copyright by minjung kim 2014

113
Copyright by Minjung Kim 2014

Upload: others

Post on 18-Nov-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

Copyright

by

Minjung Kim

2014

The Dissertation Committee for Minjung Kimcertifies that this is the approved version of the following dissertation:

Ab initio simulation methods for the electronic and

structural properties of materials applied to molecules,

clusters, nanocrystals, and liquids.

Committee:

James R. Chelikowsky, Supervisor

Alexander A. Demkov

John G. Ekerdt

Gyeong S. Hwang

Brian A. Korgel

Ab initio simulation methods for the electronic and

structural properties of materials applied to molecules,

clusters, nanocrystals, and liquids.

by

Minjung Kim, B.S.

DISSERTATION

Presented to the Faculty of the Graduate School of

The University of Texas at Austin

in Partial Fulfillment

of the Requirements

for the Degree of

DOCTOR OF PHILOSOPHY

THE UNIVERSITY OF TEXAS AT AUSTIN

May 2014

To my parents,

Kim Dong-Ju and Kim Kun-Hea.

Acknowledgments

It is a great pleasure to acknowledge the ones who have shared my time throughout

my PhD journey.

First of all, I would like to express my profound gratitude to my advisor, Dr. Jim

Chelikowsky. He has provided tremendous support and given me freedom to explore

a wide range of interesting problems. He has never made me feel as if there were a

barrier between him and me. His kindness and valuable advice will not be forgotten.

I would also like to thank our research group members: Grady Schofield, Ben Garrett,

Alex Lee, Charles Lena, Scotty Bobbitt, and Jaime Souto. I must also acknowledge

two previous postdoctoral researchers, Dr. Khoong Hong Khoo and Dr. Noa Marom.

Part of my work could not have been completed without their help.

There are many names that I would like to acknowledge outside of the research group,

but I will limit myself to just a few:

To Greg and Mary Jane Grooms at the Hill House for their love and prayers.

To my best friends, Jisun Kim, So Youn Kim, Hee Jeong Oh, Szu-Hua Chen, Shruthi

Viswanath, Myoung Ji Jang, and Rachel Breeding.

Very special thanks to Katelyn Bobbitt for her close friendship and encouragement.

I will never forget the time spent with her (and her husband as well).

To my parents and brother, and in-laws. Their love and faith have made this thesis

v

possible.

Lastly, I would like to thank Hyunwook Kwak, for his tremendous support and en-

couragement.

vi

Ab initio simulation methods for the electronic and

structural properties of materials applied to molecules,

clusters, nanocrystals, and liquids.

Publication No.

Minjung Kim, Ph.D.

The University of Texas at Austin, 2014

Supervisor: James R. Chelikowsky

Computational approaches play an important role in today’s materials science

owing to the remarkable advances in modern supercomputing architecture and algo-

rithms. Ab initio simulations solely based on a quantum description of matter are

now very able to tackle materials problems in which the system contains up to a

few thousands atoms. This dissertation aims to address the modern electronic struc-

ture calculation methods applied to a range of various materials such as liquid and

amorphous phase materials, nanostructures, and small organic molecules. Our simu-

lations were performed within the density functional theory framework, emphasizing

the use of real-space ab initio pseudopotentials. On the first part of our study, we per-

formed liquid and amorphous phase simulations by employing a molecular dynamics

technique accelerated by a Chebyshev-subspace filtering algorithm. We applied this

technique to find l- and a- SiO2 structural properties that were in a good agreement

with experiments. On the second part, we studied nanostructured semiconducting

oxide materials, i.e., SnO2 and TiO2, focusing on the electronic structures and opti-

cal properties. Lastly, we developed an efficient simulation method for non-contact

vii

atomic force microscopy. This fast and simple method was found to be a very powerful

tool for predicting AFM images for many surface and molecular systems.

viii

Table of Contents

Acknowledgments v

Abstract vii

List of Tables xii

List of Figures xiii

Chapter 1. Introduction 1

Chapter 2. Theoretical and Computational Backgrounds 5

2.1 Electronic structure calculations . . . . . . . . . . . . . . . . . . . . . 5

2.1.1 Born-Oppenheimer approximation . . . . . . . . . . . . . . . . 5

2.1.2 Density Functional Theory . . . . . . . . . . . . . . . . . . . . 6

2.1.3 Pseudopotentials . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.2 Computational approach . . . . . . . . . . . . . . . . . . . . . . . . . 9

2.2.1 Real-space method . . . . . . . . . . . . . . . . . . . . . . . . . 9

2.2.2 Chebyshev iteration algorithm . . . . . . . . . . . . . . . . . . 11

Chapter 3. Ab initio molecular dynamics study for disordered system:The case of SiO2 14

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

3.2 Computational Details . . . . . . . . . . . . . . . . . . . . . . . . . . 16

3.3 Born-Oppenheimer molecular dynamics techniques . . . . . . . . . . . 17

3.4 Liquid simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3.5 Amorphous simulations . . . . . . . . . . . . . . . . . . . . . . . . . . 23

3.6 Defect structure analysis . . . . . . . . . . . . . . . . . . . . . . . . . 28

3.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

ix

Chapter 4. Electronic and structural properties of nanocrystals andclusters 36

4.1 SnO2 nanocrystals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

4.1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

4.1.2 Computational details . . . . . . . . . . . . . . . . . . . . . . . 38

4.1.3 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . 40

4.1.3.1 Quantum confinement effect in Sb-doped SnO2 nanocrys-tals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

4.1.3.2 Antimony vs. Fluorine dopant atoms . . . . . . . . . . 44

4.1.3.3 Higher doping concentration . . . . . . . . . . . . . . . 47

4.1.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

4.2 TiO2 clusters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

4.2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48

4.2.2 Global minimum searching methods . . . . . . . . . . . . . . . 49

4.2.2.1 Simulated-annealing technique . . . . . . . . . . . . . . 49

4.2.2.2 First-principles basin-hopping technique . . . . . . . . 50

4.2.3 Computational details . . . . . . . . . . . . . . . . . . . . . . . 53

4.2.4 Structural analysis of the low-energy clusters found in basin-hopping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

4.2.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

Chapter 5. Noncontact atomic force microscopy study 59

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

5.2 Framework for simulating noncontact atomic force microscopy images 62

5.2.1 Forces between the tip and sample . . . . . . . . . . . . . . . . 62

5.2.2 Derivation of expressions for the frequency shift calculations . . 63

5.2.3 An efficient method for force calculations . . . . . . . . . . . . 65

5.3 Two-dimensional structures . . . . . . . . . . . . . . . . . . . . . . . . 67

5.3.1 GaAs(110) surface . . . . . . . . . . . . . . . . . . . . . . . . . 67

5.3.1.1 Computational details . . . . . . . . . . . . . . . . . . 67

5.3.1.2 Results and discussion . . . . . . . . . . . . . . . . . . 68

5.3.2 Graphene and its defect structures . . . . . . . . . . . . . . . . 73

5.3.2.1 Computational details . . . . . . . . . . . . . . . . . . 74

5.3.2.2 Results and discussion . . . . . . . . . . . . . . . . . . 75

5.4 Small molecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

5.4.1 Computational details . . . . . . . . . . . . . . . . . . . . . . . 79

5.4.2 Results and discussion . . . . . . . . . . . . . . . . . . . . . . . 82

5.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

x

Bibliography 87

Vita 98

xi

List of Tables

3.1 Peak positions in partial pair correlation function (See Fig. 3.2 andtext). Units are A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

3.2 Diffusion constants at several temperatures. . . . . . . . . . . . . . . 25

3.3 Average bond lengths and bond angles of a-SiO2. Our work is com-pared to CPMD, empirical potential MD (EPMD) and experiments.Full width at half maximum is indicated in parenthesis. . . . . . . . . 27

4.1 The number of atoms and diameter of the nanocrystal. . . . . . . . . 40

xii

List of Figures

2.1 Schematic of the SCF cycle using the CheFSI algorithm . . . . . . . . 13

3.1 Temperature (upper) and evolution of atomic mean square distancesfrom the original position (lower) during the randomization and theannealing process of the model amorphous silica structure. The blackline depicts the targeted temperature and the dahsed line shows theactual temperature of the simulation box. . . . . . . . . . . . . . . . 19

3.2 Partial pair correlation function of liquid silica at 3,120 K(dashed line)and 3,700 K. The peak positions are tabulated in Table 3.1. . . . . . 21

3.3 Bond angle distribution function for liquid silica. 2 A was chosen forcutoff radius. Red dots are result of 72 atoms CPMD simulation. . . 22

3.4 Concentration of Si and O atom as a function of distance from atomcenter. Our results are compared with the Car-Parrinello MD (CPMD)simulations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

3.5 The circles indicate this work, triangles are CPMD [23], diamonds areclassical CHIK potential [35], and squares are classical BKS potential [35]. 25

3.6 Total static structure factor of amorphous silicon dioxide from a 192atom simulation (line) and from experiment (circles) [38]. . . . . . . . 26

3.7 Bond angle distribution function in amorphous silicon dioxide. . . . . 28

3.8 Partial pair correlation function. For the comparison, several pointsfrom the CPMD simulation results [23] are indicated as red dots. . . . 29

3.9 Relaxed structure total energy. This graph shows the strain energy oftwo-membered rings do not affect to the total energy of the system. . 30

3.10 Clusters used to calculate cohesive energy. (a) corner-sharing cluster.(b) two-membered ring cluster. . . . . . . . . . . . . . . . . . . . . . 31

3.11 One more layered cluster of Fig. 3.10 (a) corner-sharing cluster. (b)two-membered ring cluster. . . . . . . . . . . . . . . . . . . . . . . . . 32

3.12 The calculated vibrational density of states (solid line) and the con-tribution of the four atoms which constitute the two-membered ring(dotted line). 32cm−1 was chosen for the gaussian broadening. Forthe comparison, experimental data (circles) and CPMD simulationdata (dashed line) were taken from Carpenter and Price [51], andPasquarello and Car [52], respectively. . . . . . . . . . . . . . . . . . 33

3.13 Density of states of amorphous structure. The X-ray photoemissionspectrum data are from Ref. [55]. . . . . . . . . . . . . . . . . . . . . 34

4.1 Band structure of bulk SnO2. A direct band gap of 1.02 eV is observed. 39

xiii

4.2 Structure of H-passivated SnO2 nanocrystals. Sizes of the nanocrystalsare: (a) 1.27 nm, (b) 1.69 nm, (c) 1.97 nm, and (d) 2.37 nm. . . . . . 41

4.3 Fundamental gap of the pure SnO2 nanocrystals (black diamonds) andelectron binding energy of the Sb-doped nanocrystals (red diamonds). 42

4.4 Ionization potential of the doped nanocrystal (blue) and electron affin-ity of the pure nanocrystal (red). . . . . . . . . . . . . . . . . . . . . 43

4.5 Formation energy for the antimony dopant atom with respect to thesize of the nanocrystal. . . . . . . . . . . . . . . . . . . . . . . . . . . 45

