You are on page 1of 25

Correlation Functions of 1D Spin Chains

M. Sc. Project (foundation) Report

Submitted to

Department of Physics,

Savitribai Phule Pune University, Pune

For the partial fulfillment of the degree of Master of Science (M. Sc.)

IN

PHYSICS

By

Dennil Joby

Under the supervision of

Dr. P. Durganandini

Internal-Guide

(Prof.Dr. P. Durganandini, Department

of Physics, SPPU)

December 2023
CERTIFICATE

This is to certify that the work incorporated in the Project (foundation)(Subject

Code: PHY-C375) Report entitled Correlation Functions of 1D Spin Chains”

submitted by Dennil Joby (PRN: 22021045 ) was carried out by the student

under my/our supervision. This project work is submitted by the student as a

part of partial fulfilment of the degree of Master of Science (Physics) course of

Savitribai Phule Pune University, Pune.This work is original and not submitted

before for any other degree at this or any other University.

Place: Pune, India.

Date: …….

Signature of Internal-Guide

(Prof.Dr. P. Durganandini, Department of

Physics, SPPU)

Head, Department of Physics,


Savitribai Phule Pune University, Pune
DECLARATION

I hereby declare that the content or any part of the project (foundation) report

entitled “Correlation Functions of 1D Spin Chains” is not copied verbatim

from any sources, such as journals, books, magazines, reports, Ph.D. / M.Phil.

Thesis, websites etc. In cases, where certain sentences / paragraphs / figures

etc. are copied verbatim, proper citation is given to the original source(s).

Name of the student: Dennil Joby

PRN of the student: 22021045

Signature of the student:


Table of Contents

1. The 1D Heisenberg Chain ......................................................................... 2


1.1 Action of Spin Operators on the Spin Chain ....................................................................... 3
1.2 Spin-1/2 Chain as a Quantum many Body Problem ........................................................... 4
1.3 The Bethe Ansatz for Arbitrary Number of Deviations ...................................................... 8

2. The DMRG Algorithm.............................................................................. 10


2.1 The Outline of the Algorithm .......................................................................................... 11
2.2 The DMRG Applied to a 4-spin Superblock .................................................................... 12
2.3 Measurements using the DMRG Algorithm .................................................................... 15

3. Correlation Functions ............................................................................... 16


3.1 Defining Correlation Functions ....................................................................................... 16
2.2 Correlation Functions in Ising Model .............................................................................. 18
2.3 Conclusion and Outlook.................................................................................................. 21

4. Bibliography .............................................................................................. 22

1
Chapter 1- The 1D Heisenberg Chain

A spin chain is defined as a 1 dimensional lattice with 𝑁 sites, where on each site we consider a
spin-1/2 particle (in this case an electron). This electron can be in spin or in spin down state,
consequently leaving each electron to be defined by a two dimensional Hilbert Space with the
basis state (10) and (01) locally. Given 𝑁 such electrons the combined Hilbert space would be
given as;

𝑉 = ℂ12 ⊗ ℂ22 ⊗ ℂ23 ⊗ ⋯ ⋯ ⊗ ℂ2N

The Heisenberg spin chain is defined as a spin-1/2 chain with 𝑁 sites on a one dimensional
lattice with the periodic boundary conditions

𝑺𝑖+𝑁 = 𝑺𝑁

And the Hamiltonian given as;


𝑁 𝑁
𝑦 𝑦
𝐻 = −𝐽 ∑ 𝑺𝑖 ∙ 𝑺𝑖+1 = −𝐽 ∑[𝑆𝑖𝑥 𝑆𝑖+1
𝑥
+ 𝑆𝑖 𝑆𝑖+1 + 𝑆𝑖𝑧 𝑆𝑖+1
𝑧
]
𝑖=1 𝑖=1

𝑥,𝑦,𝑧
Where 𝑆𝑖 are the usual spin operators operating on the 𝑖𝑡ℎ spin and satisfying the
commutation relations

[𝑆𝑖𝑎 , 𝑆𝑗𝑏 ] = 𝛿𝑖𝑗 𝜖 𝑎𝑏𝑐 𝑆𝑖𝑐

for 𝑖 = 1, 𝑁; 𝑖 ≠ 𝑗 and 𝑎, 𝑏, 𝑐 ≡ 𝑥, 𝑦, 𝑧

and 𝐽 is the Heisenberg exchange interaction, which basically measures the strength of the
interaction between two neighboring spins, where if 𝐽 > 0 the lowest energy occurs when all the
spins point in the same direction;

giving rise to a ferromagnetic arrangement. In the opposite case if 𝐽 < 0 the lowest energy would
occur for the anti-ferromagnetic arrangement; i.e. when each pair of neighboring spins is anti-
parallel to the other;

2
Where in either of the cases the total energy of the system is given as

𝑁𝐽
𝐸=−
4
for both the ferromagnetic and the anti-ferromagnetic cases.

1.1 Action of Spin Operators on the Spin Chain

Let there be a two spin system with the basis states {|↑↑⟩, |↑↓⟩, |↓↑⟩, |↓↓⟩}. Which gives us
a 𝑁 = 2 1D spin chain, with the Hamiltonian;
𝑁=2
𝑦 𝑦
𝐻 = −𝐽 ∑[𝑆𝑖𝑥 𝑆𝑖+1
𝑥
+ 𝑆𝑖 𝑆𝑖+1 + 𝑆𝑖𝑧 𝑆𝑖+1
𝑧
]
𝑖=1
𝑦
Defining ladder operators as 𝑆𝑖± = 𝑆𝑖𝑥 ± 𝑖𝑆𝑖 , the hamiltonian becomes,
𝑁=2
1
𝐻 = −𝐽 ∑ [ (𝑆𝑖+ 𝑆𝑖+1

+ 𝑆𝑖−𝑆𝑖+1
+ )
+ 𝑆𝑖𝑧 𝑆𝑖+1
𝑧
]
2
𝑖=1
We will now consider the action of each of the terms in the Hamiltonian on the basis
states
A) 𝑆1𝑧 𝑆2𝑧
Given the state |↑↑⟩;
1 1 1
𝑆1𝑧 𝑆2𝑧 |↑↑⟩ = ( ) ( ) |↑↑⟩ = |↑↑⟩
2 2 4
Similarly for the rest of the basis states;
1 1 1
𝑆1𝑧 𝑆2𝑧 |↑↓⟩ = ( ) (− ) |↑↓⟩ = − |↑↓⟩
2 2 4
1 1 1
𝑆1𝑧 𝑆2𝑧 |↓↑⟩ = (− ) ( ) |↓↑⟩ = − |↓↑⟩
2 2 4
1 1 1
𝑆1𝑧 𝑆2𝑧 |↓↓⟩ = (− ) (− ) |↓↓⟩ = |↓↓⟩
2 2 4
(Note: we have set ħ = 1)

