Download as pdf or txt
Download as pdf or txt
You are on page 1of 64

Dynamics of Interacting Particles in Diamond

Lattices

A Thesis Submitted
in Partial Fulfilment of the Requirements
for the Degree of

MASTER OF SCIENCE

by

Saumya Prakash Sharma


(Roll No. 1911151)

under the supervision of

Dr. Tapan Mishra

to
SCHOOL OF PHYSICAL SCIENCES
NATIONAL INSTITUTE OF SCIENCE EDUCATION
AND RESEARCH
BHUBANESWAR - 752 050, INDIA
May 2024
DECLARATION

I, Saumya Prakash Sharma (Roll No: 1911151), solemnly affirm that the
document titled ”Dynamics of Interacting Particles in Diamond Lattices,”
presented to the National Institute of Science Education and Research Bhubaneswar
as part of the requirements for a Master of Science degree in Physical Sci-
ences, represents my original research work. This work was conducted under the
guidance of Dr. Tapan Mishra and has not been previously submitted for any
academic qualification. I have adhered to academic integrity and honesty through-
out this endeavor. Any external sources or findings referenced in this report have
been appropriately credited and cited.

Bhubaneswar - 752050 Saumya Prakash Sharma

May 2024

ii
CERTIFICATE

This is to certify that the work contained in this project report entitled “Dynamics
of Interacting Particles in Diamond Lattices” submitted by Saumya Prakash
Sharma (Roll No: 1911151) to National Institute of Science Education and Re-
search, Bhubaneswar towards the partial requirement of Master of Science in
Physical Sciences has been carried out by him under my supervision and that it
has not been submitted elsewhere for the award of any degree.

Bhubaneswar - 752 050 Dr. Tapan Mishra

May 2024 Project Supervisor

iii
ACKNOWLEDGEMENT

I express my heartfelt gratitude to my research supervisor, Dr. Tapan Mishra, for


his exceptional mentorship, unwavering support, and invaluable guidance throughout
this project. His profound expertise in the field of quantum simulations and non-
equilibrium many-body systems, along with his constant encouragement, has been
instrumental in shaping my understanding and fostering my growth as a researcher.
The freedom to explore ideas and the warmth and constant encouragement provided
by Dr. Mishra have significantly contributed to the quality and depth of this work.

I am deeply thankful to the members of the Quantum Simulations Group, NISER,


especially Biswajit Paul, Soumya Ranjan Padhi, Rajashri Parida, and San-
chayan Banerjee, for their collaborative spirit and intellectual camaraderie. En-
gaging in insightful discussions and receiving constructive feedback from them has
not only enriched my understanding of the subject but has also fostered a stimulating
research environment, propelling the progress of this work.

Furthermore, I am profoundly grateful to my family for their unwavering support


and encouragement throughout my academic journey. Their love, understanding,
and sacrifices have been the foundation upon which I’ve built my aspirations and
achievements. Their constant belief in my abilities has provided me with the strength
and motivation to overcome challenges and pursue excellence.

I extend my sincere appreciation to Mubashirah, and my friends Sanyam and


Saurabh, for their unwavering friendship, encouragement, and moral support. Their
presence in my life has brought joy, laughter, and solace during both the highs and
lows of my academic endeavors. Their belief in my capabilities and their readiness
to lend an ear or a helping hand have been invaluable throughout this journey.

iv
I would also like to express my appreciation for the challenging aspects of this
project, which have served as valuable learning experiences. Overcoming these chal-
lenges has not only strengthened my problem-solving skills but has also instilled in
me a deeper sense of resilience and determination.

Bhubaneswar - 752050 Saumya Prakash Sharma

May 2024

v
ABSTRACT

The presence of artificial gauge fields in low-dimensional lattices unveils exotic


phenomena dependent on lattice geometry, observed in both interacting and non-
interacting systems. Among such lattice structures, the quasi-1D rhombus lattice
stands out, showcasing phenomena like Aharonov-Bohm Caging and inverse An-
derson localization. Motivated by these discoveries, we investigate the behavior of
such systems under external magnetic flux to understand the interplay among flux,
interactions, and disorder in shaping their electronic structure. We begin our inves-
tigation by studying our model under a non-interacting limit. Our analysis reveals
the emergence of robust edge states appearing on both edges within a specific range
of magnetic flux. We extend the study to interacting limits, we observe the forma-
tion and subsequent breaking of caging due to particle interactions, shedding light
on lattice dynamics. In the interacting case, we observe the restoration of caging in
the limit where the interaction strength matches the rung potential at π flux. This
observation highlights the intricate interplay between flux and interactions in deter-
mining the system’s behavior. By elucidating these phenomena, our work contributes
to a deeper understanding of quantum transport under magnetic flux, offering valu-
able insights into particle behavior within confined geometries. These insights not
only advance our comprehension of fundamental quantum principles but also lay the
groundwork for exploring exotic quantum states and their potential applications in
quantum technologies.

vi
Contents

List of Figures x

1 Introduction 1

1.1 Flat band systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Bose-Hubbard model . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.3 Superconducting circuits . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.4 Flat band diamond lattices . . . . . . . . . . . . . . . . . . . . . . . . 3