4.6 Dopant level wave function isosurface plot. Red and purple indicateoxygen and tin, respectively. Blue sphere indicates the surface hydro-gen. Wave function is localized around the dopant antimony atom. . . 45

4.7 Defect wave function isosurface plot. Antimony and fluorine dopantsare indicated in grey and yellow color, respectively. . . . . . . . . . . 46

4.8 Electron binding energy (diamond) and formation energy (square) fordifferent doping concentration. . . . . . . . . . . . . . . . . . . . . . . 48

4.9 Annealing schedule. The initial temperature was set to 3000K andthe temperature was decreased at every 100 steps until when temper-ature reaches to 300K. For (TiO2)4 cluster simulations, we chose thetemperature step of 250K instead of 500K. . . . . . . . . . . . . . . . 50

4.10 Illustration of basin-hopping optimization process. E({Rm}) respre-sents the original potential-energy surface and E ({Rm}) is transformedpotential-energy surface. Adapted from Ref. [85]. . . . . . . . . . . . 51

4.11 n=2-3 isomers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

4.12 n=4 isomers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

4.13 lumo energies for neutral clusters (upper), and homo energies foranion clusters (lower). Courtesy of Noa Marom. . . . . . . . . . . . . 55

4.14 Structural analysis for the cluster size of n=2-6. . . . . . . . . . . . . 56

4.15 Structural analysis for the cluster size of n=7-10. . . . . . . . . . . . 57

5.1 Basic AFM set-up. Adapted from Ref. [95] . . . . . . . . . . . . . . . 60

5.2 Tip motion in nc-AFM. The equilibrium tip-surface position is at q = 0,and d is the closest distance between the tip and the surface. . . . . . 63

5.3 A side view (a) and a top view (b) of the relaxed GaAs(110) surface.Magenta, yellow, and blue indicate Ga, As, and H, respectively. . . . 68

5.4 Simulated AFM images with respect to the tip turning point (d). ∆ isset to be 1 A for (a)-(c), and the values for d are: (a) 3 A, (b) 4 A, and(c) 5 A. The images are overlaid with the surface Ga (magenta) andAs (yellow) atom. Black and white indicate low and high frequencyshift values, respectively, and the gray scale is adjusted independently.(d)-(f): Noncontact AFM images of GaAs(110) from experiment [113].The frequency shift is -137 Hz, -188 Hz, -218 Hz for (d), (e), and (f). 70

xiv

5.5 Comparison of tip-surface forces. (a) A top view of the GaAs(110)surface and the black color indicates top layer atoms. Dashed-line Aand B correspond to graph (b) and (c). Top panels in graph (b) and(c) show our results calculated from Eq. (5.10). Other three panels areprevious ab initio results simulated by Si-cluster tip with Si, Ga, andAs apexes (Ref. [114] and [115]). The tip-surface distances are 3.41 Aand 4.21 A for (b) and (c), respectively. . . . . . . . . . . . . . . . . . 71

5.6 (a) The dangling bonds of the surface As atom. The electron densitywithin 1 eV energy window below the Fermi level is visualized. Blackand light gray represent Ga and As, respectively. (b) The empty dan-gling bonds of the surface Ga atom. (1 eV energy window above theFermi level.) (c) Ga and As signals from AFM experiments. (Adaptedfrom Ref. [113]) Dashed and solid lines indicate X-X′ and Y-Y′, respec-tively. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

5.7 Simulated nc-AFM images for defect-free graphene structure. Tip-turning point was set to 2 A (left) and 3 A (right). Smaller d yieldsbright spots at carbon atom site whereas larger dts yields bright spotsat hollow site. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

5.8 ∆f vs. tip-sample distance plot. Left and right show the results fromthe metal tip (Ir) and the CO-terminated tip, respectively. The metaltip shows the inversion of the image contrast, i.e., carbon is visiblewhen the tip-sample is relatively close. Adapted from Ref. [122]. . . 76

5.9 Frequency shift with respect to the tip-graphene distance obtained byEq. (5.10) and Eq. (5.4) . . . . . . . . . . . . . . . . . . . . . . . . . 77

5.10 nc-AFM simulation results for two graphene defect structures. Thetip-sample turning point is set to 3 A based on our results from theprevious section. Yellow dots indicate the carbon atoms around thedefects. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80

5.11 Electron density plots for the single vacancy (a)-(c), and the Stone-Wales defect (d)-(f). From the left column to the right column, isosur-faces are taken from the graphene surface at 1.5 A, 2.5 A, and 3.5 Adistances from the graphene sheet. . . . . . . . . . . . . . . . . . . . 81

5.12 (a) 8-hydroxyquinoline molecule. (b) AFM experiment from Ref. [127].(c) Electron density contour plot at 3.4 A above from the molecule.(d) Simulated AFM image without the explicit model for the tip. Tipheight is set to 3.4 A. (e) Simulated frequency shift map by using COtip. Tip height is set to 3.4 A. . . . . . . . . . . . . . . . . . . . . . 84

5.13 (a) Dibenzo(cd,n)naphtho(3,2,1,8-para)perylene molecule. (b)-(c) AFMexperiment from Ref. [129]. CO tip provides much higher resolutionfor C-C bond than the Xe tip. (d) Simulated AFM image without theexplicit model for the tip. Tip height is set to 3.4 A. (e) Simulatedfrequency shift map by using the CO tip. Tip height is set to 3.4 A.(f) Electron density contour plot at 3.4 A above from the molecule. . 86

xv

Chapter 1

Introduction

The design and discovery of advanced materials is one of the most important

topics in science and engineering. One can never overemphasize that developing new

materials is crucial to tackling many challenges of our time such as clean energy

solutions. To overcome limitations of conventional devices and to develop novel prop-

erties, controlling electronic and structural properties of materials at the nanoscale

is critical as numerous desirable properties for new materials can be observed at the

1–100 nm length scale.

All physical matter is composed of combinations of 114 elements. The basic

law that governs matter should therefore start from the understanding of behavior of

individual atom that consists of electrons and nuclei, specifically for nanoscale ma-

terials. Since the discovery of electrons by J. J. Thomson in the late 19th-century,

Newtonian physics, which had successfully produced rich descriptions of nature’s phe-

nomena, failed to explain the stability of atoms. Many efforts toward understanding

the behavior of electrons and nuclei introduced a new branch of science, i.e., quantum

mechanics, by many pioneer scientists such as Max Plank, Erwin Schrodinger, Werner

Heisenberg, and Paul Dirac. The emergence of quantum mechanics is considered an

important paradigm shift in the history of science.

Within the quantum mechanical framework, motions of particles are governed

by a simple equation, HΨ = i~∂Ψ∂t , known as a Schrodinger wave equation. It is

1

one of the fundamental equations of quantum mechanics that describes how parti-

cles behave with time at small length scales. If we knew the exact solution of this

equation, we could in principle predict any properties of matter. However, even for

a simple molecule that contains just a couple of atoms, this equation becomes ex-

tremely complicated since the number of unknown variables increases exponentially:

the Schrodinger wave equation is not tractable for most systems of the interest.

A few decades later, density functional theory (DFT), invented by L. H.

Thomas and E. Fermi, shed light on a practical way of solving the Schrodinger equa-

tion by Hohenberg, Kohn, and Sham [1, 2]. Rather than dealing the many-body wave

functions, DFT focuses on the electron density that has only one physical variable, a

position. Following in the footsteps of computational scientists such as J. C. Slater,

who developed the workable computational method utilizing DFT, it has become the

most practical and widely used electronic structure method in computational physics,

chemistry, and material science [3]. Moreover, recent advances in supercomputing ar-

chitectures and algorithms enable us to utilize DFT-based computational material

science to solve materials problems that contain thousands of atom.

My thesis focuses on applying the electronic structure calculation methods to

a broad range of materials and developing advanced simulation techniques within the

DFT framework. The remainder of this dissertation is outlined as follow:

Chapter 2 reviews the fundamental concepts of density functional theory and

practical approaches to solve the Kohn-Sham equation. A few major numerical tech-

niques that have been implemented in our electronic structure code are also summa-

rized.

Chapter 3 reports ab initio molecular dynamics simulations for liquid and

2

amorphous SiO2 systems. In general, ab initio molecular dynamics simulations are

very limited with regard to the system size and simulation time due to the high

computational costs. We implemented a new algorithm that significantly reduces the

time required to obtain a self-consistent field solution of the Kohn-Sham equation.

We used this method to simulate liquid and amorphous SiO2. Detailed structural

properties and defect structure analysis is found in this chapter.

Chapter 4 describes electronic and structural properties of nanostructured ma-

terials. The first part of this chapter consists of electronic structure calculations for

the SnO2 nanocrystals. One of the well-known phenomena of nanocrystals is that

its electronic properties are tunable by varing the nanocrystal size. We address the

quantum confinement effect, and how the electronic structures change with the addi-

tion of a few dopant atoms. In the second part, global minimum structure searching

methods for small TiO2 clusters are presented. For the small clusters, optical proper-

ties strongly depend on its structure, which makes finding the most stable structure

so important. To find energetically stable structures, we employ two different global

minimum searching methods, i.e., simulated-annealing and basin-hopping. Detailed

procedures for both simulation methods are explained in this chapter. By comparing

several experiments, we suggest how to predict the most likely found structure of

clusters in the photoemission experiment.

Chapter 5 illustrates noncontact atomic force microscopy (nc-AFM) simula-

tions for various surface and molecular systems. nc-AFM is one of the most widely

used techniques in nanoscience and engineering because of its high-resolution atomic-

scale images. However, it is not apparent how one can interpret nc-AFM results

as there always exists uncertainty in the AFM experiment. Theoretical studies are

useful to this end. In this chapter, we introduce an efficient simulation method for

3

nc-AFM. We show our method can be applied to various materials system, such as

semiconducting surfaces, graphene, and small molecules.

4

Chapter 2

Theoretical and Computational Backgrounds

2.1 Electronic structure calculations

2.1.1 Born-Oppenheimer approximation

The first step in understanding atomic systems containing electrons and nuclei

is to write a Hamiltonian:

H = −~

2

2me

i

∇2i +

i,I

ZIe2

|ri − RI |+

1

2

i6=j

e2

|ri − rj|−

I

~2

2MI∇2

I +1

2

I 6=J

ZIZJe2

|RI − RJ |,

(2.1)

where ri and RI are the positions of electrons and nuclei of mass MI , and ZI is the

charge of the nuclei. The Born-Oppenheimer approximation is based on the fact that

the masses of nuclei are much larger than those of electrons. If we assume the masses

of nuclei to be infinity, the kinetic energy of the nuclei can be ignored. The positions

of the nuclei are now considered as classical parameters. With this assumption, the

Hamiltonian becomes

H = −~

2

2me

i

∇2i +

i,I

ZIe2

|ri −RI |+

1

2

i6=j

e2

|ri − rj |+

1

2

I 6=J

ZIZJe2

|RI −RJ |

= T + Vion + Vint + EII ,

(2.2)

where T is the kinetic energy operator for the electrons, Vext is the potential acting

on the electrons owing to the nuclei, Vint is the electron-electron interaction, and EII

is the interaction energy of nuclei.

