 0 0

0 1 2 1 . . 0 0
A= (1.75)
. . . . . . . .

0 0 0 . . 1 2 1
0 0 0 . . 0 1 2
Br :
2 1 0

. . . 0 0

1 2 1 . . . 0 0

0 1 2 1 . . 0 0
A= (1.76)

. . . . . . . .

0 0 0 . . 1 2 2
0 0 0 . . 0 1 2
Cr :
2 1 0

. . . 0 0

1 2 1 . . . 0 0

0 1 2 1 . . 0 0
A= (1.77)

. . . . . . . .

0 0 0 . . 1 2 1
0 0 0 . . 0 2 2
Dr :
2 1 0

. . . 0 0

1 2 1 . . . 0 0

0 1 2 1 . . 0 0
A=
. . . . . . . .
(1.78)

0 0 0 . . 2 1 1

0 0 0 . . 1 2 0
0 0 0 . . 1 0 2
E6 :
2 1 0

0 0 0

1 2 1 0 0 0

0 1 2 1 0 1
A= (1.79)

0 0 1 2 1 0

0 0 0 1 2 0
0 0 1 0 0 2

11 FS2016
Part 1. Complex semi-simple Lie Algebras

E7 :
2 1 0

0 0 0 0

1 2 1 0 0 0 0

0 1 2 1 0 0 1

A=
0 0 1 2 1 0 0 (1.80)

0 0 0 1 2 1 0
0 0 0 0 1 2 0
0 0 1 0 0 0 2
E8 :
2 1 0 0 0 0 0 0


1 2 1 0 0 0 0 0


0 1 2 1 0 0 0 0

0 0 1 2 1 0 0 0
A= (1.81)

0 0 0 1 2 0 0 1


0 0 0 0 1 2 1 0

0 0 0 0 0 1 2 0
0 0 0 0 1 0 0 2
F4 :
2 1 0

0
1 2 2 0
A=
0 1 2 1 (1.82)
0 0 1 2
G2 :  
2 3
A= (1.83)
1 2
Table 1.1 shows the Dynkin diagrams of the finite-dimensional simple Lie algebras.

1.9 Fundamental weights and Dynkin labels


Weights and roots live in the same r-dimensional vector space. The weights can be expanded
in the basis of simple roots. However, for irreducible finite-dimensional representations,
their coefficients are not integers. A more convenient basis is the dual of the simple coroot
basis. It is denoted by {i } and defined by the relation

(i , j ) = ij . (1.84)

The i are called the fundamental weights and the basis

{i , i = 1, . . . , r } (1.85)

is called the Dynkin basis. The components i of a weight in the Dynkin basis are called
the Dynkin labels,
r
= i i i = (, i ). (1.86)
i =1

The Dynkin labels of weights in finite-dimensional irreducible representations are always


integers. Such weights are called integral. Whenever we write a weight in component

FS2016 12
Part 1. Complex semi-simple Lie Algebras

Ar 1 2 3 r1 r E6
1 2 3 4 5

Br 1 2 r1 r E7
1 2 3 4 5 6

Cr 1 2 r1 r E8
1 2 3 4 5 6 7

Dr r2 F4
1 2 1 2 3 4
r1

G2 1 2

Table 1.1: Dynkin diagrams of the finite-dimensional simple Lie algebras

13 FS2016
Part 1. Complex semi-simple Lie Algebras

form = (1 , . . . , r ), we understand these components to be Dynkin labels. Note that the


elements of the Cartan matrix are the Dynkin labels of the simple roots:

i = Aij j , (1.87)
j

i.e. the ith row of A is the set of Dynkin labels for the simple root i . The Dynkin labels are
the eigenvalues of the Chevalley generators of the Cartan subalgebra:

hi |i = (hi )|i = (, i )|i, (1.88)

that is,
h i | i = i | i. (1.89)
Note the role of the position of the index:

hi : eigenvalue of hi (Dynkin label), i : eigenvalue of H i . (1.90)

A weight of special importance is the one for which all Dynkin labels are equal to one:

= i = (1, . . . , 1). (1.91)


i

This is the Weyl vector or principal vector and is alternatively defined as

= 1
2 . (1.92)
+

The scalar product of weights can be expressed in terms of a symmetric quadratic form
matrix Fij :
(i , j ) = Fij . (1.93)
Fij is the transformation matrix relating the Dynkin basis {i } and the simple coroot basis
{i }:
i = Fij j . (1.94)
j

The relation between the quadratic form and the Cartan matrix is given by

| j |2
Fij = ( A1 )ij . (1.95)
2
The scalar product between the weights = i i and = i i takes the form

(, ) = i j (i , j ) = i j Fij . (1.96)
i,j i,j

1.10 The Weyl group


The root system of a simple Lie algebra has a high degree of symmetry. Also, there are
many equivalent choices for the basis of simple roots. The symmetries of the root system of
a Lie algebra form a group, the automorphism group Aut(). A particularly interesting
subgroup of Aut() is the Weyl group W which we will study in detail in the following.
A distinguished element of W can be inferred from the fact that if is a root, so is :
the mapping
7 , (1.97)
is clearly a symmetry of the root system.

FS2016 14
Part 1. Complex semi-simple Lie Algebras

Example: Automorphisms of A1 . For A1 , there are only two roots, and , each of
which can take the role of the simple root. Therefore, Eq. (1.97) is the only non-trivial map.
For any weight in a finite-dimensional representation of A1 , also is a weight of that
representation. This reflection therefore also leaves the weight system invariant. Together
with 1, Eq. (1.97) forms the group

Z2 = {1}, (1.98)

hence Aut( A1 ) is isomorphic to Z2 .

A map on the set of roots must fulfill the following two conditions in order to qualify as
a symmetry of the root system:

the map must be linear and invertible

the map must be a permutation of the roots.

Therefore, Aut() Sdim gr , the symmetric group with dim g r elements. In general,
the automorphism group is however much smaller than Sdim gr , as arbitrary permutations
cannot be described by a linear map.
It is however not difficult to give elements of W explicitly. Consider the reflection s
with respect to the hyperplane (through 0) in root space perpendicular to a fixed root :

( , )
s = 2 , (1.99)
(, )

i.e. we subtract from each root twice the component in the -direction. This is
indeed a permutation of the roots. The set of all such reflections with respect to roots forms
the Weyl group W. The product of two Weyl reflections is given by the composition of
maps
s s 0 = s 0 s . (1.100)
Since s is a reflection, it is its own inverse, and the unit element of the composition is
the identity map. Note that the composition of reflections leads to reflections as well as
rotations. A generic w W is therefore not of the form (1.99).
By linearity, the action of the Weyl group extends naturally to the weight space of g:

s = ( , ). (1.101)

Since an arbitrary element w W is a product of reflections, it leaves the inner product


invariant:
(s , s ) = (, ). (1.102)
Thus any w W is an isometry.
The Weyl group is generated by a small number of specific reflections of type (1.99),
namely the reflections corresponding to the simple roots. These are the simple or funda-
mental Weyl reflections:
si si , i = 1, . . . , r. (1.103)
Every w W can be written as a word (Weyl word)

w = si s j . . . s k (1.104)

15 FS2016
Part 1. Complex semi-simple Lie Algebras

in the letters si . This decomposition is however not unique. The length l (w) of w is the
minimum number of si among all possible decompositions of w. The signature of w is
defined as
e(w) = (1)l (w) . (1.105)
The si fulfill the relations
s2i = 1, si s j = s j si , if Aij = 0. (1.106)
These relations generalize to
(
2 if i = j
(si s j )mij = 1, where mij = (1.107)

ij if i 6= j,

and ij is the angle between the simple roots i and j . Eq. (1.107) can be used as the
defining relation of the Weyl group. Any group having such a representation is called a
Coxeter group.
On the simple roots, the action of si takes the simple form
si j = j A ji i . (1.108)
We have seen, that W maps into itself. In consequence, it provides a simple way to
generate the full root system from the simple roots by acting with all the elements of W
on s :
= {w1 , w2 , . . . , wr | w W }. (1.109)
This makes it clear that any set {w0 i } for fixed w0 could serve as a basis of simple roots.
The Weyl group induced a natural splitting of the rdimensional weight vector space
into a fan of open cones Cw . These cones, whose number is equal to the order of W, are
defined as
Cw = {|(w, i ) 0, i = 1, . . . , r }, w W (1.110)
and are called Weyl chambers. They intersect only at the reflecting hyperplanes, their
boundaries (w, i ) = 0. One of the chambers (depending on the choice of simple roots) is
distinguished: the unique chamber whose points have only positive Dynkin labels i Z.
This is the fundamental or dominant Weyl chamber C0 (also called fundamental Weyl
domain). A weight in this chamber is said to be dominant. The highest root is an example
of a dominant weight.
W acts transitively and freely on the Weyl chambers. When acting with W on C0
including its boundary, one obtains the whole root space. Conversely, for any weight
/ C0 , there exists a unique w W such that w C0 . The W orbit of a weight is given
by {w|w W }. The W orbit of every weight has exactly one point in C0 .
We can label the Weyl chambers by the elements of W. The fundamental chamber C0
corresponds to the identity element of the Weyl group.

1.11 Normalization convention


Until now, all normalizations have been fixed with respect to the square lengths of the
roots. To fully fix the notation, we need to give a specific value to these lengths. In the
standard convention, the square length of the long roots is set equal to 2. We can fix our
normalization completely by setting
| |2 = 2, (1.111)
since is necessarily a long root.

FS2016 16
Part 1. Complex semi-simple Lie Algebras

1.12 Examples: rank 2 root systems and their symmetries


Example 1: A2 . As we have seen, A2 has Cartan matrix
 
2 1
(1.112)
1 2
and two simple roots 1 , 2 of the same length. Using Eq. (1.87), we find
1 = (2, 1), 2 = (1, 2). (1.113)
Making use of Eq. (1.95), we can write down its quadratic form matrix:
 
1 2 1
F= . (1.114)
3 1 2
The Weyl group of A2 contains for sure the elements 1, s1 and s2 . We use Eq. (1.107) with
12 = 2
3 to find the relation
(s1 s2 )3 = 1. (1.115)
From this, we find s1 s2 s1 = s2 s1 s2 , meaning that no words with more than three elements
can appear in W. The full Weyl group is thus given by
W = {1, s1 , s2 , s1 s2 , s2 s1 , s1 s2 s1 }. (1.116)
Using the Weyl group, we can find the full root system of A2 by acting with the elements of
W on the two simple roots. We find
= { 1 , 2 , 1 + 2 , 1 2 , 1 , 2 }. (1.117)
The highest root is = 1 + 2 . The roots are shown in Figure 1.2. The Weyl chambers
and their labels are indicated in red.

Example 2: C2 . As we have seen, C2 has Cartan matrix


 
2 1
(1.118)
2 2
and two simple roots 1 , 2 of the different length. Using Eq. (1.87), we find
1 = (2, 1), 2 = (2, 2). (1.119)
Making use of Eq. (1.95), we can write down its quadratic form matrix:
 
1 1 1
F= . (1.120)
2 1 2
The Weyl group of A2 contains for sure the elements 1, s1 and s2 . We use Eq. (1.107) with
12 = 3
4 to find the relation
(s1 s2 )4 = 1. (1.121)
From this, we find s1 s2 s1 s2 = s2 s1 s2 s1 . The full Weyl group is thus given by
W = {1, s1 , s1 , s1 s2 , s2 s1 , s1 s2 s1 , s2 s1 s2 , s1 s2 s1 s2 }. (1.122)
This time, we start from the highest root = 21 + 2 and construct the root system by
repeated subtraction of the simple roots. We find
= { 21 + 2 , 1 + 2 , 1 , 2 , 1 , 2 , 1 2 , 21 2 , }. (1.123)
The root system is shown in Figure 1.3. The Weyl chambers and their labels are indicated
in red.

17 FS2016
Part 1. Complex semi-simple Lie Algebras

2 1 1 + 2 =
s1

s1 s2 s2
1 1

s1 s2 s1 s1 s2
1 2 2

Figure 1.2: Root system and Weyl chambers of A2

C2
s1 1
2
1 + 2
21 + 2

s1 s2 s2
1 1

s1 s2 s1 s2 s1
2
21 2 1 2
s1 s2 s1 s2 s2 s1 s2

Figure 1.3: Root system and Weyl chambers of C2

FS2016 18
Part 1. Complex semi-simple Lie Algebras

Figure 1.4: Root systems of the rank 3 Lie algebras

1.13 Visualizing the root system of higher rank simple Lie alge-
bras
The root systems of the rank two cases can be easily visualized via root diagrams in the
plane. Also the rank three cases are still amenable to visualization, see e.g. Fig. 1.4.
Higher dimensional cases however, must make recourse to some form of projection to two
dimensions. The projection of choice for root systems is the one onto the Coxeter plane.
A Coxeter element of a Lie group is the product of all simple Weyl reflections. Changing
the order of the reflections leads to a conjugate Coxeter element. The Coxeter number
h is the number of roots divided by the rank of the algebra. For a given Coxeter element
w, there is a unique plane (the Coxeter plane) on which w acts by rotation by 2/h. The
Coxeter plane is used to draw diagrams of root systems: the vertices and edges roots are
orthogonally projected onto the Coxeter plane, yielding a polygon with h-fold rotational
symmetry. No root maps to zero, so the projections of orbits under w form h-fold circular
arrangements and there is an empty center. The root systems of F4 (h = 12), E6 (h = 12),
E7 (h = 18) and E8 (h = 30) in the Coxeter plane projection are displayed in Figures 1.5 to
1.8 (images sources: Wikimedia Commons).

1.14 Lattices
Let (e1 , . . . , ed ) be a basis of d-dimensional Euclidean space Rd . A lattice is the set of all
points whose expansion coefficients (in terms of the specified basis) are all integers:

Ze1 + Ze2 + + Zed , (1.124)

i.e. the Zspan of {ei }. There are three lattices that are of importance for Lie algebras:

the weight lattice:


P = Z1 + Z2 + + Zr , (1.125)

the root lattice:


Q = Z1 + Z2 + + Zr , (1.126)

the coroot lattice:


Q = Z1 + Z2 + + Zr . (1.127)

19 FS2016
Part 1. Complex semi-simple Lie Algebras

Figure 1.5: Root system of F4 in the Coxeter plane projection with 12-fold symmetry.

Figure 1.6: Root system of E6 in the Coxeter plane projection with 12-fold symmetry.

FS2016 20
Part 1. Complex semi-simple Lie Algebras

Figure 1.7: Root system of E7 in the Coxeter plane projection with 18-fold symmetry.

Figure 1.8: Root system of E8 in the Coxeter plane projection with 30-fold symmetry.

21 FS2016
Part 1. Complex semi-simple Lie Algebras

s1

2
1
s1 s2 s2
1

s1 s2 s1 s2 s1

Figure 1.9: Root and weight lattices of A2 .

Weights in finite-dimensional representations have integer Dynkin labels, hence they belong
to P. The integers specifying the position of a weight in P are the eigenvalues of the Cheval-
ley generators hi . The effect of the remaining generators of g is to shift the eigenvalues by
an element of Q, the root lattice. Since roots are weights in a particular finite-dimensional
representation, Q P. This means that by acting with E , a point in P is translated to
another point in P.
For the algebras G2 , F4 and E8 , we have Q = P. In all other cases, Q is a proper subset
of P and the quotient P/Q is a finite group. Its order | P/Q| is equal to det A.
The distinct elements of the coset P/Q define the so-called congruence classes or
conjugacy classes. A weight lies in exactly one congruence class. For any algebra g, the
congruence classes take the form
r
= i i mod | P/Q| (mod Z2 for g = D2l ), (1.128)
i =1

where the vector (1 , . . . , r ) is called the congruence vector.


Since the bases {i } and {i } are dual, P and Q are dual lattices. A lattice is said to
be self-dual if it is equal to its dual. For simple Lie algebras, only E8 has a self-dual weight
lattice.

1.15 Highest weight representations


Lie algebras play their role in physics not as abstract algebras, but through their represen-
tations, which act on suitable representation spaces. The highest weight representations

FS2016 22
Part 1. Complex semi-simple Lie Algebras

are a particularly interesting subclass. any finite-dimensional representation of a simple


Lie algebra belongs to this class. As we will see, the key idea for the analysis of these
representations is to reduce the problem to the representation theory of sl (2) which we
have studied in Sec. 1.4. There we have seen that to each (simple) root, an sl (2) subalgebra
spanned by Ei , Ei , H i belongs. It follows that any representation space has a basis on
which the whole Cartan subalgebra h acts diagonally.
Any finite-dimensional irreducible representation has a unique highest weight state
|i, which is completely specified by its Dynkin labels (hi ) = i . Among all the weights
in the representation, the highest weight is the one for which the sum of its coefficients is
maximal when expanded in simple roots. Thus for any > 0, + cannot be a weight in
the representation, so that
E |i = 0, > 0. (1.129)
From Eq. (1.52) (in this case, p = 0), it is clear that the highest weight of a finite-
dimensional representation is necessarily dominant. An irreducible finite-dimensional
representation space of a semi-simple Lie algebra is characterized by the fact that it has a
highest weight of multiplicity one and that this weight is dominant integral. Conversely,
each dominant integral weight is the highest weight of a unique irreducible finite-
dimensional representation L . By abuse of notation, representations are often specified by
their highest weight. The highest weight of the adjoint representation is .

