You are on page 1of 74

Integrable Quantum Spin Chains

Usman Kayani
March 28, 2012
Candidate number: P04349
Supervisor: Dr B Doyon
Kings College London
MSci in Mathematics
Abstract
Overview of Integrable Quantum Spin Chains focusing in particular on the
XXX Heisenberg spin chain with topics covering the coordinate Bethe
Ansatz, the algebraic Bethe Ansatz, integrability, higher conserved charges
and correlation functions. I have adapted and lled in some of the proofs
which are commonly left out in the literature. Material for students at a
masters level.
Contents
1 Introduction 3
1.1 Integrability . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 What is Spin? . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Spin Chains? . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Notation, Conventions and Concepts from QM . . . . . . . . 6
2 Heisenberg Spin Chain 9
2.1 Basis for observables . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Derivation of Hamiltonian . . . . . . . . . . . . . . . . . . . . 10
2.2.1 One-particle wave function N = 1 . . . . . . . . . . . 11
2.2.2 Two-particle wave function N = 2 . . . . . . . . . . . 12
2.2.3 N Spin chain . . . . . . . . . . . . . . . . . . . . . . . 14
2.3 Energy spectrum of H . . . . . . . . . . . . . . . . . . . . . . 17
3 Bethe Ansatz 19
3.1 Coordinate BA . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.1.1 Symmetries of Heisenberg Model . . . . . . . . . . . . 20
3.1.2 Eigenstates of H . . . . . . . . . . . . . . . . . . . . . 22
3.2 BAE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.3 Algebraic BA . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.3.1 Integrability . . . . . . . . . . . . . . . . . . . . . . . . 34
3.3.2 Yang-Baxter equation . . . . . . . . . . . . . . . . . . 34
3.3.3 Spectrum of H . . . . . . . . . . . . . . . . . . . . . . 42
3.3.4 Bethe Ansatz equations . . . . . . . . . . . . . . . . . 45
3.4 Thermodynamic limit N - String Hypothesis . . . . . . 48
3.4.1 XXZ Model . . . . . . . . . . . . . . . . . . . . . . . . 54
4 Higher Conserved Charges of the Heisenberg Hamiltonian 57
4.1 Logarithmic derivative of the transfer matrix . . . . . . . . . 58
5 Correlation functions 62
5.1 EFP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
5.1.1 Integral representation of EFP . . . . . . . . . . . . . 64
5.1.2 Satos formula . . . . . . . . . . . . . . . . . . . . . . 66
1
6 Conclusion 68
References 70
2
Chapter 1
Introduction
The career of a young theoretical physicist consists of treating
the harmonic oscillator in ever-increasing levels of abstraction.
Sidney Coleman
H =
1
2m
p
2
+
1
2

2
q
2
We will begin by briey talking about the quantum harmonic oscillator - a
model which is easy to understand, familiar amongst most graduate students
and will help us understand the idea behind a quantum integrable system.
The quantum harmonic oscillator is one of the most important models in
quantum mechanics, simply because it is one of the few systems for which
an exact, analytic solution is known. Though solving this system directly is
rather straightforward as you essentially have to solve a dierential equation,
it can get quite tedious since we will have to deal with Hermite polynomials
which get increasingly complicated
1
for energy eigenstates above the ground
state. Fortunately there exists another method to extract the energy eigen-
values without solving the dierential equation known as the ladder operator
method. The familiar terminology creation and annihilation operator
that appears in quantum eld theory and other areas of physics comes from
generalizing this simple and elegant concept. It will also be essential in
the Algebraic Bethe Ansatz, where the action of the creation operator is
thought of as a excitation in the chain with respect to the reference ferro-
magnetic ground state. The quantum harmonic oscillator is one example of
an integrable system.
1.1 Integrability
The concept of integrability hails from our old friend classical mechanics
and we will later give a more precise denition of this in terms of conserved
1
Increase in the polynomial order
3
quantities. For now, we will think of integrability as simply meaning we
can solve the system in question exactly i.e. given the Hamiltonian, which
encodes the dynamics of the system we can nd the entire energy spectrum
and other physical quantities such as correlation functions. Other examples
of integrable systems include the hydrogen atom, which we will briey meet
in the next chapter, the sine-Gordon model and of course the model of focus
for the project, the Heisenberg spin chain. The concepts that are important
in Quantum integrable models are the Coordinate Bethe Ansatz and the
Quantum Inverse Scattering Method also known as the Algebraic Bethe
Ansatz due to the strong algebraic framework. The CBA was originally
developed by Bethe in 1931 and it will be the rst method we will use to
nd the energy spectrum, after which we will use the ABA and prove its
integrability.
1.2 What is Spin?
In classical mechanics, we have the notion of orbital angular momentum
associated with the motion of the center of mass. In quantum mechanics, the
orbital angular momentum has a simular structure as in classical mechanics.
There is however another type of angular momentum which arises known as
spin. A fundamental dierence between the two types that the spin can have
both integer and half integer values while only integer values of the orbital
angular momentum are allowed. Spin is a characteristic property of ele-
mentary particles (fermions and bosons), composite particles (hadrons) and
atomic nuclei. The direction of spin can change but an elementary particle
cannot be made to spin faster or slower. For our consideration we will be
looking at Spin-1/2 particles i.e. the electron. Particles (like the electron)
with Spin posses a magnetic dipole moment, just like a rotating electrically
charged body in classical electrodynamics. As a result, the movement of
individual electrons in their atomic nuclei will generate a weak magnetic
eld.
Conventionally the direction chosen for the quantization axis is the z-
axis, but this choice is arbitrary and chosen for convenience as we shall see
shortly. The spin projection operator S
z
aects a measurement of the spin
in the z direction and the eigenstates [1, 0]
t
and [0, 1]
t
form a complete basis
for the Hilbert space describing a spin-1/2 particle, which we can call up
4
and down. This means that we can essentially write any eigenvector by
taking linear combinations of the above eigenstates and hence describe all
possible states of the spin. It is important to note that by the postulates of
quantum mechanics, an experiment designed to measure the electron spin
on the x, y or z axis can only yield an eigenvalue of the corresponding spin
operator (S
x
, S
y
or S
z
) on that axis. A measurement of the z-component
of spin destroys any information about the x and y components that might
previously have been obtained.
1.3 Spin Chains?
A quantum spin chain is a particular example of an integrable system usu-
ally in 1+1 space-time dimensions for the purposes of integrability. Picture
a ring of atoms each which posses a quantum degree of freedom (spin), which
with our convention can point in two directions, up or down and which form
a complete basis for the Hilbert space describing the spin-1/2 particle. The
study of spin chains naturally arises in the study of magnetism (particularly
ferromagnetism). Ferromagnetism is the basic mechanism which certain ma-
terials (such as Iron) form permanent magnets or are attracted to magnets.
An example of ferromagnetism is a refrigerator magnet used the hold notes
on the refrigerator door. It is the only type that creates forces strong enough
to be felt and is responsible for the common phenomena of magnetism en-
countered in everyday life. Though the study of spin chains arose to solve
particular problems in condensed matter physics, its application has also
reached String theory particularly in the AdS/CFT correspondence and in
high-energy QCD scattering.
In ferrous materials such as Iron, the individual electrons that generate
their weak elds will combine to generate a strong magnetic eld and thus
what we experience as ferromagnetism. At absolute zero temperature, a
ferromagnet reaches the state of lowest energy (ground state energy) in
which all the atomic spins align. As the temperature increases more and
more spins deviate from the common direction which can be thought of
excitations from the ground state caused by a magnon.
2
Thus there is an
increase in the internal energy and reducing the net magnetization. Thus
2
The concept of a magnon was introduced in 1930 by Felix Bloch in order to explain
the reduction of spontaneous magnetization in a ferromagnet. A magnon can be thought
of as a quantized spin wave. We will introduce the concept of a magnon in chapter 3 in
terms of the low-lying excitations from the ground state.
5
magnetism arises because the state in which all atomic spins align has less
energy than disaligned spins. We will begin by considering the diatomic
Hydrogen Molecule H
2
which contains two Hydrogen atoms
3
. What we
nd is that the total internal energy of the system is actually greater than
the sum of the individual atomic energies. This extra energy comes from
the Coloumb interaction between the protons and electrons. By invoking
the Pauli exclusion principle for the total wave function and making a few
arguments, we will then derive the main system of study in this project -
the XXX Heisenberg Hamiltonian. This Hamiltonian describes the nearest-
neighbour interaction between electron spins of a linear chain of N atoms.
1.4 Notation, Conventions and Concepts from QM
Unless stated otherwise we will use natural units i.e. set = 1.
For Spin-1/2 particles, we can represent the Spin operators in terms of
Pauli matrices:
S

=
1
2

, x, y, z

x
=
_
0 1
1 0
_
,
y
=
_
0 i
i 0
_
,
z
=
_
1 0
0 1
_
[
a
,
b
] = 2i
abc

c
, (
abc
= 1)

a
,
b
= 2
ab
I

b
=
ab
I +i
abc

c
tr

= 0
It is convenient to specify the eigenkets of

S
2
and S
z
, given a state with
spin s and z-component m of the spin:

S
2
[s, m = s(s + 1)[s, m
S
z
[s, m = m[s, m
The possible values for s and m are:
s = 0,
1
2
, 1,
3
2
, , m = s, s + 1, , s 1, s
In many applications there is no confusion about the spin s of the particles
and therefore the index is dropped. For a spin-1/2 particle s = 1/2, we have
3
Each with one electron and one proton
6
two possible states [1/2 and [ 1/2 which is commonly denoted as [
and [ .
Tensor product
Let A and B be two n n matrices with elements A
i
j
, B
k
l
. The upper
index denotes a row and the lower index a column. The tensor product
AB is a n
2
n
2
matrix elements:
(AB)
ik
jl
= A
i
j
B
k
l
Where i, j are the block indices and k, l the intrinsic indices. For example
if A and B are 2 2 matrices with
A =
_
a
11
a
12
a
21
a
22
_
, B =
_
b
11
b
12
b
21
b
22
_
Then:
AB =
_
a
11
B a
12
B
a
21
B a
22
B
_
=
_
_
_
_
a
11
b
11
a
11
b
12
a
12
b
11
a
12
b
12
a
11
b
21
a
11
b
22
a
12
b
21
a
12
b
22
a
21
b
11
a
21
b
12
a
22
b
11
a
22
b
12
a
21
b
21
a
21
b
22
a
22
b
21
a
22
b
22
_
_
_
_
Properties:
det (AB) = (det (A))
2
(det (B))
2
tr (AB) = tr (A) tr (B)
(AB)

= A

(A+C) B = AB +C B
A(B +C) = AB +AC
(AB)(C D) = AC BD
We also have the notion for a vector tensor product, given x, y V = C
2
:
x y =
_
x
1
x
2
_

_
y
1
y
2
_
=
_
x
1
y
x
2
y
_
=
_
_
_
_
x
1
y
1
x
1
y
2
x
2
y
1
x
2
y
2
_
_
_
_
which is an element of the tensor product space V V = C
2
C
2
Let V
1
V
2
V
N
be a d
N
- dimensional vector space. We can
interchange V
j
with V
k
by means of the permutation matrix P
j,k
:
P
j,k
(V
j
V
k
) = (V
k
V
j
)
7
with property P
2
j,k
= 1. When V
j
are a 2-dimensional space e.g. C
2
as will
be the case for the discussion in this project, the permutation matrix can
be represented by:
P =
1
2
(I I +

)
which is a permutation in C
2
C
2
i.e. Given x, y C
2
we have:
P(x y) = y x
Now if we use a basis for C
2
C
2
which we will dene in the next section,
we can see that the permutation operator P can be written explicitly as:
P =
_
_
_
_
1 0 0 0
0 0 1 0
0 1 0 0
0 0 0 1
_
_
_
_
Note: P
i,j
permutes the spin at site i and j. We also have the following
properties:
tr
a
P
a,b
= tr
b
P
a,b
= I
P
a,b
= P
b,a
P
n,p
P
n,q
= P
p,q
P
n,p
= P
n,p
P
p,q
P

= P P
2
= I
The notation for any operator A End( ), where is the tensor
product of many vector spaces, means that the operator acts as A in the
tensor product space and trivially in any other space.
It will be useful to dene the Riemann zeta function (s) since it will be
used for the correlation functions in Chapter 5. The Riemann zeta function
is dened as:
(s) =

n=1
1
n
s
, 1(s) > 1
Alternative the Riemann zeta function can be expressed in terms of the
alternating zeta series
a
as:
(s) =
1
1 2
1s

a
(s), s ,= 1,
where

a
(s) =

n>0
(1)
n1
n
s
.
The alternating zeta function is well dened for s = 1 and has value
a
(1) =
ln 2 unlike the Riemann zeta function which has a pole at s = 1.
8
Chapter 2
Heisenberg Spin Chain
In this chapter, we shall derive the Hamiltonian for the Heisenberg Spin
Chain by considering the Coloumb Hamiltonian. First we will state the set
of observables and the vector space on which the operators act. The material
in this chapter is adapted from [1],[2],[3] and [4]. An observable is, roughly
speaking, any measurable property of a physical system such as: position,
spin, energy, momentum or angular momentum. In Quantum Mechanics,
each observable is represented by a maximally Hermitian linear operator
acting on the state space. Each eigenstate of an observable corresponds to
an eigenvector of the operator, and the associated eigenvalue corresponds to
the value of the observable in that eigenstate. The observable in question
for the next part is given by the spin operators. For example, for a spin-1/2,
S
z
aects the measurement in the z direction and has eigenvalues /2,
where if we set = 1 simply gives 1/2. These correspond the eigenstates
pointing up or down, as we shall see shortly.
2.1 Basis for observables
For N = 1, the case where there is one particle in the chain is trivial in the
context of the Spin Chain. It is still important to state the set of observables
as we will use this to construct the set of observables for N > 1.
The basis for observables for N = 1 are given by the identity I and the
rmilar Pauli Matrices , that is: I, = I,
x
,
y
,
z

These operators act on a two-dimensional complex vector space V



= C
2
,
for our consideration these will operate on the space spanned by the up
and down Spinors:
[ =
_
1
0
_
, [ =
_
0
1
_
For N = 2 we need to construct our basic observables by using the notion
of a tensor product. The basic observables at each site are:
1
I and
9

2
I . The basis for the observables at each site are therefore given by:
I
2
,
i
for i = 1, 2. where I
2
= I I.
These act on the tensor product space V
1
V
2

= C
2
C
2
, again for our
consideration these correspond to the space spanned by the vectors:
[ = [ [ , [ = [ [ , [ = [ [ , [ = [ [
For general N the basic observables
n
, n = 1 , 2 , . . . , N are dened by:

n
=
1

I
I
n

I
N

I
,
which acts nontrivially on the n
th
space and trivially on the rest. A basis
for the observables at each site are therefore given by: I
N
,
i
for i =
1, 2, , N. These are operators which act on the Hilbert (tensor product)
space:
1
N
=
1