1.5 Quantum walks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

1.5.1 Discrete time quantum walk: . . . . . . . . . . . . . . . . . . . 6

1.5.2 Continuous time quantum walk: . . . . . . . . . . . . . . . . . 7

1.6 Localization induced caging in diamond lattices . . . . . . . . . . . . 8

2 Single particle dynamics 10

2.1 Flat bands leading to localization . . . . . . . . . . . . . . . . . . . . 12

2.2 Emergence of edge states . . . . . . . . . . . . . . . . . . . . . . . . . 16

vii
3 Interacting particles: bound pair dynamics 19

3.1 Delocalization due to onsite interaction, U . . . . . . . . . . . . . . . 20

3.2 Delocalization due to rung interaction, VL . . . . . . . . . . . . . . . 25

3.3 Re-emergence of localization: U = VL . . . . . . . . . . . . . . . . . . 27

3.3.1 Dynamics in weakly interacting regime: U = 0 . . . . . . . . . 29

3.3.2 Strongly interacting regime: U = 15 . . . . . . . . . . . . . . . 31

3.3.3 Establishment at U = 5 . . . . . . . . . . . . . . . . . . . . . 34

3.4 Explanation for the restoration of caging . . . . . . . . . . . . . . . . 35

4 Discussions & Conclusion 37

Appendices 42

A Exact Diagnolization 42

A.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42

A.1.1 Brief Overview . . . . . . . . . . . . . . . . . . . . . . . . . . 42

A.2 Exact Diagonalization Process . . . . . . . . . . . . . . . . . . . . . . 43

A.2.1 Overview of the Steps Involved . . . . . . . . . . . . . . . . . 43

A.2.2 Construction of Basis . . . . . . . . . . . . . . . . . . . . . . . 43

A.2.3 Application of Hamiltonian to Chosen Basis . . . . . . . . . . 44

A.2.4 Formation of Matrix Representation of Hamiltonian . . . . . . 44

viii
A.2.5 Solution of Eigenvalue Problem . . . . . . . . . . . . . . . . . 45

A.3 Hamiltonian Formulation . . . . . . . . . . . . . . . . . . . . . . . . . 45

A.3.1 Introduction to the Bose-Hubbard Hamiltonian . . . . . . . . 45

A.4 Memory Requirement in Exact Diagonalization . . . . . . . . . . . . 46

A.5 Reduced Hilbert Space Dimension . . . . . . . . . . . . . . . . . . . . 46

A.5.1 Explanation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

A.6 Quantum Walk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

Bibliography 49

ix
List of Figures

1.1 a) Schematic of a parallel LC-oscillator (also known as a quantum


harmonic oscillator, QHO), consisting of an inductance L in parallel
with capacitance C.(b)The energy potential diagram for the QHO,
where energy levels are evenly spaced approximately ωr apart. (c)
Circuit diagram of a Josephson qubit, featuring the representation of
nonlinear inductance through the Josephson-subcircuit (d) The pres-
ence of Josephson inductance modifies the quadratic energy potential
(dashed red curve) into a sinusoidal shape (solid blue curve), resulting
in non-uniformly spaced energy levels. Adapted from [9]. . . . . . . . 4

1.2 Schematic representation of a flat band diamond lattice. . . . . . . . 5

1.3 Quantum Walk on a one-dimensional node. a) Walk takes place on


discrete nodes. b) Walk on a continuous line, where the particle’s
width does not overlap. c) Walk when widths overlap with each other. 6

x
1.4 Aharonov-Bohm caging in all bands Flat diamond lattices. (a) Band
diagram of the lattice with the compact localized state. (b) We see
all bands becoming flat at π-flux magnetic field. (c) Eigenbasis for a
single plaquette in zero-flux (left) and (right) Eigenenergies and states
calculated using exact diagonalization. Adapted from A. Houck et. al
[11] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.1 A single particle is initialized (in blue)in the bulk of the lattice. . . . 10

2.2 Band structure for a single particle system. We see two dispersive
bands at a non dispersive band at E = 0. All bands are flat at flux
ϕ = π. This indicates the localization of the particle. . . . . . . . . . 12

2.3 Population for first four lattice sites for a single particle. The particle
is initialized in the left most site in the plaquette. . . . . . . . . . . . 13

2.4 Aharonov-Bohm caging of a single particle. The Zero population for


site 3 shows localization of the particle in the other three sites. . . . . 14

2.5 Quantum walk in the bulk of lattice. The particle is initialized in


the bulk (middle) of the lattice. Localization due to flat bands is
observed at Flux ϕ = π. A transition from the ballistic regime to
diffusive regime to finally localized regime is also observed. . . . . . 15

2.6 Quantum walk when particle is initialized at the edge of the particle.
Again, localization is observed at flux ϕ = π and the particle is caged.
Additionally, signatures of localization at ϕ = 0.75π is seen. . . . . . . 16

xi
2.7 a) Energy spectrum of a single particle with varying flux. b) Energy
spectrum at non-π flux case, ϕ = π. The energy spectrum against the
site index shows degenerate edge states visible in the band gaps. c) &
d) Particle density corresponding to the lower (and upper) degenerate
edge modes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

3.1 Caging breaks in the presence of strong onsite interaction (U = 15)


between the particles. There is no rung interaction here. Localization
at ϕ = π/2 is due to the strong interaction which makes the bound
pair to move as an effective single particle . . . . . . . . . . . . . . . 21

3.2 Population dynamics for interaction bounded doublon with ϕ = 0


across all sites. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

3.3 Population dynamics for interaction bounded doublon with ϕ = π


flux. The non-zero population on site 3 shows the bound pair escapes
the caging. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

3.4 Survival Probabilities for r = 2 with U = 40. ϕ is varied from 0 to π . 23

3.5 Survival Probabilities for r = 3 with U = 40. ϕ is varied from 0 to π.


The constant probability of 1 for ϕ = 0.5π shows the localization of
the particles in this region in the lattice. . . . . . . . . . . . . . . . . 24

3.6 Diamond lattice with rung interaction. . . . . . . . . . . . . . . . . . 25

3.7 Caging breaks due to rung interaction, VL = 15. There is no onsite


interaction here and this is effectively single particle dynamics, as the
pair is not bounded by interaction. . . . . . . . . . . . . . . . . . . . 26

xii
3.8 Both particles occupy the middle site in the bulk, in the presence of
onsite and rung interaction. . . . . . . . . . . . . . . . . . . . . . . . 28

3.9 Particles occupy the top and bottom of a rung in the middle of the
lattice. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

3.10 Particles are on the first two sites of the lattices. . . . . . . . . . . . . 28

3.11 Caging restored at ϕ = π for VL = U when U = 0. Both particles


occupy the same site in the middle of the lattice. This is effectively the
single particle case as studied in chapter 1. Dynamics for flux other π
remains unaffected here and shows strong delocalization as before. . . 29

3.12 Restoration of Caging at ϕ = π for VL = U = 0. Particles occupy the


top and bottom sites of the middle rung. . . . . . . . . . . . . . . . . 30

3.13 Restoration of Caging at ϕ = π for VL = U = 0. Particles occupy the


first two sites on the edge at t = 0. In the case of VL = 15. there
is weak localization due to the strong value of interaction, which is
in accordance of the fact that very strong interaction leads to bound
pair formation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

3.14 Particle dynamics for U = VL = 15. The restoration of caging at ϕ = π


is evident, with particles predominantly localized at the middle site.
The presence of additional localization at ϕ = π/2 is attributed to
bound pair hopping induced by strong onsite interaction, as discussed
in Chapter 1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

3.15 Caging restoration at ϕ = π for U = VL = 15. Particles predominantly


occupy the rung sites, demonstrating the cooperative effect of onsite
and rung interactions in confining the particles. . . . . . . . . . . . . 33

xiii
3.16 Caging restoration at ϕ = π for U = VL = 15. Particles predomi-
nantly occupy the first two sites on the lattice edge, demonstrating
the robustness of caging restoration even in edge cases. . . . . . . . . 33