Weights and multiplicities. Starting from the highest-weight state |i, all states in the
representation space L can be obtained by the action
E E . . . E | i for , , . . . , + . (1.130)
The set of eigenvalues of all states in L is the weight system . Any 0 is such
that
0 + . (1.131)
All the weights of a given representation lie therefore in exactly one congruence class.
For the weight 0 = ni i , ni Z+ , we call ni the level or depth. Whenever two
weights , satisfy = with of the form = ni i , we say . This provides
a partial ordering of . For any irreducible highest weight representation L , there is a
unique weight of depth 0, namely the highest weight .
In order to find all the weights 0 , we use again the representation theory of
sl (2), rewriting Eq. (1.52) as
(0 , i ) = i0 = ( pi qi ), p i , q i Z+ . (1.132)
0 is necessarily of the form
0 = ni i , n i Z+ . (1.133)
We now proceed level by level, using a simple algorithm which stems from the fact that
all weights of a finite-dimensional irreducible sl (2)-representation are obtained from the
highest weight by subtracting the positive root :
From the highest weight = i i , all weights of level 1 are obtained by subtracting
those simple roots i for which i > 0.
More generally, for an arbitrary weight 0 of L , when i0 > 0, we can subtract the
simple root i i0 times from 0 , producing weights that are all part of .
Proceeding recursively, we can produce in this way all distinct weights of L until
we arrive at a depth where all weights have negative Dynkin labels and the process
terminates.

23 FS2016
Part 1. Complex semi-simple Lie Algebras

(1, 1)
1 2

( 1, 2) (2, 1)

2 1
(0, 0)
1 2

( 2, 1) (1, 2)

2 1
( 1, 1)
Figure 1.10: Weight diagram for the adjoint representation of A2 .

Example: the adjoint representation of A2 . The highest weight of the adjoint represen-
tation of A2 corresponds to the highest root, = (1, 1). As we have learned before, the
two simple roots are 1 = (2, 1) and 2 = (1, 2). The weight diagram resulting from
the above algorithm is given in Figure 1.10.

Example: the adjoint representation of C2 . The highest weight of the adjoint represen-
tation of C2 corresponds again to the highest root, = 21 + 2 . So = (2, 0). The two
simple roots are 1 = (2, 1) and 2 = (2, 2). The weight diagram resulting from the
above algorithm is given in Figure 1.11.

Example: a representation of D5 . Let us study a representation of D5 with highest


weight (1, 0, 0, 0, 0). The roots can be read off from the Cartan matrix given in Eq. (1.78).
The weight diagram resulting from the above algorithm is given in Figure 1.12.

The above examples show that the multiplicity of a (non-highest) weight might be
larger than one. In general, the "quantum numbers" i do not characterize a weight vector
completely. A complete specification requires

1
(dim g r ) = |+ | (1.134)
2

labels, i.e. 12 (dim g 3r ) quantum numbers in addition to the Dynkin labels. To compute
these multiplicities, we can use the Freudenthal recursion formula, which gives the

FS2016 24
Part 1. Complex semi-simple Lie Algebras

(2, 0)
1

(0, 1)
1 2

( 2, 2) (2, 1)

2 1
(0, 0)
1 2

( 2, 1) (2, 2)

2 1
(0, 1)
1

( 2, 0)

Figure 1.11: Weight diagram for the adjoint representation of C2 .

(1, 0, 0, 0, 0)
1
( 1, 1, 0, 0, 0)
2
(0, 1, 1, 0, 0)
3
(0, 0, 1, 1, 1)
4 5
(0, 0, 0, 1, 1) (0, 0, 0, 1, 1)
5 4
(0, 0, 1, 1, 1)
3
(0, 1, 1, 0, 0)
2
(1, 1, 0, 0, 0)
1
( 1, 0, 0, 0, 0)

Figure 1.12: Weight diagram for the representation of D5 with highest weight (1, 0, 0, 0, 0).

25 FS2016
Part 1. Complex semi-simple Lie Algebras

multiplicities of the weight 0 in the representation in terms of all the weights above it:

| + |2 |0 + |2 mult (0 ) = 2 (0 + k, )mult (0 + k).
 
(1.135)
>0 k =1

Example: multiplicity of the weight (0, 0) in the adjoint representation of A2 . We


apply the recursion formula (1.135) to compute the multiplicity of (0, 0) from the last
example. We know that k = 1 and that the multiplicities of all three weights above are each
one. For the three positive roots, (0 + k, ) = 2. Using = = = 1 + 2 , we find

(8 2)mult (0, 0) = 2(2 + 2 + 2), (1.136)

therefore mult (0, 0) = 2.

Generally, the weights of the adjoint representation of any semi-simple Lie algebra g
are the groots, each occurring with multiplicity one, and in additions the weight = 0
with multiplicity r.
Note that all weights in a given Worbit have the same multiplicity:

mult (w0 ) = mult (0 ) w W. (1.137)

Finally, note that a finite-dimensional irreducible representation L is always unitary.

( H i ) = H i , ( E ) = E (1.138)

results in the norm of any state |i L being positive definite.

1.16 Conjugate representations


In a finite-dimensional irreducible representation, there is obviously also a unique lowest
weight state. It lies in the Worbit of the highest weight , in the Weyl chamber exactly
opposite to C0 . This chamber corresponds to the longest element w0 of W. The lowest
weight state is thus given by w0 . The conjugate representation has for its highest
weight the negative of the lowest weight state of L :

= (w0 ). (1.139)

The Weyl vector is the highest weight of a self-conjugate representation:

= (w0 ). (1.140)

More generally, all the weights in are the negatives of those in .


The conjugation is related to the reflection symmetries of the Dynkin diagram. For Ar ,
the conjugation amounts to reversing the order of the finite Dynkin labels. A representation
is self-conjugate if and only if its weight system is invariant under the change of sign.
As for any root , also is a root, the adjoint representation of any Lie algebra is self-
conjugate. The Dynkin diagrams with reflection symmetry are Ar , Dr and E6 . The operation
of conjugation corresponds to this reflection, which is an automorphism of weight space
(the nodes of the Dynkin diagrams corresponding to the fundamental weights). In order for

FS2016 26
Part 1. Complex semi-simple Lie Algebras

complex algebra compact form real form


sl (n, C) su(n) sl (n, R)
so (2l + 1, C) so (2l + 1) so (l + 1, l )
sp(n, C) sp(n, 0) sp(n, R)
so (2l, C) so (2l ) so (l, l )

Table 1.2: Compact and normal real forms for the classical Lie algebras

representations of these three cases to be self-conjugate, the Dynkin labels of the highest
weights must be invariant under the reflections, i.e. they have to satisfy
A r : i = r i +1 , (1.141)
Dr (r even) : r = r1 , (1.142)
E6 : 1 = 5 , 2 = 3 . (1.143)

1.17 Remark about real Lie algebras


So far, we have always assumed that the base field of g is C, making use of the fact that
C is algebraically closed. This is not the case for R. This means in particular that the
eigenvalue equation (1.6) which determines the roots and therefore encodes the abstract
structure of a Lie algebra need not have a solution in R. The theory of real Lie algebras is
therefore much more involved. It is however possible to construct real forms of the complex
Lie algebras we have studied. A simple Lie algebra has several non-isomorphic real forms.
There are two standard real forms one can construct for any complex simple Lie algebra,
based on the observation that in a CartanWeyl basis, all structure constants are real. As a
consequence, the real vector space spanned by all real linear combinations

i H i + E (1.144)
i

is a real Lie algebra. This Lie algebra is the normal or split real form. It is the least compact
real form.
On the other hand, the compact real form is spanned with real coefficients by {iH i }
for the Cartan subalgebra and
q q
{(i (, )/2)( E + E ) ( (, )/2)( E E )} (1.145)

for its orthogonal complement. All other real forms can be obtained from the compact
form by multiplying suitable generators by a factor of i. Table 1.2 collects the compact and
normal real forms for the classical Lie algebras.

1.18 Characteristic numbers of simple Lie algebras


In this section, we collect the characteristic numbers of the simple Lie algebras we have
encountered, see Tables 1.3 and 1.4.

1.19 Relevance for theoretical physics


In the first part of this course, we have studied the structure of abstract simple Lie algebras.
In physics, Lie algebras and in particular their representations are omnipresent. In particle

27 FS2016
Part 1. Complex semi-simple Lie Algebras

Ar Br Cr Dr
dimension dim( g) r2 + 2r 2r2 +r 2r2 +r 2r2 r
dual Coxeter number g r+1 2r 1 r+1 2r 2
order of Weyl group |W | (r + 1) ! 2r r! 2r r! 2( r 1)r!
highest root (1, 0, . . . , 1) (0, 1, . . . , 0) (2, 0, . . . , 0) (0, 1, . . . , 0)
weight/root lattice P/Q Zr + 1 Z2 Z2 Z4 r odd,
Z2 Z2 r even
congruence vector (1, 2, . . . , r ) (0, . . . , 0, 1) (1, 2, . . . , r ) (2, 4, . . . , r 1, r ) r odd
(0, . . . , 0, 1) r odd

Table 1.3: Characteristic numbers of the classical Lie algebras

E6 E7 E8 F4 G2
dimension dim( g) 78 133 248 52 14
dual Coxeter number g 12 18 30 9 4
order of Weyl group |W | 51840 2903040 696729600 1152 12
highest root (0, . . . , 0, 1) (1, 0, . . . , 0) (1, 0, . . . , 0) (1, 0, 0, 0) (1, 0)
weight/root lattice P/Q Z3 Z2 1 1 1
congruence vector (1, 2, 0, 1, 2, 0) (0, 0, 0, 1, 0, 1, 1) - - -

Table 1.4: Characteristic numbers of the exceptional Lie algebras

physics, we mostly deal with the A-series (e.g. when studying the gauge groups of the
standard model, SU (3) SU (2) U (1)). For spin representations, orthogonal and special
orthogonal Lie algebras play a role.
Also in integrable systems such as spin chains, Lie algebras are of great importance.
In the simplest model, the XXX spin chain, each lattice point carries a representation of
su(2) (corresponding to a spin pointing either up or down), but in more general models,
representations of any simple Lie algebra are admitted.
In modern theoretical physics beyond the standard model, the whole machinery of
simple Lie algebras is of paramount importance. Lie algebras show up in many contexts,
such as grand unified gauge groups (e.g. SU (5), SO(10), SU (8), E6 , O(16)), and global
symmetries. In string theory, gauge groups of the A and Dseries can be created via
Dbrane constructions in type II superstring theory, whereas the heterotic string has gauge
group SO(32) or E8 E8 .

In quantum systems, however, classical symmetries do not carry over directly, and the
concept of the central extension becomes necessary. Other generalizations of Lie algebras
that lead to infinite-dimensional algebras also appear naturally in physical contexts. In
integrable models, super algebras and infinite dimensional Yangian algebras appear. The
second part of this course will therefore be dedicated to extensions and generalizations of
simple Lie algebras.

Literature
This part of the course presents material that is contained both in [DMS97] and [FS97].
The notation of [DMS97] is used throughout, but some sections follow more closely the
exposition of [FS97]. [DMS97] is sometimes a little terse, while [FS97] takes a more
mathematical approach.

FS2016 28
Part 1. Complex semi-simple Lie Algebras

References
[DMS97] P. Di Francesco, P. Mathieu, and D. Snchal. Conformal field theory. Graduate texts
in contemporary physics. New York: Springer, 1997. ISBN: 0-387-94785-X. URL: http:
//opac.inria.fr/record=b1119694.
[FS97] J. Fuchs and C. Schweigert. Symmetries, Lie Algebras and Representations: A Graduate
Course for Physicists. Cambridge Monographs on Mathematical Physics. Cambridge:
Cambridge University Press, 1997. ISBN: 978-0521541190.

29 FS2016
Part 2

Generalizations and extensions:


Affine Lie algebras

When considering quantum systems in physics, we often reach the limits of applicability
of (semi-)simple Lie algebras. While at the level of classical mechanics or field theory,
the symmetries of a physical systems are described by a Lie algebra g, in the quantum
description of the same system, the Lie brackets are not recovered completely. Additional
terms appear in the commutation relations (in physics, this phenomenon is referred to
as a quantum anomaly). In order to describe the quantum theory in Lie algebraic terms,
we must interpret these new constant terms as the eigenvalues of some new operators
which have constant eigenvalues on any irreducible representation space of g. These new
operators extend g to a closely related algebra g, via a so-called central extension. Not every
given Lie algebra admits however a non-trivial central extension. The (semi-)simple Lie
algebras we have studied so far, in particular, do not.
Our aim in this part of the course is to construct (untwisted) affine Lie algebras g.
In order to do so, we need to extend our familiar simple Lie algebras g to an infinite-
dimensional loop algebra which in turn receives a central extension and is supplemented
by a derivation. We will associate to each finite-dimensional g an affine extension g by
adding an extra node related to the highest root to the Dynkin diagram of g. The
introduction of this extra simple root will make the root system and the Weyl group of g
infinite dimensional. Also the highest weight representations become infinite dimensional,
however, they can be organized in terms of a new parameter, the level. The discussion
in Part 2 of this course will follow the one of Part 1 very closely, pointing out important
differences to the finite-dimensional case.

2.1 From simple to affine Lie algebras


Let us consider a generalization of g in which the elements of the algebra are also Laurent
polynomials in some variable t. The set of these polynomials is denoted by C[t, t1 ]. The
generalization
g = g C[t, t1 ] (2.1)

is called the loop algebra g. By a loop in a topological space M, one means the smooth
embedding of a circle into M, together with a chosen parametrization. The loop algebra
associated to g consists of the space of analytic mappings from the circle S1 to g via the
map
t = ei , R. (2.2)

FS2016 30
Part 2. Generalizations and extensions: Affine Lie algebras

The generators of g are given by J a tn . The algebra multiplication rule is the natural
extension from g to g:

[ J a tn , J b tm ] = i f abc J c tn+m . (2.3)


c

A central extension is obtained by adjoining to g a central element:

[ J a tn , J b tm ] = i f abc J c tn+m + k n K ( J a , J b )n+m,0 , (2.4)


c

where k commutes with all J a s and K is the Killing form of g. Assuming again the generators
J a to be orthonormal with respect to the Killing form and using the notation

Jna J a tn , (2.5)

we can rewrite the commutation relations as

[ Jna , Jmb ] = i f abc Jnc +m + k n a,b n+m,0 , (2.6)


c

These relations must be supplemented by

[ Jna , k] = 0. (2.7)

While seemingly ad hoc, the central extension is actually unique, as we will see in the
following. Let us start with the generic cocommutatormutator
l
[ Jna , Jmb ] = i f abc Jnc +m + ki (diab )nm , (2.8)
c i =1

containing l central terms. Except for when n + m = 0, the central terms can be eliminated
by a redefinition of the generators. So

[ J0a , Jnb ] = i f abc Jnc , (2.9)


c

i.e. the generators { Jna } transform in the adjoint representation of of g (ad( J0a )). Since
the central extensions commute with all the generators { Jna }, they are invariant tensors of
the adjoint representation. There is, however only one (up to normalization) such tensor,
namely the Killing form itself. Therefore, only one central element can be added to the
loop extension of a simple Lie algebra. The only central extension of a simple Lie algebra
compatible with the antisymmetry of the commutators and the Jacobi identity is indeed
the one of Eq. (2.6).
As in the simple case, we want to rewrite everything in the (affine) CartanWeyl basis.
With the non-zero Killing norms
2
K ( H i , H j ) = ij , K ( E , E ) = , (2.10)
| |2
the commutation relations become
j
[ Hni , Hm ] = k n i,j n+m,0 , (2.11)
[ Hni , Em

] = i En+m (2.12)
+

N,En+m  if +


2
[ En , Em ] = |2 | Hn+m + k n n+m,0

if = (2.13)


0 otherwise.

31 FS2016
Part 2. Generalizations and extensions: Affine Lie algebras

The set of generators { H01 , . . . , H0r , k } is manifestly Abelian. In the adjoint representation,
the eigenvalues of ad( H0i ) and ad(k ) on the generators E are, respectively, i and 0. Being
independent of n, the eigenvector (1 , . . . , r , 0) is the same for all Em , m = 0, . . . , , i.e.

it is infinitely degenerate. { H01 , . . . , H0r , k } is therefore not a maximal Abelian subalgebra.