V
1

n

V
n

N

V
N

= C
2
N
,
with dimension dim(1
N
) = 2
N
. This corresponds to the space spanned by
the 2
N
elements:
[ , [ , [ , [ , , [ ,
which consists of the complex linear superpositions of all the dierent pos-
sible states.
2.2 Derivation of Hamiltonian
In this section, we will derive the Hamiltonian for the Heisenberg Spin Chain
based on considerations of the Coulomb interaction Hamiltonian between
idealized charged bodies. There are many variants of this derivation (See
[1],[2],[3] for a formal treatment) that can be found in numerous papers
and textbooks and as such I will give quick overview based on my own
understanding. Since we are talking about spin-1/2 particles, it is useful to
consider the electron. We will particularly be interested in how the total
energy is aected by the electron dynamics. By denition, atoms have no
overall electrical charge and so must contain an equal number of protons and
electrons.
1
For this reason, when we study electron systems in atoms and
solids we must also take into account the various electrostatic interactions
between the electrons and protons, we will see how this comes into play
when we deal with two-particle wave function (atoms with one un-paired
electron next to each other). Well-separated atoms are described by atomic
1
An atom can gain or lose electrons, becoming an ion which is nothing more than an
electrically charged atom. Adding or removing electrons from an atom does not change
the element, just its net charge.
10
wave functions, but in a molecule or solid the wave functions overlap. The
amount they overlap will depend on the spin orientation of the electrons,
if they posses the same spin then by the Pauli exclusion principle the two
electrons cannot be in the same quantum state and so their wave functions
will not remix as much. If the spin is dierent between the electrons, their
wave functions will overlap more and the electrons have a greater probability
of being near to each other, this means that electrons with opposite spins
posses a higher potential energy. The more the wave functions overlap, the
greater the Coloumb repulsion between them - this is why the ground state
for a ferromagnetic tends to have all parallel spins as they tend to align.
2.2.1 One-particle wave function N = 1
Firstly, let us consider a one-electron system (i.e. the Hydrogen atom). This
is a famous integrable model in physics for which the solution can be found
exactly. The Hamiltonian will be given by the kinetic energy of the electron
and the attractive Coloumb interaction between the proton and the electron.
We will neglect the kinetic energy associated with the motion of the protons,
assuming that they are relatively stationary because of their large mass.
Figure 1: The Hydrogen atom
We will introduce some labelling which will be useful later on. For each
electron we will associate a number (n E = 1, 2, ) and for each
proton we will associate a letter ( P = A, B, ). We will denote
the position of each electron by r
n
R
3
, and the position of each proton
by r

. We will be particularly working with the magnitude of the distance


between the particles i, j, this will be denoted by r
ij
= [r
ij
[ = [r
i
r
j
[ where
i, j E, P. The kinetic energy for each electron will be given by
h
2
2m

2
n
while the protons will be assumed to be stationary (relative to the electrons),
the coloumb interaction term between two such particles will be given by
k
e
2
r
ij
where the sign depends on whether it is repulsive or attractive. For
example, for the one-particle wave function we have one-electron and one-
proton with positions r
1
and r
A
, with relative magnitude r
1A
= [r
1
r
A
[,
in this case since there are no other particle labels to consider we can simply
call this r = r
1A
for convenience. See Figure 1. The Hamiltonian for this
11
system will be given by:
H
0
(r) =
h
2
2m

2
k
e
2
r
Where k = 1/4
0
. The wave function
2
describing this system will be given
as a solution to H
0
= E
0
and (r) L
2
(R
3
).
The total wave function
tot
for the particle is an element of the Hilbert
space given by the tensor product of the spatial and spin wave function
spaces, i.e. the Hilbert Space is: H
n
= L
2
(R
3
) V
n
. Here V
n
is a vector
space spanned by [ and [ as in the previous section.
2.2.2 Two-particle wave function N = 2
Figure 2: The Helium atom
Now we wish to consider a two-electron system. One such system is
the Helium atom, which contains a doubly positively charged nucleus. See
Figure 2.
Figure 3: The Hydrogen
Molecule (H
2
), we will assume

R
AB
to be xed.
For our discussion, since we would like to study the nearest neighbour
interaction between a linear chain of atoms which will give rise to magnetism,
we will consider the more general diatomic Hydrogen molecule (H
2
). It is
the smallest molecule that consists of two Hydrogen atoms. Like Helium, a
Hydrogen molecule also has two electrons, and so the intermolecular forces
are going to be small - but not as small as Helium. In the Hydrogen molecule,
you have two atoms that you can distribute the charge over.
The Hamiltonian will be given by the sum of the Hamiltonians for the
one-electron system (for each of the two Hydrogen atoms) and an additional
2
The wave function can be determined by the time-independent Schrdinger equation,
which is the eigenvalue equation for the Hamiltonian
12
term known as the interaction or exchange Hamiltonian, which will encode
the electrostatic Coloumb interaction between the electrons and protons, and
each proton with each corresponding neighbouring electron. See Figure 3.
The Hamiltonian is:
H = H
0
(r
1A
) +H
0
(r
2B
) +H
exch
where
H
0
(r
n
) =
h
2
2m

2
n
k
e
2
r
n
and
H
exch
= k(
e
2
r
1B

e
2
r
2A
+
e
2
r
12
+
e
2
r
AB
)
and the unperturbed hydrogen atom wave functions
1
and
2
satisfy
the equations:
H
0
(r
n
)
n
(r
n
) = E
0

n
(r
n
)
We can use these as building blocks to construct symmetric and an-
tisymmetric wave functions using the Heitler-London approximation
3
(see
[3]):

S
(r
1
, r
2
) =
1

2
(
1
(r
1
)
2
(r
2
) +
2
(r
1
)
1
(r
2
))

A
(r
1
, r
2
) =
1

2
(
1
(r
1
)
2
(r
2
)
2
(r
1
)
1
(r
2
))
or in compacted notation:

(r
1
, r
2
) =
1

2
(
1
(r
1
)
2
(r
2
)
2
(r
1
)
1
(r
2
))
where
1
and
2
are wave functions in L
2
(R
3
) determined by H
0
. We can
calculate this energy by using:
E

[H

We can calculate this by plugging in Hamiltonian and the energy for the
unperturbed wave functions. Firstly, calculating the normalization integral
in the dominator of E

_
=
__

(r
1
, r
2
)

(r
1
, r
2
) d
3
r
1
d
3
r
2
= 1
2
3
Based on the Calculation made by Walter Heitler and Fritz London on the hydrogen
molecule (H
2
) in 1927, it plays an important role in the context of quantum chemistry.
13
Where we have used:
_
[
i
(r)[ d
3
r = 1 i = 1, 2
and is dened to be the overlap integral:

_

1
(r)
2
(r) d
3
r
The numerator in expression for E

can be found in a simular manner,


after some manipulation:

_
=
__

(r
1
, r
2
) H

(r
1
, r
2
) d
3
r
1
d
3
r
2
= 2E
0
(1
2
) +V U
E
0
is the one- particle energy dened as before. V and U are called the
Coulomb integral and exchange integral, respectively, dened by:
V
_
H
exch
[
1
(r
1
)[
2
[
2
(r
2
)[
2
d
3
r
1
d
3
r
2
U
_

1
(r
2
)

2
(r
1
) H
exch

1
(r
1
)
2
(r
2
) d
3
r
1
d
3
r
2
The energy is therefore given by:
E

= 2E
0
+
V U
1
2
Note that , V and U need to be calculated. This result shows that
the total energy of the system is in fact greater than the individual atomic
energies.
2.2.3 N Spin chain
Since we are only considering the nearest neighbour interaction, it suces to
consider the interaction for N = 2, then to generalise for any N by summing
up the interaction for all possible nearest neighbours. The Hilbert space
for the total wave function of the Spin Chain will be given by the tensor
product of all the individual Hilbert spaces for each particle, i.e. for N = 2
the total wave function
tot
is an element of H
tot
= H
1
H
2
= h
spatial

h
spin
, where h
spatial

= L
2
(R
3
) L
2
(R
3
) is the spatial component of H
tot
,
and h
spin

= V
1
V
2

= C
4
is the spin component of H
tot
.
The Pauli exclusion principle states that the total wave function for two
identical fermions is anti-symmetric with respect to the exchange of the
particles.
4
This means when we interchange H
1
with H
2
the total wave
function is replaced by its negative. The total wave function
tot
is an
4
Denition taken from wiki
14
eigenstate of the total Hamiltonian H
tot
and therefore we can express it
as the product
spatial

spin
, where
spatial
is clearly an eigenstate of
H
tot
in h
spatial
and
spin
a vector in h
spin
. We will have symmetric and
antisymmetric wave functions for both the spatial and spin components,
we have already introduced the symmetric and antisymmetric spatial wave
functions from the Heitler-London approximation.
The symmetric and antisymmetric spin wave functions are given by:

sym
spin
=
S
(s
1
, s
2
) = [ ,
1

2
([ +[ ), [

anti
spin
=
A
(s
1
, s
2
) =
1

2
([ [ )
These are also called the triplet and singlet states. Note these spin wave
functions are eigenstates of the total spin operator

S
2
, with eigenvalues 2
and 0 respectively.
It is clear from the Pauli exclusion principle that the symmetric spatial
wave function will be associated to the antisymmetric spin wave function and
vice versa, in order for the total wave function to always be antisymmetric
with respect to the exchange. This will allow us to associate each spin wave
function with an energy (calculated from the corresponding spatial wave
function) and hence construct the spin Hamiltonian H
spin
. This will be
formally equivalent to the Coloumb Hamiltonian H and will give the same
energy spectrum.
We can calculate the energy of the spatial wave functions, this will give
E
+
for the symmetric and E

for the antisymmetric spatial wave function as


we have shown. Note these are also the energies of the total wave function,
since H acts only on the spatial wave functions. We can dene the exchange
energy by J = E
+
E

. We can now dene the spin Hamiltonian to be:


H
spin
= E
+
+
1
2
(E

E
+
)

S
2
= E
+

J
2

S
2
Where
J = 2
V
2
U
1
4
Let us comment on H
spin
, it will have energy E
+
for the antisymmetric spin
wave function and E

for the symmetric spin wave function. As required,


this Hamiltonian will produce the same energy spectrum as H but is inde-
pendent of the spatial wave function and hence of the coordinate system for
the spatial component. We have formally replaced a Hamiltonian that de-
pends only on the spatial component by a Hamiltonian which depends only
on the spin component - by virtue of the Pauli exclusion principle with the
Coloumb potential. We can expand the operator

S
2
= (S
1
+S
2
) (S
1
+S
2
),
15
using the properties of the spin operator and the fact that we have s = 1/2
for our consideration we obtain:
H
spin
=
1
4
(E
+
+ 3E

) JS
1
S
2
Since we measure energy dierences (energies are relative), we can ignore
the rst term that appears in the Hamiltonian as it is a constant.
H
spin
= J

S
1


S
2
Now we consider the Spin chain. In a magnetically ordered solid with
no external magnetic eld, we will sum this Hamiltonian over all pairs of N
particles. We thus obtain the famous Heisenberg spin Hamiltonian:
H =

i,j
J
ij

S
i


S
j
The Heisenberg model works in higher dimensions as well such as in a lattice
of spins but we restrict our attention to 1 + 1 space-time dimensions for
the purposes of integrability. More generally, a spin chain is a chain of
local degrees of freedom with a local Hilbert space of low dimension, with
a Hamiltonian that represents the interaction between a few neighbours
everywhere along the chain. The Heisenberg spin chain is a special case of
this where we consider the interaction between nearest neighbouring spins.
One can thus simplify the Hamiltonian by assuming the exchange integral
J
ij
decreases rapidly with increasing distance between the two electrons,
essentially being zero except for nearest neighbours. We will also assume
that the exchange integral will be equal for all nearest neighbour pairs of
electrons. Note we can expand

S
j


S
j+1
= S
x
j
S
x
j+1
+ S
y
j
S
y
j+1
+ S
z
j
S
z
j+1
=
S

j
S

j+1
, giving:
H = J
N

j=0

S
j

S
j+1
= J
N

j=0
(S
x
j
S
x
j+1
+S
y
j
S
y
j+1
+S
z
j
S
z
j+1
) = J
N

j=0
S

j
S

j+1
This is called the XXX Heisenberg Spin chain for N particles. Note, one
can also write this in terms of the Pauli matrices as S

n
=

n
/2 for = x, y, z
if one wishes, The Hamiltonian then becomes
H =
J
4
N

j=0

n+1
We will however stick to writing the Hamiltonian in terms of the spin op-
erator S

n
in what follows. The sign of the exchange integral J indicates
whether parallel or antiparallel spins are energetically preferred, that is if
J > 0 means we have a parallel (ferromagnetic) spin coupling and for J < 0
16
means we have a antiparallel (antiferromagnetic) coupling. The ground state
for a ferromagnetic will be given by all spins being parallel (i.e. all spins
pointing in the same common direction e.g. up or down, or a linear com-
bination) and the ground state for an antiferromagnet will be the highest
excited state of the corresponding ferromagnetic. One can generalise the
Hamiltonian to an anisotropic arrangement, where the magnetization is dif-
ferent in all directions. This gives the generalized anisotropic Heisenberg
Hamiltonian:
H =
N

j=0
(J
x
S
x
j
S
x
j+1
+J
y
S
y
j
S
y
j+1
+J
z
S
z
j
S
z
j+1
)
Which is also known as the XY Z Heisenberg Spin Chain. When there is
one preferred direction, we align this direction with the z axis (J
x
= J
y
) and
we obtain the XXZ Heisenberg Spin Chain. The XXX Spin Chain is an
isotropic special case of this Hamiltonian (J
x
= J
y
= J
z
), in this report I will
mainly be discussing the XXX Spin Chain and briey stating results for
the XXZ that can be obtained in a simular manner to the XXX spin chain
via the Algebraic Bethe Ansatz. For my project, I will be using the above
representation of the Hamiltonian for the XXX spin chain. It is sometimes
common however in the literature to subtract a constant proportional to the
identity matrix, this will not change the dynamics of the system and merely
brings shifts the energy, in particular the energy of the ground state to 0.
H = J
N

j=0
(S
x
j
S
x
j+1
+S
y
j
S
y
j+1
+S
z
j
S
z
j+1
1)
Also, when we talk about a closed chain we will also impose the periodic
boundary condition S