5

2.1.2 Density Functional Theory

Density functional theory is one of the most widely used methods for solv-

ing the electronic structure problem. Even though the complexity of the problem is

reduced by the Born-Oppenheimer approximation, it is not easy to solve the Hamil-

tonian in Eqn. (2.2) because six independent variables are involved with only one

electron; namely, the positions and the momentums in 3-dimensional space. Spin

also increases the degrees of freedom. Consequently, it is not feasible to obtain an

exact solution for systems containing more than a few dozen electrons.

An approach for solving many-body problems was proposed by Hohenberg and

Kohn in 1964 [1]. Key points of this approach are (a) the external potential, Vext, is

uniquely determined by a ground state density n0(r) and (b) the ground state density

is the density which minimizes the total electronic energy, E[n].

In 1965, Kohn and Sham suggested a practical method to find a solution for

many-body systems using density functional theory [2]. They argued that the solution

of the Hamiltonian of an original system that contains correlated electrons can be

interpreted by a solution mapped on to non-interacting system, which is solvable.

Once the density is obtained from the solution of non-interacting system, all of the

interaction terms are integrated within the exchange-correlation functional of the

density. In atomic unit,1 the total energy of the system is written as

EKS = Ts[n] +

drVext(r)n(r) +

dr

dr′n(r)n(r′)

|r − r′|+ Exc[n], (2.3)

where Ts[n] is the kinetic energy of electrons, and n(r) is the density of the non-

1Atomic units, ~ = me = e = 1, are used in the rest of this thesis.

6

interacting system defined by

n(r) =occ∑

i

|ψi(r)|2. (2.4)

The ground state energy of the functional of Eqn. (2.3) can be obtained by

using the variational principle with orthonormalization constraints and conservation

of particles. The Kohn-Sham equation is written:

−1

2∇2ψi + Veffψi = εiψi (2.5)

Veff = Vion(r) +

dr′n(r′)

|r− r′|+δExc

δn(r). (2.6)

Exchange interactions and correlations of electrons are essential to describing

the energy of the system since electrons are fermions whose wave functions must

be antisymmetric. In DFT, these interaction terms are handled as a functional of

the density, Exc[n]. However, the exact form of the exchange-correlation functional

is unknown. To calculate the exchange-correlation energy, several approximations

have been proposed, e.g., local density approximation (LDA), generalized-gradient

approximation (GGA), orbital dependent functionals, and hybrid functionals [4].

LDA is based on the homogeneous electron gas model. Within the uniform

electron gas, the exchange and correlation effects are known to be local. Accordingly,

we may write the total exchange-correlation energy as a simple integration of the

exchange-correlation density ǫxc:

Exc[n] =

drn(r)ǫxc(n(r)). (2.7)

The exchange-correlation density is normally written separately: ǫxc = ǫx + ǫc. The

exchange energy density is derived from the uniform gas and the correlation energy

density has been calculated with Monte Carlo methods by Ceperley and Alder [5].

7

LDA is a simple and general approach, and is known to provide an appropriate

ground-state structure with a reasonable computational cost. We used the LDA

exchange-correlation functional parametrized by Perdew and Zunger [6] in most of

our work.

2.1.3 Pseudopotentials

The idea of pseudopotentials is a very powerful concept in solving the electronic

structure problem. Pseudopotentials treat the core electrons and the valence electrons

separately so that only the valence electrons, which are relevant to the chemical

environment, are included in the calculations. By using pseudopotentials, we avoid

calculating the highly localized and chemically inert core states. The resulting Kohn-

Sham equation has energy and length scales fixed by the valence states alone.

Pseudopotentials can be generated in several different ways. In our simula-

tions, we employed norm-conserving pseudopotentials that generally follow three con-

straints: 1) the eigenvalues are the same with all-electron wave functions, 2) pseudo

wave functions are identical to all-electron wave functions outside of the core region

(φp(r) = ψAE(r) for r > rc where rc defines the size spanned by the nucleus and

core electrons), and 3) pseudo wave functions and their derivatives must be contin-

uous. With these conditions, the pseudo wave functions can be obtained from the

all-electron Kohn-Sham equation for each isolated atom,

(

−1

2∇2 −

Z

r+ VH [r] + Vxc[ρ; r]

)

ψAE,n(r) = EnψAE,n(r), (2.8)

where Z is the ion charge and ρ is the valence density. Within the Troullier-Martins

8

formalism [7], the pseudo wave function inside of the core region is written as

φp(r) = rl exp(p(r)) for r < rc

and p(r) = co +6

n=1

c2nrn.

(2.9)

The coefficients of the polynomial in Eqn. (2.9) are determined by the norm-conserving

constraint and fixing derivatives at rc.

Once the pseudo wave functions are obtained, ionic pseudopotentials can be

generated by inverting the Kohn-Sham equation (2.8),

V pion(r) = En − VH(r) − Vxc(ρ; r) +

∇2φp,n

2φp,n, (2.10)

where φp,n is the pseudo wave function.

2.2 Computational approach

2.2.1 Real-space method

Within pseudopotential theory, the ionic potential in equation (2.6) can be

written:

V pion(r) =

a

V pion,a (2.11)

As Vhart and Vxc potentials are obtained by the density that depends on the wave func-

tion, ψi, Kohn-Sham equation can be considered as a nonlinear eigenvalue problem.

A common practice to solve this equation is finding a self-consistent field.

The most widely used techniques to solve K-S equation is using plane wave

basis. It is especially useful for crystalline matter. The basis for the plane wave can

be written:

ψk(r) =∑

G

α(k,G) exp(i(k + G) · r)), (2.12)

9

where k is a wave vector, G is a reciprocal lattice vector, and α(k,G) are coefficients of

the basis. To calculate the density, the plane wave basis method employs fast Fourier

transforms (FFTs). Generally, FFTs require numerous global communications in a

parallel computing environment, which makes the plane wave method less efficient

for high-performance computing [8].

An alternative to a plane wave basis is to solve the Kohn-Sham equation in

real-space. In this method, the wave function is represented as a single vector whose

components are the values of the wave function at each real-space grid (xi, yj, zk).

Some major advantages of using the real-space method are 1) ease of implementa-

tion in parallel computing, 2) avoiding artificial periodicity for non-periodic systems

such as molecules in contrast to the plane wave method, and 3) flexible boundary

conditions.

The real-space method uses a finite-difference discretization over the domain of

interest. An important numerical technique for successfully calculating the Laplacian

operator is a higher-order, finite-difference method. This method approximates the

second order derivatives of (∂2ψ/∂x2) at (xi, yj, zk) by

∂2ψ(xi, yj, zk)

∂x2=

M∑

n=−M

Cnψ(xi + nh, yj, zk) +O(h2M+2), (2.13)

where h is the grid spacing, M is a positive integer, and Cn is a coefficient given by

Fornberg [9]. Using a uniform grid in each x, y, and z direction, the Kohn-Sham

equation over the gird points can be computed with the following equation [10]:

10

−~

2

2m

[

M∑

n1=−M

Cn1ψn(xi + n1h, yj, zk) +

M∑

n2=−M

Cn2ψn(xi, yj + n2h, zk) +

M∑

n3=−M

Cn3ψn(xi, yj, zk + n3h)

]

+[

Vion(xi, yj, zk) + VH(xi, yj, zk) + Vxc(xi, yj, zk)]

ψn(xi, yj, zk) = Enψn(xi, yj, zk).

(2.14)

If the domain contains n grid points, the size of the Hamiltonian matrix be-

comes n × n. This matrix size can be much larger than that of the plane wave

method. However, the Hamiltonian matrix in real-space is extremely sparse since the

Laplacian operator is a simple stencil and all local potential elements reside on the

diagonal. Consequently, the n × n Hamiltonian matrix does not have to be stored.

In the discrete form, the nonlocal part of the ion core pseudopotential is a sum over

all atoms, a, and quantum number, (l,m), of rank-one updates:

Vion =∑

a

Vloc,a +∑

a,l,m

ca,l,mUa,l,mUTa,l,m (2.15)

where Ua,l,m are sparse vectors which are only nonzero in a localized region around

each atom, and ca,l,m are normalization coefficients [11].

Another advantage of this method is its good scalability. Since the main

bottleneck in high performance computing is the communication operations between

processors, improved scalability can be obtained by reducing global communications.

The real-space method requires few global communications compared to those of the

plane waves since the Hamiltonian matrix is very sparse and localized in real-space.

2.2.2 Chebyshev iteration algorithm

Solving the nonlinear Kohn-Sham equation involves constructing self-consistent

field (SCF). SCF calculations require an explicit matrix diagonalization at each it-

11

eration step, which is the most expensive computational operation. To reduce the

computational load, Zhou and coworkers proposed a Chebyshev-filtered subspace iter-

ation (CheFSI) technique [12]. Within the CheFSI, only one explicit diagonalization

is required in the first SCF step. This diagonalization provides a good initial subspace.

After the first iteration cycle, a new subspace is obtained by mth order of Chebyshev

polynomial {ψi} = Pm(H){ψi}, rather than performing a matrix diagonalization.

The goal of this filtering algorithm is not to find accurate eigenvectors for each itera-

tion cycle since the Hamiltonian is only approximate in the intermediate SCF steps;

rather, it is designed to approximate progressively the desired eigen-subspace of the

final Hamiltonian when self-consistency is reached. The filtering method significantly

reduces computational time compared to the highly efficient eigensolvers [13]. Fig-

ure 2.1 shows the algorithm for the SCF calculations using CheFSI. Technical details

for CheFSI and parallel implementations are provided in the literature [12, 14, 15].

12

Figure 2.1: Schematic of the SCF cycle using the CheFSI algorithm

13

Chapter 3

Ab initio molecular dynamics study for disordered

system: The case of SiO2

3.1 Introduction

Silicon dioxide is a very abundant material on earth’s crust. Forms of silica

exist in many allotrope forms with varying temperature and pressure conditions. Be-

cause of this, silica is considered a fundamental oxide system and an archetype of

tetrahedral structures, which are thought to include amorphous, and liquid phases.

During the last few decades, silica has played a crucial role in development of elec-

tronic devices and technologies. For example, amorphous silica is commonly used in

electronic devices such as MOSFET as a dielectric material. It forms an electronically

passive interface with silicon, and it can be precisely patterned in the construction

of nano-scale devices. Silica also is used in optical fibers as a transparent mate-

rial [16, 17].

Owing to the fundamental importance of silica in earth and materials science,

and its many technological uses, numerous studies of silica structures have been car-

ried out, both theoretical and experimental. In contrast to the well-defined structural

properties of crystalline forms [18], structural details of amorphous and liquid silica

are problematic. For instance, the details of Si-O-Si bond angle distributions of amor-

phous silica obtained from various experiments using x-ray, neutron diffraction, and

NMR analysis are not in general agreement with recent simulations [19].