It must be augmented by a new grading operator L0 , whose eigenvalues in the adjoint
representation depend on n:
d
L0 = t . (2.14)
dt
In the mathematics literature, usually the operator D = L0 is used instead, it is called a
derivation. The action of L0 on the generators is

ad( L0 ) J a tn = [ L0 , J a tn ] = nJ a tn [ L0 , Jna ] = n Jna . (2.15)

The element L0 measures therefore the mode number n of the generators Jna , i.e. the
degrees with respect to the gradation are given by the eigenvalues of L0 . The maximal
Cartan subalgebra is generated by

{ H01 , . . . , H0r , k, L0 }. (2.16)

The other generators, En for any n and Hni for n 6= 0, play the role of ladder operators.
With the addition of L0 , the resulting algebra is denoted as g:

g = g Ck CL0 . (2.17)

It is referred to as an untwisted affine Lie algebra.


The addition of L0 has made the Killing form on g non-degenerate and allows thus for a
non-degenerate inner product on g. Having an infinite number of generators { Jna }, n Z,
it is clearly an infinite-dimensional algebra. From the perspective of g, g is referred to as
the corresponding finite algebra, generated by the zero modes { J0a }.
In the physics literature, affine Lie algebras are often referred to as KacMoody algebras.
However the name KacMoody refers to a more general construction.

Example: the Heisenberg algebra. An already familiar infinite-dimensional example is


the algebra generated by the modes of a free boson:

[ an , am ] = n n+m,0 . (2.18)

This is the so-called Heisenberg algebra. It is the affine extension of the u(1) algebra
generated by the element a0 . Comparing to the commutation relations Eq. (2.6), the level
appears to be 1, however, the central terms can be changed arbitrarily by rescaling the
modes. For the case u(1), the level has no meaning.

2.2 The Killing form


In parallel to Part 1 of this course, we want to equip g with an inner product. This means
that we need to extend the Killing form from g to g. We use again the identity Eq. (1.36).
With X, Y { Jna }, Z = L0 , we get

K ( Jna , Jm
b
) = 0, unless n + m = 0. (2.19)

FS2016 32
Part 2. Generalizations and extensions: Affine Lie algebras

For n + m = 0, the t-factors disappear, leaving us with the Killing form of g, implying

K ( Jna , Jm
b
) = a,b n+m,0 . (2.20)

Note that the affine Killing form is still orthonormal with respect to the finite algebra
indices. The choice X, Z { Jna }, Y = k yields

K ( Jna , k ) = 0, K (k, k ) = 0, (2.21)

whereas Y = L0 leads to
K ( Jna , L0 ) = 0, K ( L0 , k ) = 1. (2.22)
The only unspecified norm is K ( L0 , L0 ), which by convention is chosen to be

K ( L0 , L0 ) = 0. (2.23)

This arbitrariness is related to the possibility of redefining

L0 L0 + ak, a constant, (2.24)

without affecting the algebra. This redefinition of L0 changes its Killing norm by 2a.
As in the finite case, the Killing form leads to an isomorphism between the elements of
the Cartan subalgebra and those of its dual and defines a scalar product for the latter. Take
a state that is a simultaneous eigenvector of all the generators of the Cartan subalgebra.
The components of the vector are given by the eigenvalues of this state:

= (( H01 ), ( H02 ), . . . , ( H0r ); (k ); ( L0 )). (2.25)

= (; k ; n ) is called an affine weight. The scalar product induced by the Killing form
is
(, ) = (, ) + k n + k n . (2.26)
As for the simple Lie algebras, affine weights in the adjoint representation are called affine
roots.
Since k commutes with all the generators of g, its eigenvalues on the states of the
adjoint representation are 0. Affine roots are therefore of the form

= ( ; 0; n). (2.27)

The scalar product between affine roots is therefore the same as of their simple counterparts,
( , ) = ( , ). The affine root associated to the generator En is

= (; 0; n), n Z, . (2.28)

If we define
= (0; 0; 1), (2.29)
then n is the root associated to Hni . In the following, we use the notation

(; 0; 0), (2.30)

so that we can write the roots in Eq. (2.28) as

= + n. (2.31)

The full set of roots is


= { + n | n Z, } {n | n Z, n 6= 0}.
(2.32)

33 FS2016
Part 2. Generalizations and extensions: Affine Lie algebras

The root is unusual as it has zero length,

(, ) = 0. (2.33)

For this reason, it is called an imaginary root. All the roots in {n} are imaginary,

(n, m) = 0, n, m. (2.34)

All the imaginary roots have multiplicity r. The other roots have multiplicity one and are
called real.

2.3 Simple roots, the Cartan matrix and Dynkin diagrams


As the next step, we want to identify a basis of simple roots for the affine Lie algebra g.
This corresponds again to splitting the root space into positive and negative roots. In such a
basis, the expansion coefficients of any root are either all positive or all negative. The basis
must contain r + 1 elements, r of which are necessarily the finite simple roots i , while the
remaining simple root must be a linear combination involving . The proper choice for the
latter is
0 (; 0; 1) = + . (2.35)
The basis of simple roots is thus given by

{ i }, i = 0, . . . , r. (2.36)

The set of positive roots is given by


+ = { + n | n > 0, } { | + }.
(2.37)

This makes sense as + n = + n0 + n = n0 + (n 1) + ( + ) and the expansion


coefficients in terms of finite simple roots of the last two factors are necessarily non-negative.
Note, that in the affine case, there is no highest root. This implies also that the adjoint
representation is not a highest weight representation.

Example: the root system of A1 . The algebra A1 has exactly two simple roots, namely
the simple root = of the simple algebra A1 and 0 = + = + . Its root system
is given by
= { + n | n Z}. (2.38)

Given a set of simple roots and a scalar product, we can now define the extended
Cartan matrix as
Abij = (i , ), 0 i, j r, (2.39)
j

where the affine coroots are given by


2 2 2
= 2
(; 0; n) = 2
(; 0; n) = ( ; 0; 2 n). (2.40)
|| || ||
As for simple roots, we omit the hat for simple coroots, such that

0 = 0 , i = (i ; 0; 0), i 6= 0. (2.41)

FS2016 34
Part 2. Generalizations and extensions: Affine Lie algebras

Compared to the finite Cartan matrix, A bij contains an extra row and column. The extra
entries are easily calculated in terms of the marks defined in Eq. (1.61). We have

(0 , 0 ) = | |2 = 2 (2.42)

and
r
(0 , j ) = (, j ) = ai (i , j ). (2.43)
i =0

The zeroth mark is defined to be a0 = 1. Since the finite part of 0 is a long root, the zeroth
comark is also 1:
| 0 |2
a0 = = 1. (2.44)
2
By construction, the extended Cartan matrix satisfies
r r
ai Abij = Abij aj = 0. (2.45)
i =0 i =0

The dual Coxeter number is given by


r
g= ai . (2.46)
i =0

Again, all the information contained in the extended Cartan matrices can be encoded in
extended Dynkin diagrams. The Dynkin diagram of g is obtained from that of g by the
addition of an extra node associated to 0 . This node is linked to the i -nodes by A
b0i A
bi0
lines.

2.4 Classification of the affine Lie algebras


Just as the simple Lie algebras, affine Lie algebras can be classified completely via their
Cartan matrices. The linear dependence between the rows of the extended Cartan matrix
means that it has one zero eigenvalue, a reflection of the semi-positive nature of the affine
scalar product. The rank of an (r + 1) (r + 1) affine Cartan matrix is thus r. As we will
see, the classification of the extended Cartan matrices relies directly on the classification of
the simple Lie algebras.

Example: The affine Lie algebras with r = 1. For r = 1, the extended Cartan matrices
are 2 2 matrices and the requirement of det A
b = 0 is enough to determine the off-diagonal
elements:
A b22 A
b11 A b21 = 4 A
b12 A b21 = 0.
b12 A (2.47)
This leads directly to the two possibilities
   
2 2 2 4
A=
b , b=
A . (2.48)
2 2 1 2

The corresponding Dynkin diagrams are

0 1, 0 1, (2.49)

35 FS2016
Part 2. Generalizations and extensions: Affine Lie algebras

so differently to the simple case, two nodes can be joined by four lines.

For higher rank cases, we start from the observation that by deleting the ith row and
column for any i {0, . . . , r } from an extended Cartan matrix, we obtain a Cartan matrix
of a (semi)-simple Lie algebra. By deleting further rows and columns, we always remain in
the class of semi-simple Lie algebras. In particular, any 2 2 matrix obtained by deleting
r 1 rows and their corresponding columns from an extended Cartan matrix A b must be
one of the rank 2 Cartan matrices we have studied. This restricts the matrix elements for
i 6= j to
A bji = 0
bij A or min{| A
bij |, | A
bji |} = 1, max{| A
bij |, | A
bji |} 3. (2.50)

We see that the rank 1 cases with A b21 = 4 are unusual.


b12 A
Implementing the above constraints, the enumeration of all possible affine Cartan
matrices of a given rank can be done by straightforward combinatorics.

Example: r = 2 with the submatrix of G2 . For illustration, we study the rank 2 case
with a 2 2 submatrix corresponding to the Cartan matrix of G2 :

2 p q
b = r 2 3 .
A (2.51)
s 1 2
The determinant is given by
b = 2 p(2r + 3s) q(r + 2s).
det A (2.52)
We find that det Ab = 0 is only fulfilled for two combinations of the integers p, q, r, s,
namely
{ p = r = 1, q = s = 0} and { p = r = 0, q = s = 1}. (2.53)
(1)
The affine Lie algebras corresponding to these two solutions are denoted G2 = G2 and
(3)
G2 .

Tables 2.1 and 2.2 show all the possible extended Dynkin diagrams resulting from this
classification. The Dynkin diagrams in Table 2.1 are all untwisted affine Lie algebras, also
(1) (1)
denoted by Ar , Br , etc. In the examples we have however seen, that we also produce
additional Cartan matrices. These additional Dynkin diagram are given in Table 2.2 and
belong to the class of twisted affine Lie algebras which we have not discussed so far. For
the latter class, a variety of naming conventions exist in the literature.
In total, we have seven infinite series of affine Lie algebras
(1) (1) (1) (1)
Ar (r 2), Br (r 3), Cr (r 2), Dr (r 4), (2.54)
(2) (2) (2)
Br (r 3), Cr (r 2), Br (r 0), (2.55)
and nine exceptional cases,
(1) (1) (1) (1) (1) (1)
A1 , E6 , E7 , E8 , F4 , G2 , (2.56)
(2) (2) (3)
A1 , F4 , G2 . (2.57)

FS2016 36
Part 2. Generalizations and extensions: Affine Lie algebras

0 6
Ar E6
1 2 3 r1 r 1 2 3 4 5

7
0

Br E7
2 3 r1 r 0 1 2 3 4 5 6
1

Cr 0 1 2 r1 r E8
0 1 2 3 4 5 6 7

0 r

Dr r2 F4
2 3 0 1 2 3 4
1 r1

A1 0 1 G2
0 1 2

Table 2.1: Dynkin diagrams of the untwisted affine Lie algebras

37 FS2016
Part 2. Generalizations and extensions: Affine Lie algebras

(2) (2) (2) 0 1 2


A1 0 1 Br , Dr+1

(2) (2) (2) (2)


Br , A2r 0 1 2 Cr , A2r1 2 3
1

(2) (2) (3) (3)


F4 , E6 G2 , D4
1 2 3 4 0 1 2 0

Table 2.2: Dynkin diagrams of the twisted affine Lie algebras

(1) (1)
For many purposes, it is natural to regard A1 as the first element of the series Ar and
(2) (2) (2)
A1 as the first element of Br .
Note, that in three lengths of simple roots occur,
Br
necessitating a new notation for the Dynkin diagram: we have, in decreasing length, the
types of nodes
. (2.58)

Inspecting Tables 2.1 and 2.2, we find that indeed, by removing any node from an extended
Dynkin diagram, we get the Dynkin diagram of a simple Lie algebra.

2.5 A remark on twisted affine Lie algebras


We have only constructed the untwisted affine Lie algebras so far, which is where our main
interest lies. Yet we have seen that the twisted cases arise naturally in the classification
of affine Lie algebras via their Cartan matrices. The twisted cases can be constructed in a
similar way to the untwisted ones, by giving up the requirement of single-valuedness of the
map from the circle S1 to g in the construction of the loop algebra, and instead imposing
twisted boundary conditions. We shall however not do this in this course.

2.6 The Chevalley basis


The commutation relations of the generators of the Chevalley basis given in Eq. (1.65) have
the following affine extension:

j 4
[hin , hm ] = (i , j )knij n+m,0 = knij n+m,0 , (2.59)
| i |2
j j
[hin , em ] = A ji en+m , (2.60)
j j
[hin , f m ] = A ji f n+m , (2.61)
j j 2
[ein , f m ] = ij hn+m + knij n+m,0 , i, j = 1, . . . , r. (2.62)
| i |2

FS2016 38
Part 2. Generalizations and extensions: Affine Lie algebras

These relation however do not involve only the generators of the r + 1 simple roots of g
and are not expressed in terms of the Cartan matrix of g. In order to construct a genuine
Chevalley basis, we need to add the generators

e0 = E1 , f 0 = E

1, h0 = k H0 (2.63)

to the set of finite generators ei and f i . e0 and f 0 are the raising and lowering operators for
0 .
From now on, we will omit the mode index 0 from the finite g Chevalley generators.
The commutation relations for the generators associated to the simple roots of g can be
written as

[hi , h j ] = 0, (2.64)
bji enj +m ,
[ hi , e j ] = A (2.65)
j
[ hi , f j ] = A
bji f n+m , (2.66)
[ei , f j ] = ij h j , i, j = 0, . . . , r. (2.67)

These commutation relations have to be supplemented by the affine Serre relations

[ad(ei )]1 A ji e j = 0, (2.68)


b

1 A
[ad( f i )]
bji j
f = 0, i 6= j. (2.69)

In this form, it is manifest that A


b encodes the whole structure of g. Its infinite-dimensional
nature is however not apparent.

Example: the Serre relations and roots of A1 . Using the Cartan matrix of A1 , we find
1Ab01 = 1 A
b10 = 3, so

[ad(e0 )]3 e1 = [e0 , [e0 , [e0 , e1 ]]] = 0, (2.70)


1 3 0 1 1 1 0
[ad(e )] e = [e , [e , [e , e ]]] = 0. (2.71)

This corresponds to the fact, that 30 + 1 and 0 + 31 are not roots, while for example
0 + 21 is a root. Using the fact, that = 0 + 1 , we can rewrite the roots. In particular,
the root system given in Eq. (2.38) can be rewritten as

= {n0 + m1 | |m n| 1, n, m Z}.
(2.72)

We see again, that unlike in the simple cases, the root system of A1 contains infinitely many
roots.

2.7 Fundamental weights


As in the simple case, the fundamental weights {i }, 0 i r are defined to be the
elements of the basis dual to the simple coroots. The fundamental weights are assumed to
be eigenstates of L0 with zero eigenvalue. For i 6= 0, the affine fundamental weights are
given by
i = (i ; ai ; 0). (2.73)

39 FS2016
Part 2. Generalizations and extensions: Affine Lie algebras

Their finite part makes them dual to the finite simple roots, while the k eigenvalue is fixed
by the condition
(i , 0 ) = 0, i 6= 0. (2.74)
The zeroth fundamental weight must have zero scalar product with all finite i s and satisfy

(0 , 0 ) = 1. (2.75)

Hence it must be
0 (0; 1; 0). (2.76)
It is called the basic fundamental weight. With

i (i ; 0; 0), (2.77)

it follows that
i = ai 0 + i . (2.78)
The scalar product between the fundamental weights is given by

(i , j ) = (i , j ) = Fij , i, j 6= 0, (2.79)
(0 , i ) = (0 , 0 ) = 0, i 6= 0, (2.80)

where Fij is the quadratic form matrix of g. Affine weights can be expanded in terms of the
affine fundamental weights and as
r
= i i + l, l R. (2.81)
i =0

Since each fundamental weight contributes to the k eigenvalue by a factor of ai , we define


r
k (k ) = ai i . (2.82)
i =0

k is called the level. This implies that the zeroth Dynkin label 0 is related to the finite
Dynkin labels {i }, i = 1, . . . , r and the level by
r
0 = (k ) ai i , (2.83)
i =1

i.e.
0 = k (, ). (2.84)
Modulo a possible factor, the relation between and its finite counterpart is simply

= k 0 + . (2.85)

Note, that roots are weights at level zero.