N+1
= S

1
.
2.3 Energy spectrum of H
Now that we have derived the XXX Heisenberg Hamiltonian, we can work
towards the problem of nding the energies of the given system. That is,
we wish to solve:
H[ = E[
This can of course be done by direct diagonalization of the matrix H,
but the matrix will become large very rapidly for increasing N. The Hilbert
space will be of dimension 2
N
and the matrix will be of size 2
N
2
N
. More
over, direct diagonalization is not possible when N when only analytic
methods will work.
17
We seek a more elegant method than direct diagonalization, one such
method is called the Bethe Ansatz which was formulated by Bethe in 1931
(See [5]).
18
Chapter 3
Bethe Ansatz
In this chapter we will examine the problem of diagnolozing the Hamiltonian
by using the Coordinate Bethe Ansatz (CBA) and the using the Algebraic
Bethe Ansatz (ABA). The material on the rst part of the chapter is based
on [5],[6] and [7].
Let us introduce the spin ip operators S

n
S
x
n
iS
y
n
, these do exactly
what it says on the tin. S
+
n
ips then
th
down spin to up (assuming it is
down in the rst place), and S

n
does the opposite. As a matrix acting on
C (i.e. N = 1):
S
+
=
_
0 1
0 0
_
, S

=
_
0 0
1 0
_
We get S

n
by applying the appropriate tensor product, as dened in Chapter
2. We will write the Hamiltonian in terms of the spin ip operators:
H = J
N

n=1
S

n
S

n+1
= J
N

n=1
_
1
2
_
S
+
n
S

n+1
+S

n
S
+
n+1
_
+S
z
n
S
z
n+1
_
We will keep the sign of J general unless specically stated otherwise, though
our considerations will be made for the ferromagnet J > 0 and our vacuum
(ground state) will be appropriately dened with this in mind. The physical
properties of the anti-ferromagnet are very dierent from the ferromagnet,
but for the purposes of calculating the eigenstates and energy eigenvalues
are interchangeable (see comments below).
[
n
[
n
S
+
n
0 [
n
S

n
[
n
0
S
z
n
1
2
[
n

1
2
[
n
Table 3.1: Action of spin operators on the basis vectors [
1
. . .
N
with
n
=, .
[
n
= [ . . . . . ., [
n
= [ . . . . . ..
19
The spin operators also satisfy the following su(2) spin algebra commu-
tation relations:
[S
a
n
, S
b
m
] = i
n,m

abc
S
c
m
, [S
z
m
, S

n
] =
n,m
S

m
, [S
+
m
, S

n
] = 2
n,m
S
z
m
with a, b, c = x, y, z. For the ferromagnet J > 0, the ground state of the
system will be given by all the spins being aligned (See Chapter 1), we will
choose this to be [0 = [ . . . =
N
n=1
[
n
. We will dene the basis
vectors for r down spins to be given by:
[m
1
, m
2
, . . . , m
r
= [
1
[
m
1
[
m
r
[
N
= S

m
1
S

m
2
. . . S

m
r
[0 = S

M
[0
Where M = m
1
, m
2
, . . . , m
r
and m
i
N i = 1, . . . , r denotes the
position of the down spin with respect to the ground state.
A general eigenstate [ will be constructed as a linear combination of
these basis vectors, we will later see this more concretely and in particular
the exact form of the general eigenstate in accordance with the Bethe Ansatz.
It can also be shown (see [7]) that the eigenvalues of H are at least JN/4
which will infact turn out to be the ground state energy of H for J > 0.
Though the sign of the coupling constant J will not aect the diagonal-
ization procedure in what follows, it is important to make a few comments
before we begin. For the anti-ferromagnet J < 0, [0 is not a true ground
state (vacuum) of the system i.e. a state with the lowest energy, it is actually
the state of highest energy due to the change of sign. Instead it is a known
as a pseudo-vacuum and it will nevertheless be useful as a reference state for
the true vacuum. The true ground state is non-trivial in terms of excitations
in the pseudovacuum and requires some work to be identied. All the eigen-
vectors remain the same for both the ferromagnet and anti-ferromagnet, but
the energy eigenvalues have opposite sign.
3.1 Coordinate BA
We will start by using the Coordinate Bethe Ansatz which was originally
developed by Bethe in 1931 (See [6]), which will essentially allow us to write
the general eigenstate of H in an elegant form. We will investigate the solu-
tions to the Heisenberg Hamiltonian, by considering the spin deviation from
the ground state. For completeness, I will go through a simular approach
taken in [6], rst by considering the low-lying states (r = 1, 2) then leading
up the general case of r down spins. I will use the same notation as in [5]
to discuss these low-lying states in terms of the spin ip operator.
3.1.1 Symmetries of Heisenberg Model
We will begin by examining the symmetries of the Heisenberg Hamiltonian,
which will be essential in the application of the Bethe Ansatz. The model has
20
a full rotational SU(2) symmetry, though we will only require the rotational
symmetry about the z-axis in spin space, which will take as the quantization
axis.
Claim: The conserved quantity which arises from this symmetry is the
z-component of the total spin S
z
= S
z
T
=

N
n=1
S
z
n
with quantum number
S
z
T
= N/2 r
Proof. We need to consider the action of S
z
on the basis vectors [m
1
, m
2
, . . . , m
r
,
by using the rules of Table 1.1:
S
z
[m
1
, m
2
, . . . , m
r
=
N

n=1
S
z
n
[m
1
, m
2
, . . . , m
r

=
N

n=1
S
z
n
S

M
[0
S
z
n
S

M
[0 =
_
S

M
[0/2 n M
S

M
[0/2 n / M
Note for n = 1, . . . , N, it is evident that n M r times since we have r
down spins by denition of the basis vector and n / M N r times.
N

n=1
S
z
n
S

M
[0 =
_
N r
2

r
2
_
S

M
[0
=
_
N
2
r
_
[m
1
, m
2
, . . . , m
r

It is now easy enough to show [H, S

] = 0, for which = z gives the result:


[H, S

] = J
N

n,m=1
[S

n
S

n+1
, S

m
] = J
N

n,m=1
_
[S

n
, S

m
]S

n+1
+S

n
[S

n+1
, S

m
]
_
= iJ
N

n,m=1
(
n,m

n
S

n+1

n+1,m

n
S

n+1
) = 0
21
One can also verify this by a direct computation, by using what we
have shown above for the quantum number of S
z
. Given an eigenstate [
constructed from the basis vectors, such that H[ = E[ and S
z
[ =
_
N
2
r
_
[:
[H, S
z
][ = (HS
z
S
z
H) [
=
_
H
_
N
2
r
_
S
z
E
_
[
=
_
E
_
N
2
r
_
E
_
N
2
r
__
[
= 0
Since [ is an arbitrary eigenstate, we conclude that [H, S
z
] = 0 as required.
The model also has a discrete translational symmetry i.e. the invariance
of H with respect to discrete translation by any number of lattice spacings.
This can be realised by considering the translation operator T which is
dened by:
T[m
1
, m
2
, . . . , m
r
= [m
1
1, m
2
1, . . . , m
r
1 = [m
2
, m
3
, . . . , m
r
, m
1

Which basically shifts the spin chain by one position. Any eigenvector [ of
T will satisfy T[ = e
ip
[ for some momentum p. In fact the interpretation
of p as momentum follows from the eigenvalue of T. It is a conserved quantity
and hence commutes with H, i.e. [H, T] = 0. These are the important
symmetries for the application of the Bethe Ansatz. The model also has a
Reection symmetry on the lattice.
3.1.2 Eigenstates of H
Let us rst discuss the case where r = 0, this will be given by the (pseudo)-
vacuum of H which dened to be [0 = [ . . . at the beginning of the
chapter. Note this is a arbitrary choice and we could have just as well chosen
all spins down which will give the same energy. This can be understood in
terms of the SU(2) symmetry of the model, since rotating the system will
leave the system invariant.
Claim: It is an eigenstate H[0 = E
0
[0, with energy E
0
= JN/4.
22
Proof. Using Table 1.1 the result is evident since we immediately have:
H[0 = J
N

n=1
_
1
2
_
S
+
n
S

n+1
+S

n
S
+
n+1
_
+S
z
n
S
z
n+1
_
[0
= J
N

n=1
S
z
n
S
z
n+1
[0
= J
N

n=1
1
4
[0 =
JN
4
[0
An excitation in the chain will be thought of as a spin deviation from
the ground state, i.e. r = 1 corresponds to an excited state with one spin
down with respect to the ground state. The N basis vectors in the invariant
subspace with one down spin are labelled by the position of the ipped spin:
[n = S

n
[0 n = 1, . . . , N
Where S

n
is the spin ip operator that changes a spin up to a spin down
at position n. These are clearly not eigenstates of H, but one can construct
a translationally invariant eigenvector from the [n by writing:
[ =
N

n=1
e
ipn
[n
With momentum p = 2m/N, m = 0, 1, . . . , N 1. This can also be
understood in terms of the discrete fourier transform of the position of the
spin down (in analogy with the continuum [p =
_
dxe
ipx
[x from quantum
mechanics).
Claim: These vectors are eigenvectors of the translation operator T
with eigenvalue e
ip
and eigenvectors of H with eigenvalue (energy) E =
E
0
+J(1 cos k).
23
Proof. First we will show the eigenvalue of T:
T[ = T
N

n=1
e
ipn
[n
=
N

n=1
e
ipn
T[n
=
N

n=1
e
ipn
[n 1
=
N

n=1
e
ip(n+1)
[n
= e
ip
N

n=1
e
ipn
[n
= e
ip
[
As required, [ is an eigenstate of T with eigenvalue e
ip
. Now we will show
the eigenvalue of H. Again, we will show this by applying the rules of Table
1.1 in a systematic manner:
H[ = J
N

n=1
_
1
2
_
S
+
n
S

n+1
+S

n
S
+
n+1
_
+S
z
n
S
z
n+1
_
[
= J
N

m=1
e
ipm
N

n=1
_
1
2
_
S
+
n
S

n+1
+S

n
S
+
n+1
_
+S
z
n
S
z
n+1
_
[m
= J
N

m=1
e
ipm
_
1
2
[m+ 1 +
1
2
[m1 +
_
N
4
1
_
[m
_
= J
N

m=1
_
1
2
e
ip(m1)
+
1
2
e
ip(m+1)
+
_
N
4
1
_
e
ipm
_
[m
= J
N

m=1
_
1
2
e
ip
+
1
2
e
ip
+
_
N
4
1
__
e
ipm
[m
= J
_
1
2
e
ip
+
1
2
e
ip
+
_
N
4
1
__
N

m=1
e
ipm
[m
= (E
0
+J(1 cos p))
N

m=1
e
ipm
[m
= (E
0
+J(1 cos p)[ = E[
By using the identity cos p = (e
ip
+ e
ip
)/2 and E
0
= JN/4. Note, we
have also used the periodicity condition which allows us to make shifts of
24
the summation variables, which in this case is simply m to bring all the
states to the uniform expression [m.
Hence we have the result:
E E
0
= J(1 cos p)
Since we have shown that [ is an eigenstate of T with eigenvalue e
ip
, by
denition it follows that the momentum of the excitation can be interpreted
as p. The expression above for the eigenvalue (energy) of H is also known as
the dispersion relation between energy E and momentum p. An excitation
of the spin chain around the (pseudo)-vacuum [0 carrying momentum p is
called a magnon, which we touched upon briey in the introduction. Thus
a magnon can be thought of a particle with momentum p = 2m/N, m =
0, 1, . . . , N 1 and energy as given above.
For r > 1 this technique of nding the translationally invariant eigenvec-
tor by exploiting the symmetries will fail as it will not completely diagonalize
the Hamiltonian. For this reason we will need to use the Bethe Ansatz as a
powerful alternative.
For a general eigenstate with r spins down, assuming that the particles
with numbers m
1
, m
2
, . . . , m
r
corresponds to xing which atoms point down
[5]:
[ =

m
1
,m
2
,...,m
r
a(m
1
, m
2
, . . . , m
r
)[m
1
, m
2
, . . . , m
r

where we can write [m


1
, m
2
, . . . , m
r
= S

m
1
S

m
2
. . . S

m
r
[0. The following
claim was included in the original Bethe paper [6], but the proof is missing
from the literature in general.
Claim: If we apply the Hamiltonian on this general eigenstate, we will
obtain the following consistency equations for the coecients a(m
1
, m
2
, . . . , m
r
):
2(E E
0
)a(m
1
, . . . , m
r
) = J

1
,...,m

r
[a(m
1
, . . . , m
r
) a(m

1
, . . . , m

r
)] (*),
where particles with numbers m
1
, m
2
, . . . , m
r
corresponds to xing which
atoms point down.
Proof. Let us introduce the short hand notation:
[ =

M
a(M)[M =

M
a(M)S

M
[0,
where M = m
1
, m
2
, , m
r
. This will be useful since considering the gen-
eral case can get rather tedious and it simplies the notation, this which will
25
become apparent during the proof. Let us begin straight away by applying
the Hamiltonian on this general eigenstate:t
H[ = J
N

n=1
_
1
2
_
S
+
n
S

n+1
+S

n
S
+
n+1
_
+S
z
n
S
z
n+1
_

M
a(M)[M
= J

M
a(M)
N

n=1
_
1
2
_
S
+
n
S

n+1
+S

n
S
+
n+1
_
+S
z
n
S
z
n+1
_
[M
To compute this we need to consider the action of the spin operators on
[M = S

M
[0:

n
S
+
n
S

n+1
[M =

n
S
+
n
S

n+1
S

M
[0 =

M
S

M
[0 =

M
[

M
Where

M = m
1
, , m
r
is obtained by displacing a toward the right
in M. We can see this by considering the object
1
(M) = S
+
n
S

n+1
S

M
[0
for dierent values of n. We see that
1
(M) vanishes for all n / M or
n + 1 M. The only terms that contribute are terms with neighbouring
opposite spins that look like in the spin chain i.e. n M and n+1 / M,
exchange if m
i+1
,= m
i
+ 1 for some i = 1, , r 1, and m
i
= m
1
+ 1.
Let
+
(M) be the set of all m
1
, , m
r
gotten by displacing a to in M
toward the right. i.e. and

M
+
(M).
Similarly, we can consider:

n
S

n
S
+
n+1
[M =

[0 =

[M

Where M

= m

1
, , m

r
is obtained by displacing a toward the left
in M. We can see this in a simular way as the above, and we see that
the object in the sum only contributes for neighbouring opposite spins that
look like in the spin chain i.e. n / M and n + 1 M, exchange if
m
i
,= m
i1
+ 1 for some i and m

i
= m
i
1. Let

(

M) be the set of all
m

1
, , m

r
gotten by displacing a in M toward the left i.e. and
M

(

M) = M :

M =
+
(M). Combining these results we obtain:
N

n=1
_
S
+
n
S

n+1
+S

n
S
+
n+1
_
[M =

M
[

M +

[M

[M

Where M

= m

1
, , m

r
can be obtained by M by exchanging oppo-
site neighbouring spins i.e. and . Now we wish to consider:
F(M)[M =
N

n=1
S
z
n
S
z
n+1
[M =
N

n=1
S
z
n
S
z
n+1
S

M
[0
26
We can denote f(n, M) by:
f(n.M) =
_
1 if n M;
1 if n / M.
Then we can write
S
z
n
S

M
[0 =
1
2
f(n, M)S

M
[0
and
S
z
n
S
z
n+1
S

M
[0 =
1
4
f(n, M)f(n + 1, M)S

M
[0
then
F(M)[M =
N

n=1
S
z
n
S
z
n+1
[M =
1
4
N

n=1
f(n, M)f(n + 1, M)[M
Now we can compute H[ as above:
H[ = J

M
a(M)
_
1
2

[M

+F(M)[M
_
Recall in the r = 1 case, we had to apply the periodicity condition inside
the sum which allowed us to make shifts of the summation variables. In this
case, we will essentially do the same in order to bring all states to the uniform
expression [M. Since we are shifting the variables M

= m

1
, , m

r
to
M in [M

, we need to compensate by also shifting a(M) appropriately (by


the periodicity condition). When we represent a shift of a to the right in
[M

(i.e. by

M), we change a(M) to represent a shift of a to the left (i.e.
by M

) and vice versa. Since M

is composed of

M and M

, it is clear that
a shift in M

in the sum would be compensated by a shifting a(M) to M

with M

(M

), hence:

M
a(M)

[M

a(M

)[M
Let us consider F(M):
F(M) =
1
4
N

n=1
f(n, M)f(n + 1, M) =
1
4
(#parallel #anti-parallel)
Since N = #parallel + #anti-parallel, this becomes:
F(M) =
1
4
(N 2#anti-parallel)
Note:
#anti-parallel =

1
27
and therefore
F(M) =
N
4

1
2

1
Putting the above together we obtain:
H[ = J

M
_
1
2

a(M

) +a(M)F(M)
_
[M
= E

M
a(M)[M
This now gives the relation between the coecients a(M) as:
J
_
1
2

a(M

) +a(M)F(M)
_
= Ea(M)
substituting our F(M):

J
2

a(M

) Ja(M)(
N
4

1
2

1) = Ea(M)
Rearranging as in the form of (), and substituting E
0
= JN/4, we obtain
the result:
2(E E
0
)a(M) = J