B) 𝑆1+ 𝑆2−
Given the state |↑↑⟩;
𝑆1+𝑆2−|↑↑⟩ = 𝑆1+𝑆2−(|↑⟩|↑⟩) = |0⟩|↓⟩ = 0
Similarly for the rest of the basis states;
𝑆1+ 𝑆2− |↑↓⟩ = 𝑆1+ 𝑆2− (|↑⟩|↓⟩) = |0⟩|0⟩ = 0
𝑆1+𝑆2− |↓↑⟩ = 𝑆1+𝑆2− (|↓⟩|↑⟩) = |↑⟩|↓⟩ = |↑↓⟩
𝑆1+𝑆2−|↓↓⟩ = 𝑆1+𝑆2−(|↓⟩|↓⟩) = |↓⟩|0⟩ = 0

3
C) 𝑆1− 𝑆2+
And following the above pattern we have;
𝑆1−𝑆2+|↑↑⟩ = 𝑆1−𝑆2+(|↑⟩|↑⟩) = |↓⟩|0⟩ = 0
𝑆1−𝑆2+ |↑↓⟩ = 𝑆1−𝑆2+ (|↑⟩|↓⟩) = |↓⟩|↑⟩ = |↓↑⟩
𝑆1− 𝑆2+ |↓↑⟩ = 𝑆1− 𝑆2+ (|↓⟩|↑⟩) = |0⟩|0⟩ = 0
𝑆1−𝑆2+|↓↓⟩ = 𝑆1−𝑆2+(|↓⟩|↓⟩) = |0⟩|↑⟩ = 0

1.2 Spin-1/2 Chain as a Quantum Many Body Problem

Imagine a system of 𝑁 spins with arbitrary alignment (spin-up or spin-down) placed on a


1D lattice with periodic boundary conditions. Let the state of this spin chain be denoted
by |𝜓⟩ where |𝜓⟩ also satisfies the eigenvalue equation;
̂|𝜓⟩ = 𝐸 |𝜓⟩
𝐻
where 𝐻 ̂ is the Heisenberg Hamiltonian defined above.
The problem at hand is to find all the possible wavefunctions of such a spin chain with an
arbitrary number of spin up (or down) at 𝑛 different lattice points (out of 𝑁).

We approach this problem by defining an “aligned” state |𝐴⟩ with all 𝑁 spins aligned, in
this case – spin up;
|𝐴⟩ = | ↑↑↑ ⋯ ⋯ ↑↑ ⋯ ⋯ ↑↑↑⟩

Based off of our results in section 1.1 we can operate the Hamiltonian on |𝐴⟩ giving;
̂ |𝐴⟩ = 𝐸𝐴 |𝐴⟩
𝐻
Or;
𝑁
1
{−𝐽 ∑ [ (𝑆𝑖+𝑆𝑖+1

+ 𝑆𝑖− 𝑆𝑖+1
+ )
+ 𝑆𝑖𝑧 𝑆𝑖+1
𝑧
]} |𝐴⟩ = 𝐸𝐴 |𝐴⟩
2
𝑖=1
Since for the aligned case both 𝑆𝑖+𝑆𝑖+1

and 𝑆𝑖−𝑆𝑖+1
+
yield a null result, the Hamiltonian
reduces to;

{−𝐽 ∑[𝑆𝑖𝑧 𝑆𝑖+1


𝑧 ] | ⟩
} 𝐴 = 𝐸𝐴 |𝐴⟩
𝑖=1
Which gives;
𝑁𝐽
̂|𝐴⟩ = −
𝐻 |𝐴⟩ = 𝐸𝐴 |𝐴⟩
4

4
Given that we have defined |𝐴⟩ to have the same number of spins as |𝜓⟩ we can interpret
|𝜓⟩ as a ladder operated version of |𝐴⟩ at the required lattice points.
For example if we were dealing with a 4-spin system |𝐴⟩ would look like;
|𝐴⟩ = | ↑↑↑↑⟩
And to obtain some |𝜓⟩ = | ↑↓↑↓⟩ from |𝐴⟩ we would simply have to operate the ladder
operators,
|𝜓′⟩ = 𝑆2− |𝐴⟩
|𝜓⟩ = 𝑆4− |𝜓′ ⟩ = 𝑆4−𝑆2−|𝐴⟩
Similarly for the 𝑁 spin case, we can obtain |𝜓⟩ from |𝐴⟩ in a similar fashion. Where we
can interpret each deviation from the initial aligned state as an independent deviated state,
where successive new deviations lead upto our desired |𝜓⟩

Keeping this in mind we now try evaluating the wavefunction of the single deviation state
or the state with exactly one spin down at some 𝑗th site.
Thus we 𝑁 such possible single deviation states labeled by the site of deviation as;
|1⟩ = |↓↑↑ ⋯ ⋯ ↑↑⟩
|2⟩ = |↑↓↑ ⋯ ⋯ ↑↑⟩
|3⟩ = |↑↑↓ ⋯ ⋯ ↑↑⟩
And for the 𝑗th case;
|𝑗 ⟩ = |↑↑↑ ⋯ ↓↑ ⋯ ↑↑⟩

Therefore, for a given state |𝑗 ⟩, if 𝑗 ≠ 𝑖, 𝑖 + 1


1
𝑆𝑖𝑧 𝑆𝑖+1
𝑧 | ⟩
𝑗 = |𝑗 ⟩
4
And if 𝑗 = 𝑖, 𝑖 + 1
1
𝑆𝑖𝑧 𝑆𝑖+1
𝑧 | ⟩
𝑗 = − |𝑗 ⟩
4
∑𝑁 [ 𝑧 𝑧 ]
Hence in all the terms occuring summation 𝑖=1 𝑆𝑖 𝑆𝑖+1 there are exactly two terms viz.
𝑗 = 𝑖 and 𝑗 = 𝑖 + 1 where the first condition is not satisfied. Therefore,
𝑁
1 1
∑[𝑆𝑖𝑧 𝑆𝑖+1
𝑧 | ⟩]
𝑗 = [(𝑁 − 2) + 2 (− )] |𝑗 ⟩
4 4
𝑖=1
𝑁
𝑁
∑ [𝑆𝑖𝑧 𝑆𝑖+1
𝑧 | ⟩]
𝑗 = ( − 1) |𝑗 ⟩
4
𝑖=1