3.17 Variation of VL from 0 to 15 with increments of 5, with ϕ = π and U


fixed at 5. The restoration of caging is observed across the range of VL . 34

3.18 Energy spectrum of the extended model, with varying U values and
VL = 10. When U = 10 localized states emerge in yellow with high
IPR values. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

3.19 Energy spectrum with eigenstate indices, a) with U = 5 and VL = 10.


Dispersion corresponding to U = 5 is observed. b) U = VL = 10.
There is a flat band corresponding to U = 10, indicating localization
in this regime. c) U = 15 and VL = 10. Dispersion corresponding to
U = 15 is observed. . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

xiv
Chapter 1

Introduction

1.1 Flat band systems

Flat band physics refers to the intriguing behavior of quantum particles in lattice
structures where certain energy bands exhibit no dispersion. In these flat bands,
particles can have unique and exotic properties, leading to phenomena such as en-
hanced correlations, localization, and unconventional quantum states.
In classical mechanics, the kinetic energy of a free particle with mass m and mo-
p2
mentum p is expressed as 2m
. When considering a crystalline solid, a concept akin
to momentum, termed crystal momentum, denoted as ℏk, is introduced to account
for momentum transfer within the lattice as a whole. The energy levels near the ex-
tremities of a band are commonly characterized by an effective mass m∗ , leading to
 2 2
the expression E(k) = E0 + ℏ2mk∗ . The entire energy band encompasses a range of
energies known as the bandwidth, ∆. When a band is described as ”flat,” it signifies
that its energy remains unaffected by changes in k, and ∆ = 0. This condition im-

1
plies an infinite effective mass, which, in a semiclassical interpretation, corresponds
to zero velocity; thereby, the electrons are said to be ”localized”, in the real space.
[1].

1.2 Bose-Hubbard model

The Bose-Hubbard Hamiltonian is defined as:

X UX X
Ĥ = −t (â†i âj + H.c.) + n̂i (n̂i − 1) − µ n̂i (1.1)
2 i i
⟨i,j⟩

Here, the terms represent the kinetic energy of particles (−t), on-site interaction
strength (U ), and chemical potential (µ), respectively. The operators â†i and âi create
and annihilate particles at lattice site i, and n̂i is the particle number operator.
P †
The tunneling (or hopping) term, −t ⟨i,j⟩ (âi âj +H.c.), describes the hopping
of particles between neighboring lattice sites. It contributes to the system’s kinetic
energy and is crucial for understanding the mobility of particles in the lattice.

U
P
The on-site interaction term, 2 i n̂i (n̂i −1), captures the repulsive interaction
between particles occupying the same lattice site. This term plays a pivotal role
in modeling strong correlations and is instrumental in the emergence of intriguing
quantum states.
P
The chemical potential term, −µ i n̂i , ensures particle number conservation.
By controlling the total particle number in the system, it influences the equilibrium
state and phase transitions.

2
1.3 Superconducting circuits

Superconducting circuits are a suitable platform to observe strongly interacting


physics. This comes because of the non-uniformly spaced energy levels and the non-
linearity in the circuit. This non-linearity is induced due to the Josephson Junctions,
as shown in Fig. 1.1

Superconducting qubits, often referred to as artificial atoms, are formed using


Josephson junctions and are distinguished by their non-linear energy levels. The
system’s behavior is described by the Hamiltonian:

Ĥtransmon = 4EC (n̂ − ng )2 − EJ cos(ϕ̂) (1.2)

In the above equation, EC represents the charging energy, EJ denotes the Josephson
energy, n̂ is the number operator, and ϕ̂ signifies the phase operator.

1.4 Flat band diamond lattices

Diamond lattices, characterized by a unique arrangement of atoms in a diamond


crystal structure, have garnered significant attention in the realm of condensed mat-
ter physics. Among the intriguing features of diamond lattices is the emergence of
flat bands, which are energy bands characterized by a constant dispersion relation,
resulting in highly degenerate states. These flat bands give rise to exotic phenomena
and novel quantum states of matter, making diamond lattices a subject of intense
theoretical and experimental investigation.

The flat band diamond lattice exhibits a highly symmetrical structure, with each

3
Figure 1.1: a) Schematic of a parallel LC-oscillator (also known as a quantum har-
monic oscillator, QHO), consisting of an inductance L in parallel with capacitance
C.(b)The energy potential diagram for the QHO, where energy levels are evenly
spaced approximately ωr apart. (c) Circuit diagram of a Josephson qubit, featuring
the representation of nonlinear inductance through the Josephson-subcircuit (d) The
presence of Josephson inductance modifies the quadratic energy potential (dashed red
curve) into a sinusoidal shape (solid blue curve), resulting in non-uniformly spaced
energy levels. Adapted from [9].

atom arranged in a tetrahedral configuration, reminiscent of the atomic arrangement


in a diamond crystal lattice. In this lattice, certain electronic energy bands form flat
regions where the dispersion relation becomes independent of momentum, leading to
an infinite density of states at specific energies. This peculiar feature of flat bands
gives rise to a variety of unconventional phenomena, including: Localization and
Correlation Effects, Inverse Anderson Transition [10], Topological Phases, Quantum
Anomalous Hall Effect, Superconductivity and Magnetism.

In summary, the study of flat band diamond lattices offers a fascinating play-
ground for exploring exotic quantum phenomena and realizing novel quantum states
of matter.

4
Figure 1.2: Schematic representation of a flat band diamond lattice.

1.5 Quantum walks

In our context, we explore the dynamics of two interacting particles through the lens
of a quantum walk, examining their behavior in the lattice structure. More is written
in the appendix.

A quantum walk can be described by the unitary evolution operator Û acting on


a quantum state |ψt ⟩ at time t:

|ψt ⟩ = Û t |ψ0 ⟩ (1.3)

Here, Û is typically a product of local unitary operators describing the dynamics


at each step.

Two-Particle State Initialization

Consider a two-particle state |ψ0 ⟩ initialized in the middle of the lattice. For a
one-dimensional lattice, this can be represented as:

1
|ψ0 ⟩ = √ (|0, L/2⟩ + |L/2, 0⟩) (1.4)
2

5
Figure 1.3: Quantum Walk on a one-dimensional node. a) Walk takes place on
discrete nodes. b) Walk on a continuous line, where the particle’s width does not
overlap. c) Walk when widths overlap with each other.

Here, |i, j⟩ denotes the state where the first particle is at site i and the second
particle is at site j.