[a(M) a(M

)]
As required.
This useful since to check that an eigenstate satises H[ = E[, we
need the coecients to satisfy (). We also have the periodicity condition
for the coecients as a(m
1
, , m
i
, , m
r
) = a(m
1
, , m
i
+N, , m
r
).
Now we will introduce the Bethe Ansatz, which is the assumed general
form of the coecients a(m
1
, m
2
, . . . , m
r
) in terms of r quasi-momenta k
j
:
a(m
1
, m
2
, . . . , m
r
) =

S
r
A

exp
_
_
i
r

j=1
k
(j)
m
j
_
_
The sum S
r
is over all r! permutations of the labels 1, 2, , r. The
coecients A

C depending on the permutation need to be determined


and will be given in terms of the scattering phase. We will work through
the examples r = 1 and r = 2 to verify the Bethe Ansatz and to give a more
explicit form for A

.
For r = 1, we obtain:
[ =
N

m=1
a(m)[m a(m) = Ae
ikm
28
If we let A = 1, then we get the same eigenstate as we discussed before. One
can verify this by () since we get:
2(E E
0
)a(m) = J[2a(m) a(m1) a(m+ 1)]
for n = 1, , N and a(n +N) = a(n). This immediately implies as before:
a(m) = e
ikm
, k =
2n
N
n = 0, , N 1
The energy of the one-magnon state is therefore:
E E
0
= J(1 cos k)
which agrees with the Bethe Ansatz.
For r = 2, we obtain:
[ =
N

m=1
a(m
1
, m
2
)[m
1
, m
2

a(m
1
, m
2
) = Ae
i(k
1
m
1
+k
2
m
2
)
+A

e
i(k
2
m
1
+k
1
m
2
)
We wish to determine the conditions on A, A

given that this Ansatz satises


(). We have two cases that we need to consider, either the two down spins
are separate from each other or nearest neighbours. Using () we get the
following condition for the former:
2(E E
0
)a(m
1
, m
2
) = J[4a(m
1
, m
2
) a(m
1
1, m
2
) a(m
1
+ 1, m
2
)
a(m
1
, m
2
1) a(n
1
, m
2
+ 1)], m
2
,= m
1
+ 1
And for the latter:
2(E E
0
)a(m
1
, m
2
) = J[2a(m
1
, m
2
) a(m
1
1, m
2
) a(m
1
, m
2
+ 1)],
m
2
= m
1
+ 1
The rst group of equations is automatically satised for arbitrary A, A

, k
1
, k
2
.
The second group can be satised by choosing A and A

such that we have


the following eigenstate conditions:
2a(m
1
, m
1
+ 1) = a(m
1
, m
1
) +a(m
1
+ 1, m
1
+ 1) (**)
which is obtained by subtracting the rst group by the second for m
2
=
29
m
1
+1. To obtain the energy, we can substitute the Ansatz in () to obtain:
2(E E
0
)
_
Ae
i(k
1
m
1
+k
2
m
2
)
+A

e
i(k
2
m
1
+k
1
m
2
)
_
= J
_
4
_
Ae
i(k
1
m
1
+k
2
m
2
)
+A

e
i(k
2
m
1
+k
1
m
2
)
_

_
Ae
i(k
1
m
1
+k
2
m
2
)
e
ik
1
+A

e
i(k
2
m
1
+k
1
m
2
)
e
ik
2
_

_
Ae
i(k
1
m
1
+k
2
m
2
)
e
ik
1
+A

e
i(k
2
m
1
+k
1
m
2
)e
ik
2
_

_
Ae
i(k
1
m
1
+k
2
m
2
)
e
ik
2
+A

e
i(k
2
m
1
+k
1
m
2
)
e
ik
1
_

_
Ae
i(k
1
m
1
+k
2
m
2
)
e
ik
2
+A

e
i(k
2
m
1
+k
1
m
2
)
e
ik
1
_
_
We see that the dependence on A and A

cancel out completely as we expect


and we get the following equation for the energy:
E E
0
= J
2

j=1
(1 cos k
j
) = J(2 cos k
1
cos k
2
)
We observe that the energy of the two-magnon state appears to be equal
to the sum of the energies of the one-magnon state. This observation is
quite remarkable as it appears that the energy is additive. This shows that
magnons essentially behave themselves as free particles in a box. We will
now use the condition () to determine the restrictions placed on A and A

Labelling n
1
= n and substituting our Ansatz into this condition we get:
Ae
i(k
1
+k
2
)n
+A

e
i(k
1
+k
2
)n
+Ae
i(k
1
+k
2
)(n+1)
+A

e
i(k
1
+k
2
)(n+1)
2
_
Ae
i(k
1
n+k
2
(n+1))
+A

e
i(k
2
n+k
1
(n+1))
_
= 0
This allows us to determine the ratio in terms of a phase angle
12
=
21

as:
A
A

e
i
=
e
i(k
1
+k
2
)
+ 1 2e
ik
1
e
i(k
1
+k
2
)
+ 1 2e
ik
2
(***)
i.e. by setting A = e
i
12
/2
and A

= e
i
21
/2
. This requirement is incorporated
into our Ansatz as phase factors:
a(m
1
, m
2
) = e
i(k
1
m
1
+k
2
m
2
+
1
2

12
)
+e
i(k
2
m
1
+k
1
m
2
+
1
2

21
)
= e
i(k
1
m
1
+k
2
m
2
+
1
2
)
+e
i(k
2
m
1
+k
1
m
2

1
2
)
Where the phase angle depends on the undetermined the as yet undeter-
mined k
1
, k
2
via:
2 cot

12
2
= 2 cot

2
= cot
k
1
2
cot
k
2
2
30
The periodicity condition requires:
a(m
1
, m
2
) = a(m
2
, m+ 1 +N)
Inserting this into our Ansatz we obtain:
e
i(k
1
m
1
+k
2
m
2
+
1
2
)
+e
i(k
2
m
1
+k
1
m
2

1
2
)
= e
i(k
2
(m
1
+N)+k
1
m
2
+
1
2
)
+e
i(k
1
(m
1
+N)+k
2
m
2

1
2
)
Since this is true for all m
1
and m
2
we have:
e
ik
1
N
= e
i
, e
ik
2
N
= e
i
If we take the logarithm of these we obtain:
Nk
1
= 2I
1
+, Nk
2
= 2I
2

Where the integers I


j
0, 1, , N 1 are called the Bethe quantum
numbers. These numbers are useful to distinguish eigenstates with dierent
physical properties. The total momentum is given by the sum of all quasi-
momenta:
p = k
1
+k
2
=
2
N
(I
1
+I
2
)
p is a true integration constant of the problem. This is the same momentum
associated to the translation operator T as we saw earlier, since displacing
a down spin by one place on the lattice does not change the physics of the
system. Now that we have discussed the case for r = 2 i.e. two down
spins, we can use the concepts that we introduced to general r down spins
in a natural way. The generalization of ( ) for the A

in terms of the
scattering phase
jl
=
lj
for each pair (k
j
, k
l
) is given by:
A

= exp
_
_
i
2

l<j

(l)(j)
_
_
i.e. the Bethe Ansatz for the coecients a(m
1
, , m
r
) can be written as:
a(m
1
, m
2
, . . . , m
r
) =

S
r
exp
_
_
i
2

l<j

(l)(j)
_
_
exp
_
_
i
r

j=1
k
(j)
m
j
_
_
=

S
r
exp
_
_
i
r

j=1
k
(j)
m
j
+
i
2

l<j

(l)(j)
_
_
If we assume that these coecients satisfy the consistency equations ()
then the energy depends on k
1
, k
2
, , k
r
given as a straightforward gener-
alization of the two-particle case:
E E
0
= J
r

j=1
(1 cos k
j
)
31
The eigenstate condition analogous to () is given by:
2a(m
1
, , m
j
, m
j
+ 1, , m
r
) = a(m
1
, , m
j
, m
j
, , m
r
)
+a(m
1
, , m
j
+ 1, m
j
+ 1, , m
r
)
These conditions relate every phase angle
jl
to the (as yet undetermined)
k
j
as a generalization of ( ):
e
i
jl
=
e
i(k
j
+k
l
)
+ 1 2e
ik
k
e
i(k
j
+k
l
)
+ 1 2e
ik
l
or in real form is given as
2 cot

jl
2
= cot
k
j
2
cot
k
l
2
, j, l = 1, , r
3.2 BAE
Here we will derive the Bethe Ansatz Equations by considering the period-
icity condition on the Ansatz given for the coecients a(m
1
, m
2
, , m
r
).
By the periodicity condition (closed chain)
a(m
1
, m
2
, , m
r
) = a(m
2
, , m
r
, m
1
+N)
Which from the general eigenstate (Bethe Ansatz) gives:
r

j=1
k
(j)
m
j
+
1
2

l<j

(l)(j)
=
1
2

l<j

(l)

(j)
2I

(r)
+
r

j=2
k

(j1)
m
j
+k

(r)
(m
1
+N)
The permutations

(j) and (j) on the left and right are related by:

(j 1) =

(j), j = 2, , r

(r) = (1)
The terms not involving the index

(r) = (1) cancel and we are thus left


with the following set of r equations:
Nk
j
= 2I
j
+

l=j

jl
, j = 1, , r (+)
Claim: If we introduce the rapidities
j
to parametrize the momenta
k
j
:

j
=
1
2
cot
k
j
2
32
then we can rewrite (+) as the Bethe Ansatz equation (BAE):
_

j
+i/2

j
i/2
_
N
=
r

l=1,l=j

j

l
+i

j

l
i
Proof. Before we begin the proof, we will need the following simple trigono-
metric identity:
e
i
(cot (/2) i) = (cot (/2) +i)
We can show this by using Eulers identity with:
e
i
=
e
i/2
e
i/2
=
cos (/2) +i sin (/2)
cos (/2) i sin (/2)
=
cot (/2) +i
cot (/2) i
and hence the result. Now if we take multiply both sides of (+) by i and
take the exponential, we get:
e
iNk
j
= e
i

l=j

jl
Since e
i2I
j
= 1, as I
j
is an integer. Now, applying the trigonometric identity
to e
ik
j
and substituting the rapidity
j
:
e
ik
j
=
2
j
+i
2
j
i
=

j
+i/2

j
i/2
and so the LHS becomes:
(e
ik
j
)
N
=
_

j
+i/2

j
i/2
_
N
We can also rewrite the RHS as:
e
i

l=j

jl
=
r

l=1,l=j
e
i
jl
Again, by applying the trigonometric identity on e
i
jl
and substituting
j
we obtain:
e

jl
=

j

l
+i

j

l
i
Which then gives using the above:
_

j
+i/2

j
i/2
_
N
=
r

l=1,l=j

j

l
+i

j

l
i
As required.
33
3.3 Algebraic BA
Here we will solve the XXX Heisenberg Spin chain using a variant of the
Bethe Ansatz known as the Algebraic Bethe Ansatz (ABA). The ABA is
more powerful since it will allow us to derive properties of the system such
as integrability, it also allows us to more easily generalise the XXZ spin
chain in which we state results for at the end of the chapter. Before we
begin by applying the ABA we will rst introduce some concepts which will
be essential in its application. The material in this section will be adapted
from [7],[8] and [9].
3.3.1 Integrability
Let us dene what it means for a Quantum system to be integrable.
A classical system with N degrees of freedom, described by a Hamilto-
nian H is Liouvelle integrable if there exists N conserved charges Q
i
with
zero poisson bracket in involution i.e. Q
i
, Q
j
= 0 and the Hamiltonian
H is one of the charges. Each of the conserved charges yields a conserva-
tion law that can be solved (integrated) to x all the independent degrees
of freedom. This notion of integrability can be carried over to Quantum
systems.
AQuantum systemwith N degrees of freedom dened by a Hamiltonian-
operator H, will be called integrable if there exists N local conserved charges
Q
i
that commute i.e. [Q
i
, Q
j
] = 0 and the Hamiltonian H is one of them.
Therefore, all the charges can be diagonlalized simultaneously and the set of
complete eigenstates with corresponding eigenvalues can be found exactly.
1
We will later see how we can compute these conserved charges.
3.3.2 Yang-Baxter equation
The Yang-Baxter equation is an essential tool in showing the integrability
of the Heisenberg model. There are dierent conventions for the form of the
equation, we will briey look at the dierent conventions and nally stick
with the most common. The denition involves the n
th
local quantum space
V
n
= C
2
and the auxiliary space A, which we will also take to be C
2
to
allow us to make sense of the permutation operator as dened above acting
on both spaces. Given , , C, and u, v , , , the invertible R
matrix R
a
1
,a
2
(
1
,
2
) End(A
1
A
2
) satises:
R
a
1
,a
2
(, )R
a
1
,a
3
(, )R
a
2
,a
3
(, ) = R
a
2
,a
3
(, )R
a
1
,a
3
(, )R
a
1
,a
2
(, )
1
This denition of quantum integrability is not entirely correct and it is still a subject
of research on how to make a better denition. For example, the XXX model actually
has 2
N
degrees of freedom since each lattice site can be occupied by a spin up or a spin
down electr on. But in this denition we only consider the lattice site of the XXX model
of having one degree of freedom.
34
On End(A
1
A
2
A
3
). The index (a
1
, a
2
) indicates that it acts nontrivially
on two copies of the auxiliary spce. Assume R
a
i
,a
j
(u, v) = R
a
i
,a
j
(u v),
then the Yang-Baxter equation can be written as:
R
a
1
,a
2
()R
a
1
,a
3
()R
a
2
,a
3
() = R
a
2
,a
3
()R
a
1
,a
3
()R
a
1
,a
2
()
If we let = 0 or by relabelling u u, we obtain:
R
a
1
,a
2
( )R
a
1
,a
3
()R
a
2
,a
3
() = R
a
2
,a
3
()R
a
1
,a
3
()R
a
1
,a
2
( )
The Lax operator L
a,n
() End(A V
n
) satises the FCR (Interwining
equation):
R
a
1
,a
2
( )L
a
1
,n
()L
a
2
,n
() = L
a
2
,n
()L
a
1
,n
()R
a
1
,a
2
( )
On End(A
1
A
2
V
n
). The index (a, n) indicates that it acts nontrivially
on the auxiliary space and the n
th
local quantum space. We also have
the commutativity of the Lax operator with completely dierent indices
(Ultralocality condition):
[L
a
1
,n
(), L
a
2
,m
()] = 0
On End(A
1
A
2
V
n
V
m
). This is obvious since L
a
i
,n
acts trivially with
L
a
j
,m
given both auxiliary spaces are distinct i.e. i ,= j and both local
quantum spaces are distinct i.e. n ,= m.
For the XXX Heisenberg model in particular, the Lax operator is given
by:
L
a,n
() = I
a
I
n
+i

a
Explicitly as a 2 2 matrix in the auxiliary space:
L
a,n
() =
_
+iS
z
n
iS

n
iS
+
n
iS
z
n
_
Which is a matrix acting in the auxiliary space A, with entries being oper-
ators on the Quantum space V
n
. Given that the operator:
P =
1
2
(I I +