14

In order to clarify the structural and dynamical properties, many theoretical

models of amorphous (a)- and liquid (l)- SiO2 have been proposed. These models

were generated by different simulation techniques: classical and ab initio molecular

dynamics, Monte-Carlo, and cluster simulations [20, 21, 22]. Among these different

simulations, ab initio molecular dynamics simulations employing periodic boundary

conditions are the most accurate as they reflect the quantum nature of the interatomic

forces. As such, they more accurately represent changes in hybridization and charge

transfer effects as bonds break and reform in dynamical simulations.

A chief drawback of ab initio simulations is that they are computationally

intensive and often limited by computational constraints to relatively small systems

and short simulation times when compared to simulations using interatomic potentials

based on fits to experiment. Previous a-SiO2 simulations have been conducted for

systems less than hundred atoms [23, 24, 25, 26] with a quenching rate of around

1015 K/s. It has been reported that while the short range of interactions are not

sensitively affected by periodic constraints, medium or long range interactions such

as ring statistics are sensitive to the size of the ensemble [27]. Of course, short

cooling rates, which are necessitated by computational constraints, can also change

the structural properties of amorphous silica [28].

Here, we present ab initio molecular dynamics simulations performed by a

filtering algorithm. The self-consistent cycle described in Fig. 2.1 is supposed to be

repeated for each MD step. To accelerate these calculations, we adopted the converged

wave function from the previous MD step as a first guess for the current step. This is

feasible as the geometries of two adjacent MD steps are not changed considerably. In

this way, we were able to reduce the cost of computational time additionally. We have

successfully applied these algorithms to liquid Al and Al1−xSix alloy system with five

15

hundred atoms [13, 29]. Here, we use the same approach for SiO2 systems containing

up to 192 atoms.

We present liquid simulations and consider quenching the liquid ensemble to

model amorphous solid. We compare structural and dynamical properties for the liq-

uid with other classical and ab initio molecular dynamics simulations. For amorphous

silica, we investigate structural properties and compare to previous simulations and

experiments. At the end of this chapter, we investigate the properties of the defect

structure of the amorphous silica.

3.2 Computational Details

All calculations we performed are based on density functional theory com-

bined with real-space pseudopotentials [30]. Convergence is determined by a single

parameter for a cubic grid, i.e., the grid spacing. We use a grid spacing of 0.35 a.u.

(1 a.u. = 0.5291 A) for our simulations, which corresponds to a ∼60 Ry plane wave

cutoff. For structural properties of amorphous silica, we use a finer grid with a spac-

ing of 0.30 a.u. to obtain highly accurate forces. We carry out our simulations in a

cubic supercell containing 192 atoms or 64 molecular units of SiO2. We assume the

experimental density of amorphous silica (2.2g/cm3) as to fix the size of the cell [26].

This constraint yields a cell edge size of 27.0 a.u. We keep the density constant during

entire simulations, which implies a change in pressure as the temperature of the cell

is altered. This is a small effect compared to other uncertainties in our simulation,

e.g., the use of density functional theory.

We employ norm-conserving pseudopotentials with a 3s23p2 atomic configura-

tion for silicon and 2s22p4 for oxygen. The silicon ionic pseudopotential was generated

with a 2.5 a.u. cutoff radius for both the s and p potentials, and the s potential was

16

chosen as the local component. For the oxygen ionic pseudopotential, a cutoff radius

of 1.45 a.u. was applied for both s and p potentials with the p potential taken as the

local component. We use the local density approximation for the exchange-correlation

functional from Ceperley and Alder [5]. Since the periodicity of the supercell has no

physical meaning in our simulation, we do not consider a sampling over different ~k-

points and consider only the ~k = 0 point. Provided the cell size is sufficiently large,

this should be an accurate approximate.

3.3 Born-Oppenheimer molecular dynamics techniques

We generate amorphous structures using simulated annealing [31, 32]. Typ-

ically, simulated annealing employs three steps. First, in order to randomize the

initial coordinates of atoms, the system is heated to a very high temperature, i.e.,

well above the melting point of the silica. Second, the system is “slowly” cooled to a

targeted temperature. Third, data is collected from microcanonical simulations using

Newtonian dynamics. We chose a Langevin equation of motion as a temperature

thermostat. The trajectories of the atomic species are

Mid~vi

dt= −γMi~vi + ~Gi(γ, T ) + ~Fi, (3.1)

where Mi is the atomic mass of the ith species, γ is a viscosity or friction coefficient,

and G is a random force appropriate for a heat bath of temperature T [31, 33]. A

time step of 165 a.u. (4 fs) was applied with the friction coefficient of 0.001 a.u.

Fig. 3.1 illustrates the details of our annealing schedule. We note that our annealing

rate, 2.5 × 1014 K/s, is significantly slower than the previous ab initio simulations

(1015 K/s (Ref. [26]), and 9 × 1014 K/s (Ref. [23]).

Various annealing schedules have been applied to previous simulations [23, 26,

17

28, 34]. As a starting point, both crystalline and random configurations of atoms can

be used. We used β-cristobalite as a starting point and set our initial temperature to

be 5,000 K. β-cristobalite is a high temperature form of crystalline silica that can be

constructed by considering a diamond crystal of silicon and placing oxygen atoms at

the bond sites. By using this form as a starting point, we can avoid unrealistically high

energy configurations that might occur by a random placement of atomic species. We

considered initial temperatures up to 7,000 K to randomize the initial geometry [34].

We found that 5,000 K is sufficient to randomize the crystalline structure and remove

any memory of the original state. Previous Car-Parrinello molecular dynamics sim-

ulations used atomic coordinates generated by empirical potential simulations and

set the initial temperature at 3,500 K [23]. Our experience is that defects existing

in the initial structure, cannot be removed by annealing even at this relatively high

temperature.

The mean square displacement shown in Fig. 3.1 was determined from

< R2α >=

1

i

[Rαi (t) −Rαi (t = 0)]2 (3.2)

where Nα is the number of atom species α in the supercell. The average displacement

during the entire simulation was 4.5 A for Si and 5.1 A for O. This displacement

is significantly larger than the Si-Si and O-O bond lengths from the initial crystal

structure: the bond lengths are 3.09 A and 2.52 A, respectively. Displacements

of this length ensure that the initial structure is sufficiently randomized to remove

correlations with the initial crystal structure.

18

0 5 10 15 200

2000

4000

6000

Tem

pera

ture

(K

)

actual temperaturetargeted temperature

0 5 10 15 20time (ps)

0

10

20

30

mea

n sq

uare

d

ispl

acem

ent (

Å2 )

SiliconOxygen

Figure 3.1: Temperature (upper) and evolution of atomic mean square distances fromthe original position (lower) during the randomization and the annealing process ofthe model amorphous silica structure. The black line depicts the targeted temperatureand the dahsed line shows the actual temperature of the simulation box.

19

3.4 Liquid simulation

Fig. 3.1 shows an annealing schedule for the simulation for an initial temper-

ature of 5,000 K and a final temperature of 300 K. Also shown is the mean square

displacement of the atomic species. Most changes in the mean square displacement

occur in high temperature region where the ensemble is removed from equilibrium.

The parabolic shape for the initial stage of the simulation is expected for ballistic

trajectories as the system has yet to thermalize. The linear regime for the first 5-6 ps

represents a liquid state. From 5 to 10 ps, the steepness of the slope is gradually de-

creased as the system attempts to solidify. After 10 ps, there is no significant changes

in the mean square displacement other than the fluctuation. In order to study the

temperature dependence of structural and dynamical properties, we performed two

liquid simulations at different temperatures.

To prepare the liquid, we extracted two snapshots at 3,000 K and 3,500 K

then ran extra 2 ps for each simulation simultaneously. The average temperatures

of liquid were 3,120 K and 3,700 K, respectively. These temperatures are well above

the experimental melting point of silica (∼2,000 K) and ensures that we are well

within the liquid regime. It is a problematic issue as to when “density functional”

silica will melt, but the nature of the mean square displacement in Fig. 3.1 indicates

solidification should not occur below these temperatures. Previous simulations have

also used this temperature region for liquid simulations [23, 35].

For the liquid state, we determined the pair correlation function and show our

results in Fig. 3.2. There are small changes in peak position and peak height between

the two temperature regimes. The first peak height of Si-O bond length increases by

∼10 %, and the entire first peak of Si-Si had been shifted towards shorter distance by

about 0.1 A as the temperature is decreased. Detailed values of the peak positions

20

Figure 3.2: Partial pair correlation function of liquid silica at 3,120 K(dashed line)and 3,700 K. The peak positions are tabulated in Table 3.1.

are given in Table. 3.1, and corresponding values of amorphous and β-cristobalite are

also given in the same table. In Table. 3.1, ‘peak1’ and ‘peak2’ indicate the position

of the first and the second peak of the partial pair correlation function, respectively.

‘Min’ is the minimum position between ‘peak1’ and ‘peak2’.

Even though only small changes were observed in partial pair correlation func-

tion, Fig. 3.3 shows significant differences in the Si-O-Si bond angle distribution

Table 3.1: Peak positions in partial pair correlation function (See Fig. 3.2 and text).Units are A

Si-O O-O Si-Sipeak1 min peak2 peak1 min peak2 peak1 min peak2

3700K 1.63 2.40 4.11 2.65 3.51 4.95 3.15 3.67 5.153120K 1.63 2.28 3.95 2.61 3.41 4.95 3.05 3.61 5.18

amorphous(300K) 1.63 4.05 2.67 5.09 3.01 5.31β-cristobalite 1.55 3.89 2.53 4.37 3.09 5.05

21

Figure 3.3: Bond angle distribution function for liquid silica. 2 A was chosen forcutoff radius. Red dots are result of 72 atoms CPMD simulation.

function between two simulations. The angle distributions for Si-O-Si and O-Si-O are

shown. The O-Si-O angle represents a tetrahedral angle (109.5◦) for crystalline silicon

whereas the Si-O-Si shows strong variations in the crystalline structure, depending

on the silica polytype. Typically the Si-O-Si bond is ∼140◦ as in quartz. In idealized

β-cristobalite, it is 180◦. The difference in the distribution occurs for the Si-O-Si

when the bond angle is smaller than 100◦. The simulation for the liquid at 3,120 K

shows a pronounced peak around 90◦, but it is not shown in the 3,700 K simulation.

In previous simulations performed for a-SiO2 simulations, the peak between 80-100◦

was regarded as evidence for the existence of two-membered ring [26]. Fig. 3.3 indi-

cates that the existence of a two-membered ring was not excluded during the cooling

process from 3,700 K to 3,120 K.