1.5.1 Discrete time quantum walk:

In a discrete time quantum walk, particles traverse a lattice in distinct steps, under-
going unitary transformations determined by their current state and lattice connec-
tivity. The evolution operator U combines shift (S) and coin (C) operators:

U = S · (C ⊗ I)

6
1.5.2 Continuous time quantum walk:

Continuous time quantum walks involve a continuous evolution, described by the


Hamiltonian:

Ĥ = −D · ∇2 + V (x) · σ̂z

Here, D is the diffusion constant, ∇2 the Laplacian operator, V (x) the potential
energy, and σ̂z the Pauli-Z matrix.

Evolution of quantum walk

The quantum walk evolves according to the unitary operator Û applied at each time
step. The dynamics of the quantum walk are influenced by the local interactions,
tunneling amplitudes, and external parameters. The resulting quantum state |ψt ⟩
captures the evolving probability distribution of the particles on the lattice.

A key feature of quantum walks is the emergence of quantum interference, lead-


ing to nontrivial probability distributions and entanglement between particles. The
entangled nature of the quantum state enables the exploration of complex quantum
phenomena, making quantum walks a valuable tool in quantum information science.

7
1.6 Localization induced caging in diamond lat-
tices

The study investigates the phenomenon of localization-induced caging in rhombus


lattices, focusing on the numerical treatment of the system starting from the initial
wave function |ψ⟩ = a† |0⟩. This initial state represents a single excitation, created
by applying the creation operator a† to the vacuum state |0⟩, which corresponds to
the absence of particles.

Figure 1.4: Aharonov-Bohm caging in all bands Flat diamond lattices. (a) Band di-
agram of the lattice with the compact localized state. (b) We see all bands becoming
flat at π-flux magnetic field. (c) Eigenbasis for a single plaquette in zero-flux (left)
and (right) Eigenenergies and states calculated using exact diagonalization. Adapted
from A. Houck et. al [11]

8
The Hamiltonian describing the system accounts for the kinetic energy of particles
hopping between lattice sites, the on-site interaction energy, and the potential energy
associated with rhombus lattice geometry. It takes the form:

X † UX
H = −t (ai aj + h.c.) + ni (ni − 1) (1.5)
2 i
⟨i,j⟩

Here, t represents the hopping amplitude, U denotes the on-site interaction


strength, Vi,i+1 accounts for the potential energy between neighboring lattice sites,
and ϵi represents the on-site potential energy.

The system’s dynamics are characterized by the interplay between hopping kinet-
ics and particle-particle interactions. In rhombus lattices, the peculiar geometry can
induce localization effects, confining particles within certain regions of the lattice.
This phenomenon, known as Aharonov-Bohm caging, arises due to the interference
of multiple scattering paths, leading to constructive interference at specific lattice
sites and destructive interference elsewhere. [8]

The introduction of the gauge field Φ introduces phase factors to the hopping
terms, leading to the Aharonov-Bohm caging effect. This effect results in the con-
finement of particles within certain regions of the lattice due to interference effects
arising from the gauge field.

9
Chapter 2

Single particle dynamics

The tight-binding Hamiltonian for the model for a single particle is given by:

X 
H = −t a†i ai+1 + H.c. (2.1)
i

where, a†i and ai denote the creation and annihilation operators, respectively,
for bosons located at site i. The parameter t signifies the hopping strength between
adjacent sites, U represents the on-site interaction energy, and µ denotes the chemical
potential. The operator ni = a†i ai determines the number of bosons present at site i.

Figure 2.1: A single particle is initialized (in blue)in the bulk of the lattice.

10
For the single particle case, there is no onsite interaction and soU = 0 . To derive
the k-space Hamiltonian, we perform a Fourier transformation. Let a†k and ak be the
creation and annihilation operators in momentum space. The Fourier transform of
the boson operators is given by:

1 X ik·ri 1 X −ik·ri †
ai = √ e ak , a†i = √ e ak (2.2)
N k N k

Substituting these expressions into the Hamiltonian and performing the sums
over i, we obtain the k-space Hamiltonian Hk :

X  X X
Hk = −t eik·rij a†k ak + H.c. + U ni (ni − 1) − µ ni (2.3)
⟨i,j⟩ i i

Here, rij is the distance between sites i and j and N is the total number of lattice
sites.

The k space Hamiltonian in matrix form is given by:

 
−ik −iϕπ −ik
0 −t(1 + e ) −t(1 + e e )
 
H(k, ϕ) =  −t(1 + eik )
 
0 0 
 
−t(1 + eiϕπ eik ) 0 0

11
2.1 Flat bands leading to localization

Figure 2.2: Band structure for a single particle system. We see two dispersive bands
at a non dispersive band at E = 0. All bands are flat at flux ϕ = π. This indicates
the localization of the particle.

In the above figure 2.1, we can see that all bands become flat at ϕ = π. This is
indicative of the localization that is predicted via Aharonov-Bohm caging.
So we will go ahead and plot the quantum walk of a single particle initialized in two
different ways: i) In the middle of the lattice. ii) At the edge of the lattice.

If |ψ0 ⟩ is a single particle state initialized in the middle of the lattice, we can
represent it using the creation operators. As the particles is initially at site i, then
|ψ0 ⟩ can be written as:

|ψ0 ⟩ = b†i |0⟩ (2.4)

where b†i and b†i are the creation operators for particles at sites i, and |0⟩ is the
vacuum state.

The time evolution of this state under a time-dependent Hamiltonian H(t) would
then be given by:

12
|ψt ⟩ = e−iHt |ψ0 ⟩ (2.5)

In the presence of the Aharonov-Bohm (AB) effect, a single particle becomes


localized within the initial three sites of the lattice. This localization emerges from
the interference of the particle’s probability amplitudes, resulting in the formation
of constructive and destructive interference patterns.

Destructive interference confines the quantum walk within an Aharonov-Bohm


cage, while constructive interference enables a fully delocalized walk across all four
sites of the plaquette with zero flux. Notably, we observe a constrained quantum walk
within the π-flux rhombus, demonstrating the influence of Aharonov-Bohm caging.

Figure 2.3: Population for first four lattice sites for a single particle. The particle is
initialized in the left most site in the plaquette.

13
Figure 2.4: Aharonov-Bohm caging of a single particle. The Zero population for site
3 shows localization of the particle in the other three sites.

We now plot the density evolution of the particles with varying magnetic flux
and interaction values.