)
is a permutation in C
2
C
2
, we can rewrite the Lax operator in terms of
P
a,n
:
L
a,n
() = (
i
2
)I
a,n
+iP
a,n
In order for this to satisfy the FCR, we must have:
R
a
1
,a
2
() = I
a
1
,a
2
+iP
a
1
,a
2
,
35
which we will refer to as our R-operator (or matrix). It is easy to see that
the Lax operator and the R-matrix are in a simular form, indeed L
a,n
() =
R
a,n
( i/2).
The R-operator given explicitly is a solution of the Yang-Baxter equa-
tion. One way to see this is explicitly, one can express R
a
1
,a
2
() as a matrix
in an appropriate basis and perform a tedious task in matrix multiplication.
Let us dene the Monodromy matrix T
a
() End(A1
N
) by:
T
a
() = T
a,N
() = L
a,N
() L
a,1
() =
N

n=1
L
a,N(n1)
Claim: This matrix also has a FCR which is in the same form as the local
FCR i.e.
R
a
1
,a
2
( )T
a
1
()T
a
2
() = T
a
2
()T
a
1
()R
a
1
,a
2
( )
On End(A
1
A
2
1
N
). We shall call this the Global FCR.
Proof. We will start with the left hand side of the equation, which we will
write in terms of the L-operators:
R
a
1
,a
2
( )L
a
1
,N
() L
a
1
,1
()L
a
2
,N
() L
a
2
,1
()
= R
a
1
,a
2
( )L
a
1
,N
()L
a
2
,N
() L
a
1
,1
()L
a
2
,1
()
Where the last line by the application of the ultralocality condition repeat-
edly. Next we will use the local FCR for the L-operators in order to transfer
the R-operator from the left to the right of the product i.e.:
= L
a
2
,N
()L
a
1
,N
()R
a
1
,a
2
( ) L
a
1
,1
()L
a
2
,1
()
= L
a
2
,N
()L
a
1
,N
() L
a
2
,1
()L
a
1
,1
()R
a
1
,a
2
( )
= L
a
2
,N
() L
a
2
,1
()L
a
1
,N
() L
a
1
,1
()R
a
1
,a
2
( )
= T
a
2
()T
a
1
()R
a
1
,a
2
( )
Where in the third line we have used the ultralocality condition again in
order to make the monodromy matrix appear.
Note: L
a,n
() is linear in i.e.
d
d
L
a,n
= I
a,n
. Since T
a
() is a
product of N of these, this implies that the power expansion for T
a
() is
nite
T
a
() =

n=1
a
n
(
0
)
n
=
N

n=1
a
n
(
0
)
n
=
N
+
N1

n=1
a
n
(
0
)
n
where
a
n
=
1
n!
d
n
d
n
T
a
()

=
0
=
T
(n)
a
(
0
)
n!
36
and
0
= i/2 is the usual parametrization, but it doesnt matter which
parametrization we use as the expansion is nite. In other words, it is
a polynomial of order N with leading coecient (of
N
) equal to unity.
This is obvious by looking at the explicit expression for T
a
() but can also
be realised by computing a
N
and using
d
N
d
N
T
a
() = N! I
a,N
. Let us
compute the coecient of the next to highest degree a
N1
:
a
N1
=
1
(N 1)!
d
N1
d
N1
T
a
()

=
0
Which is the coecient of
N1
, here we will take
0
= 0 for convenience.
If f
n
(x) is linear i.e.
d
dx
f
i
(x) = 1 i f
n
(x) = x +r
n
d
N1
dx
N1
_
N

n=1
f
n
(x)
_
= (N 1)!
N

n=1
f
n
(x)
From this we can compute:
d
N1
d
N1
T
a
()

=0
=
d
N1
d
N1
(L
a,N
L
a,1
)

=0
= (N 1)!
N

n=1
L
a,n
()

=0
= (N 1)!
N

n=1
L
a,n
(0)
L
a,n
(0) = i

(S

a
)
The total spin is dened as:
S

= S

T
=
N

n=1
S

n
Therefore we have:
a
N1
= i

n=1
(S

)
= i

(S

)
T
a
() =
N
+i
N1

(S

) +
37
This operator can also be represented by the following 2 2 matrix in
the auxiliary space A, with entries being operators on the Quantum space
V
n
:
T
a
() =
_
A() B()
C() D()
_
The transfer matrix is dened by tracing over the auxiliary space:
t() = tr
a
T
a
() = A() +D()
Claim: The transfer matrix consists of a one-parameter family of com-
muting operators:
[t(), t()] = 0
Proof.
t()t() = (tr
a
1
T
a
1
())(tr
a
2
T
a
2
())
= tr
a
1
,a
2
[T
a
1
()T
a
2
()]
= tr
a
1
,a
2
[R
a
1
,a
2
( )
1
R
a
1
,a
2
( )T
a
1
()T
a
2
()]
= tr
a
1
,a
2
[R
a
1
,a
2
( )
1
T
a
2
()T
a
1
()R
a
1
,a
2
( )]
= tr
a
1
,a
2
[T
a
2
()T
a
1
()]
= (tr
a
2
T
a
2
())(tr
a
1
T
a
1
())
= t()t()
Where we have taken the double trace over both auxiliary spaces and have
used the cyclic property of the trace.
Note since this holds for , C, we can take the series expansion
t() =

i
Q
i
and by the commutation relation:

i,j

j
[Q
i
, Q
j
] = 0
This automatically tells us that we have
[Q
i
, Q
j
] = 0 i, j
Let us compute the trivial terms in the power expansion for the transfer ma-
trix t() = a
N
+b
N1
+ , the transfer matrix will be a nite polynomial
38
of order N, the coecients are given by:
a =
1
N!
d
N
d
N
t()
=
1
N!
tr
a
(
d
N
d
N
T
a
())
=
1
N!
tr
a
(N! I
a,N
)
=
1
N!
tr
a
(N! I
a
I
N
)
= tr
a
(I
a
) tr
a
(I
n
) = 2I
N
This is true for all , we shall now compute the
N1
term, we will nd that
the
N1
term vanishes since the Pauli matrices

are traceless:
b =
1
(N 1)!
d
N1
d
N1
t()

=0
=
1
(N 1)!
tr
a
_
d
N1
d
N1
T
a
()
_

=0
=
1
(N 1)!
tr
a
_
i(N 1)!

a
_
= i tr
a
_

a
_
= i

tr
a
(S

) tr
a
(

a
)
= i

tr
a
(

a
) = 0
Where we have used the identity tr(AB) = tr Atr B.
This tells us that the non-trivial expansion begins with power
N2
:
t() = 2
N
+
N2

n=0
Q
n

n
and produces N 1 commuting operators Q
i
(conserved charges). As we
shall see in Chapter 4, these charges will turn out to be intractable and non-
local, it will therefore be useful to consider the expansion of the logarithm
of the transfer matrix since it has been shown by Lusher [17] to produce
local charges which will be algebraically related to the coecients Q
n
. The
Hamiltonian will be related to the rst term of the expansion (the rst
derivative) which we can write in terms of t() as:
d
d
ln t() = t

()t()
1
39
Claim: The Hamiltonian H for the XXX model can be generated by the
coecients Q
n
as:
H = J
_
i
2
Q
1
0
Q
1

N
4
_
Proof. Let us consider the point = i/2:
L
a,n
(i/2) = iP
a,n
we also have for any :
d
d
L
a,n
() = I
a,n
This makes it easy to control the expansion of t() in the neighborhood of
= i/2. We have:
Q
0
= t(i/2) = tr
a
(L
a,N
(i/2) L
a,1
(i/2))
= tr
a
((iP
a,N
) (iP
a,1
))
= i
N
tr
a
(P
a,N
P
a,1
)
These string of permutations inside the trace can be transformed by the
identities from Chapter 1 to isolate a single permutation carrying the index
of the auxiliary subspace:
P
a,N
P
a,1
= P
N,N1
P
N,N2
P
N,1
P
a,N
= P
1,2
P
2,3
P
N1,N
P
a,N
Recall that tr
a
P
a,N
= I
N
, therefore we have:
t(i/2) = i
N
P
1,2
P
2,3
P
N1,N
We can dene the shift (translation) operator T as:
T = i
N
t(i/2) = P
1,2
P
2,3
P
N1,N
The operator T is unitary T

T = TT

= I because the permutations have


the properties P

= P and P
2
= I, it generates a shift along the chain as:
T
1
X
n
T = X
n1
By denition an operator of the innitesimal shift is the momentum on the
lattice dened as
T[ = e
ip
[
where p is the momentum as we used in the CBA for the translation operator.
40
Now let us dierentiate the logarithm of the transfer matrix:
Q
1
= t

(i/2) =
d
d
t()

=i/2
=
d
d
(tr
a
[L
a,N
() L
a,1
()])

=i/2
= tr
a
d
d
([L
a,N
() L
a,1
()])

=i/2
Let us compute the object inside the trace:
d
d
([L
a,N
() L
a,1
()])

=i/2
= i
N1

n
(P
a,N


P
a,n
P
a,1
)
Thus we get:
t

(i/2) = i
N1

n
tr
a
(P
a,N


P
a,n
P
a,1
)
= i
N1

n
(P
1,2
P
n1,n+1
P
N1,N
)
Putting the above together, we are now in a position to compute the loga-
rithmic derivative:
d
d
ln t()

=i/2
=
_
d
d
t()
_
t()
1

=i/2
= i
N
Q
1
(P
N,N1
P
N1,N2
P
2,1
)
=
1
i
N

n=1
P
n,n+1
Recall that:
H =
J
4
N

n=1

n+1
= J
N

n=1
S

n
S

n+1
= J
_
1
2
N

n=1
P
n,n+1

N
4
_
Hence we can write the Hamiltonian in terms of the transfer matrix as:
H = J
_
i
2
Q
1
0
Q
1

N
4
_
= J
_
i
2
d
d
ln t()

=i/2

N
4
_
Therefore, H belongs to a family of N 1 commuting operators. If we
add the total spin operator S
z
T
to this family we get N commuting operators
and hence the XXX-model is integrable.
41
3.3.3 Spectrum of H
Now that we have proven the integrability of the model, we need to solve it
i.e. nd the spectrum of the Hamiltonian H. It is clear that [H, t()] = 0,
this implies that eigenstates of H will also be eigenstates of t(), hence in
order to solve the model it suces to diagonalize t(). Given [
n
V
n
and v a, the action of the Lax operator L
a,n
() : A V
n
A V
n
on
arbitrary element of v [
n
AV
n
:
L
a,n
()(v [
n
) =
_
( +iS
z
n
)[
n
iS

n
[
n
iS
+
n
[
n
( iS
z
n
)[
n
_
v
=
_
( +
i
2
)[
n
i[
n
0 (
i
2
)[
n
_
v
The action on v is that of matrix multiplication on V
a
, in what follows we
can omit it. Let us now compute the action of L
a,n
() on the ground state
[0 = [ =
N
n=1
[
n
1
N
:
L
a,n
()[0 =
_
+iS
z
n
iS

n
iS
+
n
iS
z
n
_
[0
=
_
( +iS
z
n
)[0 iS

n
[0
iS
+
n
[0 ( iS
z
n
)[0
_
Note (from the table at the beginning of the section):
S
z
n
[0 =
1
2
[0 S
+
n
[0 = 0
The action of S

n
on the ground state merely changes the nth spin in the
chain to a spin down, the explicit form will be unimportant in what follows
since we will be considering the action of the trace of the Monodromy matrix
(i.e. the transfer matrix) on the ground state. Denoting any unimportant
items with a i.e., we get from the above:
L
a,n
()[0 =
_
( +
i
2
)[0
0 (
i
2
)[0
_
=
_
( +
i
2
)
0 (
i
2
)
_
[0
Using this, we can now compute the action of the Monodromy matrix
42
T
a
() on [0 by the repeated application of the above:
T
a
()[0 = L
a,N
() L
a,1
()[0
=
_
+iS
z
N
iS

N
iS
+
N
iS
z
N
_

_
+iS
z
1
iS

1
iS
+
1
iS
z
1
_
[0
=
_
+iS
z
N
iS

N
iS
+
N
iS
z
N
_

_
( +
i
2
)[0
0 (
i
2
)[0
_
=
_
( +
i
2
)
N
[0
0 (
i
2
)
N
[0
_
=
_
( +
i
2
)
N

0 (
i
2
)
N
_
[0 =
_
A() B()
C() D()
_
[0
More importantly, what this tells us is that the ground state [0 is an
eigenstate of A() and D() simultaneously i.e.:
A()[0 = a()[0 D()[0 = d()[0
Where a() = ( +
i
2
)
N
and d() = (
i
2
)
N
. Consequently, it is also an
eigenstate of the transfer matrix t() = A() +D(). This is not surprising
since we know that [H, t()] = 0 and we have also proved that [0 is an
eigenstate of H earlier in the section, this immediately implies that it is also
an eigenstate of t(). It is also clear from the above that [0 is annihilated
by C() i.e.:
C()[0 = 0
The algebraic Bethe Ansatz (ABA) is the assumption that C() and
B() are annihilation and creation operators. A general state with r down
spins will be given by choosing such that:
[
1
, ,
r
= B(
1
) B(
r
)[0 =
r

n=1
B(
n
)[0
The conditions on for the above to be an eigenstate of t() will produce
the BAEs, which we have also derived using the CBA. The
i
, the so-called
Bethe roots are comparable to the k
i
(momenta) from the CBA and we
shall see how they are related to each other. To calculate the eigenvalue for
the above eigenstate will require us to establish the commutation relations
between [A(), B()], [B(), B()] and [D(), B()]. These will result from
the RTT relations (the Global FCR).
Claim: Applying the RTT relations we get the following:
[B(), B()] = 0
A()B() = f( )B()A() +g( )B()A()
D()B() = h( )B()D() +k( )B()D()
43
Where:
f() =
i

, g() =
i

, h() =
+i

, k() =
i

Proof. We will do this by direct computation of the RTT relations and


comparing coecients. We will use the explicit matrix representation of the
FCR in End(A
1
A
2
). All objects are 4 4 matrices in A
1
A
2
with the
basis given by 1
2
. The R-matrix in this representation looks like:
R
a
1
,a
2
() = ( +i)
_
_
_
_
1 0 0 0
0

+i
i
+i
0
0
i
+i

+i
0
0 0 0 1
_
_
_
_
The matrices T
a
1
() and T
a
2
() in this representation are given by:
T
a
1
() = T
a
() I
T
a
2
() = I T
a
()
Explicitly:
T
a
1
() =
_
_
_
_
A() B()
A() B()
C() D()
C() D()
_
_
_
_
T
a
2
() =
_
_
_
_
A() B()
C() D()
A() B()
C() D()
_
_
_
_
Thus:
T
a
1
()T
a
2
() =
_
_
_
_
A()A() A()B() B()A() B()B()
A()C() A()D() B()C() B()D()
C()A() C()B() D()A() D()B()
C()C() C()D() D()C() D()D()
_
_
_
_
and
T
a
2
()T
a
1
() =
_
_
_
_
A()A() B()A() A()B() B()B()
C()A() D()A() C()B() D()B()
A()C() B()C() A()D() B()D()
C()C() D()C() C()D() D()D()
_
_
_
_
By direct computation of the RTT relation:
R
a
1
,a
2
( )T
a
1
()T
a
2
() T
a
2
()T
a
1
()R
a
1
,a
2
( ) = 0
By looking at the 1st row and 4th column of the matrix on the LHS and
equating it to 0, we obtain the rst relation. Doing the same with the 1st row
and 3rd column and interchanging with we obtain the second relation,
the third is obtained by looking at the 3rd row and 4th column.
44
3.3.4 Bethe Ansatz equations
Now we are in a position to derive the BAE from the above considerations,
these will follow from the following claim.
Claim: The condition that [
1
, ,
r
is an eigenstate of the trans-
fer matrix t() will lead to a set of algebraic relations on the parameters
;
i
=
1
, ,
r
, these will be the BAE.
Proof. To this end, let us compute the action of A() and D() on [
1
, ,
r
:
A()[
1
, ,
r
= A()B(
1
) B(
r
)[0
= a()
r

n=1
f(
n
)B(
1
) B(
r
)[0
+
r

n=1
M
n
(,
j
)B()
r

j=1,j=n
B(
j
)[0
Where
M
n
(,
j
) = g(
n
)a(
n
)
r

j=1,j=n
f(
n

j
)
We can see this explicitly by using the commutation relations repeatedly
to move A() to the right, lets compute the coecient for the nth term. Note
we can write [
1
, ,
r
in the following way:
[
1
, ,
r
= B(
n
)
r

j=1,j=n
B(
j
)[0 n = 1, , r
since the operators B() commute with each other, using this along with
the second commutation relation, we can move the A() past the B(
n
) i.e.:
A()[
1
, ,
r
= A()B(
n
)
r

j=1,j=n
B(
j
)[0
= f(
n
)B(
n
)A()
r

j=1,j=n
B(
j
)[0
+g(
n
)B()A(
n
)
r

j=1,j=n
B(
j
)[0
From this it is clear that the rst term will not contribute to M
n
(,
i
)
since it contains B(
n
), only the second term will contribute. If we now move
A(
n
) past the B(
j
) we see that the only way to avoid the appearance of
45
B(
n
) is by using the rst term of the A-B commutation relation. The
resulting term will be:
g(
n
)a(
n
)
r

j=1,j=n
f(
n

j
)B()
r

j=1,j=n
B(
j
)[0
i.e. we have:
M
n
(,
i
) = g(
n
)a(
n
)
r

j=1,j=n
f(
n

j
)
Similarly we have:
D()[
1
, ,
r
= D()B(
1
) B(
r
)[0
= d()
r

n=1
h(
n
)B(
1
) B(
r
)[0
+
r

n=1
N
n
B()
r

j=1,j=n
B(
j
)[0
Where
N
n
= N
n
(,
j
) = k(
n
)d(
n
)
r

j=1,j=n
h(
n

j
)
Combining these results we get for the transfer matrix t():
t()[
1
, ,
r
= (A() +D())[
1
, ,
r