Similarly for the eigenstates of 𝑆𝑖+𝑆𝑖+1



, we have for 𝑖 ≠ 𝑗
𝑆𝑖+ 𝑆𝑖+1
− | ⟩
𝑗 =0
And for the case when 𝑖 = 𝑗;
𝑆𝑖+𝑆𝑖+1
− | ⟩
𝑗 = |𝑗 + 1⟩

5
Yielding the summation equation as;
𝑁

∑[𝑆𝑖+𝑆𝑖+1
− | ⟩]
𝑗 = |𝑗 + 1⟩
𝑖=1
And in a similar fashion the summation of eigenstates of 𝑆𝑖− 𝑆𝑖+1
+
would yield;
𝑁

∑[𝑆𝑖−𝑆𝑖+1
+ | ⟩]
𝑗 = |𝑗 − 1⟩
𝑖=1
Hence the combined hamiltonian eigenvalue equation would look like,
𝑁
1
{−𝐽 ∑ [ (𝑆𝑖+ 𝑆𝑖+1

+ 𝑆𝑖−𝑆𝑖+1
+ )
+ 𝑆𝑖𝑧 𝑆𝑖+1
𝑧
]} |𝑗 ⟩ = 𝐸 |𝑗 ⟩
2
𝑖=1
Which when simplified would like;
1 𝑁
−𝐽 [ (|𝑗 + 1⟩ + |𝑗 − 1⟩) + ( − 1) |𝑗 ⟩] = 𝐸 |𝑗 ⟩
2 4
1
−𝐽 [ (|𝑗 + 1⟩ + |𝑗 − 1⟩) − |𝑗 ⟩] = (𝐸 − 𝐸𝐴 )|𝑗 ⟩
2

Now we can define an eigenbasis by defining the eigenvector |𝜙 ⟩ as;


𝑁
1
|𝜙 ⟩ = ∑ 𝑒 𝑖𝑘𝑗 |𝑗 ⟩
√𝑁 𝑗=1
2𝜋𝑚
where 𝑘 = , 𝑚 = 0, ⋯ , 𝑁 − 1.
𝑁
Combining this with the orthormanility property of the basis states we get;
1
−𝐽 [ (𝑒 𝑖𝑘 + 𝑒 −𝑖𝑘 ) − 1] = 𝐸 − 𝐸𝐴 = 𝜀𝑘
2
Or
𝜀𝑘 = 𝐽(1 − cos 𝑘)
Which gives us the energy of each eigenstate.

Here, the vectors |𝜙 ⟩ represent magnon excitations (∆𝑆 = 1 excitations) or spin waves, in
which the complete spin alignment of the aligned state |𝐴⟩ is periodically disturbed by a
2𝜋
spin wave with wavelength 𝜆 = . Given the nature of the states |𝜙 ⟩ (plane-wave), we
𝑘
can interpret the 𝑘 values as a momentum analog for magnons. Note that the 𝑘 = 0 state
is degenerate with |𝐴⟩ .

Operating the lowering operator at two spin sites (𝑗1 and 𝑗2 ) of the original aligned state
𝑁(𝑁−1)
we obtain set of possible two deviation states with state lables |𝑗1 , 𝑗2 ⟩ (we
2
𝑁(𝑁−1)
consider 𝑗2 > 𝑗1 ). Among these possible spin configuations we can distinctly
2

6
categorize the chains into two types, viz. the two down spins adjacent to each other (𝑗2 =
𝑗1 + 1) or the two down spins not adjacent to each other.

Extending the same set of operations as in the two deviation case we end up with two
simultaneous plane wave equations;
𝐽
(𝜀 − 2𝐽 )𝑓𝑙1 ,𝑙2 + (𝑓𝑙1 −1,𝑙2 + 𝑓𝑙1+1,𝑙2 + 𝑓𝑙1 ,𝑙2 −1 + 𝑓𝑙1 −1,𝑙2 ) = 0
2
𝑖𝑓 𝑙2 ≠ 𝑙1 + 1
And,
𝐽
(𝜀 − 𝐽 )𝑓𝑙1 ,𝑙2+1 + (𝑓𝑙1 −1,𝑙1 +1 + 𝑓𝑙1 ,𝑙1+2 ) = 0
2
𝑖𝑓 𝑙2 = 𝑙1 + 1
where
𝑓𝑙1 ,𝑙2 = 𝑒 𝑖𝑘1𝑙1 𝑒 𝑖𝑘2𝑙2
which when further simplified gives;
𝜀 = −𝐽(cos 𝑘1 + cos 𝑘2 − 2)
Which satisfy the following algebra,
2𝜋 𝜃
𝑘1 = 𝜆1 +
𝑁 𝑁
2𝜋 𝜃
𝑘2 = 𝜆2 −
𝑁 𝑁
𝜃 𝑘1 𝑘2
2 cot = cot − cot
2 2 2
Coming together to give the eigenstates fo the form,
𝑁
|𝜙 ⟩ = ∑ 𝑓𝑙1 ,𝑙2 |𝑗1 , 𝑗2 ⟩
1<𝑗1<𝑗2<𝑁
where,
𝑓𝑙1 ,𝑙2 = 𝐶(𝑒 𝑖𝑘1𝑙1 𝑒 𝑖𝑘2𝑙2 + 𝑒 𝑖𝜃 𝑒 𝑖𝑘2𝑙1 𝑒 𝑖𝑘1𝑙2 )

Here the states |𝜙 ⟩ represent interacting magnons where the interaction is reflected by the
phase shift 𝜃 and labeled by the deviation of the quasi-momenta 𝑘1 and 𝑘2 from the
single (free) magnon wave numbers (Franchini, 2017; Parkinson, 2016).