Affine weights are generally given in terms of Dynkin labels of the form

= [0 , 1 , . . . , r ], (2.86)

e.g.
0 = [1, 0, . . . , 0], 1 = [0, 1, 0, . . . , 0], r = [0, . . . , 0, 1]. (2.87)

FS2016 40
Part 2. Generalizations and extensions: Affine Lie algebras

Note, however, that this notation does not keep track of the eigenvalue of L0 . The Dynkin
labels of simple roots are given by the rows of the affine Cartan matrix:

i = [ A
bi0 , A bir ].
bi1 , . . . , A (2.88)

The affine Weyl vector is defined as


r
= = [1, 1, . . . , 1], (2.89)
i =0

and
(k ) = g, (2.90)
the dual Coxeter number. Note, that unlike the simple case, cannot be written as 1/2 the
sum of positive affine roots.
As in the simple case, affine weights whose Dynkin labels are all non-negative integers
play a special role, they are called dominant. This property is however level-dependent, as
0 is fixed by k and i , i 6= 0, see Eq. (2.84). The set of all dominant weights of level k is
denoted by P+k . The finite part of an affine dominant weight is itself a dominant weight,

P+k P+ . (2.91)

2.8 The affine Weyl group


The Weyl reflection with respect to the real affine root is defined in complete analogy to
the case of the simple Lie algebra:

s = (, ), (2.92)

and the set of all such reflections generates the affine Weyl group W. b Just like the Weyl
group of a simple Lie algebra, it acts on the weight space by linear maps. Many of its
properties are analogous to the simple case, but we will see that important differences
arise. Wb is generated by the reflections si with respect to the simple roots. Each of these
elementary reflections permutes the set of positive roots.
The new feature in the affine case are related to the existence of the imaginary roots.
As (, ) = 0, the imaginary roots are unaffected by the affine Weyl reflections:

s = (, ) = . (2.93)

Thus any Weyl reflection acts on the set of imaginary roots {n | n 6= 0} as the identity
map.
For i = 1, . . . , r, the si act precisely as the simple Weyl reflections of g. We will see
however, that the reflection s0 with respect to 0 acts differently, namely as a reflection
supplemented by a translation. The affine Weyl group acts therefore on the weight space of
g as an affine mapping hence the name affine for the affine algebras.
With = (; k; n) and = (; 0; m), we find

2m
s = (s ( + km ); k; n [(, ) + km] ). (2.94)
| |2

To analyze the structure of W,


b we want to rewrite this as

s = s (t )m , (2.95)

41 FS2016
Part 2. Generalizations and extensions: Affine Lie algebras

2 1 1 + 2 =
s1
k=3

s1 s2 s2
k=2
1 1

k=1

s1 s2 s1 s1 s2
1 2 2

Figure 2.1: Shifted hyperplanes for s0 in the case A2 .

with t defined as
t = s + s = s s + , (2.96)
i.e.
t = ( + k ; k; n + [||2 | + k |2 ]/k2 ). (2.97)
The action of t on the finite part of corresponds thus to a translation by the coroot .
Since
(t )(t ) = t + , (2.98)
the set of all t generates the coroot lattice Q . An affine Weyl reflection is thus a product
of a finite Weyl reflection times a translation by an appropriate coroot. As the group of such
translations is infinite, the affine Weyl group is infinite-dimensional. Q is an invariant
subgroup of W: b
w(t )w1 = tw , w W. b (2.99)
As Q and W only have the identity in common, W b is isomorphic to the semi-direct product
of W and the Abelian group T of translations by k-multiples of elements of Q :
b
W = W n T = W n kQ . (2.100)

While the affine Weyl group is independent of the level, its action on those weights which
have a definite value k of the level depends in a non-trivial manner on k.
Let us now study the action of s0 in Eq. (2.94):

s0 = ( + k (, ); k; n k + (, )) = s t (). (2.101)

With s = , the finite part of s0 is s + k. This mapping can be described as a


reflection with respect to an appropriately shifted hyperplane, see the example of A2 shown
in Figure 2.1. Just as in the simple case, the affine Weyl chambers are defined as those
open subsets of the vector space of affine weights which are obtained by removing all
hyperplanes that are left invariant by some Weyl reflection,

Cw = { | (w, i ) 0, i = 0, 1, . . . , r }, w W.
b (2.102)

FS2016 42
Part 2. Generalizations and extensions: Affine Lie algebras

2
1
1

Figure 2.2: Affine Weyl chambers of A2 at for the level k = 2.

Due to the presence of the subgoup T W,b the Weyl chambers are now polytopes of finite
volume rather than infinite cones. In particular, they contain a finite number of weights.
They are also referred to as Weyl alcoves. Note that because of the level dependence
of T = kQ , the size of the chambers depends on the level. Figure 2.2 shows the Weyl
chambers of A2 at for the level k = 2.
The fundamental or dominant chamber corresponds to the element w = 1. Weights
in this chamber have all Dynkin labels positive:
r
= i i + l, l R, i 0. (2.103)
i =0

To any weight which does not lie on the boundary of some chamber, there is a unique
Weyl transformation w associated such that w () lies in the fundamental affine Weyl
chamber.
The affine Weyl group preserves the scalar product,

2m
(s , s ) = (s ( + km ), s ( + km )) + 2k(n [(, ) + km] )
| |2
= (, ) + 2kn = (, ). (2.104)

Thus, all weights in a given Weyl orbit have the same length. A W
b orbit contains infinitely
many weights and has a unique weight in he fundamental Weyl chamber.

43 FS2016
Part 2. Generalizations and extensions: Affine Lie algebras

k=4
k=3
s1 s0 s1 s0 s1 s1 1 s0 s1 s0
k=2
k=1
k=0
!4 !3 !2 !1 !1 !2 !3 !4

Figure 2.3: Affine Weyl chambers of A1 at levels k = 0, . . . , 4.

Example: the Weyl group of A1 . The affine Weyl group of A1 is generated by the
reflections s0 , s1 . Their actions on a weight = [0 , 1 ] are given by
s0 = 0 0 = [0 , 1 + 20 ], (2.105)
s1 = 1 1 = [0 + 21 , 1 ]. (2.106)
Let the level of be k. Using 0 = k 1 , we can rewrite the simple affine Weyl reflections
as
s0 = [k + 1 , 2k 1 ], (2.107)
s1 = [k + 1 , 1 ]. (2.108)
We find
s0 s1 = [k 1 , 2k + 1 ]. (2.109)
This means that s0 s1 translates the finite part of by 2k 1 = k1 = k1 . This means that
s0 s1 corresponds to the basic translation operator t1 . The structure of Wb is

b = {(s0 s1 )n , s1 (s0 s1 )n | n Z}.


W (2.110)
The translation s0 s1 has no finite order. The above expressions for the actions of the group
elements can be used together with Eq. (2.102) to determine the boundaries of the affine
Weyl chambers, which are displayed in Figure 2.3. We see that the size of chambers
increases with the level.

Example: the Weyl group of A2 . The affine Weyl group of A2 is generated by the
reflections s0 , s1 , s2 . Their actions on a weight = [0 , 1 , 2 ] are given by
s0 = [0 , 0 + 1 , 0 + 2 ], (2.111)
s1 = [0 + 1 , 1 , 1 + 2 ], (2.112)
s2 = [0 + 2 , 1 + 2 , 2 ]. (2.113)
Using again 0 = k 1 2 , we find
t1 = s2 s0 s2 s1 , t2 = s1 s0 s1 s2 . (2.114)

The Weyl chambers of A2 change in size with the level. For k = 2, they are shown in
Figure 2.2.

FS2016 44
Ar 0 Ar E6 60 E6 6
Ar B
ArE
r E6 E7
1 2 3 r 1 r 1 2 2 13 3 2 r r31 1r 4r 5
6
0
1 1
2 2
3
1 2 3 r 1 r Part 2. Generalizations r
11 2 2 3 3 rand1 extensions: 1 2
Affine Lie algebras 3 4
1 4 5

Part 2. Generalizations and Extensions 0 7


0 0 7 7
A0r Br 0
0 6
Br Brr
C E7 E
Br Ar 6 2 1 3 2
E7 E0
Br r 1 r E7 E78
2 3 r 1 r 2 03 1 r r2 1 1r3r 4 5 6 0 01 12 23
21 3 r 1 rr 10 1 2 3 4 5 6 0 1 2
1 1
2 3 r 1 1 1 2 3 4 5

D2l+1
Extensions Cr 0 8
7r 8
0
C DrEr 00E8 11 22
C r 2 F48
E8 E
Cr r
Br 0
0 1 12 2 r r1 r1 r 8
E7 2 3
r r 1 1r r 0 1 2
2 3 r 1 r 10 0 10 121 23 2 r34 31 45 4 56 5 67 6 70 01 12 23
1 0
D2l D2l
0 0 r r 00 r r
A1 0 61 8 G2
Ar E6 DFr r 2 F
DrDr r r22 Dr 4 F042 13 2 3 4r 2 4
F4 0 01 12 23
23 2r 3 13 r 2 03 1 2 3 4 0 1 2
1 C2r 1 2 r 1 r E
1 10 r r1 1 1 18 2
1
3 4 r5 1
r 1
0 1 2 3 4 5 6 7
0 Table 2.1: Dynkin diagrams of the untwisted affine
A1E AE
12 0
G 1 7 G2
0 A1 06 00 1 1 6 r A1 7 0G2 1 G20 1 2
0 1 2
E6 0 1 2 0 1 2
Br Dr r 2 E7 F4
1 r 0 1 2 3 4
r 2 3 1 2 r2 3 13 r 4 5 0 1 2 3 4 5 6
1 Table 2.1: Dynkin 1
r diagrams Table
of the untwisted 2.1:Lie
affine Dynkin diagrams of the untwisted affine Lie
algebras
1
Table 2.1: Dynkin diagrams of the untwisted
Table 2.1:affine
Dynkin
Lie algebras
diagrams of the untwisted affine
Figure 2.4: Outer automorphisms 7 of the Dynkin diagrams of the untwisted affine Lie
algebrasA1 0 1 G2 8
0 1 2
E7
r 1 r Cr 2.90 Outer1 20 automorphisms
1 r 21 3 r 4 5 E
68
FS2016 0 1 2 3 4 5 6 736
Table 2.1: Dynkin diagrams of the untwisted affine Lie algebras
Let D ( g) be the symmetry group of the g Dynkin diagram and D ( g) the symmetry group of
the g Dynkin diagram. These are the sets of symmetry transformations of the simple roots
that0 preserve the scalarr product and 8hence the Cartan matrix. Since the scalar product
of the affine roots only depends on their finite parts, it is enough to consider 36
FS2016 the finite
r 1 FS2016
r projectionE8of the systemr of 2simple roots. 36
A simple root is thus mapped into another simple
Dr F4
2 3 0 1 2 3 4
FS2016 0
root with the same 1mark2 and3 comark. 5
4FS2016 6 367 36
1We define the groupr of1 outer automorphisms of the Dynkin diagram of g as

r O( g) = D ( g)/D ( g). (2.115)


A1 0 1F G2
r 2 This quotient 4 makes sense as D ( g) is a subgroup of D ( g). O( g) is thus the set of symmetry
0 1 2 3 4 0are not 2
1 symmetry
transformations
FS2016 of the Dynkin diagram of g
36that transformations of the
r 1 Dynkin diagram of g. The outer automorphisms of the Dynkin diagrams of the untwisted
affine Lie algebras are given in Figure 2.4.
Table
In terms of 2.1: Dynkin
the action of diagrams of the element
their generating untwisted
on anaffine Lie algebras
arbitrary weight = [0 , . . . , r ],
G
the outer automorphism
2 groups of the affine Lie algebras are given in Table 2.3 (for the
cases, where O( 0 g)16= 12). Note that the action of O( g) does not change the level since
every fundamental weight is mapped into another fundamental weight of the same comark.
It is clear that the set of dominant weights P+k is mapped into itself. Thus, O( g) preserves
2.1: Dynkin diagrams
the of the untwisted
fundamental Weylaffine Lie algebras
chamber.

45 FS2016
Part 2. Generalizations and extensions: Affine Lie algebras

g O( g) action of generators
Ar Zr + 1 a = [r , 0 , . . . , r2 , r1 ]
Br Z2 a = [1 , 0 , . . . , r1 , r ]
Cr Z2 a = [r , r1 , . . . , 1 , 0 ]
D2l Z2 Z2 a = [1 , 0 , 2 . . . , r , r1 ]
a = [r , r1 , . . . , 1 , 0 ]
D2l +1 Z4 a = [r1 , r , r2 , . . . , 1 , 0 ]
E6 Z3 a = [1 , 5 , 4 , 3 , 6 , 0 , 2 ]
E7 Z2 a = [6 , 5 , 4 , 3 , 2 , 1 , 0 , 7 ]

Table 2.3: Action of the generating element of the outer automorphisms

Let A be a generic element of O( g). Its action on an affine weight is given by

r
A = kA0 + i A(i ai 0 ), (2.116)
i =1

where k is the level of . This follows directly from Eq. (2.83). The second term on the
r.h.s. of Eq. (2.116) acts on the finite part of like an automorphism on the finite weight
lattice that leaves the origin fixed. It is, in fact, an element w A of the finite Weyl group.
The sum in Eq. (2.83) is the affine extension of w A at level zero, w A k 0 . Therefore,

A = k ( A 1)0 + w A . (2.117)

In general, w A can be characterized as follows. Let wi be the longest element of W(i) , the
subgroup of the finite Weyl group generated by all s j , j 6= i. Then,

w A = wi w0 for i such that A0 = i , (2.118)

where w0 is the longest element of W.


Note, that outer automorphisms must preserve the commutation relations of the algebra.

Example: the outer automorphism group of A1 . For A1 , the only non-trivial outer
automorphism is
a := 0 1 . (2.119)

Since W = {1, s1 }, wa = s1 . Comparing

a [ 0 , 1 ] = [ 1 , 0 ] = [ 1 , k 1 ] (2.120)

to

a[0 , 1 ] = k ( a 1)0 + s1 [0 , 1 ] (2.121)


= k(1 0 ) + [0 + 21 , 1 ] = [1 , k 1 ],

we find that this is correct.

FS2016 46
Part 2. Generalizations and extensions: Affine Lie algebras
3.1 Hasse diagrams 55

(1,1)

(1,0) (0,1)

Figure 3.2: Hasse diagram of the positive roots of A2 . The numbers (m1 , m2 ) denote the
root vector.
Figure 2.5: Hasse diagram for the roots of A2 .

The the
Example: firstouter
step isautomorphism
to distribute thegroup
simpleof
roots
A2 . evenly on2 , athe
For A horizontal line around
basic element a maps
the origin. This is achieved by the following horizontal projection Px of the simple
roots i : a := 0 1 2 0 . (2.122)
i 1 1
Px (i ) = xi , (3.3)
We find that i = 1 and the longest element n of1W (1
2 ) is s . Using w = s s s = s s s , we
2 0 2 1 2 1 2 1
find where n is the rank of the Lie algebra. The horizontal position of a generic root
= mi i can now be defined as
w a = s2 s2 s1 s2 = s1 s2 . (2.123)
Px ()
We find indeed by direct calculation = mi Px (i ) = mi xi .
that (3.4)

Note that the explicit summation of the index i has been dropped. From now on,
a[0 , index
any contracted 1 , 2 ]will
= kbe (a 1)0 +over.
summed s1 s2 [We
0 , 2]
1 , formalize
can the above a bit by(2.124)
introducing a projection=vector that
k ( 1 ' 0 ) + [0 + 22 + 1 , 1 2 , 1 ]
satisfies
= [2 , k (
1 , ] = [ 2 , 0 , 1 ].
i |')2= x1i . (3.5)

Expanded in the basis of simple co-roots, the projection vector ' explicitly reads
F
' = (A 1 ij
) xj i_ . (3.6)

When we take its inner product with a generic root , we see that it indeed gives us
2.10 theVisualizing the
desired projection root
(3.4): systems
(|') of complete
= mi xi . The affine projection
Lie algebras
P = (Px , Py )
can then be written as
We have seen how higher rank root systems of simple Lie algebras can be visualized in
Px () = necessitate
Sec. 1.13. Affine root systems, being infinite, (|') , (3.7a)
other means of visualizations.
_
A good means of visualization arePHasse y () = (| ),
diagrams. A Hasse diagram is a(3.7b) graph that
displays the ordering between the different elements of a set, in our case, the roots. We
where the projection in the vertical coordinate y is just the height of the root.
have seenNote
that that
we can define ancoordinate
the horizontal ordering (3.3)on the
of aroots
simplebyroot
saying one root
i strongly is larger
depends on than
the other, if their difference is a positive root. Additionally, we need
its number i. If the order of the simple roots is changed, the Hasse diagram changes
, to introduce a
cover relation:
shape too.aThe
rootbest
islooking
said todiagrams
cover , are produced
, if there
when is no
theroot smaller
ordering of nodesthan
in and
the Dynkin
bigger than . Wediagram
can now (and thusa the
draw ordering
Hasse of simple
diagram usingroots) matches the
the following connections
rules:
between the nodes. See also Figure 3.3.
If , the vertical coordinate for is less than that for .

If  there is a straight line connecting and .