=
_
a()
r

n=1
f(
n
) +d()
r

n=1
h(
n
)
_
[
1
, ,
r

+
r

n=1
(M
n
+N
n
)B()
r

j=1,j=n
B(
j
)[0
The rst term has a desirable form and if the second term was absent
then we would immediately have that [
1
, ,
r
is an eigenstate of t() for
all . Since it is present however, it will produce restrictions on which we
will see will lead to the BAE. We can now make the following statement: The
general eigenstate with r spins down, produced by the creation operators
B(
i
), [
1
, ,
r
is an eigenstate of t() (and consequently of H) with
eigenvalue:
() =
_
a()
r

n=1
f(
n
) +d()
r

n=1
h(
n
)
_
46
provided that the second term vanishes i.e. M
n
+N
n
= 0 n. This means
one needs to satisfy the condition:
g(
n
)a(
n
)
r

j=1,j=n
f(
n

j
) = k(
n
)d(
n
)
r

j=1,j=n
h(
n

j
)

g(
n
)a(
n
)
k(
n
)d(
n
)
=
r

j=1,j=n
h(
n

j
)
f(
n

j
)
Substituting the explicit expressions for f(), g(), h(), k() and the eigen-
values of the ground state a(), d(), we obtain the Bethe Ansatz Equa-
tion:
_

n
+i/2

n
i/2
_
N
=
r

j=1,j=n

j
+i

j
i
As required.
We will now look at the expressions for the eigenvalues of the components
of spin on [
1
, ,
r
. To this end, we will look at the limit in
the RTT relation, using the power expansion of T
a
() and the R- matrix
representation in terms of the permutation matrix P. We get:
_
( ) +
i
2
[I
a
1
I
a
2
+

a
1

a
2
)
__

N
+i
N1

(S

) +
_
T
b
()
= T
b
()
_

N
+i
N1

(S

) +
__
( ) +
i
2
[I
a
1
I
a
2
+

a
1

a
2
)
_
_

a
I
b
S

)
_
(I
a
T
b
()) +
1
2

b
I
h
)(I
a
T
b
())
= (I
a
T
b
())
_

a
I
b
S

)
_
+ (I
a
T
b
())
1
2

b
I
h
)

a
[T
b
(), I
b
S

+
1
2
(

b
I
h
)] = 0
_
T
b
(), I
b
S

+
1
2
(

b
I
h
)
_
=
__
A() B()
C() D()
_
,
_
S

0
0 S

_
+
1
2
(

b
I
h
)
_
= 0
Taking = z:
__
A() B()
C() D()
_
,
_
S
z
0
0 S
z
_
+
1
2
_
I 0
0 I
__
= 0
47
The (1, 2) matrix entry of this equation gives us
[S
z
, B] = B
Taking = x, y we get the relation:
[S
+
, B] = AD
For our ground state [0 we have:
S
+
[0 = 0 S
z
[0 =
N
2
[0
Which tells us that [0 is a highest weight vector for the spin algebra. Let
us look at the eigenstates [
1
, ,
r
. Using the above relations we get:
S
z
[
1
, ,
r
=
_
N
2
r
_
[
1
, ,
r

which conrms the result we got in the rst part of the section. One can
show in a simular fashion to the proof of the BAE that:
S
+
[
1
, ,
r
= 0
This shows that the eigenstates obtained from the ABA are also the highest
weight vectors of the spin algebra.
We will now calculate the energy for the corresponding eigenstates using
the logarithmic derivative of the eigenvalue of the transfer matrix. We obtain
after straightforward dierentiation:
E = J
_
i
2
d
d
ln ()

=i/2

N
4
_
= E
0
+
J
2
N

i=1
1

2
i
+ 1/4
If we identify
i
as the rapidity as shown in the CBA and use the parame-
terization
k
=
1
2
cot(
k
j
2
) we get:
E E
0
= J
N

i=1
2
cot(k
j
/2)
2
+ 1
= J
N

i=1
2 sin
2
(
k
j
2
) = J
N

i=1
(1 cos (k
j
))
3.4 Thermodynamic limit N - String Hy-
pothesis
We will use our analysis on the CBA and ABA in order to analyse the be-
haviour at N . The mathematical and physical features of the N
are completely dierent for the cases the cases J > 0 and J < 0. We will
consider these separately corresponding to the ferromagnetic and antifer-
romagnetic phases respectively. We saw earlier that the rapidities
j
can
48
have real and complex solutions, so far our discussion has been general but
it turns out that the majority of the states are characterized by complex
rapidities. In general these have to found numerically but they can be ex-
tremely hard to nd. We would like to understand the structure of the
complex rapidities to develop more suitable ways to determine them, but
for nite N this can be intractable. The advantage of analysing N is
that a simple structure emerge, called the string hypothesis which allows us
to analyse these complex rapidities in more detail. We will begin by talking
about the FM case J > 0, though some of the deductions will allow us to
explore the physics of the AFM J < 0. The material in this section will be
based on [8] ,[11] and [12].
Recall the BAE:
_

n
+i/2

n
i/2
_
N
=
r

j=1,j=n

j
+i

j
i
Rapidities
j
to parameterize the quasi-momenta k
j
as:

j
= cot
k
j
2
or
k
j
=
1
i
ln

j
+i

k
i
=
1
(
j
),
where
n
() = 2 arctan

n
.
Taking the Log of the BAE or using what we obtained from the CBA
we obtain (k
j
= k(
j
)):
Nk(
j
) = 2I
j
+

l=j

jl
, j = 1, , r
In the BAE, N enters the exponent of the LHS. For real
1
, ,
r
, both
sides of the BAE are functions on the circle with the LHS oscillating wildly
when N is large. For large N and I with r kept xed, the second term of
the above is negligible thus we have
Nk
j
2I
j
= k
j

2I
j
N
Which is simply the momentum of a free particle on the chain. When J > 0,
the energy of this particle is (p) = 1 cos p. The correction caused by
the second term of the RHS encodes the information about the scattering
of these particles.
The BAE also allows for
j
C, which correspond to bound states for
J > 0. Let us look at the case r = 2 (two down spins), the BAE are:
_

1
+i/2

1
i/2
_
N
=

1

2
+i

2
i
(1)
49
_

2
+i/2

2
i/2
_
N
=

2

1
+i

1
i
(2)
=
_

1
+i/2

1
i/2
_
N
_

2
+i/2

2
i/2
_
N
= 1
Since we require that the total momentum k(
1
) + k(
2
) be real. For
Im
1
,= 0 the LHS of (1) grows (or decreases) exponentially when N
and thus the RHS must have Im(
1

2
) = i.
1
and
2
can be
interchanged and hence their form is given by:

1
=
1/2
+
i
2
,
2
=
1/2

i
2
The momentum and energy of this state are:
e
ip
1/2
()
= e
ip
0
(+i/2)+ip
0
(i/2)
=
+i
i

1/2
() = J
d p
1/2
d
=
0
( +i/2) +
0
( i/2) =
4J

2
+ 4
Which gives the dispersion relation:

1/2
(p) =
J
2
(1 cos p
1/2
)
For J > 0 we have p, p

< 2:

1/2
(p) <
0
(p p

) +
0
(p

)
Bound states are energetically favoured compared to the real solutions in
the ferromagnetic case.
For r > 2 more complex solutions can appear and we can describe them
in a simular way. Roots
r
are combined with complexes (or strings) of type
M, where M can have half-integer values M = 0, 1/2, 1, , we can have
complexes of 2M+1 rapidities characterized by the same real value
M
, but
dierent imaginary parts.

M;m
=
M
+im M m M
Where m is a integer or half-integer together with M. Counting all the
complexes of length (type) M, we can denote the number of complexes by

M
. The set of integers
M
denes the conguration space of Bethe roots.
For a state with a given magnetization we can dene the partition:
r =

M
(2M + 1)
M
50
It is important to stress that this string structure is only value in the ther-
modynamic limit N . The energy and momentum of a M-complex is
given by:
p
M
(
M
) =
1
i
ln

M
+i(M + 1/2)

M
i(M + 1/2)

M
(
M
) =
2J(M + 1/2)

2
M
+ (M + 1/2)
2
=
J
2M + 1
(1 cos p
m
)
For a ferromagnetic coupling the completely aligned state [0 can be
taken as the ground state. Excited states above the ground state can be
constructed in terms of magnon excitations, keeping into consideration that
bound states corresponding to 2M + 1 complexes have lower energies com-
pared to states with 2M + 1 real magnons. For the Anti-Ferromagnetic
(J < 0) case we will have to do more work in order to determine the phys-
ical spectrum as the completely polarized [0 used as our reference state
is clearly dierent from the anti-ferromagnet ground state. In the ferro-
magnetic regime, string solutions have lower energy than unbound, purely
real ones. The opposite is true for the AFM, the string solutions will have
higher energy. From here on we will let J = 1 for the anti-ferromagnet
without loss of generality. Let us assume that N is even in what follows. The
ground state conguration for the AFM will be composed of single particle
excitations:

0
=
N
2
;
M
= 0 M
1
2
For this conguration we have r = N/2, this implies that the spin of the
states vanishes. Excited states over the ground state can be characterised
by with:

0
=
N
2
;
M
= 0 M
1
2
We will now proceed to approximate the distribution of the solutions of the
Bethe equation with their continuous distribution. We will start with the
ground state, for which the quantum numbers I
0,j
ll the allowed interval
of vacancies without holes. Let us assume that N/2 is odd (with the even
case requiring a minor modication), so that:
I
0,j
=
N
4
+
1
2
,
N
4
+
3
2
, ,
N
4

1
2
We can write the Bethe equations as:

1
(
j
) =

N
I
0,j
+
1
N

2
(
j

k
)
Explicitly this is:
arctan
j
=

N
I
0,k
+
1
N

k
arctan
_

j

k
2
_
51
In the limit N , the variable x = I
0,k
/N becomes continuous and
limited in the range 1/4 x 1/4. The set of roots
j
becomes a
function (x) and we obtain:
arctan (x) = x +
_
1/4
1/4
arctan
_
(x) (y)
2
_
dx
Using:
_
1/4
1/4
f((x)) dx =
_

f()
0
() d
where the change of variables x (x) maps interval 1/4 x 1/4 into
the real line < < due to the monotonicity of (x), i.e. explicitly:

0
() =
dx
d
=
1

(x)

x=
1
()
Now dierentiation this wrt to , the Bethe equations for the ground state
become:

0
() =
1

1
1 +
2

1

2
( )
2
+ 4

0
() d
In the thermodynamic limit, the energy and momentum can also be ex-
pressed in terms of the density function
0
() as:
E
AFM
E
0
+N
_

0
()
0
() d
K
AFM
N
_
p
0
()
0
() d
We can solve
0
() from the above integral equation by using the Fourier
transform, we get (see [11]):

0
() =
1
4 cosh (

2
)
And we thus obtain the anti-ferromagnet (J = 1) ground state energy:
E
AFM
= N
_
1
4
ln 2
_
and
K
AFM
=
N
2
mod 2
Now we turn to the state with
0
= N/2 1 and
M
= 0 for M 1/2.
The state is characterized by two wholes which we place at j
1
and j
2
:
I
0,j
= j +H(j j
1
) +H(j j
2
),
52
where H(x) is the usual Heaviside step-function. The integral equation for
the rapidity density of the real roots
t
() (where t stands for triplet) is:

t
() =
1

1
1 +
2

1

2
( )
2
+ 4

t
() d
1
N
[(
1
) +(
2
)],
where
i
, i = 1, 2 are the images of x
i
= j
i
/N under the map x (x).
Since we are dealing with linear equations, we can write the solution for

t
() as:

t
() =
0
() +
1
N
[(
1
) +(
2
)],
where () solve the integral equation:
() +
1

2
( )
2
+ 4

(
) d +()
As before, solving this integral equation using the fourier transform as in
[11], we obtain the total momentum and energy as:
K = N
_
p
0
()
t
() d = K
AFM
+(
1
) +(
2
),
E = N
_

0
()
t
() d = E
AFM
+(
1
) +(
2
),
where
()

2
arctan sinh

2
, ()

2 cosh

2
This is a state with two-particle excitations (spinons) and (), () are
their dressed energy and momentum. Combing the two, we see that these
excitations are characterized each by the dispersion relation:
(k) =

2
sin k,

2
k

2
Therefore each hole in the quantum numbers generates a quasi-particle ex-
citation, which is called a spinon, i.e. a spin- 1/2 excitation.
We see that if we follow the same procedure for the state with
0
=
N/2 2,
1/2
= 1 and
M
= 0 for M 1. We can compute rapidity density
for the real roots
s
() (s is for singlet) by solving its integral equation using
a simular procedure to the above given in [11], after which we obtain for the
energy:
E = E
AFM
+N
_