7
1.3 The Bethe Ansatz for Arbitrary Number of Deviations

In 1931 Hans Bethe a German-American theoretical physicist suggested and


subsequently proved (Bethe, 1947) that for an arbitrary number of deviations 𝑟 from the
aligned state |𝐴⟩, the solution could be basically obtained by a generalization of the three
deviation form, wherein we have the magnons labeled by the quasi-momenta 𝑘1 , 𝑘2 and
𝑘3 with the phase shift between each interacting magnon 𝜃𝑖𝑗 (𝑖 ≠ 𝑗), combining to give
the phase factors;
𝛩1 = 𝜃12 + 𝜃23 + 𝜃13
𝛩2 = 𝜃23 + 𝜃21 + 𝜃31
𝛩3 = 𝜃31 + 𝜃32 + 𝜃12
where 𝜃𝑖𝑗 and 𝑘𝑖 satisfy;
𝜃𝑖𝑗 𝑘𝑖 𝑘𝑗
2 cot = cot − cot
2 2 2
𝑁𝑘1 = 2𝜋𝜆1 + 𝜃12 + 𝜃13
𝑁𝑘2 = 2𝜋𝜆2 + 𝜃21 + 𝜃23
𝑁𝑘3 = 2𝜋𝜆3 + 𝜃31 + 𝜃32
Generalizing the results of three deviation states to the arbitrary deviation states, we have
a total of 𝑀 = 𝑟! permutations of possible states each permutation labeled as 𝑃1 , … , 𝑃𝑀 .
As before we expand the eigenstates into the natural basis;
𝑁
|𝜓⟩ = ∑ 𝑓𝑗1 ,𝑗2 ,..,𝑗𝑟 |𝑗1 , 𝑗2 , … , 𝑗𝑟 ⟩
1<𝑗1<⋯<𝑗𝑟 <𝑁
Where
𝑀
𝑝 𝑝 𝑝
𝑓𝑗1 ,𝑗2,..,𝑗𝑟 = ∑ 𝐴𝑝 (𝑒 𝑖𝑘1 𝑙1 +𝑖𝑘2 𝑙2+⋯+𝑖𝑘𝑟 𝑙𝑟 )
𝑝=1
Where the summation is over the 𝑀 permutations. Now for the phase terms we define 𝐴𝑝
as;
𝑟 𝑟
𝑝
𝐴𝑝 = 𝐶 exp [∑ ∑ 𝜃𝑖𝑗 ]
𝑖=1 𝑗=𝑖+1
𝑝
Where 𝜃𝑖𝑗 = ±𝜃𝑖𝑗 (the –ve terms occur in the permutations where the order of 𝑘𝑖 and 𝑘𝑗
is reversed.) Therefore we end up with the energy equation;
𝑟

𝜀 = 𝐽 ∑(1 − cos 𝑘𝑖 )
𝑖=1
With the equations relating the wave-vectors and phase shifts as

8
𝜃𝑖𝑗 𝑘𝑖 𝑘𝑗
2 cot = cot − cot
2 2 2
And

𝑁𝑘𝑖 = 2𝜋𝜆𝑖 + ∑ 𝜃𝑖𝑗


𝑗≠𝑖

These are the set of equations that form what we know as the Bethe Ansatz, published in his set
of papers “On the Theory of Metals ”. Bethe showed that these equations are correct for the spin
- ½ Heisenberg chain however they do not apply for the 𝑆 > 1/2 cases. The fact that the Bethe
Ansatz works for the spin - ½ case categorizes this model as an example of what we know as an
integrable system, subsequently the 𝑆 > 1/2 systems are therefore not integrable.

9
Chapter 2 – The DMRG Algorithm

In order to approach such the spin chains discussed so far to obtain measurable quantities or
observables we first and foremost need to be able to understand how the Hamiltonian is
constructed diagonalized. In the beginning of this report we mentioned that for a spin chain with
𝑁 spins arranged in a 1-dimensional lattice, the Hilbert space associated with the system would
be 2𝑁 dimensional. Consequently the Hamiltonian describing the said system must also be 2𝑁 ×
2𝑁 dimensional. For the Heisenberg chain we already know that the Hamiltonian is defined as;
𝑁

𝐻 = −𝐽 ∑ 𝑺𝑖 ∙ 𝑺𝑖+1
𝑖=1
2𝑁
In light of the dimensional Hilbert space the spin operator, operating on the 𝑖th spin would be
defined as the tensor product;
1 𝑖 𝑖+1
𝑺𝑖 = [𝟙12×2 ⊗ 𝟙22×2 ⊗ ⋯ ⋯ 𝟙𝑖−1 𝑁−1
2×2 ⊗ 𝝈2×2 ⊗ 𝟙2×2 ⊗ ⋯ ⋯ 𝟙2×2 ⊗ 𝟙2×2 ]
𝑁
2
𝑗
where 𝟙2×2 (𝑗 = 1, 𝑁) is the 2 × 2 identity matrix labeled by the spin index in the superscript,
1 1
and 𝝈𝑖2×2 the Pauli spin operator operating on the 𝑖th spinor where for 𝑠 = ; ↑ → ( ) and ↓ →
2 0
0
( ), yielding a 2𝑁 × 2𝑁 dimensional matrix operator which when combined with the respective
1
𝑺𝑖+1 operator and summed over all the spins forms a 2𝑁 × 2𝑁 dimensional Hamiltonian.

While the Bethe ansatz does supplement us with the eigenkets |𝜓⟩ of the above mentioned
Hamiltonian, using the said eigenfunctions to obtain expectation values of the required
observables (for example 𝑄̂ ) serves as an enormous task in on itself. Right from the beginning
we are served with constructing the matrix for obtaining the said quantity, which would involve
explicitly evaluating 22𝑁 matrix elements and then using the matrix thus obtained to find the
expectation value of 𝑄̂ defined as;
〈𝑄̂ 〉 = ⟨𝜓|𝑄̂|𝜓⟩
which would further demand calculations in the multiples of 22𝑁 . Therefore, it becomes much
more convenient to approach such quantum many body problems, numerically instead of
analytically. Such numerical problems often demand high computational resources which is
usually tackled using highly optimized algorithms and softwares and one of the most sought after
algorithms among them is Density Matrix Renormalization Group algorithm.

The Density Matrix approach or the DMRG algorithm was introduced as an enhancement to a
preexisting renormalization group procedure used by Wilson, by correcting the errors popping up
in the algorithm, given that Wilson didn’t consider the interactions between successive

10
collections of spins (or more concretely defined as “blocks” later in this chapter), making
quantitatively accurate results impossible for most problems.

2.1 The Outline of the Algorithm

The Density Matrix approach or the DMRG algorithm was introduced as an enhancement
by Steven White in 1992 (White, 1992) to a preexisting renormalization group procedure
proposed by Keneth G. Wilson in 1975 (Wilson, 1975), where the initial procedure had
errors popping up given that the algorithm didn’t account for the interactions between
successive collections of spins (or more concretely defined as “blocks” later in this
chapter), making quantitatively accurate results impossible for most problems.