Of course, we can also use Hasse diagrams for finite root systems, resulting in a finite
graph. As a first simple example, the roots of A2 are displayed in Figure 2.5. The lines
joining the roots correspond to the Weyl reflections with respect to simple roots which turn
the simple root below into the one above. In the figures, each color and angle of a line
encodes a Weyl reflection, see e.g. in Fig. 2.6 for the roots of A4 . We can also use a Hasse
diagram to visualize the Serre construction, where all elements of the algebra are produced
via commutators of the Chevalley generators, see Figure 2.7 for the Serre construction of
A4 . While the root systems of simple Lie algebras can be visualized by other means, the
true power of the Hasse diagrams is their ability to depict the infinite affine root systems.
Figure 2.8 shows the roots systems of A1 , C2 , D4 , A8 , D7 and E7 . The graphs display the

47 FS2016
56and extensions: Affine Lie algebras
Part 2. Generalizations Chapte

(1,1,1,1) (1,1,1,1)

(1,1,1,0) (0,1,1,1) (1,1,1,0) (1,1,0,

(0,1,1,0)

(1,1,0,0) (0,0,1,1) (1,1,0,0) (1,0,1,0)

(1,0,0,0) (0,0,0,1) (1,0,0,0)

(0,1,0,0) (0,0,1,0) (0,1,0,0) (0,0

1 2 3 4 3 1 2

58
Figure 3.3:
Figure Two Dynkin
2.6: Hasse diagrams
diagram (below)
for the roots of A4 . and Hasse diagrams (above
Chapter 3 Visualizations
algebra, A4 . The ordering of nodes in the left Dynkin diagr
numbers below the nodes, is canonical. The ordering of
ade
Dynkinaddiagramaddoes not match ade
the connections between th
e 1 e 2 3 4
Hasse diagram with crossing lines.
(a) Legend

[e1 , e2 ] [e2 , e3 ] [e3 , e4 ]

The lines drawn in a Hasse diagram represent the Weyl reflecti


[e1 , [e
e21 , e3 ]]e2
e [e , [e
e , e ]]
roots. Say there 3
is a root projected to the point (x, y). Then
2 3
4 4

connected to it by the line


(b) Positive Chevalley generators of the
(c) Single fundamental Weyl reflection wi
commutators

the point (x + xi , y + 1). The [e , [e line of a fundamental reflection is th


, [e , e ]]] 1 2 3 4

an angle given by
[e1 , [e2 , e3 ]] [e2 , [e3 , e4 ]]

1 1
wi = tan .
[e1 , e2 ] [e3 , e4 ] xi

Because xi is unique for all i, the n distinct fundamental reflections


at dierent angles, and reflections
(d) Double commutators in the same simple root are dr
(e) Triple commutators
distinguish between them even further they will get drawn in diere
Figure 3.5: The Serre construction for A4 .
from blue
Figure (the diagram
2.7: Hasse first fundamental of A4 . to red (the nth ).
reflection)
for the Serre construction

3.1.1 The Hasse


Visualizing diagram
the Serre of the full root system is symmetric around th
construction
FS2016 of thecanChevalley
Hasse diagrams involution
serve as a neat tool (2.21).
48 to visualize It isoftherefore
the results customary to draw
the Serre con-
struction, roots in a Hasse
the step-by-step diagram.
construction of the full algebra from the Cartan matrix
(see Example 2.2). One then has to interpret the points in the diagram not as roots,
Following
but as the generator theto. above
they belong procedure
Furthermore, itthen
the lines can is straightforward,
be interpreted though so
Part 2. Generalizations and extensions: Affine Lie algebras

Figure 2.8: Hasse diagram for the roots systems of A1 , C2 , D4 , A8 , D7 and E7 .

49 FS2016
Part 2. Generalizations and extensions: Affine Lie algebras

symmetry structure of the various affine Lie algebras and give a more intuitive idea of their
structure. All illustrations in this section are reproduced with permission from [Nut10],
where the visualization via Hasse diagrams is explained in more detail.

2.11 Highest weight representations


In contrast to the case of simple Lie algebras, all non-trivial representations of an affine
Lie algebra are infinite-dimensional. Again, the most interesting representations are the
highest weight representations, as they allow for an easy investigation of unitarity and
have many other similarities to the finite-dimensional case.
Highest weight representations are characterized by a unique highest state |i which is
annihilated by the action of all ladder operators for positive roots,

E0 |i = En |i = Hni |i = 0, n > 0, > 0. (2.125)

The eigenvalue of this state, , is the highest weight of the representation

H0i |i = i |i, k |i = k|i, L0 |i = 0, i 6= 0. (2.126)

Setting the L0 eigenvalue to zero is a matter of convention; a redefinition of L0 would yield


any desired value. In the Chevalley basis, the eigenvalues are the Dynkin labels:

h0i |i = i |i, i = 0, . . . , r. (2.127)

All states in this representation are generated by the action of the lowering operators on
|i. Since k commutes with all the generators, these states all have the same k eigenvalue.
From now on, k will be identified with its eigenvalue k, the level. In most applications, k is
fixed from the onset.
The analogues of the irreducible highest weight representations of g are those representa-
tions whose projections onto the sl (2) subalgebra associated to any real root are finite. It is
sufficient to concentrate on simple roots.
The weight system is the set of all weights in the representation of the highest-
weight state |i. An analysis parallel to the simple case shows that any weight 0
satisfies
(0 , i ) = ( pi qi ), i = 0, . . . , r (2.128)
for some positive integers pi , qi . This implies that

i0 Z, i = 0, . . . , r. (2.129)

For the highest weight , pi = 0 for all i, therefore

i0 Z+ , i = 0, . . . , r. (2.130)

This requires in particular that

0 = k (, ) Z+ . (2.131)

Since (, ) Z+ , this means that k must be a positive integer, bounded from below by
(, ):
k Z+ , k (, ). (2.132)
A far-reaching consequence of this constraint is that for a fixed value of k, there can only
be a finite number of dominant highest-weight representations. For example, for k = 1, the

FS2016 50
Part 2. Generalizations and extensions: Affine Lie algebras

only such representations are those with highest weight i such that the corresponding
simple root i has unit comark. Since a0 = 1 for all g, 0 is always dominant. The level-1
representation with highest weight 0 is called the basic representation.
For Ar , all comarks are one. There are thus r + 1 dominant highest-weight representa-
tions at level 1 whose highest weights are the i , i = 1, . . . , r.
In the following, we will use the notation gk for the algebra g at level k.

Example: Dominant highest weight representations of A2 at level 2. For A2 , all the


comarks are one, so the set of all dominant highest-weight representations at level two is
given by
[2, 0, 0], [0, 2, 0], [0, 0, 2], [1, 1, 0], [1, 0, 1], [0, 1, 1]. (2.133)

Example: Dominant highest weight representations of G2 at level 2. For G2 , the


comarks are a0 = a2 = 1, a1 = 2, which leads to the set of all dominant highest-weight
representations at level two being

[2, 0, 0], [0, 0, 2], [1, 0, 1], [0, 1, 0]. (2.134)

Representations that decompose into finite irreducible representations of sl (2) and can
further be written as a direct sum of finite-dimensional highest-weight spaces are called
integrable. The adjoint representation, although not a highest weight representation is
integrable. The first requirement is obviously met, and the direct sum decomposition is the
root space decomposition into finite roots and imaginary roots. Dominant highest-weight
representations are also integrable.
Moreover, if

( Jna ) = J
a
n, or ( Hni ) = Hni , ( En ) = E n, (2.135)
dominant highest-weight representations are easily checked to be unitary. For instance,

2 2
| E

n | i| = h | En En | i = [nk (, )]h|i 0 (2.136)
| |2

since for n > 0, any and dominant

nk (, ) = [k (, )] + (n 1)k + ( , ) 0. (2.137)

The condition for the simple case given in Eq. (1.52) is for dominant highest weights
equivalent to the existence of the singular vectors

E0i |i = E1 |i = 0 (2.138)

and
( E0i )i +1 |i = ( E

1)
k (, )+1
|i = 0, i 6= 0. (2.139)
In the Chevalley basis, these vectors read

ei |i = ( f i )i +1 |i = 0, i = 0, . . . , r. (2.140)

51 FS2016
Part 2. Generalizations and extensions: Affine Lie algebras

In sharp contrast to the simple Lie algebras, the representation space L resulting from
quotienting out these singular vectors is not finite dimensional. The imaginary root can
be subtracted from any weight without leaving the representation:

If 0 , then 0 n n > 0. (2.141)

The source of the infinity lies in the absence of a singular vector similar to Eq. (2.139), but
involving .
We will now study how the various weights in can be obtained. The algorithm we
used for g also works for g, just involving one additional Dynkin label. In the affine case,
however, the algorithm never terminates.
We define the grade to be the L0 eigenvalue shifted such that L0 |i = 0 for the highest
state |i. At grade zero, the states are obtained from |i by applying the finite Lie algebra
generators, as they do not change the L0 eigenvalue. The finite projection of all the weights
at grade zero are all the weights in the g irreducible finite-dimensional representation with
highest weight . The weights at grade one are obtained from those at grade zero that
have 0 > 0 by subtraction of 0 , followed again by the subtraction of all possible finite
simple roots. The analysis of the higher grades follows the same pattern.
An important point is that the finite projections of the affine weights at a fixed value of
the grade are organized into a direct sum of finite-dimensional weight spaces. This shows
that dominant highest-weight representations are integrable.
Finally, we must give the multiplicity of each weight. When the L0 eigenvalue is taken
into account, the multiplicities of the weights are clearly finite. It can be calculated from a
modified version of the Freudenthal recursion formula Eq. (1.135) which keeps track of
the root multiplicities:

| + |2 |0 + |2 mult (0 ) = 2 (0 + p, )mult (0 + p).
 
mult() (2.142)
>0 p =1

Recall that real roots have multiplicity one, while imaginary roots have multiplicity r.
Using our convention for the L0 eigenvalue of the highest weight states, the scalar
product of two affine highest weights does not differ from its finite form:

(, ) = (, ) for ( L0 ) = ( L0 ) = 0. (2.143)

Thus, with = (; k; 0) and = (; g; 0),

| + |2 = | + |2 . (2.144)

However, at grade m, 0 = (; k; m) and

|0 + |2 = |0 + |2 2m(k + g). (2.145)

Multiplicity calculations using the Freudenthal recursion formula are rather involved.
The constancy of the weight multiplicities along W-orbits
b greatly simplifies the analysis.
The generating function for such multiplicities is called a string function. Let be a weight
in such that +
/ and denote the set of such weights as max
. The multiplicity
of the various weights in the string , , 2, . . . is given by the string function

(q) = mult ( n)qn . (2.146)
n =0

For more complicated representations, several string functions are required. The complete
information about the multiplicities of all the weights in the representation is contained in

FS2016 52
Part 2. Generalizations and extensions: Affine Lie algebras

[1, 0]1

0


1
[ 1, 2]1 [1, 0]1 [3, 2]1

[ 1, 2]1 [1, 0]2 [3, 2]1

[ 1, 2]2 [1, 0]3 [3, 2]2

[ 3, 4]1 [ 1, 2]3 [1, 0]5 [3, 2]3 [5, 4]1

Figure 2.9: Weights at the first few grades of the basic representation of ( A2 )1 . The
multiplicity is given in the subscript. The colors encode the orbits under the Weyl group.

the set of string functions (q) for all max



. However, since weight multiplicities are
constant along Weyl orbits, that is,

w (q) = (q), (2.147)


it is sufficient to know the string functions for those weights in max

that are also dominant
(recall that a Weyl orbit contains exactly one element in the fundamental chamber). We
note further, that all the weights in must also be in the same congruence class as .
The number of independent string functions required to fully specify the representation
of highest weight is thus equal to the number of integrable weights at level k that are
in the same congruence class as . For example, in ( A1 )2 , there are three integrable
weights, [2, 0], [0, 2], [1, 1]. The first two are in the same conjugacy class, therefore two
string functions are needed in this case.

Example: the basic representation of ( A2 )1 . Let us consider the basic representation


of ( A2 )1 with highest weight [1, 0]. Using the algorithm described above, it is easy to write
down the weights in the first few grades, see Figure 2.9. The first weight with non-trivial
multiplicity is [1, 0] at grade 2. With = (0; 1; 0) and 0 = (0; 1; 2), we find | + |2 = 21
and |0 + |2 = 12 12. To calculate the r.h.s. of Eq. (2.142), we must consider all the
weights 0 + p for p, > 0, up to grade zero. The positive roots of A2 are 1 , 1 + n, n
for n > 0, they all have multiplicity one. Finally, we find that the r.h.s. of Eq. (2.142) is
equal to 24, resulting in multiplicity 2 for [1, 0] at grade 2.
We have seen that we can simplify the calculation of the multiplicities by taking into
account the Weyl orbits. We have already studied the action of the Weyl group A1 on the
Dynkin labels in an earlier exercise. We start with the dominant weight [1, 0] at grade 0.
We find the orbit
s0 s
1 0 s 1 s
[1, 0]
[1, 2]
[3, 2]
[3, 4]
[5, 4] (2.148)
It is important to take into account that s0 increases the L0 eigenvalue of the weight it acts
on and thus the grade by 0 . Thus, the second weight in the above sequence is at grade 1

53 FS2016
Part 2. Generalizations and extensions: Affine Lie algebras

and the fourth and fifth are at grade 4. s1 does not change the grade. In Figure 2.9, the
W-orbits
b are color coded. We see that each of the weights [1, 0] = (0; 1; m) represents one
orbit. The orbits are consistent with the multiplicities of the weights given in the figure. We
[1,0]
see that the first few coefficients of the string function [1,0] are given by 1, 1, 2, 3, 5, . . . .
This is the number p(n) of inequivalent decompositions of n into positive integers (number
of integer partitions) for which a closed formula exists:

1
1 qn ,
[1,0]
[1,0] (q) = mult ( n)qn = p(n)qn = (2.149)
n =0 n =0 n =1

it is the inverse of the Euler function.

Literature
The discussion on affine Lie algebras follows again [DMS97] and [FS97], using the notation
of [DMS97] and supplementing some material from [FS97]. [Fuc92] treats affine Lie
algebras in more mathematical detail, in particular the twisted case. The discussion of the
visualization of affine root systems via Hasse diagrams as well as the illustrations can be
found in [Nut10].

References
[Nut10] T. Nutma. Kac-Moody symmetries and gauged supergravity. PhD thesis. Rijksuniver-
siteit Groningen, Dec. 2010. URL: http://www.rug.nl/research/portal/files/
14628607/15_thesis.pdf.
[DMS97] P. Di Francesco, P. Mathieu, and D. Snchal. Conformal field theory. Graduate texts
in contemporary physics. New York: Springer, 1997. ISBN: 0-387-94785-X. URL: http:
//opac.inria.fr/record=b1119694.
[FS97] J. Fuchs and C. Schweigert. Symmetries, Lie Algebras and Representations: A Graduate
Course for Physicists. Cambridge Monographs on Mathematical Physics. Cambridge:
Cambridge University Press, 1997. ISBN: 978-0521541190.
[Fuc92] J. Fuchs. Affine Lie Algebras and Quantum Groups: An Introduction, with Applications
in Conformal Field Theory. 1st. Cambridge Monographs on Mathematical Physics. Cam-
bridge University Press, 1992. ISBN: 0521415934.

FS2016 54
Part 3. Advanced topics: Beyond affine Lie algebras

Part 3

Advanced topics: Beyond affine Lie


algebras

In this last part of the course, we are entering advanced territory of further extensions of
simple Lie algebras beyond the affine case or in different directions. The motivation for
these extensions comes again from physics, they are the mathematical answers to some
necessities that arise naturally in the physics context. Due to the advanced nature of these
topics, this part of the course will have more the character of an overview. We will first
visit the Virasoro algebra which arises naturally in two-dimensional conformal field theory
and is a further extension of the untwisted affine Lie algebras. Next, we will consider Lie
super-algebras, which bring the physical concept of fermions to Lie algebras by the inclusion
of fermionic Dynkin nodes. Applications of this structure can be found for example in the
study of integrable spin chains (the tJ model). Lastly, we will have a look at so-called
quantum groups that solve the problem of how to deform a Lie algebra by a new parameter
q. The simplest example of a supergroup is motivated by the XXZ spin chain, where the
spins on the chain are subject to an anisotropy in the form of an external magnetic field
in the Z-direction. Another important example of a quantum group is the Yangian, which
appears in the algebraic Bethe ansatz resp. quantum inverse scattering method to solve
integrable spin chains.