The population density (ρi ) at site i is given by the expectation value of the
number operator ni = b†i bi . For our quantum state |ψt ⟩, the population density is
calculated as:

ρi = ⟨ψt |ni |ψt ⟩ (2.6)

14
Figure 2.5: Quantum walk in the bulk of lattice. The particle is initialized in the
bulk (middle) of the lattice. Localization due to flat bands is observed at Flux ϕ = π.
A transition from the ballistic regime to diffusive regime to finally localized regime
is also observed.

In figure 2.3, we see that there is a sharp localization at ϕ = π, which is due


to the Aharonov-Bohm caging. Similarly, in the fig. 2.4, we can see the same
phenomena when the particle is initialized at the edge. This indicates the single
particle localization is robust, irrespective of the initialization on the edge or bulk.

15
Figure 2.6: Quantum walk when particle is initialized at the edge of the particle.
Again, localization is observed at flux ϕ = π and the particle is caged. Additionally,
signatures of localization at ϕ = 0.75π is seen.

2.2 Emergence of edge states

The indication of weak localization on the edge at ϕ = 0.75π in fig. 2.4 compels us
to further investigate and study the density at flux other than π. Therefore, we plot
the energy spectrum with varying flux from 0 to π.

To analyze the density dynamics, we utilized the inverse participation ratio (IPR),
which quantifies the spatial localization of quantum states. The IPR is defined as:

L
X
IPR = |ψi |4 (2.7)
i=1

where ψi represents the probability amplitude of the quantum state at site i. A


larger value of IPR indicates more localized states, while a smaller value signifies
delocalization.

From Fig. 2.7, we observe an edge mode emerging from the energy spectrum.

16
Further investigation at ϕ = 0.75 confirms the localization at the extreme edges.
It is essential to note that this localization should preserve chiral symmetry, as the
localization is not symmetric for any given site index.

17
Figure 2.7: a) Energy spectrum of a single particle with varying flux.
b) Energy spectrum at non-π flux case, ϕ = π. The energy spectrum against the site
index shows degenerate edge states visible in the band gaps.
c) & d) Particle density corresponding to the lower (and upper) degenerate edge
modes.

18
Chapter 3

Interacting particles: bound pair


dynamics

It has been shown that the caging at π flux breaks down in the presence of onsite
interaction between two particles [11]. Recently, it has also been shown that this
caging can be broken if instead of onsite interaction, there is a long range nearest
neighbour interaction [12]. We will see the results for different kinds of interactions
in the sections below.

To analyze the delocalization phenomenon, we employ exact diagnolization tech-


nique to study the eigenstates and eigenenergies of the Hamiltonian. Through a
comprehensive analysis of the interacting case, we aim to elucidate the underlying
physics of bound pair dynamics on Aharonov-Bohm caging under interaction effects,
providing valuable insights for future experimental investigations and theoretical
studies in quantum many-body systems.

19
3.1 Delocalization due to onsite interaction, U

The system now has an additional interaction term and onsite potential and the
hamiltonian can be written as:

X  UX
Ĥ = −t eiϕ b̂†i b̂j + H.c. + n̂i (n̂i − 1) (3.1)
2 i
⟨i,j⟩

where the terms represent:

• The kinetic energy associated with boson hopping between nearest-neighbor


sites.

• The on-site interaction energy term.

• The on-site potential energy term.

Here, the operators and parameters used in this Hamiltonian are defined as fol-
lows:

b̂†i creates a boson at site i,

b̂j annihilates a boson at site j,

n̂i is the number operator at site i,

t represents the hopping amplitude between nearest-neighbor sites,

U denotes the on-site interaction energy,

20
Figure 3.1: Caging breaks in the presence of strong onsite interaction (U = 15)
between the particles. There is no rung interaction here. Localization at ϕ = π/2
is due to the strong interaction which makes the bound pair to move as an effective
single particle

U
PL
In the presence of interaction characterized by the term 2 i=1 n̂i (n̂i − 1), the
dynamics of the bound pairs are altered. The interplay between the kinetic energy
associated with hopping, the potential energy due to interaction, leads to the delo-
calization of bound pairs.

21
Figure 3.2: Population dynamics for interaction bounded doublon with ϕ = 0 across
all sites.

Figure 3.3: Population dynamics for interaction bounded doublon with ϕ = π flux.
The non-zero population on site 3 shows the bound pair escapes the caging.

22
Survival probability

At flux π2 , we see there is linear spreading and hence localization of the states. To
quantify this localization we compute the survival probability which is defined as

r
1 X
SPr (t) = ⟨ψ(t) |ni | ψ(t)⟩ (3.2)
N i=−r

Where, N is the total number of particles, |ψ(t)⟩ is the time evolved state at
time t, and n̂i is the number operator. The quantity SPr (t) when calculated within
a range of sites r around the initial position of the particles in the dynamics, tends
to one if there is localization in the lattice.

Figure 3.4: Survival Probabilities for r = 2 with U = 40. ϕ is varied from 0 to π

23
Figure 3.5: Survival Probabilities for r = 3 with U = 40. ϕ is varied from 0 to π.
The constant probability of 1 for ϕ = 0.5π shows the localization of the particles in
this region in the lattice.

The disruption of caging due to strong onsite interaction (U = 15) signifies the
dominance of particle-particle interactions over the kinetic energy associated with
hopping. This results in the bound pairs behaving as quasi-particles with effectively
enhanced mass, leading to their confinement within specific regions of the lattice.
The emergence of such localized behavior has implications for various physical phe-
nomena, including correlated states, quantum transport, and the formation of exotic
quantum phases.

24
3.2 Delocalization due to rung interaction, VL

Here, we will add a different kind of nearest neighbor interaction. We turn on a


potential in the diagonally opposite points (we call it a rung or a leg) of the lattice
and create an interaction between the top and bottom sites of the rung.

Figure 3.6: Diamond lattice with rung interaction.

The system is an extended Bose-Hubbard model and the Hamiltonian is given


by:

X  X
Ĥ = −t eiϕ b̂†i b̂j + H.c. + VL n̂i n̂j (3.3)
⟨i,j⟩ ⟨i,j⟩rung

where the terms represent:

• The kinetic energy associated with boson hopping between nearest-neighbor


sites.

• The potential energy term for rung hopping, denoted by VL .

• The on-site potential energy term.

25
Figure 3.7: Caging breaks due to rung interaction, VL = 15. There is no onsite
interaction here and this is effectively single particle dynamics, as the pair is not
bounded by interaction.

PL−1
In the presence of the rung potential i=1 Vl n̂i n̂i+1 , the dynamics of the bound
pairs are altered. The interplay between the kinetic energy associated with hopping,
and the confinement induced by the rung potential leads to the delocalization of
bound pairs.