0
()
S
() d +
1/2
(
1/2
)
= E
AFM
+(
1
) +(
2
)
We see that the contributions from the string cancel exactly and this state
has the same momentum, energy and (dispersion relation) as the one without
53
complexes given by the state with
0
= N/21 and
M
= 0 for M 1/2. We
see that the excitations in the anti- ferromagnetic ground state [AFM can
be interpreted as spinon excitations over the spinon vacuum, this point of
view is particularly suitable to desribe the anti-ferromagnetic case only. For
the ferromagnet case the ground state is a magnon-vacuum and excitations
are interpreted as magnons.
3.4.1 XXZ Model
We will briey state the results for the XXZ-model that can be obtained by
the ABA through a very simular method to the XXX-model. The results
presented here will be based on [12] and [13]. The generalized anisotropic
Heisenberg Hamiltonian as introduced in Chapter 2 is given by:
H =
N

j=0
(J
x
S
x
j
S
x
j+1
+J
y
S
y
j
S
y
j+1
+J
z
S
z
j
S
z
j+1
)
The XXZ model in particular corresponds to J
x
= J
y
= J and J
z
= J,
where is the measure of anisotropy:
H = J
N

j=0
(S
x
j
S
x
j+1
+S
y
j
S
y
j+1
+ S
z
j
S
z
j+1
),
where = 1 corresponds to the XXX model. For the XXZ model, the Lax
operator is given by:
L
a,n
() =
_
sinh ( +S
z
n
) i2S

n
sinh()
S
+
n
sinh () sinh ( S
z
n
)
_
,
where is the spectral parameter. The relationship between and is given
by = cosh . This will satisfy the Yang-Baxter algebra, in particular the
FCR relation as given before:
R
a
1
,a
2
( )L
a
1
,n
()L
a
2
,n
() = L
a
2
,n
()L
a
1
,n
()R
a
1
,a
2
( ),
on End(A
1
A
2
V
n
). This will be given as a direct consequence by the
following R-matrix (R() acting on C
2
C
2
) which is a solution to the
Yang-Baxter equation as before:
R() =
1
sinh ( +)
_
_
_
_
sinh ( +) 0 0 0
0 sinh sinh 0
0 sinh sinh 0
0 0 0 sinh ( +)
_
_
_
_
.
Indeed we have the R-matrix and the Lax operator in a simular form as
L
a,n
() = R
a,n
( /2) which is simular to what we had in the XXX
54
model. As with the XXX model, we have the Monodromy matrix:
T
a
() = L
a,N
() L
a,1
() =
_
A() B()
C() D()
_
Transfer matrix:
t() = tr
a
T
a
() = A() +D()
It follows as from the isotropic case that these transfer matrices generate
higher conserved charges, in particular the Hamiltonian which can be written
as:
H = (2J sinh ) t
1
(/2)t

(/2) + const
= 2J sinh

ln t()

=/2
+ const
The eigenvalues of t() are again given as excitations from the pseudo-
vacuum [0,
[
1
, ,
n
=
r

i=1
B(
i
)[0,
where B() and C() are assumed to be creation and annihilation operators
as before.
i
also satisfy the BAEs which in the XXZ model are given by:
_
sinh (
j
+/2)
sinh (
j
/2)
_
N
=
r

l=1,l=j
sinh (
j

l
+i)
sinh (
j

l
i)
, j = 1, , r
For the XXZ model, the energy is given by:
E =
JN
4
+J
r

j=1
(cos k
j
),
which for the isotropic limit 1 reduces to the energy that we calculated
before for the XXX model.
The construction of [0
AFM
= [AFM (J = 1) is credited due Hulthen
[23] in the thermodynamic limit N . For 1 < < 1 the value in the
thermodynamic limit for the AFM (J = 1):
E
AFM
= N

4
N
sin
2
_

dz
sinh (1 )z
sinh z cosh z
= N

4
+N
sin
2
_
+i/2
+i/2
dx
1
sinh x
cosh x
sinh x
Where the anisotropy parameter is parameterized in [23] as = cos .
In the XXX limit 1, the ground state energy is more simply given by:
E
AFM
= N
_
1
4
ln 2
_
,
55
which agrees with the result that we obtained in the previous section in the
thermodynamic limit N > for the anti-ferromagnet ground state. It
is useful to introduce the ground state energy per site which will simply be
dened as:
e
0
= lim
N
E
AFM
N
For the above we obtain:
e
0
=

4
+
sin
2
_
+i/2
+i/2
dx
1
sinh x
cosh x
sinh x
,
and in particular for the XXX spin chain:
e
0
=
_
1
4
ln 2
_
56
Chapter 4
Higher Conserved Charges of
the Heisenberg Hamiltonian
In this chapter we will be looking at the Higher Conserved charges of the
Heisenberg Hamiltonian which will produce the local conservation laws of
the XXX-model. The material in this chapter will be based on [14],[15] and
[16].
In the previous chapter we saw that the transfer matrix t() produced
a mutually commuting set Q
n
of N 1 independent conserved quantities,
essentially given as the coecients of the power series of t(). We also
saw that we could generate the Hamiltonian from the transfer matrix, in
particular it was related to second term in the expansion (coecient of the
term) via:
H = J
_
i
2
t(i/2)Q
1

N
4
_
and
Q
1
=
d
d
t()

=i/2
A local conserved charge is a sum over the whole chain of terms that are
supported (i.e. non-trivial) only on a nite number of sites, which makes
sense as a denition in the large N limit. For nite N, these number of sites
will be suciently small compared to N, we will give a more mathematical
description in the next section. In order to satisfy the locality conditions, a
Hamiltonian should be expressible as a sum of terms each one only involving
spins close to each on the chain. The Heisenberg Hamiltonian in particular
is a local conservation law, but the other coecients Q
n
are in general
not. These conserved quantities are rather intractable due to their non-
locality, we wish to consider only local conservation charges since they are
the characteristic of integrability. If we look at the Hamiltonian, we see that
in the left hand side we notice that t(i/2)Q
1
= t(i/2) t

(i/2) corresponds
57
to the rst derivative of the logarithm of the Transfer matrix evaluated at
= i/2. This provides us a clue to what we must do in order to generate
local conserved quantities which we address in this section.
4.1 Logarithmic derivative of the transfer matrix
A more useful set of operators to consider are given by coecients from
the expansion of P() = ln (t()), given essentially as derivatives of P()
evaluated at some point =
0
. It can be shown that the higher derivatives
of this logarithm commute and so form a set of conserved quantities, taking
the logarithm ensures locality but it is very non-trivial to see this explicitly.
We can write P() as a power series in the spectral parameter around

0
= i/2 as:
P() = ln (t()) =

n=1
(
0
)
n
p
n
n!
Notice that the expansion of P() will provide innitely many coe-
cients, but these coecients will be expressed as functions of the coecients
of t() (i.e. Q
n
) - so they are not all algebraically independent, there are
non-linear algebraic relations between them. In general if Q
1
, , Q
n
are
conserved then so is p
i
= f(Q
1
, , Q
n
), so these coecients constructed
from other conserved charges do not produce any new conservation laws.
It is clear from the expansion of t() in the previous section that we
cannot have more than N1 conserved quantities. Since t() is a polynomial
of order N, we have the following properties which we have been shown or
are inferred, true :
t
(N1)
() = 0, t
(N)
() = 2N!, t
(i)
() = 0 i > N
This controls the expansion of P() in some sense and prevents higher
than N order coecients from being algebraically independent. The coe-
cient of P() at order higher than N will therefore not produce new charges,
these will be algebraically related to the local conserved charges at lower or-
ders and will lose locality since it is focused around a site of extent of order
N.
The Hamiltonian is related to p
1
by:
H = ap
1
+b
Where a and b are constants, in this case given by a = J/2i, b = JN/4
Since p
1
corresponds to the Hamiltonian, the coecients of the above
expansion will form a commuting set of local conservation laws, the locality
was shown by Luscher [17]. In the previous section when we showed that we
can generate the Hamiltonian from the derivative of the transfer matrix, it
was explicitly clear that the Hamiltonian was local. When we multiplied the
58
derivative of t() by t
1
(), we cancelled out the most of the permutation
matrices and we were left with identities everywhere except for a small bit of
size n ( P
n,n+1
). This is simular to what happens in higher local charges,
admitting an expansion as:
p
n
=
d
d
ln (t())

=
0
= t
1
(
0
)t
(n)
(
0
) + polynomials,
where the polynomials depend on lower order local charges. It can be shown
in [21] that the expectation value of such higher charges is actually equal to
the expectation value of the rst term i.e. t
1
(
0
)t
(n)
(
0
).
Higher charges p
i
for i > 1 correspond to Hamiltonians with more neigh-
bours interacting. Since these charges are local operators, they can be put
in the form [17]
1
:
p
n
=

{i
1
, ,i
n1
}
G
T
n1
(i
1
, , i
n1
)
Where the summation is over the ordered subsets i
1
, , i
n1
1, , N
and G
T
is a translationally covariant and totally symmetric function, obey-
ing the locality property:
G
T
n
(i
1
, , i
n
) = 0
for [i
n
i
1
[ n. For the innite XXX chain, further properties of the
conserved charges, including their completeness, have been proved in [22].
It is very dicult to extract the explicit form of the charges directly by
computing higher derivatives of P(), this is because the size for the transfer
matrix grows exponentially with the length of the chain. Fortunately there
exists a shortcut which will involve an alternative method of constructing
higher charges with a boost operator. See [14]. The Boost operator is dened
as the rst moment of the Hamiltonian i.e.:

B =
N

n=1
n

S
n


S
n+1
We use the fact that its commutator with the transfer matrix is equal to the
derivative i.e.:
[

B, t()] =

t()
A consequence of this is that up to some constant terms, the boost operator
generates conserved quanties recursively i.e.:
p
n+1
= [

B, p
n
]
1
Shown by Luscher for the XYZ model
59
Recall
p
1
=
d
d
ln (t())

=i/2
=
1
i
N

n=1
P
n,n+1
=
J
2a
N

n=1
S

n
S

n+1

b
a
We can omit the constants in this expression for convenience, since multiply-
ing the charge by a constant will not change its conservation or locality and
[

B, k] = 0 for any constant k. This allows us to concentrate mainly on the
part with the spin matrices i.e. p
1


N
n=1
S

n
S

n+1
. Using our knowledge
of p
1
and the above, we will now proceed to calculate p
2
and p
3
:
p
2
= [

B, p
1
]
=

n,m
n[S
a
n
S
a
n+1
, S
b
m
S
b
m+1
]
In order to compute this, we will need the following identities for commuta-
tors and the knowledge of the su(2) algebra of spins (see appendix):
[A, BC] = [A, B]C +B[A, C]
[AB, C] = A[B, C] + [A, C]B
[AB, C D] = A[B, C]D + [A, C]BD +C A[B, D] +C[A, D]B
[S
a
n
, S
b
m
] = i
n,m

abc
S
c
m
p
2

n,m
m
abc
(
m+1,n
S
a
m
S
c
i
S
b
i+1
+
m,n
S
c
n
S
a
m+1
S
b
n+1
+
m+1,n+1
S
b
n
S
a
m
S
c
n+1
+
m,n+1
S
b
n
S
c
n+1
S
a
m+1
)
=
N

n=1

abc
[(n 1)S
a
n1
S
c
n
S
b
n+1
+nS
c
n
S
a
n+1
S
b
n+1
+nS
b
n
S
a
n
S
c
n+1
+ (n + 1)S
b
n
S
c
n+1
S
a
n+2
]
Since
abc
is antisymmetric, terms with two Ss acting on the same site
vanish, we obtain:
p
2
=
N

n=1

abc
[(n 1)S
a
n1
S
c
n
S
b
n+1
+ (n + 1)S
b
n
S
c
n+1
S
a
n+2
]
If we shift the rst term by 1 then recombine both terms we get:
p
2
=
N

n=1

abc
[nS
a
n
S
c
n+1
S
b
n+2
+ (n + 1)S
b
n
S
c
n+1
S
a
n+2
]
=
N

n=1
(n
acb
+ (n + 1)
cab
)S
a
n
S
b
n+1
S
c
n+2
=
N

n=1

abc
S
a
n
S
b
n+1
S
c
n+2
60
By the denition of the cross product:
(AB)
i
=
ijk
A
j
B
k
Finally we have:
p
2
=
N

n=1
(

S
n


S
n+1
)

S
i+2
Indeed this has a local form that acts on the nearest and next-to- nearest
neighbouring sites. In the same way iteratively we will obtain p
3
:
p
3
= [

B, p
2
] =

m,n
m
abc
[S

m
S

m+1
, S
a
i
S
b
i+1
S
c
i+2
]
By repeatedly using the commutator identities, we get:
[AB, C DE] = A[B, C]DE + [A, C]BDE +C A[B, D]E
+C AD[B, E] +C[A, D]E B +C D[A, E]B
p
3
=

n,m
m
abc
(S

m
[S

m+1
, S
a
n
]S
b
n+1
S
c
n+2
+ [S

m
, S
a
n
]S

m+1
S
b
n+1
S
c
n+2
+S
a
n
S

m
[S

m+1
, S
b
n+1
]S
c
n+2
+S
a
n
S

m
S
b
n+1
[S

m+1
, S
c
n+1
]
+S
a
n
[S

m
, S
b
n+1
]S
c
n+2
S

m+1
+S
a
n
S
b
n+1
[S

m
, S
c
n+1
]S

m+1
)
Inserting the commuting relation for the spin algebra:
p
3

n,m
m
abc
(
ad

m+1,n
S

m
S
d
n
S
b
n+1
S
c
n+2
+
ad

m,n
S
d
n
S

m+1
S
b
n+1
S
c
n+2
+
bd

m+1,n+1
S
a
n
S

m
S
d
n+1
S
c
n+2
+
cd

m+1,n+2
S
a
n
S

m
S
b
n+1
S
d
n+2
+
bd

m,n+1
S
a
n
S
d
n+1
S
c
n+2
S

m+1
+
cd

m,n+2
S
a
n
S
b
n+1
S
d
n+2
S

m+1
)
Using the same procedure as before and shifting appropriate terms we
nally get:
p
3
= 2
N

n=1
[(

S
n


S
n+1
)

S
n+2


S
n+3
+

S
n


S
i+2
] 4p
1
61
Chapter 5
Correlation functions
The material based here will be from [17], [18] and [19].
The matrix element computed by inserting a product of operators be-
tween two states, usually the vacuum states, is called a correlation function.
Consider an operator acting on the j-th site, S
a
j
j
S
x
n
, S
y
n
, S
z
n
. The aver-
age of this operator with respect to the ground state [0 is:
0[
n

j=1
S
a
j
j
[0 =
n

j=1
S
a
j
j

We will particularly be looking at the thermodynamic limit for the anti-
ferromagnet (J < 0) since the correlation functions for the ferromagnet
(J > 0) are trivial since the ground state has all the spins aligned. The
XXZ model for the antiferromagnetic phase (J = 1), the energy is given
by:
E =
N
4
+
M

j=1
(cos k
j
)
The construction of [0
AFM
(see [19]) is credited to Hulthen [23] in the
thermodynamic limit N , which we also stated in Section 3. For
1 < < 1 the value in the thermodynamic limit the ground state energy
per site is given by:
e
0
= lim
N
E
AFM
N
=

4

sin
2
_

dz
sinh (1 )z
sinh z cosh z
=

4
+
sin
2
_
+i/2
+i/2
dx
1
sinh x
cosh x
sinh x
Given
H =
N

j=1
(S
x
j
S
x
j+1
+S
y
j
S
y
j+1
+ S
z
j
S
z
j+1
)
62
Where = 1 corresponds to the XXX model.
e
0
() =
0[H[0
N
= (2a +b)
Where a, b are given by:
a = S
x
j
S
x
j+1
= S
y
j
S
y
j+1
, b = S
z
j
S
z
j+1

By the Hellmann Feynman theorem:


S
z
n
S
z
n+1
=
d
d
e
0
()

=1
,
i.e we can compute the nearest-neighbour correlation directly from our ex-
pression for the anti-ferromagnetic ground state energy. It it well known
that the nearest neighbour correlator can be obtained from the ground state
energy per site by Hulthen [23], what is less obvious is calculating higher
neighbouring correlators. I will state a few results that have been obtained:
S
z
n
S
z
n+1
=
1
12

1
3
ln 2
=
1
12

1
3

a
(1) = 0.1477157168
which is obtained as above from the ground state energy per site.
S
z
n
S
z
n+2
=
1
12