In order to understand coordination changes in configurations at different tem-

peratures, we examined coordination number as a function of coordination radius in

22

Fig. 3.4. At high temperatures, Si and O atoms are often miscoordinated. For

example, 20 % of Si atoms are coordinated with only three O atoms even at the

2 A coordination radius, and few Si atoms are coordinated with five O atoms at

3,700 K. However, these coordination errors were significantly reduced at 3,120 K.

We also display results from simulations using Car-Parrinello molecular dynamics

(CPMD) [23] for comparison. The temperature for the CPMD simulation as taken

to be 3,500 K. Their coordination number statistics are similar to our simulation

performed at 3,120 K.

Diffusion coefficients are an important measure for quantifying liquid behav-

ior [27]. We employed the Einstein relation [36] to calculate diffusion constant:

Dα = limt→∞

< [Rα (t)]2 >

6t(3.3)

The calculated diffusion constants for Si and O in liquid silica within the temperature

range from 3,000 K to 3,700 K are tabulated in Table. 3.2 and also shown in Fig. 3.5

as are previous results. We also indicated the diffusion coefficient ratio between Si

and O in Table. 3.2. As is expected, the mobility of the lighter oxygen is always

higher than silicon.

3.5 Amorphous simulations

To obtain statistical average for the amorphous structure, we carried out 400

steps of molecular dynamics simulations at 300 K. We compared several structural

properties with experimental data and previously performed simulation results.

The total static structure factor of neutron scattering experiment is available

for a-SiO2 [37]. Since silica is a heterogeneous system, the structure factor can be

calculated by weighted sum of partial structure factors, Sαβ.

23

Figure 3.4: Concentration of Si and O atom as a function of distance from atomcenter. Our results are compared with the Car-Parrinello MD (CPMD) simulations.

24

Table 3.2: Diffusion constants at several temperatures.

temperature DSi(cm2/s) DO(cm2/s) DSi/DO

ab initio MDPARSEC

3120K 6.7×10−6 7.9×10−6 0.853700K 1.0×10−5 1.5×10−5 0.67

CPMD [23] 3500K 5±1×10−6 9±1×10−6 0.56

classical MDBKS

[35] 3000K 9.5×10−7 1.9×10−6 0.503580K 1.8×10−5 2.8×10−5 0.64

CHIK[35] 3000K 4.6×10−6 6.6×10−6 0.72

3580K 6.0×10−5 8.3×10−5 0.72

CPMD: Car-Parrinello molecular dynamicsBKS: Beest-Kramer-Santen potentialCHIK: Carre-Horbach-Ispas-Kob potential

Figure 3.5: The circles indicate this work, triangles are CPMD [23], diamonds areclassical CHIK potential [35], and squares are classical BKS potential [35].

25

0 2 4 6 8q(Å

-1)

0

0.5

1

1.5

2

S(q

)

experimental result

Figure 3.6: Total static structure factor of amorphous silicon dioxide from a 192 atomsimulation (line) and from experiment (circles) [38].

S (q) =

α,β bαbβ (cαcβ)1/2 [Sαβ (q) + 1]∑

α cαb2α

(3.4)

cα,β is the concentration of silicon and oxygen, and bα,β is a scattering length. (bSi=4.149

fm, bO= 5.803 fm). Sαβ is obtained by a Fourier transform of the partial pair corre-

lation function, gαβ.

Sαβ (q) = δαβ + 4πρ (cαcβ)1/2

∫ ∞

0

r2 sin (qr)

qr(gαβ (r) − 1) dr (3.5)

The structure factor only depends on the magnitude of wave vector q owing to the

isotropic character of the amorphous systems. The calculated static structure factor

fits very well with the experimental data (See Fig. 3.6). Our simulations accurately

reproduce the position of the first three peaks in static structure factor. The partial

pair correlation functions, gαβ(r), that were used in Eq. (3.5) are shown in Fig. 3.8.

26

Table 3.3: Average bond lengths and bond angles of a-SiO2. Our work is comparedto CPMD, empirical potential MD (EPMD) and experiments. Full width at halfmaximum is indicated in parenthesis.

This work CPMD [24] EPMD [34] EXP [37, 39, 40, 41]

d(Si-O) 1.63 1.62 1.61 1.610(0.09) (0.08) (0.08) ±0.050

d(Si-Si) 3.01 2.98 3.07 3.080(0.36) (0.25) (0.21) ±0.100

d(O-O) 2.67 2.68 2.76 2.632(0.26) (0.21) (0.25) ±0.089

Si-O-Si 138 136 148 140-150(24) ±14 (27)

O-Si-O 110 109 109 109.4-109.7(10) ±6 (15) (15)

We calculated the average values of the bond angle and the bond length in

Table 3.3. For comparison, Car-Parrinello MD and classical MD simulation results

are also tabulated. Our simulation shows good agreement with the experimental

data. Details of the short-range bond angle distributions, Si-O-Si and O-Si-O, are

shown in Fig. 3.7. Previous CPMD results are indicated by a dashed line in the

same figure. A noticeable difference between two simulations is a peak below 100◦

in the Si-O-Si bond angle distribution function. This peak suggests the existence

of two-membered (2m) ring (an edge-sharing pair) as we noted in discussing our

liquid simulations. The relatively small Si-O-Si angle comes from the geometry of

a quadrangular configuration of the 2m ring as shown in Fig. 3.10. This peak also

contributes to a slightly smaller value for average bond angle of Si-O-Si relative to

the experiments in Table 3.3, since even a single occurance of a 2m ring makes a

considerable change the bond angle distribution function owing to the size of the

supercell. The character of 2m ring structure is also detected in partial pair correlation

function in Fig. 3.8. We note the small peak in gSiSi(r) around 2.4 A, which is not

27

60 90 120 150 180

Si-O

-Si

60 90 120 150 180angle (degrees)

O-S

i-O

Figure 3.7: Bond angle distribution function in amorphous silicon dioxide.

shown in the compared CPMD data. Typically, the distances between Si-Si of 2m

rings is in the range of 2.3–2.5 A, which is comparable value to the small peak in

gSiSi(r) figure.

3.6 Defect structure analysis

Two-membered rings have been observed in some previous simulations [26, 28];

however, these rings are absent in other simulations [23, 34]. To understand the

origin of these differences, we performed a variety of different preparations for our

simulations, i.e., we examined cells containing 8, 32, and 64 unit of SiO2 with cooling

28

Figure 3.8: Partial pair correlation function. For the comparison, several points fromthe CPMD simulation results [23] are indicated as red dots.

29

0 0.2 0.4Relative Energy (eV/atom)

00.20.40.60.8

11.21.41.61.8

22.22.42.62.8

33.2 2-membered ring

without 2-membered ring

Figure 3.9: Relaxed structure total energy. This graph shows the strain energy oftwo-membered rings do not affect to the total energy of the system.

rates ranging from 2.5 × 1014 to 1015K/s. Among them, only the simulation for

the 8 units of SiO2 with a cooling rate of 2.5 × 1014K/s did not result in a 2m ring

configuration. In general, increasing the size of the system allowed the two membered

ring configuration even for the slowest cooling rate.

The same phenomenon was reported by Binder et al. [28] They tested several

cooling rates with 1,002 atoms and the slowest cooling rate was 4.4×1012K/s, which is

two orders of magnitude slower than most ab initio simulations. Their study showed

evidence for a 2m ring even for the slowest cooling rate.

In order to determine the possibility of existence of 2m ring in amorphous

structures, we compared total energy of thirteen different systems prepared as men-

tioned above. We picked one snapshot in each simulation at 300 K and performed

structural relaxation for each cell. Among them, we chose the lowest energy as the

zero reference energy. We indicated total energy per atom instead of total energy

since cells contain different number of atoms. The energy differences are not signifi-

cant between the two simulation groups as illustrated in Fig. 3.9.

However, since the 2m ring population is a small fraction of the entire system,

30

Figure 3.10: Clusters used to calculate cohesive energy. (a) corner-sharing cluster.(b) two-membered ring cluster.

total energy comparisons may not accurately reflect the presence of 2m rings. Rather

than performing total energy comparison of the entire cell, we attempted to exam-

ine the energy of the local structure. Extracting an energy representing a localized

configuration is not a trivial exercise within density functional theory as contrasted

with a classical simulation. We estimated the energy cost for a 2m ring and corner-

sharing (see Fig. 3.10) by considering clusters of bulk amorphous silica. Hydrogen

was used to passivate our model clusters. We considered small clusters as it is shown

in Fig. 3.10 and calculated the cohesive energy of two configurations labeled by (a) for

a corner sharing geometry and (b) for a two membered ring cluster. In (a) clusters,

the average value of cohesive energy was -6.17 eV/atom while for (b) clusters was

-6.43 eV/atom, implying the 2m ring clusters are favorable structures compared to

the corner-sharing clusters. To check the applicability of this approach to a bulk envi-

ronment, we considered clusters with a second-shell of SiO2 (Fig. 3.11). The cohesive

energies of both (a) and (b) cluster in Fig. 3.11 were similar, that is -6.56 eV/atom

for (a) and -6.52 eV/atom for (b). Adding second-shell atoms to the cluster results in

reducing the energy difference. This explains why Fig. 3.9 does not show a difference

in total energy between two groups and suggests that there is not a significant energy

disadvantage of generating 2m rings in amorphous silica.

31

Figure 3.11: One more layered cluster of Fig. 3.10 (a) corner-sharing cluster. (b)two-membered ring cluster.

One can argue that a 2m ring may result in a large strain than larger rings,

e.g., three-membered or four-membered ring. However, previously performed sim-

ulations reported the calculated strain energies of 2m ring were in range of 1.23–

1.85 eV/Si2O4 [42, 43, 44] which is smaller than formation energy of oxygen vacan-

cies frequently observed in silica. Boureau et al. [45] discussed the thermodynamical

lower bound of formation energy of the oxygen vacancy in β-cristobalite SiO2 is 7.3

eV/defect and ab initio studies showed the defect formation energy is between 5–9

eV per a defect [46, 47, 34]. These values support the idea that the strain energy of

2m ring may not be a considerable barrier of generating this configuration in silica if

one compares to the oxygen defects.

The formation of the 2m ring on silica surfaces has been also discussed in the

previous infrared studies [48, 49, 50]. In order to understand the vibrational spectrum

of the 2m ring, we calculated the vibrational density of states with the 24-atom

system. Figure 3.12 shows our calculations in comparison to the CPMD simulation

(dashed line) and experiment (circles). The overall agreement with experiment is

good despite the small simulation cell. The dotted line indicates the contribution

32

0 200 400 600 800 1000 1200 1400wavenumber(1/cm)

0

0.05

0.1

0.15

0.2

0.25

VD

OS

Figure 3.12: The calculated vibrational density of states (solid line) and the con-tribution of the four atoms which constitute the two-membered ring (dotted line).32cm−1 was chosen for the gaussian broadening. For the comparison, experimentaldata (circles) and CPMD simulation data (dashed line) were taken from Carpenterand Price [51], and Pasquarello and Car [52], respectively.

of the 2m ring atoms which is extended in 200–1000 cm−1 range. Two distinctive

and broad peaks between 250–450 and 700–900 cm−1 result from the 2m ring, and

the position of these two peaks are similar to the previous theoretical calculations

by Bendale and Hench [48], who showed several sharp peaks between 200–400 and

740–1100 cm−1. In IR experiments on dehydroxylated a-SiO2 surface, two unique

peaks at 888 and 908 cm−1 have been reported [49]. These two peaks are regarded

to be a strained defect, i.e., the edge-sharing structure on the surface. We note that

the IR experiments focused on the surface that was thermally treated. Therefore, the

population of the edge-sharing structure on the surface may have been very dense.