3.1 The Virasoro algebra

In the second part of this course, we have constructed the untwisted affine Lie algebras
by extending the simple Lie algebras by a central element k and a derivation L0 . An
important extension of the affine Lie algebras arises when we introduce another central
element c and infinitely many generators Ln , n Z. Their commutation relations are given
by

c
[ Ln , Lm ] = ( n m ) Ln+m + n(n2 1)n+m,0 , (3.1)
12
[ Ln , Lm ] = 0, (3.2)
c
[ Ln , Lm ] = (n m) Ln+m + n(n2 1)n+m,0 , (3.3)
12
[ Ln , c] = 0. (3.4)

55 FS2016
Part 3. Advanced topics: Beyond affine Lie algebras

These bracket relations define a Lie algebra, the so-called Virasoro algebra. The relations
involving both the Ln and the generators of g read

[ Lm , Jna ] = n Jna+m , (3.5)


[ Ln , k] = [c, Jna ] = [c, k] = 0. (3.6)

Comparing with the commutation relations of L0 in Eq. (2.15) in the affine case, the
naming convention for the derivation suddenly makes sense, as it is now identified with
the Virasoro generator L0 .
Let us state a few important facts about the Virasoro algebra Vir. Vir possesses a
triangle decomposition,
Vir+ Vir0 Vir (3.7)
into the subalgebras generated by the positive, zero , and negative modes

Vir+ = span{ Ln | n > 0}, Vir0 = span{ L0 , c}, Vir = span{ Ln | n < 0}. (3.8)

Vir0 is maximal Abelian subalgebra. Owing to the existence of a triangular decomposition,


the representation theory of the Virasoro algebra parallels to a large extent the one of affine
Lie algebras. There are, in particular, highest-weight representations.
We will briefly study the representations with respect to the Ln . The representations of
their anti-holomorphic counterparts are constructed along the same lines. As we have seen,
the holomorphic and anti-holomorphic components of the overall algebra decouple.
Since no pairs of generators in (3.1) commute, we choose a single operator, L0 , which
will be diagonal in the representation space, which in this context is also called a Verma
module. We denote the highest-weight state by |hi, with eigenvalue h of L0 :

L0 | h i = h | h i. (3.9)

Since [ L0 , Lm ] = mLm , Lm for m > 0 is a lowering operator for h, and Lm , m > 0 a


raising operator. We adopt the convention,

Ln |hi = 0, n > 0. (3.10)

A basis for the other states of the representation, the descendant states, is obtained by
applying the raising operators in all possible ways:

L k 1 L k 2 . . . L k n | h i, 1 k1 k n , (3.11)

where, by convention, the Lki appear in increasing order of k i . This state is an eigenstate
of L0 with eigenvalue
h0 = h + k1 + k2 + + k n = h + N, (3.12)
where N is the level of the state (which differs from what we called the level in the affine
case, as there, it referred to the eigenvalue of the operator k).

The lowest levels of a Verma module. The first few levels of the representation with
highest state |hi are spanned by the states given in Table 3.1. Looking at the number
of distinct, linearly independent states at level N, we find again the coefficients of the
series of number of partitions p( N ) of the integer N that we have already encountered in
Eq. (2.149).

FS2016 56
Part 3. Advanced topics: Beyond affine Lie algebras

level number of states states


0 1 |hi
1 1 L 1 | h i
2 2 L21 |hi, L2 |hi
3 3 L31 |hi, L1 L2 |hi, L3 |hi
4 5 L1 |hi, L1 L2 |hi, L1 L3 |hi, L22 |hi, L4 |hi
4 2

Table 3.1: The first few levels of the Verma module generated by |hi

Using Lm = Lm , we define an inner product on the Verma module. The inner product
of two states Lk1 Lk2 . . . Lkn |hi and Ll1 Ll2 . . . Lln |hi is

h h | L k m . . . L k 1 L l1 . . . L l n | h i , (3.13)

where the dual state hh| satisfies

hh| Ll = 0, j < 0. (3.14)

Note that the inner product of two states vanishes unless they belong to the same level. In
general, two eigenspaces of a Hermitian operator (here, L0 ) having different eigenvalues
are orthogonal.
Since c is a central generator, all vectors in the Verma module generated by |hi have
the same eigenvalue of c, called the central charge.
The Virasoro algebra is intimately linked with conformal transformations in 2d and
therefore of paramount importance in the study of 2d conformal field theory (CFT), i.e. a
quantum field theory which is covariant under conformal transformations.
A conformal transformation of the coordinates in d spacetime dimensions is an in-
vertible mapping x x 0 which leaves the metric tensor g invariant up to a scale:
0
g ( x 0 ) = ( x ) g ( x ). (3.15)

A conformal transformation is locally equivalent to a (pseudo-)rotation and a dilation. The


name conformal comes from the fact that the conformal group preserves angles.
In d = 2, conformal invariance takes a new meaning. The reason is that here, there is an
infinite variety of locally conformal transformations, namely the holomorphic mappings of
the complex plane onto itself. Consider the coordinates (z0 , z1 ) on the plane. The covariant
metric tensor g transforms under a change of the coordinate system given by z w ( x )
as   
w w
g


g . (3.16)
z z
The condition that the above transformation is conformal is g0 (w) g(z), in other words,
2 2  1 2  1 2
w0 w0
 
w w
+ = + , (3.17)
z0 z 1 z 0 z1
w0 w1 w0 w1
+ = 0. (3.18)
z0 z0 z1 z1
This corresponds in turn to

w1 w0 w0 w1
= and = , (3.19)
z0 z1 z0 z1

57 FS2016
Part 3. Advanced topics: Beyond affine Lie algebras

which are the CauchyRiemann equation for holomorphic maps. The result of the above is
that CFTs in 2d can be solved exactly, which is why they hold such an important place in
theoretical physics.
Any holomorphic infinitesimal transformation can be expressed as

z 0 = z + e ( z ), e(z) = c n z n +1 (3.20)
n=

for the complex variable z = z0 + iz1 , where by hypothesis, the infinitesimal mapping
admits a Laurent expansion around z = 0. The effect of such a mapping on a spinless and
dimensionless field (z, z) living on the complex plane is

=
= e(z) e(z) [cn ln (z, z) + cn ln (z, z)], (3.21)
n

where we have introduced the generators

l n = z n +1 z , ln = zn+1 z . (3.22)

These generators obey the following commutation relations:

[ ln , lm ] = ( n m ) ln+m , (3.23)
[ln , lm ] = 0, (3.24)
[ln , lm ] = (n m)ln+m . (3.25)

Thus the conformal algebra is the direct sum of two isomorphic algebras (generated
by the ln and ln , respectively). It is also called the Witt algebra. Each of these two
infinite-dimensional algebras contains a subalgebra generated by l1 , l0 and l1 . This is the
subalgebra associated with the global conformal group. Indeed, from Eq. (3.20) we see
that l1 = z generates translations on the complex plane, l0 = zz generates scale
transformations and rotations, and l1 = z2 z generates special conformal transformations.
In CFT, however, one needs unitarizable representations of the symmetry algebra.
However, just like in the case of the loop algebra, the Witt algebra does not have any
non-trivial unitary representations. Therefore, we must again introduce a central extension.
It can be shown on quite general grounds that it must just consist of a complex number on
the right hand side of the bracket relation (the central charge). In physics language, we say
that the conformal symmetry develops an anomaly.

3.2 Lie superalgebras


Superalgebras are related to the concept of particles of different statistics, i.e. fermions and
bosons. They contain even and odd generators, the odd ones being of fermionic nature.
Mathematically speaking, these algebras have a Z2 grading.
Lie superalgebras present a different kind of generalization of complex semi-simple Lie
algebras from the ones we have studied so far. While they share many of the key concepts
of simple Lie algebras such as the Cartan subalgebra, roots, weights, the Cartan matrix and
Dynkin diagrams, new features arise due to the possibility of vanishing Killing forms. The
latter give rise to a host of pathological possibilities, which makes many definitions involved
in the study of superalgebras more bulky as they must either encompass or exclude them.
We only have time for a brief look at simple Lie superalgebras and, apart from the
basic definitions, will focus on their classification. We will concentrate on one class of

FS2016 58
Part 3. Advanced topics: Beyond affine Lie algebras

complex simple Lie superalgebras, namely the basic classical ones, which are nearest to the
simple case we are familiar with. They also hold the most interest from the point of view
of mathematical physics.
The Lie superalgebras are however not "the superalgebra" that appears in particle
physics beyond the standard model, which is an extension of the Poincar algebra. Of
course, all superalgebras share the basic Z2 grading. In the context of theoretical physics,
Lie superalgebras appear as symmetry algebras of integrable spin chains, the simplest
example being the tJ model.

First of all, we need to introduce the grading. Inspired by the properties of integer
numbers, we introduce the following structure with product rules

even even = even, (3.26)


even odd = odd, (3.27)
odd odd = even. (3.28)

The above grading is also called a Z2 grading. Let us first define a graded vector space
V. Let V be a complex vector space of dimension m + n, m, n Z+ , and let A1 , . . . , Am+n
be a basis of V. Then any A V can be written as
m+n
A= aj Aj, a j C. (3.29)
j =1

V can be graded by saying that


m
A= aj Aj, is even and (3.30)
j =1
m+n
A= aj Aj, is odd. (3.31)
j = m +1

Thus the even elements only involve the first m basis elements, while the odd elements
involve only the remaining n basis elements. Any A V that is either even or odd is said
to be homogeneous, and the degree (or parity) of such elements is defined as
(
0 if A is even
deg A = (3.32)
1 if A is odd.

The set of even elements of V forms the even subspace V0 , the odd elements form the odd
subspace V1 :
V = V0 V1 . (3.33)
If we supplement the graded vector space V by an associative product, the resulting
structure is an associative superalgebra.
Let gs be a graded vector space with g0 and g1 being its even, respectively odd subspaces
and
dim g0 = m, dim g1 = n, m 0, n 0, m + n 1. (3.34)
Assume that A, B gs , there exists a generalized Lie product or supercommutator
[ A, B] with the following properties:

[ A, B] gs A, B gs ,

59 FS2016
Part 3. Advanced topics: Beyond affine Lie algebras

A, B gs , a, b C,

[ aA + bB, C ] = a[ A, C ] + b[ B, C ], (3.35)

for any two homogeneous A, B gs , also [ A, B] is homogeneous with degree

deg([ A, B]) = (deg A + deg B) mod 2. (3.36)

for any two homogeneous A, B gs ,

[ B, A] = (1)(deg A)(deg B) [ A, B], (3.37)

for any three homogeneous A, B, C gs ,

[ A, [ B, C ]](1)(deg A)(deg C) + [ B, [C, A]](1)(deg B)(deg A)


+ [C, [ A, B]](1)(deg C)(deg B) = 0. (3.38)

(generalized Jacobi identity)


Then gs is called a complex Lie superalgebra with even dimension m and odd dimension
n. We choose a homogeneous basis of gs in which the basis elements A1 , . . . , Am+n are
pq
such that A1 , . . . , Am g0 and Am+1 , . . . , Am+n g1 . Then, the structure constants cr can
be defined by
m+n

pq
[ A p , Aq ] = cr Ar . (3.39)
r =1

Any generalized Lie product can be evaluated from the knowledge of the structure constants.
The grading implies that
pq p q qp
cr = (1)(deg A )(deg A ) cr . (3.40)
For m 1, the even subspace g0 is an ordinary Lie algebra. For m 1, n 1, the
odd subspace g1 of gs is a carrier space for a representation of the Lie algebra g0 . This
representation is called the representation of g0 on g1 . Every Lie algebra can be regarded
as a special case of a Lie superalgebra for m 1, n = 0.
When represented as matrices, the elements of a superalgebra are block matrices
 
A B
M= . (3.41)
C D

If M g0 , it has the form  


A 0
M= , (3.42)
0 D
while if M g1 , it has the form  
0 B
M= . (3.43)
C 0
As in the simple case, we define ad( X )Y = [ X, Y ], but now using the supercommutator.
The Killing form is defined as

K ( X, Y ) = str(adX adY ), X, Y gs , (3.44)

where str is the supertrace, defined as

strM = Tr A Tr D (3.45)

FS2016 60
Part 3. Advanced topics: Beyond affine Lie algebras

for the block matrix M.


A Lie superalgebra is said to be Abelian or commutative if

[ A, B] = 0 A, B gs . (3.46)

A subalgebra g gs is a subset of elements of gs that form a vector subspace of gs and


that is closed under the generalized Lie product,

[ A, B] g for A, B g. (3.47)

A graded subalgebra gs0 gs is itself a Lie superalgebra and its even subspace g00 is a
subspace of g0 and its odd subspace g10 is a subspace of g1 . gs0 is said to be a proper graded
subalgebra of gs if at least one element of gs is not contained in gs0 . A graded subalgebra gs0
is invariant if
[ A, B] gs0 A gs0 , B gs . (3.48)
A Lie superalgebra is said to be simple if it is not Abelian and does not possess a proper
invariant graded subalgebra. If gs is a simple Lie superalgebra, its Killing form K ( , ) is
either non-degenerate or identically zero. We see that the situation is very different from
the case of Lie algebras, where a Lie algebra with a zero Killing form cannot be semi-simple.
Unlike in the case of Lie algebras, for Lie superalgebras, there do exist semi-simple Lie
superalgebras that are not expressible as the direct sum of simple Lie superalgebras. In
order to define them correctly, we first need to introduce the concept of solvability. Let
(0) (k) ( k 1) ( k 1)
gs = gs and gs = [ gs , gs ] for each k = 1, 2, 3, . . . . Then the Lie superalgebra gs is
(k)
solvable if there exists a value of k for which gs = {0}.
A Lie superalgebra gs is said to be semi-simple if it does not possess a solvable invariant
graded subalgebra.
Just as for the complex finite-dimensional simple Lie algebras, we want to classify the
simple Lie superalgebras. We will however only look at the so-called classical ones, which
are the only ones of interest for mathematical physics. As before, we will only consider
complex Lie superalgebras.
A simple Lie superalgebra is said to be classical if the representation of its even part g0
on its odd part g1 is either irreducible or if it is reducible, it is completely reducible, i.e. can
be written as a direct sum of irreducible representations. If gs is a complex classical simple
Lie superalgebra, its even part has the form

g0 = g0A g0ss , (3.49)

where g0A is an Abelian complex Lie algebra and g0ss is a semi-simple complex Lie algebra.
Let h0ss be a Cartan subalgebra of g0ss . Then,

hs = g0A h0ss (3.50)

is a Cartan subalgebra of gs . The rank r of a classical simple complex Lie algebra is the
dimension of its Cartan subalgebra. We can define the roots analogously to the case of
simple Lie algebras. If ( H ) is a linear functional on hs , we can find at least one element
E gs such that
[ H, E ] = ( H ) E H hs , (3.51)
and is called a root of gs , and the set of all elements E that satisfy Eq. (3.51) forms the
root subspace gs, . If E g0 , is said to be an even root of gs , while if E g1 , is said
to be an odd root of gs . The set of all distinct non-zero even roots is denoted by 0 , and

61 FS2016
Part 3. Advanced topics: Beyond affine Lie algebras

the set of all distinct odd roots is denoted by 1 . The set of distinct roots of gs contained
either in 0 or 1 or both is denoted by s .
hs can be regarded as the subspace of gs corresponding to zero even roots.

The set of classical simple Lie superalgebras can be divided into

1. Basic classical simple Lie superalgebras which posses a non-degenerate bilinear


supersymmetric consistent invariant form K 0 . This set can in turn be divided into

(a) those for which the Killing form K is non-degenerate, so K = K 0 .


(b) those for which the Killing form is identically zero, K 6= K 0 .

2. Strange classical simple Lie superalgebras, which do not posses any non-degenerate
supersymmetric consistent invariant form.

The following is a complete list of the complex classical simple Lie superalgebras:

1. Basic classical simple Lie superalgebras

(a) with non-degenerate Killing form:


i. simple complex Lie algebras
ii. A(r |s), r > s 0
B(r |s), r > 0, s 1
C ( s ), s 2
D (r |s), r 2, s 1, r 6= s + 1
F (4)
G (3)
(b) with zero Killing form:
A (r |r ), r 1
D ( s + 1| s ), s 1
D (2|1; ), C \ {0, 1, }

2. Strange classical simple complex Lie algebras

P (r ), r2
Q (r ), r 2.

In the following, we will collect the necessary material in order to define the Cartan
matrices and Dynkin diagrams of basic complex classical simple Lie superalgebras.
For semi-simple complex Lie algebras, we have learned that (, ) > 0 for all non-zero
roots . This is not the case for basic complex classical simple Lie superalgebras:

1. If 0 is an even, non-zero root of gs , (, ) 6= 0, but it need not be real and


positive. If gs has non-degenerate Killing form, (, ) R, but can be positive or
negative.

2. If 1 is an odd root of gs , it is possible to have (, ) = 0, even when is not


identically zero.

FS2016 62
Part 3. Advanced topics: Beyond affine Lie algebras

Example: generators and roots of gs = A(1|0). A(1|0) is the complex Lie superalgebra
sl (2|1; C). Its even part is given by

g0 = g0A g0ss , (3.52)

where g0A is a one-dimensional Abelian Lie algebra and g0ss = A1 . The Cartan subalgebra
is two-dimensional with basis elements H11 and C, where C is the generator of g0A and H11
is the generator of h0ss = h A1 . Let E pq be the 3 3 matrix with entry ( p, q) = 1 and all
others zero. The generators of the Cartan subalgebra are given by H11 = E11 E22 and
C = E11 E22 2E33 . E12 , E21 , E13 , E31 , E23 and E32 generate the root subspaces. It is
easy to work out the commutation relations, we have e.g.