4
3
ln 2 +
3
4
(3)
=
1
12

4
3

a
(1) +
a
(3) = 0.06067976995 ,
which is obtained from the half-lled Hubbard model by Takahashi [25].
The next nearest correlators and other type of correlation functions up to
8 lattice sites have been calculated in many papers (see [18]) and it has
the same pattern structure as above, given in terms of the Riemann zeta
functions
1
. We generally would like to compute S
z
j
S
z
j+n
n.
5.1 EFP
An important class of correlation functions called the Emptiness Formation
Probability (EFP) which is the probability to nd the formation of a ferro-
magnetic string of length n in the anti- ferromagnetic ground state [AFM.
The EFP is dened as:
P(n) = AFM[
n

j=1
P
j
[AFM =
n

j=1
(
1
2
+S
z
j
)
1
This was dened in the introduction
63
Where P
j
= 1/2 + S
z
j
is the projector on the state with the spin up in the
j-th lattice site. The importance of the EFP was realised in [28]. It was
conjectured in [18] that P(n) is always in terms of ln 2 and Riemann zeta
functions (n) with odd arguments and rational coecients. This is also
the case for S
z
j
S
z
j+n
since it was also conjectured that we can write these
correlators in terms of P(n). In particular, we can recover the rst and
second-neighbour correlators by the following relations
P(2) = (
1
2
+S
z
j
)(
1
2
+S
z
j+1
) = S
z
j
S
z
j+1
+
1
4
P(3) = S
z
j
S
z
j+1
+
1
2
S
z
j
S
z
j+2
+
1
8
Though it is important to note that S
z
j
S
z
j+3
cannot solely be determined
from P(4). It was also shown in [26] that P(n) exhibits interesting Gaussian
decay at n . The rst four values of the EFP look as follows:
P(1) =
1
2
= 0.5
P(2) =
1
3
(1 ln 2) = 0.102284273
P(3) =
1
4
ln 2 +
3
8
(3) = 0.007624158
P(4) =
1
5
2 ln 2 +
173
60
(3)
11
6
(3) ln 2
51
80

2
(3)

55
24
(5) +
85
24
(5) ln 2
The value of P(1) is evident from the symmetry, P(2) can be extracted
from the explicit expression of the ground state energy from S
z
j
S
z
j+1
as we
have shown above. P(3) can be extracted from the results M. Takahashi in
[25] on the calculation of the next to nearest neighbour correlation, as we
discussed above. P(4) can be extracted from the integral representation of
the EFP as in [26].
5.1.1 Integral representation of EFP
There is an explicit formula for P(n) which was determined in [28] and in
the XXX limit reads:
P(n) =
n

j=1
_
C
d
j
2i
U
n
(
1
, ,
n
) T
n
(
1
, ,
n
)
Where:
U
n
(
1
, ,
n
) =
n(n+1)/2

1k<jn
sinh (
k

j
)

n
j=1
sinh
n

j
64
T
n
(
1
, ,
n
) =

n
j=1

j1
j
(
j
+ 1)
nj

1k<jn
(
j

k
i)
The contour C goes parallel to the real axis with the imaginary part
conned between 0 and i for each integral, though in principle it can be
chosen arbitrary between the two points.
It can be shown in [29] that the expression for P(2) reduces to the one
dimensional integral for arbitrary as:
P(2) =
1
2
+
1
2
2
sin

_
sin
_

sinh (1 )x
sinh xcosh x
d x
_
,
which leads to a one-dimensional integral representation for S
z
n
S
z
n+1
, since
we have S
z
n
S
z
n+1
= P(2) 1/4. This reduces to the same form as the
ground-state energy per site e
0
shown above by the Feynmann Hellman
theorem.
One can compute these type of integrals in general by transforming the
integrand to a canonical form and thereby reducing it in such a way that
the integral doesnt change, then one can perform the integral simply using
the residue theorem. This method of evaluating integrals was introduced
by Boos and Korepin (see [26],[27]). I will briey go through the details for
P(3) as in [27].
P(3) =
3

j=1
_
i/2
i/2
d
j
2i
U
3
(
1
,
2
,
3
) T
3
(
1
,
2
,
3
)
U
3
(
1
,
2
,
3
) =
6

1k<j3
sinh (
k

j
)

3
j=1
sinh
3

j
T
3
(
1
,
2
,
3
) =

3
j=1

j1
j
(
j
+ 1)
3j

1k<j3
(
j

k
i)
=
(
1
+i)
2
(
2
+i)
2
3
(
2

1
i)(
3

1
i)(
3

2
i)
By utilizing the anti-symmetry and some analytic properties of U
3
(
1
,
2
,
3
),
we can replace the integrand T
3
(
1
,
2
,
3
) by a certain canonical form
T
C
3
(
1
,
2
,
3
), the details of the transformation to canonical form are given
in [26],[27], which was derived as:
T
3
(
1
,
2
,
3
) T
C
3
(
1
,
2
,
3
) = P
(0)
3
+
P
(1)
3

1
Where P
(0)
3
= 2
2

2
3
and P
(1)
3
= 1/3i
1
i
2
i
3
2
1

3
. Performing
the integration by the residue theorem, we obtain:
J
(0)
3
=
3

j=1
_
i/2
i/2
d
j
2i
U
3
(
1
,
2
,
3
) P
(0)
3
=
1
4
65
J
(1)
3
=
3

j=1
_
i/2
i/2
d
j
2i
U
3
(
1
,
2
,
3
)
P
(0)
1

1
= ln 2 +
3
8
(3)
and hence the result for P(3) is given by:
P(3) = J
(0)
3
+J
(1)
3
=
1
4
ln 2 +
3
8
(3),
which agrees with the result as above. This result also allows us to com-
pute the second-neighbour correlation function since we have S
z
n
S
z
n+2
=
2(P(3) P(2)) + 1/4 which is obtained by rearranging the expression for
P(3) in terms of S
z
n
S
z
n+2
and substituting the expression for P(2).
5.1.2 Satos formula
The exact calculation of correlation functions has been a non-trivial task,
except in the limit 0 for the XX Ising model. In this case an exact
formula can be derived in terms by the use of Wicks theorem as in [30],[31]:
S
z
n
S
z
n+k
=
(1 (1)
k
)
2
2
k
2
.
As we have mentioned before, it was conjectured that any correlation func-
tions of the XXX spin chain can be represented in terms of ln 2 and Riemann
zeta functions (n). This conjecture was actually proved in 2006 by H.boos,
M. Jimbo, T. Miwa, F smirmnov and Y.Takeyama [24]. It is reasonable
therefore to ask whether there is an explicit expression for all the rational
coecients that enter the expressions for the correlation function, this does
not exist yet and is an open problem. Fortunately, a recent development by
Juw Sato [18]
2
, who obtained a formula for the linear terms and some of the
higher terms for the coecients. In order to present the formula let
a
(s)
be the alternating Riemann zeta function, we will dene:
c
0
=
1/4
a
(1)
3
c
n
=

a
(2n 1)
a
(2n + 1)
3
Then Satos formula is dened by:
S
z
j
S
z
j+n
=
n

k=0
(1)
k
_
n 1
k
__
2k + 1
k
_
c
k
24
_
n 1
2
_
c
1
(c
0
+c
1
)+higher terms
The higher terms represent terms which contribute for n > 7. This for-
mula can reproduce the result up to n = 7, though the calculation itself can
2
Though this result was not directly obtained in [18], the author has listed the reference
as a private communication, as such no other references exist to this
66
be rather tedious. For the nearest-neighbour correlator, we immediately
have according the formula:
S
z
n
S
z
n+1
=
1

k=0
(1)
k
_
0
k
__
2k + 1
k
_
c
k
= c
0
=
1
12

1
3

a
(1),
which is the same result obtained by the ground state energy and the integral
formulation.
67
Chapter 6
Conclusion
We started our discussion by deriving the Heisenberg Hamiltonian through
considerations of the Hydrogen Molecule and the Pauli exclusion principle.
We then examined some of the symmetries of the model and saw that it
possed full rotational, translational and reection symmetry. The key be-
hind the CBA is through exploiting these known symmetries . We then
solved the model via the CBA and obtained the energy spectrum by making
an Ansatz about the coecients for the general eigenstate of the model.
By considering the periodicity condition on these coecients led to the for-
mulation of the Bethe Ansatz equation (BAE). In the context of the ABA
framework, we then introduced the mathematical description of integrabil-
ity. We thus showed from that the transfer matrix we can produce the
Hamiltonian, and generate local conserved quantities that commute with
the Hamiltonian from which we were able to deduce the integrability of the
model. The Ansatz for the ABA is the assumption that B() and C()
through the entries of the transfer matrix are creation and annihilation op-
erators respectively, this lead to conditions on the parameter which led to
same BAEs as from the CBA. By using the fact that the (pseudo)-vacuum
[0 is a highest weight vector for the global spin algebra su(2), we then
showed that this was the case for any general eigenstate. We then looked
at the thermodynamic limit N and in particular examined the string
hypothesis. This lead us to calculate the non-trivial anti- ferromagnetic
ground state energy.
We then stated some results for the XXZ-model in the context of the
ABA and the thermodynamic limit, which immediately reduced to the XXX-
model in the limit 1. We then examined the higher conserved charges,
where we had to take the charges hailing from the expansion of the loga-
rithm transfer matrix in order to maintain locality. We thus showed that
we can produce a family of N 1 indepedant local charges from the trans-
fer matrix for the XXX Hamiltonian, which when by adding the total spin
operator S
z
(which is also local) produces N such charges and thus proves
68
the integrability of the model. Finally, we looked at correlation functions
for the XXX anti-ferromagnetic spin chain, rst the nearest neighbouring
correlator which we saw can be calculated directly from the ground state
energy per site. Other correlators as we saw are not as obvious to compute,
we then stated some known results in and in particular stated modestly ways
to compute such correlators.
Some comments have to made on the ABA and CBA: Though the CBA
and ABA led to the same spectra, the ABA has the advantage of an algebraic
framework easily generalizable to other models such as the XXZ spin chain,
it allows us to prove the integrability of the model which the CBA lacks and
it also allows us to more readily compute other quantities such as correlation
functions, using methods such as the quantum inverse scattering method.
The CBA does provide a picture of the excitations in the spin chain in
terms of the concept of a magnon (for the ferromagnet magnon-vacuum),
with which its description leads natrually to establish the momentum of
the magnon. We also saw that the energy was additive and the that the
total momentum of the system was conserved, this description is harder to
imagine in the ABA because of its purely algebraic nature.
Though the Heisenberg XXX spin chain is one of the bestl understood
models in terms of its integrability, open problems within the area still exist
such as to produce a general formula for any correlation function. As we saw
in Chapter 5, these correlators have been proven to be expressible in terms
of polynomials of the alternating zeta function, with a regular pattern as can
be directly observed. It is reasonable to assume that such a formula does
exist, perhaps as a generalization of Satos formula given in Chapter 5. It is
advantageous to therefore to try to determine the coecients either by brute
force in trying to recognize a pattern by nding higher correlators, or by a
more elegant approach utilizing mathematical tools from number theory in
the description of the Riemann zeta function. I will personally look into such
a problem during my spare time as I feel a more mathematically rigorous
treatment is needed in order to realize the full potential of the connection
between the Riemann zeta function and the Heisenberg spin chain.
69
Bibliography
[1] Spin Waves, Theory and Applications. Stancil, D.D.;Prabhakar A.
(2009), XII, 348 P.20, ISBN:978-0-387-77864-8
[2] Derivation of model Hamiltonians. Gagliano, R A.A.; Gustavo A.
(1985). Phys Rev, 32(1)
[3] Simple Models of Magnetism. Skomski, R. (2008) (Oxford Graduate
Texts). ISBN: 978-0198570752
[4] A Spin Chain Primer. Nepomechie, R. I. (1998) arXiv:
hep-th/9810032v1
[5] Introduction to the Bethe ansatz I. Karbach, M.; Muller, G. (1998).
arXiv: cond-mat/9809162
[6] On the Theory of Metals. H.Bethe.(1931) Z.Phys 71 205
[7] The many-body problem: an encyclopedia of exactly solved models in
one dimension. (1993). Mattis C.D. World Scientic Publishing Co Pte
Ltd ISBN: 978-9810209759
[8] How Algebraic Bethe Ansatz works for integrable model. Faddeev, L.D.
(1996) arXiv: hep-th/9605187v1
[9] Bethe Ansatz for Heisenberg XXX Model. Lin, S. (1995) arXiv:
cond-mat/9509183v1
[10] Review of AdS / CFT Integrability, Chapter III.1: Bethe Ansatze and
the R-Matrix Formalism. Staudacher, M. (2010) arXiv: 1012.3990v2
[11] Notes on Bethe Ansatz Techniques. Franchini, F. (2010) arXiv:
hep-th/9605187v1
[12] Anisotropic (XXZ) Heisenberg Chain. Steenstra, J. (2011)URL:
http://tinyurl.com/7oc4pld
[13] Quantum Magnetism. Lecture Notes in Physics. Board, E. (2004)
ISBN: 3540214224
70
[14] Ladder operator for the one-dimensional Hubbard model. Links, J.;
Zhou, H.; Mckenzie, G. (2001) arXiv: cond-mat/0011368v2
[15] Integrability test for spin chains. Grabowski, M.P.; Mathieu, P.
(1994) arXiv: hep-th/9412039v1
[16] Quantum Integrals of Motion for the Heisenberg Spin
Chain.Grabowski, M.P.; Mathieu, P. (1994). arXiv: hep-th/9403149v1
[17] Dynamical Charges in the Quantized, Renormalized Massive Thirring
Model. Lusher, M. (1976). Nucl. Phys. B 117, 475.
[18] Some Open Problems in Exactly Solvable Models. Korepin, V. E.
(2008). URL: http://insti.physics.sunysb.edu/ korepin/montreal.pdf
[19] Exact Calculation of Correlation Functions for Spin-1/2 Heisenberg
Chain. Shiroishi, M. (2004). doi: 10.1143/JPSJS.74S.47
[20] Integrability of quantum chains: theory and applications to the
spin-1/2 XXZ chain. Klmper, A. (2005) arXiv: cond-mat/0502431v1
[21] Integrable quantum spin chains and their classical continuous
counterparts. Avan, J.; Doikou, A.; Sfetsos, K. (n.d.). (2010)
arXiv:1102.0425v1
[22] Babbit, D.; Thomas. (1979). L. J. Math. Anal. 72, 305
[23] Hulthen, L. (1938). Arkiv Mat. Astron. Physik A 26 1.
[24] Density matrix of a nite sub-chain of the Heisenberg
anti-ferromagnet. Boos, H.; Jimbo, M.; Miwa, T.; Smirnov, F.;
Takeyama,Y. (2006). Lett. Math. Phys. 75, 201 arXiv:hep-th/0506171
[25] Half-lled Hubbard model at low temperature. (1977). Takahashi M.
J. Phys. C 10, 12891301.
[26] Quantum spin chains and Riemann zeta function with odd arguments.
Boos H.E.; Korepin V.E. (2001). J. Phys. A: Math. Gen. 34, 53115316.
arXiv: hep-th/0104008
[27] Evaluation of integrals representing correlations in XXX Heisenberg
spin chain. Boos H.E.; Korepin V.E. (2001). Math. Phys., Vol. 23,
65108. arXiv: hep-th/0105144
[28] Korepin, V.E.; Izergin, A.G.; Essler, F.; Uglov, D.B. (1994). Phys.
Lett. A 190, 182-184.
[29] Jimbo, M.;Miwa, T. (1996) J. Phys. A 29 2923.
71
[30] E. Lieb, T. Schultz and D. Mattis. (1961). Ann. Phys. (N.Y.) 16 407.
[31] McCoy, B.M. (1968) Phys. Rev. 173531.
72

You might also like