The initial structure of the algorithm proposed by Wilson is described as follows - for a
given system of 𝑁 spins on a 1𝐷 lattice we divide the entire chain into a set of identical
blocks “𝐴” as;

We now start from a single block 𝐴, and evaluate its corresponding Hamiltonian 𝐻𝐴 and
other such operators relevant to our study. Once we have obtained the Hamiltonian for a
single block we now try to form a larger block 𝐴𝐴, made of two neighboring blocks
considered together, with interaction terms kept in mind.

One diagonalizes the Hamiltonian of this new block 𝐻𝐴𝐴 and uses the 𝑚 lowest lying
eigenstates (out of a total of 𝑙 eigenstates) obtained through the diagonilization to form a
new simpler Hamiltonian 𝐻𝐴′ defined as;
[𝐻𝐴′ ]𝑚×𝑚 = [𝑂]𝑚×𝑙 [𝐻𝐴𝐴 ]𝑙×𝑙 [𝑂† ]𝑙×𝑚
where 𝑂 is defined as the matix whose rows are the lowest lying eigenstates of 𝐻𝐴𝐴 . Once
we have successfully combined the first two blocks in this way, we repeat this procedure
over and over again for the whole chain.

11
The main assumptions here being that any contribution from the interaction of
neighboring blocks is negligible, and the states with the most statistical weight in the
formation of the bigger blocks are the lowest 𝑚 states.

To counter this issue the DMRG approach, instead of diagonalizing two blocks at a time
diagonalizes an entire “super-block” composed of three or more blocks at a time. Here
instead of truncating the combined Hamiltonian using the lowest 𝑚 states, we instead
truncate using the low lying states of the superblock, projected onto the combined block
𝐴𝐴.

In effect let’s assume that we have diagonalized the superblock Hamiltonian and
extracted the ground state |Ψ⟩ . Now let |𝑖⟩ , 𝑖 = 1, … , 𝑛 be a complete set of states
required to describe the system 𝐴𝐴 and let |𝑗 ⟩, 𝑖 = 1, … , 𝐽 be the states of the rest of the
superblock. We can then express the ground state |𝜓⟩ as;
|Ψ⟩ = ∑ 𝜓𝑖𝑗 |𝑖⟩|𝑗 ⟩
𝑖,𝑗
In effect to describe |Ψ⟩ completely we would need all 𝑛 existing states of |𝑖 ⟩. However,
during each iteration we only keep 𝑚 of these states (𝑛 > 𝑚), rendering an exact
representation of |Ψ⟩ implausible. Let there be an orthogonal set of states |𝑢 𝛼 ⟩, 𝛼 =
1, … , 𝑚, approximating |Ψ⟩ as;
|Ψ⟩ ≈ |𝜓⟩ = ∑ 𝑎𝑗,𝛼 |𝑢 𝛼 ⟩|𝑗 ⟩
𝑗,𝛼
The accuracy of the expansion defined above would depend upon the magnitude of the
quantity;
2

2
𝑆 = ||Ψ⟩ − |𝜓⟩| = ||Ψ⟩ − ∑ 𝑎𝑗,𝛼 |𝑢 𝛼 ⟩|𝑗 ⟩|
𝑗,𝛼

To minimize our error in this representation we would have to demand for 𝜖 to be


minimized with respect to both 𝑎𝑗,𝛼 and |𝑢 𝛼 ⟩. Given the density matrix for the combined
block 𝐴𝐴 is defined as;
𝜌𝑖𝑖′ = ∑ 𝜓𝑖𝑗 𝜓𝑖′ 𝑗
𝑗
We find that the |𝑢 𝛼 ⟩ are
the eigenfunctions of 𝜌 with the greatest eigenvalues 𝜔𝛼 , when
𝜌 is defined in terms of |𝑢 𝛼 ⟩ = ∑𝑖 𝑢𝑖𝛼 |𝑖⟩ as;
𝜌 = ∑ 𝜔𝑖 |𝑢 𝑖 ⟩⟨𝑢 𝑖 |
𝑖
where 𝜔𝑖 can be interpreted as the probability of the system 𝐴𝐴 to be in the state |𝑢 𝑖 ⟩,
and 𝜔𝛼 as mentioned above are the 𝑚 greatest eigenvalues of 𝜌 from among all 𝜔𝑖 s.

12
From this, it is clear that best states to keep for truncating the Hamiltonian 𝐻𝐴𝐴 in each
iteration are the eigenstates corresponding to the 𝜔𝛼 eigenvalues of the density matrix 𝜌
for the newly combined block in every iteration.

Now that we have decided which states to keep and which to reject in each iteration, we
try to resolve the next issue in the original algorithm. Consider for example two
neighboring blocks on the Heisenberg chain 𝐴 and 𝐵. The combined Hamiltonian for the
new block 𝐴𝐵 when the interaction between the blocks is taken into account would like;
1
𝐻𝐴𝐵 = 𝐻𝐴 ⊗ 𝟙𝐵 + 𝟙𝐴 ⊗ 𝐻𝐵 + [𝑆𝐴𝑧 ⊗ 𝑆𝐵𝑧 + (𝑆𝐴+ ⊗ 𝑆𝐵− + 𝑆𝐴− ⊗ 𝑆𝐵+)]
2
which would contain matrix elements resulting from the composite block 𝐴𝐵. In the next
section we go on to show how this is carried out in practice for the anti-ferromagnetic
Heisenberg chain.

2.2 DMRG Applied to a 4-spin Superblock

For the 4 spin case construct an initial superblock consisting of four spin sites as follows;

labeled as sites 1, 2, 3 and 4.


To proceed with the DMRG algorithm, we first need to find the superblock Hamiltonian
by combining relevant operators. If we initially think of each spin in the 4-spin
superblock as an individual blocks 𝐴, 𝐵, 𝐶 and D the composite Hamiltonian for 𝐻𝐴𝐵
would be expressed as;
1
𝐻𝐴𝐵 = 𝑆𝐴𝑧 ⊗ 𝑆𝐵𝑧 + (𝑆𝐴+ ⊗ 𝑆𝐵− + 𝑆𝐴− ⊗ 𝑆𝐵+)
2
𝑧,+,−
where the 𝑆 operators are defined as;
1
0 0 1 0 0
𝑆𝑧 = (2 1); 𝑆+ = ( ); 𝑆− = ( )
0 − 0 0 1 0
2
Explicitly 𝐻𝐴𝐵 then becomes;