[ H11 , E12 ] = 2 E12 , [ H11 , E13 ] = E13 , [ H11 , E23 ] = E23 , (3.53)
[C, E12 ] = 0, [C, E13 ] = E13 , [C, E23 ] = E23 , (3.54)
[ E12 , E13 ] = 0, [ E13 , E23 ] = 0, [ E12 , E23 ] = E13 . (3.55)

A(1|0) has the even roots 1 associated with E12 , E21 and the odd roots 2 , 3 associ-
ated to E23 , E32 and E13 , E31 . We see that 3 = 1 + 2 .
Taking the basis elements of A(1|0) in the order

{C, H11 , E12 , E21 , E13 , E31 , E23 , E32 }, (3.56)

we can calculate adC, which is a diagonal matrix with elements {0, 0, 0, 0, 1, 1, 1, 1} and
adH11 which is again diagonal with elements {0, 0, 2, 2, 1, 1, 1, 1}. From here, it is easy
to find the Killing forms K (C, C ) = 4, K ( H11 , H11 ) = 4 and K (C, H11 ) = 0.
The roots have the scalar products

(1 , 1 ) = 1, (2 , 2 ) = (3 , 3 ) = 0, (3.57)
( 1 , 2 ) = ( 2 , 3 ) = 21 , ( 1 , 3 ) = 1
2. (3.58)

In the following, we need to define the concepts of positive, negative and simple roots
such that as many results from simple Lie algebras carry over as possible. One main
difference however remains. For simple Lie algebras, all choices for the set of positive roots
are equivalent, which is not the case for simple Lie superalgebras.
Let h0ss be the Cartan subalgebra of the semi-simple Lie algebra g0ss , and suppose a
choice of positive roots of g0ss has been made. Then the subalgebra b0ss of g0ss that consists of
all elements of h0ss together with the root subspaces corresponding to + of g0ss forms a
maximal solvable subalgebra of g0ss . If g0 contains a one-dimensional Abelian Lie algebra
g0A , then we define the subalgebra b0 as

b0 = b0ss g0A , (3.59)

but if g0A = {0}, then


b0 = b0ss . (3.60)
In both cases, b0 is a maximal solvable subalgebra of g0 and is known as a Borel subalgebra
of g0 . Let b be a maximal solvable subalgebra of the Lie superalgebra gs and b0 b. As
the Cartan subalgebra is contained in b0 , also hs b, and so a subspace N + gs can be
defined such that
b = hs N + . (3.61)

63 FS2016
Part 3. Advanced topics: Beyond affine Lie algebras

We can introduce a further subspace N via

gs = N h s N + . (3.62)

A root gs is said to be positive if the intersection of the root subspace gs, with N + ,
gs, N + , is non-trivial and negative if gs, N is non-trivial. This definition implies
that if is positive, is negative and vice versa. Moreover, every positive even root is
an extension of a positive root of g0ss , and every positive root of g0ss extends to a positive
even root of gs . In the exceptional case gs = A(1|1), dimgs, = 2 for every odd root , and
gs, N + and gs, N are both non-trivial, implying that for A(1|1), every odd root is
both positive and negative.
A non-zero root of gs is said to be simple if is positive, but cannot be expressed as
= + , with , positive roots of gs .

Example: Positive, negative and simple roots of gs = A(1|0). We have seen that
s = {1 , 2 , 3 }. The corresponding root subspaces are all one-dimensional and
generated by the Ei .
1. One choice of b has basis {C, H11 , E1 , E2 , E3 }, so that N + has basis { E1 , E2 , E3 }.
With this choice, 1 , 2 , 3 are all positive roots and 1 , 2 , 3 are negative roots.
Since 3 = 1 + 2 , the corresponding simple roots are 1 and 2 .

2. Another inequivalent choice of b has basis {C, H11 , E1 , E2 , E3 }, so that N + has


basis { E1 , E2 , E3 } and hence N has basis { E1 , E2 , E3 }. However, now
3 = 1 + 2 is no longer a sum of positive roots and is hence simple. We rename the
positive roots to 10 = 3 , 20 = 2 , 30 = 1 . The simple roots are now 10 and 20 ,
both being odd. Note that (10 , 10 ) = (20 , 20 ) = 0, (10 , 20 ) = 21 .

The number of simple roots R is related to the rank r of gs by


(
r + 1 if gs = A( p, p), p > 0,
R= (3.63)
r for any other basic gs .

Moreover, the set of simple roots 1 , . . . , R is always linearly independent except for
A( p, p), p > 0.
We can define the Cartan matrix A of a basic classical simple complex Lie superalgebra
gs to be a R R matrix with matrix elements A jk defined in terms of the simple roots as

2( j , k )
A jk = if ( j , j ) 6= 0, (3.64)
( j , j )
( j , k )
A jk = if ( j , j ) = 0, (3.65)
( j , j0 )

where j0 is another simple root such that ( j , j0 ) 6= 0. In the first case, A jj = 2 and the
only possible values of A jk , j 6= k are 0, -1, -2 and -3. In the second case, A jj = 0 and
A jj0 = 1, which is clearly very different from the allowed values of a Cartan matrix of an
ordinary simple Lie algebra.
A is an r r matrix except for A( p, p), p > 0. For D (2|1; ) with / R, at least one
off-diagonal element of A is not real.

FS2016 64
Part 3. Advanced topics: Beyond affine Lie algebras

Example: Cartan matrices of of gs = A(1|0). With choice (1) for the simple roots, we
had (2 , 2 ) = 0, but (2 , 1 ) 6= 0, so we can take 20 = 1 . Then,
 
2 1
A= . (3.66)
1 0
With choice (2) for the simple roots, taking 20 = 1 and 10 = 2 , we find
 
0 1
A= . (3.67)
1 0
F

We want to again associate to each Cartan matrix a generalized Dynkin diagram.


They are constructed according to the following rules:
1. assign to each simple root j a vertex j which is drawn as
if j is even. This is called a white node.
if j is odd and ( j , j ) = 0. This is called a grey node.
if j is odd and ( j , j ) 6= 0. This is called a black node.

2. draw l jk lines from the vertex j to the vertex k, where


l jk = max{| A jk |, | Akj |}. (3.68)

3. add an arrow pointing from the j vertex to the k vertex if | Akj | > 1 (except for
D (2|1; ).

Example: Generalized Dynkin diagrams of gs = A(1|0). The generalized Dynkin dia-


gram corresponding to the choice (1) is

1 2 (3.69)
The generalized Dynkin diagram corresponding to the choice (2) is

1 2 (3.70)
F

While each simple complex Lie algebra has a unique Dynkin diagram, we see that this
uniqueness is lost in the superalgebra case. Due to the anomalous cases D (2|1; ) and
D (2|s), s > 0, the above prescription cannot be reversed to determine the Cartan matrix
from the Dynkin diagram. Therefore, the Dynkin diagrams play a less important role for
superalgebras.
For each basic classical simple complex Lie superalgebra gs with non-trivial odd part,
there is a distinguished choice of simple roots, which consists of one odd root, all other
R 1 simple roots being even. The corresponding Dynkin diagram therefore has R 1
white vertices and one grey or black vertex. In Table 3.2, all the Dynkin diagrams of the
basic classical simple complex Lie superalgebras corresponding to the distinguished choice
of the Borel subalgebra are depicted.

65 FS2016
Part 3. Advanced topics: Beyond affine Lie algebras

1 2 r r+1 r+s+1
A (r | s ) r s > 0

1 2 r r+1 2r + 1
A (r |r ) r 1

1 2 r r+1 r1 r
B(r |s), r > 0, s > 0

1 2 s1 s
B ( s ), s > 0

C ( s ), s > 0 1 2 s1 s

r+s
D (r |s), r > 2, s > 0 r+s2
1 2 s1 s
r+s1
s+2

D (2| s ), s > 0 s
1 2
s+1
3

D (2|1; ), 6= 1, 0, 1

F4
1 2 3 4

G2
1 2 3

Table 3.2: Basic classical simple complex Lie superalgebras corresponding to the distin-
guished choice of the Borel subalgebra

FS2016 66
Part 3. Advanced topics: Beyond affine Lie algebras

3.3 Quantum groups


In this last section of the course, we touch on a subject which requires the highest level
of abstraction encountered so far, namely quantum groups. They act as a generalized
symmetry and arise naturally in several independent contexts in theoretical physics, one of
which is the study of integrable models.
The concept of quantum group is mathematically speaking equivalent to the one of
the Hopf algebra, which is an associative algebra with a host of extra structure. We will
concentrate on a particular class of Hopf algebras, namely Uq ( g), a deformation of the
enveloping algebra U ( g) of a complex simple Lie algebra g. Uq ( g) depends in a natural
way on the algebra g alone, so the theory of Uq ( g) can be viewed as a self-contained
outgrowth of Lie algebra theory. It should be pointed out, however, that the theory of
general quantum groups is much richer and encompasses new phenomena that have no
analogy in Lie algebra theory.

The fundamental new concept to be introduced is the one of a Hopf algebra. Let a be
a vector space and F the base field of the vector space (in the following, we take again
F = C). A Hopf algebra a is a vector space endowed with five operations
M: aa a (multiplication) (3.71)
: Fa (unit map) (3.72)
: a aa (co-multiplication) (3.73)
e: aF (co-unit map) (3.74)
: aa (antipode), (3.75)
which have the following properties:
M (id M) = M (M id) (associativity) (3.76)
M (id ) = id = M ( id) (existence of unit) (3.77)
(id ) = ( id) (co-associativity) (3.78)
(e id) = id = (id e) (existence of co-unit) (3.79)
M (id ) = e = M ( id) (3.80)
M = (M M) ( ) (connecting axiom). (3.81)
At first sight, this structure might seem quite complicated, but we will see that it arises
quite naturally.
Vector spaces with a multiplication satisfying Eq. (3.76) are called associative algebras
and are very common. Many of them have a unit (unital algebras).
By formalizing the notion of unmultiplying objects, one gets the operations of co-
multiplication and co-unit. The co-multiplication of an element X a is the sum of all
those things in a a which could give X when combines according to an underlying group
structure.
Vector spaces endowed with a co-multiplication are called co-algebras. the theory of
co-algebras is dual to the one of algebras, hence it is not necessary to consider them in their
own right. The requirement of co-associativity is also natural. Having co-multiplication as
a dual operation to multiplication, it makes sense to require also the existence of a co-unit
as the dual operation to the unit. An algebra a that possesses the four operations M, ,
and e is called a bi-algebra.
Finally, we need a way of connecting the operations of multiplication and co-multiplication
non-trivially. We need to construct a map as the composition of multiplication and co-
multiplication, which leads to the concept of the antipode, and hence to Hopf algebras. The

67 FS2016
Part 3. Advanced topics: Beyond affine Lie algebras

antipode is a weaker structure than the inverse, it provides a nonlocal linearized inverse. It
means that now not individual elements, but certain linear combinations are invertible.
We can express the defining properties Eq. (3.76)Eq. (3.81) more explicitly in terms of
the elements X a. If we use  for the multiplication M,

M : X Y 7 M( X Y ) X  Y, (3.82)

Eq. (3.76) reads now

X  (Y  Z ) = ( X  Y )  Z X, Y, Z a. (3.83)

Similarly, the existence of a unit means that there is an element E a (the unit element)
such that
E  X = X = X  E X a. (3.84)
The map is then given by
: 7 E F. (3.85)
Another way of expressing the properties Eq. (3.76)Eq. (3.81) in a more explicit way is
via commutative diagrams. A diagram is said to be commutative iff the composite maps
which are obtained by following the arrows are independent of the path used to link any
two given spaces in the diagram.
The following diagram shows the associativity property Eq. (3.76).

aa
id M M

aaa a (3.86)
M id M

aa

The existence of the unit (Eq. (3.77)) is shown below, with the map s : a F a the scalar
multiplication.
s
aF a s Fa
id M M id (3.87)
M
aa a aa
M

Co-associativity (Eq. (3.78)) is shown below:

aa

id
aaa a (3.88)

id
aa

Finally, the existence of the co-unit (Eq. (3.79)) is expressed as follows, with the map
i : a a F the inclusion X 7 X 1, where 1 is the multiplicative unit of F.
i
aF a Fa
i
ide eid (3.89)

aa a aa

FS2016 68
Part 3. Advanced topics: Beyond affine Lie algebras

We see that the third and fourth diagrams are obtained from the first and second by
reversing the arrows, in other words, the rules for are just the same as the rules for
multiplication, with the arrows reversed.
When a basis { J a } of a is fixed, then multiplication, co-multiplication and antipode can
be expressed in terms of structure constants, just like the bracket relations of a Lie algebra:

Ja  Jb = cab J c , (3.90)
c
( J c ) = abc J a J b , (3.91)
a,b
( J c ) = ac J a . (3.92)
a

The associativity property Eq. (3.76) is then expressed as

bc ad ab dc
d f = d f . (3.93)

Similarly, the structural properties Eq. (3.78) and Eq. (3.80) read
a b a b
cd e f = ec f d, (3.94)
a c bd a d bc
bc d e = dc b e . (3.95)

Example: Universal enveloping algebra U ( g) of a complex Lie algebra g. The uni-


versal enveloping algebra U ( g) of a Lie algebra g consists of all finite formal power series
in the elements of g. U ( g) is an associative algebra with the product  given by termwise
formal multiplication. As a vector space, it is generated by all monomials in the generators
of g, identifying however all monomials which become equal to each other upon use of the
bracket relations of g.
U ( g) becomes a Hopf algebra by taking M as the usual formal multiplication on U ( g),
and defining the unit element by

( ) = 1 C. (3.96)

Co-multiplication, co-unit and antipode are defined by

( X ) = X 1 + 1 X, (3.97)
e( X ) = 0, (3.98)
( X ) = X (3.99)

and

(1) = 1 1, (3.100)
e(1) = 1, (3.101)
(1) = 1. (3.102)

Note, that in this case, = id. In particular, for any X g, C, the formal element
X = eX U ( g) obeys

(eX ) = eX eX , (3.103)
(X ) = eX . (3.104)

As a consequence, this element satisfies

X  ( X ) = 1 = ( X )  X, (3.105)

69 FS2016
Part 3. Advanced topics: Beyond affine Lie algebras

which justifies the description of the antipode as the analogue of an inverse.

Some general properties of Hopf algebras are the following:

For a given multiplication and co-multiplication, the co-unit is unique.

If ( a, M, , , e, ) is a Hopf algebra, then the dual vector space a inherits a Hopf


algebra structure by interchanging M, with , e.

The co-multiplication and co-unit are homomorphisms of a, i.e. preserve the multipli-
cation. For the co-unit, this means

e ( X  Y ) = e ( X ) e (Y ) , X, Y a, (3.106)

and for the co-multiplication,

( X  Y ) = ( X )  (Y ) , X Y a a. (3.107)

The antipode is an anti-homomorphism, i.e.

( X  Y ) = (Y ) ( X ) , X, Y a. (3.108)

The antipode is an anti-cohomomorphism, i.e.

( ) = , (3.109)

where is the permutation map

: aa aa (3.110)
X Y Y X. (3.111)

The map 0 := is also a co-associative multiplication.