Since the 2m ring is not predicted to be abundant in bulk a-SiO2, we do not expect

contributions from 2m rings to result in vibrational features such as the D1 and D2

defect bands in Raman spectrum, which have been shown to be correlated with a

breathing mode of the 4m ring and 3m ring structure, respectively.

The density of states (DOS) for our simulated amorphous silica is given in

33

-20 -10 0 10energy (eV)

0

1

2

DO

S (

stat

es/e

V)

PARSECEXPCPMD

Figure 3.13: Density of states of amorphous structure. The X-ray photoemissionspectrum data are from Ref. [55].

Fig. 3.13. The dashed-line comes from x-ray photoemission experiments [53]. Each

peak of the DOS can be characterized by the atomic nature of the corresponding

electronic states in the energy region of interest [54, 55]. The states above -5 eV

correspond to lone pair, nonbonding 2p orbitals of O and the energies from -6 to

-11 eV is strong bonding of Si sp3 hybrid orbital and O p orbital. The states in the

region -15 to -20 eV are primarily O 2s orbitals. While the DOS ranging from -5–

0 eV is accurately predicted by our simulation, there is a disagreement around -10 eV

peak between calculated DOS and X-ray photoemission spectra (XPS). According

to Pantelides et al., the disagreement between XPS and the theoretical prediction

is casued by a matrix element effects since XPS is determined by not only valence

electron density of states, but also interaction between bond orbitals at different bond

sites [56]. Owing to this effect, we observed the same disagreement near -10 eV peak

in previous studies [55, 23].

3.7 Summary

In summary, we have performed ab initio molecular dynamics simulations for

both liquid and amorphous silicon dioxide including 64 units of SiO2 using real-

34

space pseudopotentials. In liquid simulations, we considered liquid systems at two

temperatures: 3,120 K and 3,700 K. We showed structural properties and dynamical

properties at each temperature, and compared our work with previously performed

Car-Parinello and classical MD simulations.

We also simulated amorphous silica. We compared several structural prop-

erties to experiments and other previous simulation results. Our calculated static

structure factor reproduced the experimental data very accurately. Bond length and

angle show similar values with comparing data except bond angle distribution. The

differences in Si-O-Si bond angle distributions between previous work and our work

are caused by a two-membered ring structure. We showed the possibility of their

existence in amorphous structure not only by performing cohesive energy calculation,

but also by calculating vibrational spectrum of the two-membered ring. Finally, we

presented the electronic structure of amorphous silica and the results were similar to

the previously performed simulation.

35

Chapter 4

Electronic and structural properties of

nanocrystals and clusters

Nanoscience and nanotechnology are among the most actively developing areas

in material sciences and engineering. Owing to their novel physical and chemical be-

havior, nanoscale materials have been designed applications in energy conversion and

storage, laser, bio-sensing, and catalysis [57, 58, 59]. Nanostructures, e.g., nanocrys-

tals, nanowires, and nanoclusters, are spatially confined in at least one direction

within a range of 1–100 nm. This results in “quantum confinement”, which alters

the electronic and optical properties of nano-scale materials compared to their bulk

counterpart [60]. To investigate nano-scale systems, quantum-mechanical based com-

putational studies are necessary. Such studies increase the level of understanding

of these materials and provides insights on how to design efficiently new materials

for industrial applications. In this chapter, we study metal oxide nanomaterials to

illustrate how electronic structure calculations can aid in understanding electronic,

structural, and optical properties of nanostructured materials. Specifically, results

are presented for SnO2 and TiO2.

4.1 SnO2 nanocrystals

4.1.1 Introduction

Transparent conducting oxides (TCOs) exhibit very interesting properties as

they are optically transparent in visible light, but electronically conductive. Owing to

36

their broad industrial applications such as optoelectronic devices and photovoltaics,

there has been much attention for TCO materials [61]. The most widely used TCO

material is In-doped tin oxide (ITO), yet In is not abundant in nature. Sb- or F-doped

tin oxide is considered to be a good alternative to ITO [62].

Although thin films are a widely used form for this material, more interesting

phenomena have been observed for nanocrystals. Successful synthesis of Sb- and F-

doped nanocrystals with a size of less than 10 nm diameter have been reported [63,

64, 65]. This broadens the opportunity to use these nanocrystals to manufacture

thin films or other nanostrucutres [66, 67]. For nanocrystals in general, controlling

electronic properties depends on not only the impurity, but also the size and the

shape of nanocrystals. Computational study plays a crucial role in providing detailed

information about these relationships. However, there are very few computational

studies reported [68] because of the complexity of the rutile structure, which is the

most stable crystal form of SnO2.

In this section, we present electronic structure calculations of Sb- and F-doped

SnO2 nanocrystals. Formation energy and electron binding energy, calculated by the

total energy difference of neutral and charged particles, for Sb and F dopants are

discussed. Our results show strong quantum confinement effects not only in the

homo-lumo1 gap of pure nanocrystals, but also in the electron binding energy of

doped nanocrystals, which is consistent with the previous nanocrystal studies [69, 70].

We also illustrate differences in structural and electronic properties between Sb- and

F-doped tin oxide nanocrystals.

1homo and lumo stand for highest occupied molecular orbital and lowest unoccupied molecularorbital, respectively

37

4.1.2 Computational details

All calculations are based on density functional theory utilizing real-space

pseudopotentials. Convergence is determined by a grid spacing, which is 0.3 a.u.

in these calculations. With this grid spacing, the total energy converges within

0.01 eV/atom. To find a minimum energy structure of each nanocrystal, we used

the BFGS method2 and all atoms were allowed to move until the largest force is less

than 0.005 Ry/a.u. The domain size was chosen to be 6.5 a.u. larger than the outer-

most atom of the nanocrystal. Outside of this spherical domain, the wave function is

set to be zero.

The pseudopotentials for Sn and O were generated with a valence configuration

of 5s25p2 and 2s22p4, respectively. We did not include 4d electrons in the valence

configuration, i.e., it is frozen into the core states, as the 4d states of Sn does not

affect the shape of the band structure except very deep levels [71]. We employed

the local density approximation (LDA) for the exchange-correlation functional from

Ceperley and Alder [5]. With these pseudopotentials, we obtained a band gap of

1.02 eV and the band structure is shown in Fig. 4.1. The experimental band gap

for rutile SnO2 is 3.6 eV which is much larger than our LDA band gap. It is a well-

known fact that the LDA underestimates the band gap. Since we are interested in

changes in the electronic structure, exact band gap calculations are not necessary for

the purposes of our study.

To construct our model for nanocrystals, we started with the rutile crystalline

SnO2 structure. We set the center atom to be Sn, then selected atoms that reside

inside a sphere with a given radius. Sn atoms with more than two dangling bonds

2Named after its inventors: Broyden, Fletcher, Goldfarb and Shanno.

38

Figure 4.1: Band structure of bulk SnO2. A direct band gap of 1.02 eV is observed.

39

Table 4.1: The number of atoms and diameter of the nanocrystal.

D(nm) Sn O H Total1.27 29 60 70 1591.69 69 140 138 3471.97 111 220 166 4972.37 191 384 262 837

and O atoms with more than one dangling bond were removed. To passivate dangling

bonds on the surface atoms, we generated fictitious hydrogens with fractional nuclear

charges and electrons. Sn and O dangling bonds were passivated with fiticious hy-

drogen atoms with 43

and 23

fractional charge, respectively [68]. In this way, we were

able to keep the same electron configuration of the rutile structure for the surface

atoms. Several sizes of nanocrystals were constructed with a diameter from 1.2 nm

to 2.5 nm. The diameter was defined by the expression d = (Ntot×Vunitcellπ )3. The

number of atoms in each nanocrystal is indicated in Table 4.1, and the shape of the

nanocrystals are shown in Fig. 4.2.

4.1.3 Results and discussion

4.1.3.1 Quantum confinement effect in Sb-doped SnO2 nanocrystals

In doped nanocrystals, the energy gap and binding energy can depend on the

size of the nanocrystals. It is a well-known fact that a strong blue shift occurs to the

energy gap as the dimension of the nanocrystals approaches the exciton bohr radius

(aB) due to the quantum confinement. Below aB, the electron and hole motion is

not treated as a correlated pair [72]. To investigate the size effect on the electronic

properties of the pure SnO2 nanocrystals, we calculated fundamental gap, defined as:

Eg = IP − EA (4.1)

40

Figure 4.2: Structure of H-passivated SnO2 nanocrystals. Sizes of the nanocrystalsare: (a) 1.27 nm, (b) 1.69 nm, (c) 1.97 nm, and (d) 2.37 nm.

where IP and EA are the ionization potential and the electron affinity, respectively.

IP and EA are calculated by the total energy difference between charged and neutral

system, defined by:

IP = E(n− 1) − E(n)

AE = E(n) − E(n+ 1).(4.2)

n indicates the number of the total electrons of the neutral system. Within the real-

space formalism, IP and EA calculations are straightfoward for the confined system

as it does not require an artificial periodicity. To calculate the energy of the charged

system in periodic boundary conditions, a mathematical trick, such as an artificial

jellium background [73], should be considered to prevent the Coulomb energy from

diverging. Figure 4.3 shows the energy gap for the undoped nanocrystals (black

diamonds). Since the exciton bohr radius for SnO2 is 2.7 nm [74], which is larger

than all of our nanocrystal sizes, we observe very steep increase in the energy gap as

41

1 1.5 2 2.5diameter (nm)

0

1

2

3

4

5

6

eV

Egap

Ebind

Figure 4.3: Fundamental gap of the pure SnO2 nanocrystals (black diamonds) andelectron binding energy of the Sb-doped nanocrystals (red diamonds).

the size of the nanocrystal decreases.

Electron binding energy is one of the key properties of doped semiconductors.

As Sb has one more valence electron than Sn, the extra electron creates a donor state,

n-type doping. In the n-type bulk semiconductor, the binding energy is defined as a

difference between the minimum conduction band energy and the donor state energy.