13
1
0 0 0
4
1 1
0 − 0
𝐻𝐴𝐵 = 4 2
1 1
0 − 0
2 4
1
(0 0 0
4)
𝐻𝐶𝐷 would similarly of the form;
1
0 0 0
4
1 1
0 0 −
𝐻𝐶𝐷 = 2 4
1 1
0 − 0
4 2
1
(0 0 0
4)
Now we must construct the superblock Hamiltonian taking the interaction terms into
account. Thus 𝐻𝐴𝐵𝐶𝐷 is defined as;
𝑧 𝑧
1 + − − + )
𝐻𝐴𝐵𝐶𝐷 = 𝐻𝐴𝐵 ⊗ 𝟙𝐶𝐷 + 𝟙𝐴𝐵 ⊗ 𝐻𝐶𝐷 + [𝑆𝐴𝐵 ⊗ 𝑆𝐶𝐷 + (𝑆𝐴𝐵 ⊗ 𝑆𝐶𝐷 + 𝑆𝐴𝐵 ⊗ 𝑆𝐶𝐷 ]
2
which explicitly came out to be;
3
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
4
1 1
0 0 0 0 0 0 0 0 0 0 0 0 0 0
4 2
1 1 1
0 − 0 0 0 0 0 0 0 0 0 0 0 0
2 4 2
1 1
0 0 0 0 0 0 0 0 0 0 0 0 0 0
4 2
1 1 1
0 0 0 − 0 0 0 0 0 0 0 0 0 0
2 4 2
1 3 1 1
0 0 0 0 − 0 0 0 0 0 0 0 0
2 4 2 2
1 1 1
0 0 0 0 0 − 0 0 0 0 0 0 0 0
2 4 2
1 1
0 0 0 0 0 0 0 0 0 0 0 0 0 0
𝐻𝐴𝐵𝐶𝐷 = 4 2
1 1
0 0 0 0 0 0 0 0 0 0 0 0 0 0
2 4
1 1 1 1
0 0 0 0 0 0 0 0 − 0 0 0 0
2 4 2 2
1 1 3
0 0 0 0 0 0 0 0 − 0 0 0 0 0
2 2 4
1 1 1
0 0 0 0 0 0 0 0 0 0 − 0 0 0
2 4 2
1 1 1
0 0 0 0 0 0 0 0 0 0 0 0 0
2 2 4
1 1 1
0 0 0 0 0 0 0 0 0 0 0 0 − 0
2 4 2
1 1
0 0 0 0 0 0 0 0 0 0 0 0 0 0
2 4
3
(0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
4)

14
Using this 16 × 16 superblock Hamiltonian matrix, we obtain the ground state |Ψ⟩ ;
|Ψ⟩ = (0 0 0 0.1493 0 0.5579 −0.4081 0 0 −0.4081 −0.5579 0 0.1493 0 0 0)
with the corresponding energy, 𝐸0 ≈ 1.606
Which we will use to obtain the density matrix for the composite block 𝐴𝐵;
0.0223 0 0 0
0 0.4777 −0.4553 0
𝜌≈( )
0 0.0223 0.4777 0
0 0 0 0.0223
In this example we keep 𝑚 = 3 states so we diagonalize density matrix 𝜌 and keep the
eigenstates with the greatest eigenvalues 𝜔 𝛼 = 0.9330, 0.0224, 0.0223 where the last
eigenvalue is discarded (𝜔 3 ≈ 𝜔 4 ), with the corresponding eigenstates;
0 −0.5499 0.8076
| 𝑢1 ⟩ ≈ (−0.7071 0.3940 0.1164
) |𝑢 2 ⟩ ≈ ( ) |𝑢 3 ⟩ ≈ ( )
0.7071 0.3940 0.1164
0 −0.6222 −0.5663
𝛼
The next step would be to form the truncation matrix 𝑂 with the 𝑢 ⟩ forming the rows of
|
the matrix;
0 −0.7071 0.7071 0
𝑂 ≈ (−0.5499 0.3940 0.3940 −0.6222 )
0.8076 0.1164 0.1164 −0.5663
The truncated Hamiltonian 𝐻𝐴′ would then take the form;
−0.7500 0 0

𝐻𝐴′ = 𝑂𝐻𝐴𝐵 𝑂 ≈ ( 0 0.2500 0 )
0 0 0.2500
To summarize the initial configuration of the system before the initiation of the DMRG
algorithm could be represented as;

whereas after the first iteration the blocks 𝐴 and 𝐵 are combined and truncated to form
the composite block 𝐴′ and the superblock is extended by one more block; in this case
since the periodic boundary comes into play right at the edge of the superblock the new
block included in the superblock would be a “reflection” of 𝐴′ ≡ 𝐴′𝑅 .

And so on so forth for the rest of the chain.


15
2.3 Measurements Using the DMRG Algorithm

As previously stated, in order to analyze a particular spin chain, we must be able to


quantify the expectation values of different operators, or observables. The modification
made to the DMRG method to account for the measurement of these expectation values is
described below.

Assume once more that we have derived and diagonalized the superblock Hamiltonian
for the 4-spin situation. By employing this, we have managed to derive the ground state
Ψ𝑖1 𝑖2 𝑖3 𝑖4 , wherein 𝑖1 designates the states of block 𝐴, 𝑖2 designates the states of block 𝐵,
and so on. Currently, we can calculate the expectation values of several operators, which
are expressed in the superblock basis as shown;
〈𝑄 〉 = ⟨Ψ| 𝑄 | Ψ⟩ = ∑ Ψ𝑖∗1 𝑖2 𝑖3 𝑖4 [𝑄]𝑖1 𝑖1′ Ψ𝑖1′ 𝑖2 𝑖3𝑖4
𝑖1 𝑖2 𝑖3 𝑖4 𝑖1′
for block A given an operator Q acting on it. With the exception of the indices being
summed over, the formulae for the expectation values of operators on other blocks
remain similar in form.
Once we have obtained the density matrix 𝜌 for the combined block 𝐴𝐵, since we have
through the properties of density matrices, for any operator 𝑄𝐴𝐵 ;
〈𝑄𝐴𝐵 〉 = 𝑇𝑟(𝜌𝑄𝐴𝐵 )
Next, following the algorithm we diagonalize 𝜌 we keep the eigenstates |𝑢 𝛼 ⟩ with the 𝑚
highest eigenvalues and discard the rest. And use these eigenstates to form the respective
truncation matrix 𝑂. Similar to the truncation of 𝐻𝐴𝐵 to 𝐻𝐴′ , the new truncated operator
matrix is defined as;
𝑄𝐴′ = 𝑂𝑄𝐴𝐵 𝑂 †
The same process is repeated for each iteration obtain the final expectation value.