An algebra a is said to be commutative iff the multiplication does not depend on the order
of the factors,
M = M. (3.112)
Analogously, a Hopf algebra is called co-commutative iff the co-multiplication satisfies

= , (3.113)

or in other words, iff 0 coincides with . An example of a co-commutative Hopf algebra


is given by the universal enveloping algebras U ( g) we have discussed above. For any
commutative or co-commutative Hopf algebra, the antipode obeys = id.
Before we can study the class of quantum universal enveloping algebras Uq ( g), we must
discuss one last property of Hopf algebras, namely quasitriangularity. A quasitriangular
Hopf algebra is a Hopf algebra for which the co-multiplications and 0 are related by
conjugation, i.e.
0 ( X ) = R  ( X )  R 1 X a (3.114)

FS2016 70
Part 3. Advanced topics: Beyond affine Lie algebras

for some element R a a which is invertible and satisfies

(id )( R) = R13  R12 , (3.115)


( id)( R) = R13  R23 , (3.116)
1
( id)( R) = R . (3.117)

Here, the inverse R1 of R a a is by definition that element of a a which satisfies

R 1  R = E E = R  R 1 . (3.118)

Generally, R has the structure


R1
(l ) (l )
R= R2 . (3.119)
l

In the above definition, R13 is meant as the identity in the second factor of a a a and as
R on the first and third factors, analogously for R12 , R23 . A quasitriangular Hopf algebra is
called triangular iff
R12  R21 = E E. (3.120)
The universal enveloping algebra U ( g) of any Lie algebra g is quasitriangular. Generally, a
quasitriangular Hopf algebra is neither commutative nor co-commutative; however, the
non-commutativity is under control, see Eq. (3.114). As a consequence, many properties of
co-commutative Hopf algebras generalize to generic quasitriangular Hopf algebras.
An immediate consequence of Eq. (3.114) is that

R12  ( id)( X Y ) = (0 id)( X Y )  R12 X Y a a. (3.121)

Taking X Y = R and using Eq. (3.115) and (3.116), this yields

R12  R13  R23 = R23  R13  R12 . (3.122)

This is the so-called YangBaxter equation which plays a fundamental role in the theory
of completely integrable systems. In this context, R is called the universal R-matrix.
The quantum universal enveloping algebra Uq ( g) is the algebra of power series in
the 3r + 1 generators
{ei , f i , hi | i = 1, . . . , r } {1} (3.123)
modulo the relations

[hi , h j ] = 0, (3.124)
[hi , e j ] = A ji e j , (3.125)
i j j
[h , f ] = A ji f , (3.126)
[ei , f j ] = ij bhi c, (3.127)

1 A ji
1 A ji
 
(1) p
p
(ei ) p (e j )(ei )1 A ji p = 0 i 6= j, (3.128)
p =0 i
1 A ji
1 A ji
 
(1) p
p
( f i ) p ( f j )( f i )1 A ji p = 0 i 6= j, (3.129)
p =0 i

1X = X = X1 X Uq ( g), (3.130)

71 FS2016
Part 3. Advanced topics: Beyond affine Lie algebras

where A ji are the elements of the Cartan matrix of g. Here, [ X, Y ] X  Y Y  X with


M( X Y ) = X  Y the formal product in U ( g). The q-number symbol b c is defined by

q X/2 qX/2
b X c b X cq := , (3.131)
q1/2 q1/2

together with

b X c i b X c qiwith qi q(i ,i )/(, ) , (3.132)


n jnk bnc!
b n c ! : = b m c, := . (3.133)
m =1
m bmc!bn mc!

The exponential functions of generators appearing in bhi c are defined via the corresponding
power series, i.e.

n n
eh = h (3.134)
n=0 n!

with hn hn defined inductively, i.e. hn = h  h(n1) , so that in particular

eh  eh = 1. (3.135)
i
Note that due to the appearance of qh /2 , we are forced to consider infinite power series
in the hi . In contrast, it is consistent to restrict to only finite power series in the ei , f i . In
the limit q 1, the relations Eq. (3.124) to (3.127) reduce to the Lie brackets of g and the
relations Eq. (3.128), (3.129) to the Serre relations of g.

Example: Serre relations for A ji = 1. In this case, the relation Eq. (3.128) is given by

ei ei e j (q1/2 + q1/2 )ei e j ei + e j ei ei = 0, (3.136)

similarly for f i . In the limit q 1, we have (q1/2 + q1/2 ) b2c = 2, so that Eq. (3.136)
reduces to
0 = ei ei e j 2 ei e j ei + e j ei ei = [ei , [ei , e j ]] = (ad(ei ))2 e j . (3.137)
F

We can define the quantum version of the operator ad( X ) as

Ad( E ) E [ E , E ]q := q(,)/4 E  E q(,)/4 E  E . (3.138)

Example: Uq ( A1 ). The relations Eq. (3.128)-(3.129) become more transparent in the


simplest example Uq ( A1 ), where there are no Serre relations. The defining relations of
Uq ( A1 ) are

[h, e] = 2 e, (3.139)
[h, f ] = 2 f , (3.140)
[e, f ] = bhc. (3.141)

The enveloping algebra U ( A1 ) is by definition associative and its unit 1 is given by the
trivial power series 1. These properties are inherited by Uq ( A1 ). Uq ( A1 ) is moreover

FS2016 72
Part 3. Advanced topics: Beyond affine Lie algebras

endowed with the structure of a quasitriangular Hopf algebra. The co-multiplication acts
on the generators h, e, f and on 1 as

(h) = h 1 + 1 h, (3.142)
(e) = e qh/4 + qh/4 e, (3.143)
h/4
( f ) = f q h/4
+q f, (3.144)
(1) = 1 1, (3.145)

and the antipode is

(h) = h, (3.146)
(e) = q1/2 e, (3.147)
1/2
( f ) = q f, (3.148)
(1) = 1. (3.149)

Finally, the co-unit is given by

e(h) = e(e) = e( f ) = 0, (3.150)


e(1) = 1. (3.151)

We need the identity

[qh , e] = (q2 1)e qh , [qh , f ] = (q2 1) f qh (3.152)

to verify the defining property of . The antipode 0 associated to 0 = is obtained


from by 1 (see Eq. (3.92)). The quantum group Uq ( A1 )0 defined by 0 and 0 is
related to Uq ( A1 ) by
Uq ( A1 )0 = Uq1 ( A1 ). (3.153)
F

The Hopf algebra structure of Uq ( A1 ) generalizes in a natural way to Uq ( g) with


arbitrary simple or even affine g. The defining formulae for the co-multiplication are

( hi ) = hi 1 + 1 hi , (3.154)
hi /4 hi /4
( ei ) = ei q +q ei , (3.155)
hi /4 hi /4
( f i ) = f i q +q f i, (3.156)
(1) = 1 1. (3.157)

For the co-unit,

e(hi ) = e(ei ) = e( f i ) = 0, (3.158)


e(1) = 1, (3.159)

and for the antipode,

( hi ) = hi , (3.160)
i h i h
(e ) = q e q , (3.161)
( f i ) = q h f i q h , (3.162)
(1) = 1, (3.163)

73 FS2016
Part 3. Advanced topics: Beyond affine Lie algebras

where
r r
1 1
h =
(, ) >0
h
(, ) >0 i=1
i hi = i hi = (, h), (3.164)
i =1
2
with := (, )
, the Weyl vector of g.
It is straightforward to check that the maps defined above satisfy the defining properties
of a Hopf algebra. To check the co-associativity of , we use
(qh ) = qh qh . (3.165)
Above, we have only presented the Hopf algebra structure for the Chevalley generators. To
identify it also for the rest of the generators, we have to use the Serre relations together
with the homomorphism property of the co-multiplication, etc.

Example: Hopf algebra action on E(1 +2 ) in Uq ( A2 ). Using Eq. (3.138), we find

E(1 +2 ) = q1/4 E1 E2 q1/4 E2 E1 . (3.166)


Using the homomorphism property of , its action on the Chevalley generators and the
commutation relations of Uq ( A2 ), we find
1 + h2 ) /4 1 + h2 ) /4
( E ( 1 + 2 ) ) = E ( 1 + 2 ) q ( h + q(h E ( 1 + 2 ) (3.167)
h1 /4 h2 /4
+ (q1/2 + q1/2 )q E 2 q E 1 ,
1 + h2 ) /4 1 + h2 ) /4
( E(1 +2 ) ) = E(1 +2 ) q(h + q(h E(1 +2 ) (3.168)
h2 /4 h1 /4
+ (q1/2 + q1/2 )q E 1 q E 2 .
Similarly, we find for the antipode
 
( E(1 +2 ) ) = q1 q1/4 E1 E2 q1/4 E2 E1 . (3.169)
F

Literature. The Virasoro algebra is briefly introduced in [FS97], and at much more length
but from a completely different starting point (i.e. conformal field theory) in [DMS97].
The section on Lie superalgebras is summarizing parts of [Cor89], which is very extensive.
The part on quantum groups is taken mostly from [Fuc92], supplemented by [Maj95].

References
[FS97] J. Fuchs and C. Schweigert. Symmetries, Lie Algebras and Representations: A Graduate
Course for Physicists. Cambridge Monographs on Mathematical Physics. Cambridge:
Cambridge University Press, 1997. ISBN: 978-0521541190.
[DMS97] P. Di Francesco, P. Mathieu, and D. Snchal. Conformal field theory. Graduate texts
in contemporary physics. New York: Springer, 1997. ISBN: 0-387-94785-X. URL: http:
//opac.inria.fr/record=b1119694.
[Cor89] J. Cornwell. Group Theory in Physics, Volume III. Supersymmetries and Infinite-Dimensional
Algebras. Techniques of Physics. Academic Press, 1989. ISBN: 0-12-189805-9.
[Fuc92] J. Fuchs. Affine Lie Algebras and Quantum Groups: An Introduction, with Applications
in Conformal Field Theory. 1st. Cambridge Monographs on Mathematical Physics. Cam-
bridge University Press, 1992. ISBN: 0521415934.
[Maj95] S. Majid. Foundations of Quantum Group Theory. Cambridge: Cambridge University
Press, 1995. ISBN: 0-521-46032-8.

FS2016 74
Part 3. Advanced topics: Beyond affine Lie algebras

Epilogue: Lie algebras in integrable


spin chains

Despite having treated a mathematical subject, this course is intended for theoretical
physicists. As this course did not contain any physics examples, let me close by giving three
examples of integrable spin chains which have a simple Lie algebra, a quantum group and
a Lie superalgebra as symmetry algebras. Of course this is just one context in which these
algebras arise, but it is a particularly simple one. We will see that small modifications of a
simple physical problem, such as turning on an external magnetic field or allowing lattice
sites to be empty, changes the structure of the symmetry algebra of the system from the
one of a simple Lie algebra to being a quantum group or a Lie superalgebra.
All these examples are completely solvable and a beautiful subject in themselves,
however we wont touch upon how to actually solve them here.

The XXX1/2 or Heisenberg spin chain. Let us consider a closed linear chain of L identical
atoms with only next-neighbor interactions. Each atom has one electron in an outer shell
(all other shells being complete). These electrons can either be in the state of spin up ()
or down (). At first order, the Coulomb- and magnetic interactions result in the exchange
interaction in which the states of neighboring spins are interchanged:

. (3.170)

In a given spin configuration of a spin chain, interactions can happen at all the anti-parallel
pairs. Take for example the configuration

. (3.171)

It contains five anti-parallel pairs on which the exchange interaction can act, giving rise to
five new configurations. The Hamiltonian of the XXX1/2 spin chain is given by
L
H = J n,n+1 , (3.172)
n =1

where J is the exchange integral1 and n,n+1 is the permutation operator of states at
positions n, n + 1. The spin operator at position n on the spin chain is given by
~Sn = (Snx , Sny , Snz ) = 1~n , (3.173)
2

where ni are the Pauli matrices for spin 1/2:


     
1 0 1 2 0 i 3 1 0
= , = , = . (3.174)
1 0 i 0 0 1
1J > 0: ferromagnet, spins tend to align, J < 0: anti-ferromagnet, spins tend to be anti-parallel.

75 FS2016
Part 3. Advanced topics: Beyond affine Lie algebras

For the closed chain, the sites n and n + L are identified:

~S L+1 = ~S1 . (3.175)

In terms of the spin operators, the permutation operator is given by

n,n+1 = 12 (1 +~n~n+1 ). (3.176)

The Hamiltonian (3.172) now becomes


L
H = J ~Sn~Sn+1
n =1
(3.177)
L
1
Sn+ Sn+1 + Sn Sn++1 + Snz Snz +1 ,

= J 2
n =1

y
where S = Snx iSn are the spin flip operators. The term in parentheses corresponds to
the exchange interaction which exchanges neighboring spin states.
The spin flip operators act as follows on the spins:

Sk+ | . . . . . . i = 0, Sk+ | . . . . . . i = | . . . . . . i,
Sk | . . . . . . i = | . . . . . . i, Sk | . . . . . . i = 0, (3.178)
Skz | . . . ...i = 1
2| . . . . . . i, Skz | . . . ...i = 12 | . . . . . . i.

The spin operators have the commutation relations



[Snz , Sm ] = Sn nm ,
[Sn+ , Sm ] = 2Snz nm , (3.179)

which we recognize as those of A1 . The Hamiltonian moreover commutes with the


generators of su(2),
[ H, S ] = [ H, Sz ] = 0, (3.180)
thus we say that su(2) is the symmetry algebra of the XXX1/2 spin chain and each lattice
site carries a spin 1/2 representation of su(2).

The XXZ1/2 or anisotropic spin chain. In the anisotropic case, a magnetic field is turned
on in the z-direction, resulting in the Hamiltonian
L
Snx Snx+1 + Sn Sn+1 + (Snz Snz +1 41 ).
y y
H = J (3.181)
n =1

The anisotropy is captured by the parameter

q + q 1
= . (3.182)
2
= 1 is the isotropic case we have treated so far.
The XXZ spin chain admits the quantum group Uq (su(2)) as symmetry algebra. Uq (su(2))
z
is generated by S+ , S and qS under the relations
z z
z z q2S q2S
q S S q S = q 1 S , [S+ , S ] = . (3.183)
q q 1

FS2016 76
Part 3. Advanced topics: Beyond affine Lie algebras

These relations reduce to the ones of A1 for q 1. For the case of spin 1/2, we find the
following representations for the operators:
z 3 /2 3
qS = q q /2 , (3.184)
| {z }
L
L L

q|
3 /2 3 3 3
S = Sn = q /2 n q /2 q /2 , (3.185)
n =1 n =1
{z } 2
n 1

where we have the Pauli matrices


     
0 1 0 0 1 0
+ = , = , 3 = . (3.186)
0 0 1 0 0 1

and  
3 q 0
q = . (3.187)
0 1/q

The tJ model. The tJ model describes a system of electrons on a lattice with a Hamiltonian
that describes nearestneighbor hopping (with coupling t) and spin interactions (with
coupling J). Consider a one-dimensional lattice of length L with periodic boundary
conditions. Each site can be either free () or occupied by a spin up () or down ()
electron. Excluding double occupancy, the Hilbert space at each point k is:

Hk = C(1|2) . (3.188)

It is convenient to introduce anticommuting creationannihilation pairs ck,s , ck,s , s = {, }


at each site, acting as

|sik = ck,s |ik , for s = {, }, (3.189)

where |ik is the vacuum, annihilated by ck,s . Let nk,s = ck,s ck,s be the number of s electrons
at position k and nk = nk, + nk, . We can further introduce su(2) spin operators at each
site:

Sk = ck, ck, , Sk+ = ck, ck, , Skz = 21 nk, nk, .



(3.190)

With these ingredients, we can write down the Hamiltonian


" #
L 1    
H = t P ck,s ck+1,s + h.c. P + J ~Sk ~Sk+1 4 nk nk+1 + 2 nk 2
1 1
,
k =1 s=,
(3.191)
where P projects out double occupancy.
For J = 2t = 2, the Hamiltonian is invariant under the action of the Lie superalgebra
sl (1|2).
The even part g0 = gl (1) sl (2) of sl (1|2) is generated by the operators S , Sz , Z with
commutation relations

[Sz , S ] = S , [S+ , S ] = 2Sz , [ Z, S ] = 0 , [ Z, Sz ] = 0 . (3.192)

There are two additional fermionic multiplets Q


s , s = {, } which transform with respect
to g0 as

[Sz , Q 1
s ] = 2 Qs , [S , Q
s ] = 0, [ Z, Q 1
] = 2 Q , [ Z, Q 1
] = 2 Q . (3.193)

77 FS2016
Part 3. Advanced topics: Beyond affine Lie algebras

The fermionic generators satisfy the following anticommutation relations:

{ Q
s , Qs } = 0 , { Q
, Q } = S , { Q z
, Q } = Z S . (3.194)

At each point k of the lattice, also the generators Q , Z can be represented in terms of
creationannihilation operators as

Q Q+ Q
  
k, = 1 nk, ck, , k, = 1 nk, ck, , k, = 1 nk, ck, , (3.195)
+
 1
Qk, = 1 nk, ck, , Zk = 1 2 nk . (3.196)

The supersymmetric Hamiltonian can be expressed in terms of these generators as

L h i
H= Q+ Q
k,s k +1,s + Q +
k +1,s k,s +
Q
k =1 s=,
L
Sk+ Sk+1 + Sk Sk++1 + 2Skz Skz+1 2Zk Zk+1 1k 1k+1 . (3.197)
 
+
k =1

Literature. The XXX1/2 spin chain was introduced by Hans Bethe [Bet31]. The quantum
group structure of the XXZ chain is discussed in [PS90]. The tJ model and its supergroup
structure are discussed in [EK92].

References
[Bet31] H. Bethe. Zur Theorie der Metalle. Zeitschr. f. Phys. 71 (1931), p. 205.
[PS90] V. Pasquier and H. Saleur. Common Structures Between Finite Systems and Conformal
Field Theories Through Quantum Groups. Nucl.Phys. B330 (1990), p. 523.
[EK92] F. H. L. Essler and V. E. Korepin. A New solution of the supersymmetric T-J model by
means of the quantum inverse scattering method (1992).

Acknowledgments
I would like to thank all those students who have made these lecture notes better by
pointing out typos. Special thanks go to Manuel Meyer who has brought me a list of typos
every week.

FS2016 78

You might also like