In a confined system, a more appropriate definition for the binding energy would

be the difference between the ionization potential of the doped nanocrystal and the

electron affinity of the pure nanocrystal [69]:

EB = IPd − AEp (4.3)

where d and p indicate doped and pure state, respectively. IPd, EAp, and EB values

with respect to the size of the nanocrystal are illustrated in Fig. 4.4. We note that

the Sb atom is located at the center of the nanocrystal. One interesting aspect of

this figure is that IPd does not change with respect to the size of the nanocrystal,

whereas the EAp has a strong size dependence. This is due to the localized donor

state as shown in the Fig. 4.6. The energy required to detach the defect electron from

42

1 1.5 2 2.5Diameter (nm)

0

1

2

3

4

5

6

eV

IPd

EAp

Ebind

Figure 4.4: Ionization potential of the doped nanocrystal (blue) and electron affinityof the pure nanocrystal (red).

the nanocrystal remains the same level regardless of the size of the nanocrystal. This

“pinned” energy level was also reported in previous nanocrystal studies for n-type

dopant systems [70].

Formation energy can be a measure of the stability of the doped nanocrystal.

It has been reported that a smaller nanocrystal tends to possess higher formation

energies [75, 76]. To determine if this is a case for SnO2 nanocrystal as well, we

calculated the formation energy of the Sb atom. In bulk systems, formation energy

depends on the chemical potential that is related to the partial pressure of the gas,

and the Fermi level. As we focus on the neutral defect in this study, the Fermi energy

is not taken into account in our analysis [75]. Therefore, the formation energy is

written:

Eform = Etot(Sb:SnO2) − Etot(SnO2) + n[µ(Sn) − µ(Sb)] (4.4)

where n is the number of defect atom of the particle and µ is the chemical potential.

Although µ is related to the different thermodynamic limits [77], we consider the

chemical potential to be the energy of the individual atom obtained from the bulk

43

crystaline structure for both Sn and Sb atom. Figure 4.5 shows a relative formation

energy of the nanocrystal. In general, more positive formation energy means less

favorable, indicating that the smaller nanocrystal requires more energy to dope the

Sb atom into the nanocrystal. To understand this phenoenon, we plot the isosurface of

the defect level orbital in Fig. 4.6. The same value of the electron density was chosen

for both isosurface plots. In small nanocrystals, the electrons are more localized

around the Sb atom. This may increase the kinetic energy of the electron, which

adds the energetically unstable character to the doped nanocrystal.

4.1.3.2 Antimony vs. Fluorine dopant atoms

As mentioned earlier, F is also used as n-type dopant for SnO2. The F-doped

SnO2 nanocrystals have been synthesized within a range of 3–10 nm by means of the

chemical vapor synthesis [64]. Owing to the smaller size of F compared to Sb, two

possibilities for the doping site can be considered, i.e., substitutional and interstitial

doping sites. Experimental studies of F:SnO2 nanocrystals show that at low F concen-

tration, F substitutes O. For high F concentration, interstitially doped F is observed,

but they tend to make substitutional-interstitial pair [64]. In this study, we consider

only one F atom in the nanocrystal, which is considered low doping concentration, to

compare the difference between Sb and F. We constructed F-doped nanocrystals that

substitute F for O. We chose the same nanocrystal size of 1.69 nm for both Sb and F

doped nanocrystals, and located Sb at the center and F at the nearest neighbor from

the center Sn atom as the size of the nanocrystal is relatively small.

Figure 4.7 shows the visualized defect state wave function for Sb (left) and

F (right). The defect state of the Sb dopant is localized on the Sb atom. For the

F-doped nanocrystal, however, the defect wave functions are lying between F and

44

1 1.5 2 2.5Diameter (nm)

-3

-2

-1

0E

form

(eV

)

Figure 4.5: Formation energy for the antimony dopant atom with respect to the sizeof the nanocrystal.

Figure 4.6: Dopant level wave function isosurface plot. Red and purple indicateoxygen and tin, respectively. Blue sphere indicates the surface hydrogen. Wavefunction is localized around the dopant antimony atom.

45

Figure 4.7: Defect wave function isosurface plot. Antimony and fluorine dopants areindicated in grey and yellow color, respectively.

Sn. This hybridized orbital feature is also observed in the bulk F:SnO2 system.

Velikokhatnyi and Kumta performed DFT calculations for Sn16O31F, and presented

total and projected density of states [78]. Rather than creating new states below the

conduction level, F-doping causes a Fermi-level shift towards to the conduction level.

Accordingly, additional electrons by F doping exhibit 5s and 5p Sb character rather

than 2p F character, which is similar to our observation.

To compare the structural difference in Sb and F doping, we calculated average

bond length changes in Sb-O and Sn-F compared to original Sn-O bond length, and

their changes was found to be small. However, the opposite trend between Sb and

F in bond length changes were observed. For the Sb doping, the Sb-O bond length

shrinks about 0.05 A which is a 2.5% decrease compared to the original Sn-O bond

length. On the other hand, the Sn-F bond increases by 0.15 A. This opposite trend

can be understood by calculating fractional charge on each atom. As Sn and O atom

have six and three bonds respectively, Sn provides −23

to O to make +4 ionic charge

because the electronegativity of O is larger than Sn. When we substitue O with F,

only one electron can fill the empty 2p state of F. Once the F 2p orbitals are filled,

46

remaining electrons act to fill 5s and 5p states of Sn rather than the 3s state of F.

This may cause the defect electron density splitting that we found in Fig. 4.7, and

the stretched bond length. The electron binding energy for Sb and F doping was

calculated as well, and their values are 2 eV and 2.8 eV, respectively. From this

result, we expect that F doping creates a deeper donor state than Sn doping.

4.1.3.3 Higher doping concentration

In the previous sections, we have considered the effect of Sb and F doping

in SnO2 nanocrystals for only one dopant in each nanocrystal. Several experimental

studies have reported that highly doped Sb:SnO2 materials exist in both thin films

and nanocrystals [79, 80]. A 2–7% doping concentration for Sb is known to produce

a degenerate semiconductor, which has a high electrical conductivity. In order to

investigate the effect of different doping concentrations on the electronic structure of

small nanocrystals, we considered four different doping concentrations (0.3, 1.2, 2.7,

and 3.9 at%) with a nanocrystal of a 1.97 nm diameter. For each concentration, we

substituted Sn to Sb and the dopant atoms were not very close to each other.

Figure 4.8 shows the binding and formation energy of the Sb dopant. Both

binding and formation energies decrease as the concentration of the antimony dopant

(cSb) increases. Our results indicate that the higher dopant concentration is more

energetically favorable than the lower concentration. This trend is oppostie compared

to ZnO nanocrystal calculations [81]. In ZnO nanocrystals, we tried substitutional

doping for Zn to Ga, and more than two Ga atoms showed much higher formation

energy than one Ga atom. It explains that why one can observe high Sb concentrations

in Sb:SnO2 system. For binding energies, as the electron affinity does not change in

Eq. 4.3, IPd decreases as a function of the cSb. This indicates that high doping

47

0 1 2 3 4Sb concentration (at%)

-4

-2

0

2

Ene

rgy

(eV

)

Ebind

Eform

Figure 4.8: Electron binding energy (diamond) and formation energy (square) fordifferent doping concentration.

concentration changes the dopant level character from deep to shallow, which makes

these particles more attractive in the applications for electric devices.

4.1.4 Summary

In summary, we presented electronic structures for SnO2 nanocrystals within

a range of 1.3–2.4 nm diameter. Strong quantum confinement effect was observed not

only for the pure nanocrystals, but also for the Sb-doped nanocrystals. The binding

energy of the Sb dopant showed a strong confinement due to the highly localized

nature of the dopant wave function. We also examined the differences in dopant level

between Sb- and F-doped nanocrystals. The higher concentration of Sb was revealed

to lower the nanocrystals’ formation energy and binding energy as well.

4.2 TiO2 clusters

4.2.1 Introduction

TiO2 is one of the most studied transition metal oxides owing to its high

potential in many industrial applications. Specifically, photocatalytic activity of TiO2

48

has attracted much interest [82] to meet demands for clean energy production. In

particular, small TiO2 clusters, containing up to a couple of tens of atoms, exhibit

exciting properties as they alter the optical properties [83]. To make this material

more attractive in such applications, it is essential to understand their electronic and

structural properties; however, due to a small dimension, there is no definitive way

to characterize the relationship between the electronic properties and their geometry.

Computational study combined with the global minimum (GM) search have been

conducted based on a conventional belief that the GM structure would be the most

probable structure in nature. Nevertheless, no such agreement with photoemission

spectroscopy (PES) has been achieved [84].

In this section, we present methods for determining the low energy structures,

and introduce many-body theory calculations that are beyond the Kohn-Sham one-

electron interpretation in order to acquire better optical properties.

4.2.2 Global minimum searching methods

4.2.2.1 Simulated-annealing technique

Simulated-annealing is a method for solving the global optimization problem.

Previous theoretical studies on clusters have shown that this method is an effective

way to determine the structures of small size clusters [31]. In our calculations, the

simulated-annealing technique was performed in two steps:

1) Molecular dynamics simulation:

We generated random configurations of titania clusters. Each (TiO2)n system was

coupled with a heat bath and the temperature was controlled via Langevin dynamics.3

3The details of Langevin dynamics were explained in chapter 3.

49

Figure 4.9: Annealing schedule. The initial temperature was set to 3000K and thetemperature was decreased at every 100 steps until when temperature reaches to300K. For (TiO2)4 cluster simulations, we chose the temperature step of 250K insteadof 500K.

Initial and final temperature and the details of annealing schedules are shown in

Fig. 4.9. A time step of 200 a.u. and a friction coefficient of 0.0005 a.u. were chosen

to integrate the equation of motion. The grid spacing and the boundary sphere radius

were 0.35 a.u. and 20 a.u., respectively.

2) Structural relaxation:

The structure of each cluster obtained from MD simulations was relaxed until the

maximum force component is less than 0.004 Ry/a.u. This was performed with a

finer grid spacing of 0.20 a.u. to ensure accurate forces. The boundary sphere radius

was chosen such that the distance between the outermost atom and boundary was

larger than 8 a.u.

4.2.2.2 First-principles basin-hopping technique

Basin-hopping is a Monte-Carlo based method for exploring potential energy

surfaces (PES) by performing consecutive jumps from one local minimum to another.

As shown in Fig. 4.13, this method provides transformed PES that maps each config-

uration space to one local minimum energy. To perform basin-hopping optimization,

50

Figure 4.10: Illustration of basin-hopping optimization process. E({Rm}) respresentsthe original potential-energy surface and E ({Rm}) is transformed potential-energysurface. Adapted from Ref. [85].

the initial random cluster is generated and the configuration of the cluster is modi-

fied by a set of trial moves (∆Rm = rme(θ, φ), where e(θ, φ) is a random unit vector

expressed in spherical coordinates and rm is a move distance for atom m. Rm is the

position of atom m) created by uniformly distributed random numbers, followed by

a local relaxation [85]. A trial move and a relaxation are indicated as green and red

arrows in Fig. 4.13. This method has been successfully applied to Lennard-Jones clus-

ters and biomolecules, but for small clusters of our interest, PES should be considered

within a quantum mechanical way. In this study, the basin-hopping optimization was

performed based on DFT with Perdew-Burke-Ernzerhof (PBE) exchange-correlation

functional [86].