The error in this process occurs due to the truncation and the deviation from the original
value is;
𝜖𝑚 = 1 − ∑ 𝜔𝛼
𝛼
where 𝛼 runs over all the 𝑚 highest eigenvalues of the density matrix and 𝜔𝛼 are the
corresponding eigenvalues. The truncation error approaches zero as 𝑚 approaches the
total number of available eigenstates of 𝜌.

16
Chapter 3 – Correlation Functions

Among all the classes of solids studied so far crystalline solids make up the most thoroughly
analyzed class of solids. In the ideal scenario, the atoms in these crystals are arranged in strictly
periodic arrays maintaining both isotropy (all directions are equivalent) and homogeneity (the
neighborhood of each and every atom is same) through space, maintained over significant
distances within the crystal grain. Such a crystal is said to posses what we term as long range
order.

In the condensed phase however, a sizable fraction of studied matter does not show long-range
order. In such sample even though the atoms may show significant correlations at microscopic
scales (of the order of a few angstroms), over macroscopic distances the atoms are more or less
uncorrelated. It is easiest to understand why these connections emerge in substances where
covalent bonding dominates. Due to the directionality of these links, the relative orientation of
the first few neighbors, who are spaced out at somewhat regular intervals, is pretty distinct
throughout the entire sample. It is possible for variations from the typical bond directions and
distances to become so significant over longer distances that correlations between atomic
positions disappear. Samples like these posses what we term as short range order.

3.1 Defining Correlation Functions

Given the long range order of crystalline materials, its structure can be entirely described
by a small number of parameters, while on the other hand simply defining the position and
expectation values of internal degrees of freedom for a large number of atoms would result
in an unmanageable and intractable amount of data, even if they were known.
Alternatively, the structure can be fully described in terms of the atomic distribution
functions.

Given the 𝑛-particle probability distribution function 𝜌𝑛 (𝒓1 , 𝒓2 , … , 𝒓𝑛 ), the probability that
the 𝑖th particle lies in a volume 𝑑𝒓𝑖 simultaneously as the 𝑗th particle lies in the volume
𝑑𝒓𝑗 for all 𝑖, 𝑗 = 1,2, … , 𝑛 is given by;
𝑑𝑃𝑛 (𝒓1 , 𝒓2 , … , 𝒓𝑛 ), = 𝜌𝑛(𝒓1 , 𝒓2 , … , 𝒓𝑛 )𝑑𝒓1 𝑑𝒓2 … 𝑑𝒓𝑛

17
For 𝑁 atoms confined in a volume 𝑉 where the position of each atom is denoted by the
vector 𝑹𝑖 , probability density function 𝜌(𝒓) for one particle is defined as;
𝑁

𝜌(𝒓) = ⟨∑ 𝛿(𝒓 − 𝑹𝑖 )⟩
𝑖=1
where ⟨… ⟩ denotes the thermal average. Extending the same definition to probability
density function for a two particles, 𝜌2 (𝒓1 , 𝒓2 ) we get the expression;
𝑁 𝑁

𝜌2 (𝒓1 , 𝒓2 ) = ⟨∑ ∑ 𝛿 (𝒓1 − 𝑹𝑖 ) 𝛿(𝒓2 − 𝑹𝑗 )⟩


𝑖=1 𝑗=1
𝑗≠𝑖
where from definition of probability density we get the usual results;

∫ 𝜌(𝒓)𝑑𝒓 = 𝑁
𝑉
Similarly,

∫ 𝜌2 (𝒓1 , 𝒓2 )𝑑𝒓2 = (𝑁 − 1)𝜌(𝒓1 )


𝑉
And subsequently;

∬ 𝜌2 (𝒓1 , 𝒓2 )𝑑𝒓1 𝑑𝒓2 = 𝑁(𝑁 − 1)


𝑉
As we mentioned above systems possessing long range order the positions of any two
atoms on the lattice is correlated even if the separation between these two atoms is large
whereas for systems with short range order these correlations decay as the separation
between the atoms increases. In other words as |𝒓𝑖 − 𝒓𝑗 | → ∞;
𝜌2 (𝒓1 , 𝒓2 ) → 𝜌(𝒓1 )𝜌(𝒓2 )
Or in simpler terms as the separation between the atoms in a weakly correlated system
increases the combined probability density function tends to achieve more and more
seperable form (becoming more “unmixed”).

Thus, one uses this fact to their advantage and defines the correlation function as (Solyom,
n.d.);
𝐶 (𝒓1 , 𝒓2 ) = 𝜌2 (𝒓1 , 𝒓2 ) − 𝜌(𝒓1 )𝜌(𝒓2 )
In fact, this correlation function shows whether the likelihood of discovering another atom
at 𝑟2 is impacted by the existence of an atom at 𝑟1 . The correlation function is exactly zero
for atomic arrangements that are fully random. For amorphous systems with short-range
order the function takes finite values at small separations and drops off exponentially at
vast distances. However, for crystalline structures the function shows a periodic behavior in

18
conjunction with the periodicity of the crystal structure which is maintained irrespective of
the separation between any two atoms.
3.2 Correlation Functions in Ising Model

The Ising Model remains one of the simplest spin lattice models among the existing
models tackling ferromagnetism in statistical mechanics. The one dimensional case
consists of a series of 𝑁 spins arranged in an ordered 1D lattice with periodic boundary
conditions where the orientation of each spin restricted to be in only two states +1 or −1.
For each pair of adjacent spins, 𝑖, 𝑖 + 1, 𝑖 = 1, … , 𝑁 the interaction is labeled by the
quantity 𝐽 which is basically a measure of the strength of interaction. For a dipole
moment 𝜇 associated with each spin the Hamiltonian for the system under the influence
of a magnetic field 𝐻 is expressed as;
𝑁 𝑁

ℋ = −𝐽 ∑ 𝑆𝑖𝑧 𝑆𝑖+1
𝑧
− 𝑚𝐻 ∑ 𝑆 𝑧
𝑖 𝑖
(here since the spins are restricted to be oriented along +1 or −1 we have chosen for the
spins to be oriented along the 𝑧 − 𝑎𝑥𝑖𝑠).

To obtain the correlation function for such a configuration we choose any site on the
lattice and fix it to be the origin or site zero. Now for any other site 𝑖 on the lattice
consider the product;
𝑆0𝑧 𝑆𝑖𝑧 = ±1
depending on whether the spin at site 𝑖 is parallel or anti-parallel to the spin at the origin.