The observed delocalization due to strong onsite and rung interaction highlights
the importance of controlling and manipulating interaction strengths in experimental
setups. By tuning the parameters governing particle-particle interactions, we can
explore and engineer novel quantum states with tailored properties, offering potential
applications in quantum information processing and quantum technologies.

26
3.3 Re-emergence of localization: U = VL

In the previous sections, we observed the breakdown of caging under the influence of
strong onsite and rung interaction, leading to the delocalization of bound pairs. In
other words, any kind of interaction present in the system will break the Aharonov-
Bohm caging.
Now, we explore the intriguing phenomenon of caging restoration when the onsite
interaction strength U is equal to the rung interaction strength VL .This investiga-
tion sheds light on the delicate balance between interaction energies and hopping
dynamics in determining the spatial localization of particles.

The extended Bose-Hubbard model hamiltonian governing the system can be


written as below. This is our model:

X  UX X
Ĥ = −t eiϕ b̂†i b̂j + H.c. + n̂i (n̂i − 1) + VL n̂i n̂j (3.4)
2 i
⟨i,j⟩ ⟨i,j⟩rung

Here, U represents the onsite interaction energy, and VL denotes the potential
energy associated with rung interactions. When U = VL = 15, the interplay between
these interaction terms and the kinetic energy of hopping leads to the restoration of
caging, as depicted in figures below.

To establish the robustness of this study, we will take the following initial states:

1. When both particles are at the center of the lattice, at the same site.

2. When a particle each are on the top and bottom of a rung in the middle of
lattice

27
3. When are particles are on the first two sites of the edge of the lattice.

Figure 3.8: Both particles occupy the middle site in the bulk, in the presence of
onsite and rung interaction.

Figure 3.9: Particles occupy the top and bottom of a rung in the middle of the
lattice.

Figure 3.10: Particles are on the first two sites of the lattices.

Now we plot the density evolution for the above three initial sites for U = 0 and
U = 15, while keeping VL = 0, 15 for all the cases. We take U = 0 and U = 15
to study the dynamics in the weakly interacting and strongly interacting regimes
respectively. Whereas, the rung potential, VL is kept 0 & 15.

28
3.3.1 Dynamics in weakly interacting regime: U = 0

We have extended the above lattice to 25 sites and initialize two particles on the
same site in the middle of the lattice.
All the below cases are for weakly interacting particles, i.e., U = 0 and show the
restoration of the caging when U = VL for all the above three initialization cases.

Figure 3.11: Caging restored at ϕ = π for VL = U when U = 0. Both particles occupy


the same site in the middle of the lattice. This is effectively the single particle case as
studied in chapter 1. Dynamics for flux other π remains unaffected here and shows
strong delocalization as before.

29
Figure 3.12: Restoration of Caging at ϕ = π for VL = U = 0. Particles occupy the
top and bottom sites of the middle rung.

The first two cases shows the density evolution in the bulk of the lattice. The
third plot shows the edge dynamics.

Figure 3.13: Restoration of Caging at ϕ = π for VL = U = 0. Particles occupy the


first two sites on the edge at t = 0. In the case of VL = 15. there is weak localization
due to the strong value of interaction, which is in accordance of the fact that very
strong interaction leads to bound pair formation.

All the above three plots show that the Aharonov-Bohm caging is restored when

30
VL = U in the weakly onsite interacting regime i.e, U = 0. The restoration of
caging at U = VL for U = 0 underscores the delicate balance between kinetic and
potential energy contributions in the system. When the onsite and rung interaction
strengths are matched, they cooperatively confine the particles within specific regions
of the lattice, effectively restoring the localized behavior observed in the absence of
interactions.

3.3.2 Strongly interacting regime: U = 15

In this section, we demonstrate the remarkable phenomenon of caging restoration


even in the presence of strong onsite interaction (U = 15). Despite the pronounced
influence of interaction energies, the delicate balance between onsite and rung inter-
actions enables the system to exhibit localized behavior under specific conditions.

We present results for U = VL = 15, where both the onsite and rung interaction
strengths are matched. In all cases considered, we observe the restoration of caging,
indicating a cooperative effect between interaction energies and hopping dynamics.

31
Figure 3.14: Particle dynamics for U = VL = 15. The restoration of caging at ϕ = π
is evident, with particles predominantly localized at the middle site. The presence
of additional localization at ϕ = π/2 is attributed to bound pair hopping induced by
strong onsite interaction, as discussed in Chapter 1.

Figure 3.14 illustrates the density evolution for U = VL = 15. Despite the strong
onsite interaction, the system exhibits caging restoration at ϕ = π, with particles
primarily localized at the middle site. The persistence of additional localization at
ϕ = π/2 highlights the intricate interplay between interaction-induced bound pair
dynamics and the restoration of caging.

32
Figure 3.15: Caging restoration at ϕ = π for U = VL = 15. Particles predominantly
occupy the rung sites, demonstrating the cooperative effect of onsite and rung inter-
actions in confining the particles.

Figure 3.15 further elucidates the restoration of caging for U = VL = 15, with
particles primarily localized at the rung sites. This observation underscores the
cooperative nature of interaction energies in maintaining spatial confinement within
the lattice structure.

Figure 3.16: Caging restoration at ϕ = π for U = VL = 15. Particles predominantly


occupy the first two sites on the lattice edge, demonstrating the robustness of caging
restoration even in edge cases.

33
Figure 3.16 showcases the restoration of caging for U = VL = 15, with particles lo-
calized primarily at the first two sites on the lattice edge. This observation highlights
the robustness of caging restoration under varying initialization conditions, further
emphasizing the cooperative influence of interaction energies in confining particle
dynamics within the lattice.

These results collectively illustrate the resilience of caging restoration phenom-


ena in the strongly interacting regime (U = 15), underscoring the intricate interplay
between interaction energies and hopping dynamics in determining the spatial local-
ization of particles.

3.3.3 Establishment at U = 5

In this section, we explore the establishment of caging under moderately strong


onsite interaction (U = 5). By maintaining a fixed value for U and varying the rung
interaction strength (VL ), we investigate the caging restoration for ϕ = π as soon as
U matches VL .

Figure 3.17: Variation of VL from 0 to 15 with increments of 5, with ϕ = π and U


fixed at 5. The restoration of caging is observed across the range of VL .

34
Figure 3.17 illustrates the variation of VL from 0 to 15 in increments of 5, with
only π flux cases, and U fixed at 5. The caging is restored as soon the U value
matches VL .

3.4 Explanation for the restoration of caging

Figure 3.18: Energy spectrum of the extended model, with varying U values and
VL = 10. When U = 10 localized states emerge in yellow with high IPR values.

35
Figure 3.19: Energy spectrum with eigenstate indices, a) with U = 5 and VL = 10.
Dispersion corresponding to U = 5 is observed. b) U = VL = 10. There is a flat
band corresponding to U = 10, indicating localization in this regime. c) U = 15 and
VL = 10. Dispersion corresponding to U = 15 is observed.

From the above figure, we can see that there is a localization signature at U = 10.
So we probe the interplay of rung potential VL and onsite interaction U, to study the
dynamics.

In the figure 3.14, we see the flat bands corresponding to U = V = 10, while
there are dispersion when for other values of U. This is the reason for studying the
dynamics when U = V . The localization reappears as an interplay of both U and V,
in the limiting case of U = V for all the values of U.

36
Chapter 4

Discussions & Conclusion

Discussion

We conducted a comprehensive study investigating the localization phenomena in


a rhombus lattice, particularly focusing on the interplay between interaction, mag-
netic flux, and lattice geometry. Initially, we successfully reproduced the flat band
localization in the lattice and observed the delocalization of interaction-induced pair
particles, indicating the intricate nature of lattice dynamics under the influence of
external parameters.

Expanding our analysis, we extended the lattice system to 25 sites to study


single-particle dynamics. Our observations revealed localization in the bulk at π flux,
consistent with experimental expectations [8], along with the emergence of robust
edge modes at the lattice periphery. This study provides insights into the role of
lattice geometry in shaping quantum dynamics and the formation of edge states.

37
Subsequently, we delved into the interacting case, considering two particles bound
by attractive interaction. In this scenario, we observed delocalization phenomena
induced separately by the onsite interaction (U ) and the rung potential (VL ). Specif-
ically, for a strong onsite interaction (U = 15), our findings showed formation of
bound pairs at ϕ = π/2 and, apart from delocalization at ϕ = π [11]. Furthermore,
the introduction of the rung potential (VL ) demonstrated similar delocalization ef-
fects, highlighting the intricate interplay between interaction energies and lattice
geometry.

Interestingly, we observed the restoration of caging in the limiting case of U = VL ,


a non-trivial result against the backdrop of delocalization observed at other arbitrary
values of U and VL . This finding underscores the subtle balance between interac-
tion energies and lattice parameters, leading to unexpected localization phenomena
ranging from weak to strongly interacting regimes.

Overall, our study unveils a rich tapestry of dynamics in the rhombus lattice,
showcasing a delicate interplay between interaction, magnetic flux, and lattice ge-
ometry. The observed localization phenomena, free of disorder and solely governed
by lattice geometry, synthetic flux and interactions, offer profound insights into the
underlying mechanisms controlling quantum dynamics in extended flat band lattice
systems.

Conclusion

Based on our analysis, we draw the following conclusions:

1. Flat band localization: Through simulations, we have successfully studied

38
flat band localization within a rhombus lattice, highlighting its significance in
quantum systems.

2. Delocalization dynamics: We observed instances of delocalization in interaction-


induced pair particles, shedding light on the nuanced dynamics within the lat-
tice.

3. Single-Particle Dynamics: Startuing our study with single-particle dynam-


ics, we noted localization within the lattice bulk at π flux, alongside robust
edge modes along the lattice periphery.

4. Interplay of interactions: Our investigation into interacting particles re-


vealed delocalization tendencies influenced by both on-site interaction (U ) and
rung potential (VL ), underscoring the interplay between various parameters.

5. Restoration of caging: Notably, we observed the restoration of caging at


U = VL , offering insights into unique system behaviors. This restoration of
caging holds promising implications for potential avenues for harnessing and
controlling quantum dynamics in real-world systems.

6. Experimental feasibility: Our work demonstrates the feasibility of experi-


mental implementation, as interactions akin to those studied can be realized in
optical lattice experiments and superconducting circuits. This paves the way
for experimental validation and further exploration of the phenomena observed
in our simulations.

7. Disorder-Free Localization: Crucially, the observed localization phenomena


are disorder-free, originating solely from lattice geometry and magnetic flux,
highlighting the pivotal role of geometry in dictating quantum dynamics.

39
In summary, our study elucidates the rich dynamics within the diamond lattices,
offering valuable insights into the interplay between various parameters and their
effects on quantum phenomena.

40
Appendices

41
Appendix A

Exact Diagnolization

A.1 Introduction

A.1.1 Brief Overview

Exact diagonalization is a powerful numerical technique used to solve the quantum


mechanical eigenvalue problem for finite-sized systems. It involves directly com-
puting the eigenvalues and eigenvectors of the Hamiltonian matrix representing the
system’s dynamics.

Importance of Understanding Quantum Systems Through Nu-


merical Techniques

Quantum systems often exhibit complex behavior that cannot be easily understood
analytically. Numerical techniques like exact diagonalization play a crucial role in

42
studying these systems, providing insights into their properties, dynamics, and phase
transitions. Understanding quantum systems is essential for various fields, including
condensed matter physics, quantum chemistry, and quantum information science.

A.2 Exact Diagonalization Process

A.2.1 Overview of the Steps Involved

Exact diagonalization is a numerical method used to solve the quantum eigenvalue


problem for finite-sized systems. The process involves the following steps:

1. Choose a suitable basis that spans the Hilbert space of the system.

2. Apply the Hamiltonian operator to the chosen basis to obtain the matrix rep-
resentation of the Hamiltonian.

3. Diagonalize the Hamiltonian matrix to find its eigenvalues and eigenvectors.

4. Use the eigenvalues and eigenvectors to compute physical observables and an-
alyze the system’s properties.

A.2.2 Construction of Basis

The basis states for N particles in a one-dimensional lattice of length L can be


constructed using tensor products of single-particle basis states. Let |ni ⟩ denote the
basis state for the ith lattice site, where ni represents the occupation number of

43
particles at site i. Then, the basis states for the entire lattice can be written as
tensor products of single-particle basis states:

|n1 , n2 , . . . , nL ⟩ = |n1 ⟩ ⊗ |n2 ⟩ ⊗ . . . ⊗ |nL ⟩

This construction allows us to represent the entire system’s Hilbert space in terms
of tensor product states, simplifying the representation and computation of many-
particle states.

A.2.3 Application of Hamiltonian to Chosen Basis

Let’s denote the chosen basis states as |n1 , n2 , . . . , nN ⟩, where ni represents the occu-
pation number of bosons at lattice site i. Applying the Bose-Hubbard Hamiltonian
ĤBH to this basis, we obtain a matrix representation H with elements:

Hmn = ⟨m|ĤBH |n⟩

where m and n are indices representing basis states.