Averaging over the entire chain, at low temperatures we can readily infer that,
⟨𝑆0𝑧 𝑆𝑖𝑧 ⟩ = ±1
where if ⟨𝑠0 𝑠𝑖 ⟩ = +1, would imply that at low temperatures the spins tend to align
themselves, thus showing ferromagnetic behavior whereas for if ⟨𝑠0 𝑠𝑖 ⟩ = −1 the system
tends to show paramagnetic behavior.
Now when we consider the product of averages instead we find that;
⟨𝑆0𝑧 ⟩⟨𝑆𝑖𝑧 ⟩ = ⟨𝑆0𝑧 ⟩2
since the lattice is homogenous both the sums would be over the same indices.
Based on these we can define the correlation function for the 1D Ising model as;
𝐶𝑖 (𝑇, 𝐻) = ⟨𝑆0𝑧 𝑆𝑖𝑧 ⟩ − ⟨𝑆0𝑧 ⟩2
To emphasize on how correlation functions can be used to evaluate various
thermodynamic quantities associated with a system, we’ll try to evaluate the magnetic
susceptibility of this system using correlation functions.
For a single configuration {𝑠1 , 𝑠2 , … , 𝑠𝑁 } the magnetization is defined as;

19
𝑁

ℳ (𝑠1 , 𝑠2 , … , 𝑠𝑁 ) = 𝑚 ∑ 𝑆𝑖𝑧
𝑖=1
And the macroscopic magnetization is therefore given as;
𝑀 = ⟨ℳ⟩ = 𝑚𝑁⟨𝑆0𝑧 ⟩
where we aim to calculate the magnetic susceptibility of the given system defined as;
𝜕𝑀
𝜒(𝑇, 𝐻) =
𝜕𝐻
We can re-write our Hamiltonian in terms of individual microstates {𝑠1 , 𝑠2 , … , 𝑠𝑁 } as;
𝑁

ℋ == −𝐽 ∑ 𝑆𝑖𝑧 𝑆𝑖+1
𝑧
− 𝐻ℳ
𝑖
Now we evaluate the sum of 𝐶𝑖 (𝑇, 𝐻) over all possible microstates and spins;
𝑁 𝑁

∑ 𝐶𝑖 (𝑇, 𝐻) = ∑[⟨𝑆0𝑧 𝑆𝑖𝑧 ⟩ − ⟨𝑆0𝑧 ⟩2 ]


𝑖=1 𝑖=1
𝑁 𝑁 2
∑𝑠 𝑆0𝑧 𝑆𝑖𝑧 𝑒 −𝛽ℋ ∑𝑠 𝑆0𝑧 𝑒 −𝛽ℋ
∑ 𝐶𝑖 (𝑇, 𝐻) = ∑ [ −( ) ]
𝑍 𝑍
𝑖=1 𝑖=1
where the partition function 𝑍 is defined as;
𝑍= ∑ ∑ … ∑ 𝑒 −𝛽ℋ
𝑠1=±1 𝑠2=±1 𝑠𝑁 =±1
And 𝛽 = 1/𝑘𝐵 𝑇.
On further simplification we find;
𝑁
𝜕 ∑𝑠 𝑆0𝑧 𝑒 −𝛽ℋ
( ) = 𝛽𝑚 ∑ 𝐶𝑖 (𝑇, 𝐻)
𝜕𝐻 𝑍
𝑖=1
However since;
∑𝑠 𝑆0𝑧 𝑒 −𝛽ℋ
⟨𝑆0𝑧 ⟩ =
𝑍
we have,
𝜕 ∑𝑠 𝑆0𝑧 𝑒 −𝛽ℋ 𝜕
( )= ⟨𝑆 𝑧 ⟩
𝜕𝐻 𝑍 𝜕𝐻 0
Or substituting the expression for macroscopic magnetization we have;
𝜕 ∑𝑠 𝑆0𝑧 𝑒 −𝛽ℋ 1 𝜕𝑀 1
( )= = 𝜒(𝑇, 𝐻)
𝜕𝐻 𝑍 𝑚𝑁 𝜕𝐻 𝑚𝑁
Therefore, we obtain the magnetic susceptibility in terms of the correlation function as;
𝑁
𝑚2𝑁
( )
𝜒 𝑇, 𝐻 = ∑ 𝐶𝑖 (𝑇, 𝐻)
𝑘𝐵 𝑇
𝑖=1
Other useful thermodynamic quantities can be obtained through a similar statistical
analysis.

20
3.3 Conclusion and Outlook

In this work we have developed a basic theoretical understanding required to tackle


quantum many body problems in the form of 1D spin chains, specifically coming under
the umbrella of Heisenberg spin-1/2 chains. Through Bethe Ansatz we were able to treat
the given many-body problem analytically by treating the required state as a culmination
of successive single spin deviations from an initial completely aligned state, to a final
spin configuration with arbitrary spin deviations. The resulting eigenstates of the system
had the form of plane waves, termed as magnons with its own discrete set of momentum
analogs, where the interaction between any two such quasi-momenta is reflected in their
respective phase factors as defined in the Bethe Ansatz.

In order to be able to study these spin chains we then shifted to a numerical approach
proposed by Steven White in 1992, viz. the Density Matrix Renormalization Group
algorithm. The algorithm not only boasts highly accurate results, it also highly cost
effective when it comes to computational resources. The DMRG algorithm presented in
this report can even be extended towards 2D or 3D lattices.

Now that we have gained a fair understanding of correlation functions and how
correlation functions apply to spin chains (The Ising Model in this case), we can now
further extend this project to a more practical phase, where we will try to evaluate the
correlation functions of 1D spin chains both time independent and time dependent, to be
able to evaluate various statistical quantities while making active use of the DMRG
algorithm.

21
Bibiliography

Bethe, H. (1947). Zur theorie der metalle. Zeitschrift FÄur Physik, 1(2), 1–33.
https://doi.org/10.1007/BF03161125

Franchini, F. (2017). The Heisenberg chain. Lecture Notes in Physics, 940(1), 47–70.
https://doi.org/10.1007/978-3-319-48487-7_3
Parkinson, J. B. ; F. D. J. J. (2016). An Introduction to Quantum Spin Systems (Vol. 01).

Solyom, J. (n.d.). Fundamentals of the Physics of Solids Volume -1 Structure and Dynamcs. In
Media. Springer.

White, S. R. (1992). Density matrix formulation for quantum renormalization groups. Physical
Review Letters, 69(19), 2863–2866. https://doi.org/10.1103/PhysRevLett.69.2863
Wilson, K. G. (1975). The renormalization group: Critical phenomena and the Kondo problem.
Reviews of Modern Physics, 47(4), 773–840. https://doi.org/10.1103/RevModPhys.47.773

22

You might also like