A.2.4 Formation of Matrix Representation of Hamiltonian

The Hamiltonian matrix H obtained from the previous step represents the system’s
dynamics in the chosen basis. This matrix is typically sparse due to the local nature
of the Bose-Hubbard Hamiltonian. In the case of finite-sized systems, H is of finite
dimension D × D, where D is the total number of basis states.

44
A.2.5 Solution of Eigenvalue Problem

Diagonalizing the Hamiltonian matrix H yields its eigenvalues Ei and correspond-


ing eigenvectors |ψi ⟩. This can be done numerically using standard linear algebra
techniques such as the Lanczos algorithm or sparse matrix diagonalization methods.
The eigenvalues represent the energy eigenstates of the system, while the eigenvectors
provide information about the system’s wavefunctions.

A.3 Hamiltonian Formulation

A.3.1 Introduction to the Bose-Hubbard Hamiltonian

The Bose-Hubbard Hamiltonian ĤBH describes interacting bosons on a lattice and


is defined as:

X UX X
ĤBH = −t (a†i aj + H.c.) + n̂i (n̂i − 1) − µ n̂i
2 i i
⟨i,j⟩

where t is the hopping parameter, U is the on-site interaction energy, µ is the chemical
potential, a†i and ai are the creation and annihilation operators at site i, and n̂i = a†i ai
is the number operator.

45
A.4 Memory Requirement in Exact Diagonaliza-
tion

The memory requirement for exact diagonalization scales with the dimensionality
of the Hamiltonian matrix, which depends on the size of the Hilbert space. For a
system of N particles in a one-dimensional lattice of length L, the dimension of the
Hilbert space is D = dL , where d is the dimension of the single-particle Hilbert space
(typically equal to the number of lattice sites). Therefore, the memory requirement
scales exponentially with system size, making exact diagonalization computationally
expensive for large systems.

For a one-dimensional lattice with L sites, the dimension d of the single-particle


Hilbert space is equal to the number of lattice sites d = L. Hence, the total dimension
D of the Hilbert space is given by:

D = dN = LN

A.5 Reduced Hilbert Space Dimension

A.5.1 Explanation

The dimension of the Hilbert space for N particles in a one-dimensional lattice


of length L can be reduced by exploiting symmetries and conservation laws. For
example, if the total number of particles N and the total momentum P are conserved,
the Hilbert space can be decomposed into smaller subspaces with fixed N and P .

46
This reduction in dimensionality significantly reduces the memory and computational
requirements for exact diagonalization.

To calculate the reduced dimension of the Hilbert space, we need to consider the
conservation laws. For instance, if the total number of particles N and the total
momentum P are conserved, the Hilbert space can be decomposed into subspaces
with fixed N and P . Let’s denote the dimension of the reduced Hilbert space as
Dred .
X
Dred = D(N, P )
N,P

where D(N, P ) is the dimension of the Hilbert space with N particles and total
momentum P .

A.6 Quantum Walk

A quantum walk on a lattice can be described by a unitary operator U that evolves


the state of a quantum walker. In one dimension, the quantum walk operator U can
be constructed using the single-step evolution operator U1 and iteratively applied to
the initial state of the walker. The evolution of the quantum walk can be analyzed
using techniques from linear algebra and quantum mechanics, providing insights into
its spreading behavior, coherence properties, and potential applications in quantum
information processing.

The single-step evolution operator U1 for a quantum walk on a one-dimensional


lattice can be expressed as a tensor product of shift and coin operators. Let S be
the shift operator and C be the coin operator. Then, U1 = S ⊗ C. By iteratively

47
applying U1 to the initial state of the walker, we can derive the evolution of the
quantum walk over multiple steps.

48
Bibliography

[1] URL: http://nanoscale.blogspot.com/2022/05/


flat-bands-why-you-might-care-and-one.html.

[2] K. Andrews and B. Rajiv. On some applications of eigenvalues of toeplitz


matrices. Journal of Mathematical Analysis and Applications, 56(2):237–239,
2007.

[3] C. C. Chang. Algebraic analysis of many valued logics. Transactions of American


Mathematical Society, 88:467–490, 1958.

[4] Abderrahim Elmoataz, Matthieu Toutain, and Daniel Tenbrinck. On the p-


laplacian and ∞ -laplacian on graphs with applications in image and data
processing. SIAM Journal on Imaging Sciences, 8:2412–2451, 10 2015. doi:
10.1137/15M1022793.

[5] B. Gerla. Automata over MV-algebras. In ISMVL ’04: Proceedings of the 34th
International Symposium on Multiple-Valued Logic, pages 49–54, Washington,
DC, USA, 2004. IEEE Computer Society.

49
[6] Mrinal Kanti Giri. Quantum dynamics of low dimensional interacting systems
using the continuous time quantum walk. 2022. URL: http://hdl.handle.
net/10603/448951.

[7] G.H. Golub and C.F. Van Loan. Matrix Computations. Second Edition. The
John Hopkins University Press, 1989.

[8] Rémy Mosseri Julien Vidal and Benoit Douçot. Aharonov-bohm cages
in two-dimensional structures. 1998. URL: https://journals.aps.org/
prl/abstract/10.1103/PhysRevLett.81.5888, doi:10.1103/PhysRevLett.
81.5888.

[9] Philip Krantz, Morten Kjaergaard, Fei Yan, Terry P Orlando, Simon Gustavs-
son, and William D Oliver. A quantum engineer’s guide to superconducting
qubits. Applied Physics Reviews, 6(2), 2019.

[10] Hang Li, Zhaoli Dong, Stefano Longhi, Qian Liang, Dizhou Xie, and Bo Yan.
Aharonov-bohm caging and inverse anderson transition in ultracold atoms.
Physical Review Letters, 129(22):220403, 2022.

[11] Jeronimo G.C. Martinez, Christie S. Chiu, Basil M. Smitham, and Andrew A.
Houck. Flat-band localization and interaction-induced delocalization of pho-
tons. 2023. URL: https://arxiv.org/abs/2303.02170, doi:10.48550/
arXiv.2303.02170.

[12] Ivan Velkovsky Tomoki Ozawa Hannah Price Jacob P. Covey Bryce Gadway
Tao Chen, Chenxi Huang. Interaction-driven breakdown of aharonov–bohm
caging in flat-band rydberg lattices. 2024. URL: https://arxiv.org/abs/
2404.00737, doi:10.48550/arXiv.2404.00737.

50

You might also like