The comparison between simulated-annealing and basin-hopping is shown in Fig. 4.11

and Fig. 4.12. The most stable structures and their relative energies are indicated.

The results of simulated-annealing and basin-hopping are consistent. The energy

differences between the global minimum cluster to other isomers of two simulations

agree within 0.4 eV.

51

Figure 4.11: n=2-3 isomers

52

Figure 4.12: n=4 isomers

4.2.3 Computational details

All electronic structure calculations were performed by using the all-electron

numerical atom-centered orbitals code fhi-aims [87]. The numerically tabulated

atom-centered orbital basis sets are grouped into a minimal basis, containing only

basis functions for the core and valence electrons of the free atom, followed by four

hierarchically constructed tiers of additional basis functions (tiers 1–4) [88]. For the

global minimum search, we used both simulated-annealing performed by the parsec

code, and basin-hopping with a tier 1 basis set. Up to n=4 (TiO2)n clusters, the same

global minimum structures were obtained from both methods. However, the basin-

hopping method provided a wider spectrum of isomers that enabled us to compare our

53

clusters to the experiment. We used basin-hopping for bigger clusters, and examined

all isomers found within 1.25 eV from the global minimum energy.

It is essential to calculate accurate values for homo-lumo energy to make a

direct comparison with electron affinity and detachment energy observed in experi-

ment. To do this, we performed many-body electronic structure calculations within

the GW method, where G is the one-particle Green’s function and W is the dynam-

ically screened Coulomb potential [89, 90]. To mitigate the computational demand,

we used a perturbation approach denoted by G0W0. Within G0W0 calculations, G

and W are obtained by the first-order corrections to the DFT eigenvalues. Our G0W0

results are based on PBE-based hybrid functional [91].

4.2.4 Structural analysis of the low-energy clusters found in basin-hopping

One might expect that the experimentally observed structure to be a global

minimum structure. Interestingly enough, this energy-structure analogy does not

apply to all of the TiO2 clusters. In Fig. 4.13, we present the lumo level of neu-

tral clusters and the homo level of anion clusters to compare to the experimentally

observed electron affinity (EA) and vertical detachment energy (VDE). Global mini-

mum isomers, expected to be the case for the best agreed structure, are illustrated as

red triangles. However, global minimum isomers do not match with either EAs and

VDEs. The best agreement is observed with highest vertical electron affinity (VEA)

clusters, i.e., the isomer that holds the lowest homo energy, except n=4 and 5. For

these clusters, slightly lower VDE isomers were selected to be a better agreement

with both EA and VDE. These clusters are indicated as green diamonds.

Highest VDE isomers also present highest VEA except n=4, 5, and 7. To

investigate the correlation between highest VEA and the structures, we performed

54

Figure 4.13: lumo energies for neutral clusters (upper), and homo energies for anionclusters (lower). Courtesy of Noa Marom.

structural analysis and found that the highest VEA clusters have three things in

common, except for the n=7 cluster:

a. The highest VEA clusters have only one tri-coordinated Ti atom, on

which the anion homo orbital is localized.

b. The oxygen atoms bound to the tri-coordinated Ti atom do not have

a dangling bond.

c. The bond angle of O-Ti-O is close to the tetrahedral angle of 109.5 ◦.

The structures that hold those three properties are shown in Fig. 4.14 and Fig. 4.15.

We listed clusters whose root-mean-square deviation is less than 10 ◦ from the tetra-

hedral angle for n ≥ 5 clusters. (For n < 5 clusters, root-mean-square deviation

becomes larger compared to the bigger clusters.) All tri-coordinated Ti atoms are

shown as green spheres. The anion homo is highly localized on that Ti atom.

55

Figure 4.14: Structural analysis for the cluster size of n=2-6.

56

Figure 4.15: Structural analysis for the cluster size of n=7-10.

57

Can one explain why the photoemission spectroscopy (PES) selectively ob-

serves the high VEA isomers rather than the global minimum structures? These

couterintuitive results were also observed in the study of SiD− clusters [92]. To an-

swer this question about the higest VEA observation, Kronik and coworkers assumed

that the neutral low energy isomers can be generated with equal probability due to

the high temperature [92]. When the clusters enter the plasma region, electrons are

selectively attached to the high VEA, in which the cluster includes Ti3+ site in our

case. After transforming to the anion, they may have a dwell-time before measure-

ment, then the clusters can relax to the metastable structure. Before these isomers

transfer to more stable structure, PES mesurement is performed. This explains how

we obtained the better agreement with the high VEA isomers, not the global mini-

mum structures.

4.2.5 Summary

We studied (TiO2)2−10 clusters by means of DFT-based basin-hopping and

simulated-annealing for the global minimum search, and the many-body perturba-

tion theory within G0W0 approximations for the electronic structure calculations.

Our simulations demonstrated the structure-selection criterion in photoemission spec-

troscopy experiments: the high vertical electron affinity clusters are selectively ob-

served rather than the global minimum clusters.

58

Chapter 5

Noncontact atomic force microscopy study

5.1 Introduction

Scanning probe microscopy is a widely used technique in surface science and

nanotechnology for studying surface properties. Since the invention of scanning

tunneling microscopy (STM) in 1982 [93] and atomic force microscopy (AFM) in

1986 [94], scanning probe microscopy has evolved not only to provide a better reso-

lution, but also to be applicable for a wide range of materials.

STM probes tunneling currents between the tip and the sample, which reflect

the distribution of electron states available for electron transfer. In order to measure

the tunneling current, STM requires conducting materials, which limits the use of

STM to various materials. Also, STM cannot be used in ambient conditions without

proper surface preparation. On the other hand, AFM probes forces between the

tip and materials, which do not require electrical conductivity. Consequently, AFM

became a popular probe microscopy.

The basic components of AFM are a light source, a cantilever containing a

sharp tip, a sample surface, and a photodetector as illustrated in Fig. 5.1. A sample

surface is mounted on a piezocrystal, which adjusts the position of the sample and

the probe. Once the position of the tip and the sample are determined, the deflection

of the cantilever is monitored by the change of the path of the laser beam. Since the

deflection is directly affected by tip-sample interactions, the tip geometry is known

59

Figure 5.1: Basic AFM set-up. Adapted from Ref. [95]

to be important for determining the interaction between the tip and the surface.

However, the tip shape dependency can be overcome by the sharpness and the aspect

ratio of the tip [95]. Most commercially manufactured AFM tips have a square-based

pyramid or a cylindrical cone shape.

The primarily AFM operation mode is a simple contact mode [95], which

utilizes the strong repulsive force between the AFM tip and the surface. Even though

the contact mode AFM showed successful images on atomic and molecular scale,

Pethica et al. [96] observed that contact AFM could not observe atomic defects since

the contact area is a limit to AFM resolution.

True atomic resolution has been achieved by noncontact AFM (nc-AFM) [97].

NC-AFM captures weak attractive forces between the tip and the sample rather than

strong forces. In nc-AFM, two operating modes are available: amplitude modula-

tion [98] and frequency modulation [99]. While amplitude-modulated AFM is known

for working well in air and liquid environments, frequency-modulated AFM works

well in ultra-high vacuum conditions, which is used in many traditional surface stud-

ies [99].

60

Even though nc-AFM has been applied to many surface structures, ambigu-

ous results are sometimes reported when the surface contains defects or reconstructed

structures. Consequently, theoretical studies have an important role in a comprehen-

sive analysis of AFM images of materials.

In general, AFM simulations include the AFM tip in order to calculate forces

between the surface and tip apex atom. To calculate the frequency shift of the AFM

tip, the forces over the range of the tip trajectory are integrated. Accordingly, the

tip-surface force must be calculated at every point within the region of interest. This

approach requires a large number of computationally intensive calculations. Another

difficulty comes from the geometry of the AFM tip since the atomic configuration of

the tip is usually unknown.

These difficulties can be overcome considering the fact that AFM images are

not sensitive to the tip. Recently, a possible solution for reducing computational

work was proposed by Chan et al. [100] They showed an effective calculation method

without including an explicit description of the tip. They successfully applied this

scheme to well-known surface structures, such as the Si(111)-(7×7) and TiO2(110)-

(1×1) surfaces. The power of the method by Chan et al. is that it only requires one

self-consistent calculation for a given a model of the surface.

In this chapter, we describe how to investigate complex nanostructured sur-

faces and two dimensional organic molecules by simulating nc-AFM images. Our

study is based on the new method that Chan et al. proposed, but also includes

computations with a physical model of the tip.

61

5.2 Framework for simulating noncontact atomic force mi-croscopy images

AFM responds to atomic forces between the tip and surfaces. As such, un-

derstanding the nature of the atomic force is a key aspect of describing the AFM

operation. Noncontact-AFM measures the change in oscillation amplitude or fre-

quency of the tip that is mainly caused by attractive forces between the tip and the

surface. In this study, we focus on describing frequency-modulated nc-AFM under

ultra-high vacuum conditions.

5.2.1 Forces between the tip and sample

In general, the forces which determine the AFM image can be classified as

long-range forces and short-range forces. Van der Waals, electrostatic and magnetic

forces are considered long-range forces and chemical forces are considered short-range

forces [101].

The van der Waals (vdW) force originates from the dipole fluctuations caused

by the correlated motion of electrons. Electrostatic forces are strong long-range forces

caused by charged objects. Two different sources are responsible for the electrostatic

force, i.e., charged defects and contact potential differences. Charged defects can be

generated during the surface preparation process. However, these charges can often

be removed by heating the sample surface [102]. The contact potential difference

arises from the difference in the work function of two different materials. When two

materials are brought into an electric contact, electrons transport from one material

with a smaller work function to the other with a higher work function until they reach

the same Fermi level. This contact potential difference is a key measurement in Kelvin

probe force microscopy, which is also a widely used scanning force microscopy [103].

62

Figure 5.2: Tip motion in nc-AFM. The equilibrium tip-surface position is at q = 0,and d is the closest distance between the tip and the surface.

In nc-AFM experiments, the contact potential difference can be cancelled out by

applying a bias voltage to the tip.

Short-range chemical forces originate from the interaction of atomic orbitals,

Pauli exclusion principles, and ion-core repulsion. The strong repulsive short-ranged

forces are dominant in contact mode AFM operation.

5.2.2 Derivation of expressions for the frequency shift calculations

Once the type of force is identified, one can set up an equation which governs

the motion of the tip. Generally, the tip motion is assumed to be a one-dimensional

oscillator with an effective mass, m:

mz(t) +mf0

Qz(t) + kz(t) − Fts(zc + z(t)) = Fext(t), (5.1)

where k, Q, and f0 are the spring constant, the quality factor, and the resonance

frequency of the tip, respectively [104]. Fext is the external force which drives the

oscillating motion. Figure 5.2 shows a schematic of the tip motion.

In the frequency-modulation mode, the tip oscillates with the resonance fre-

quency, f0, which is a material-dependent parameter. To keep a constant amplitude,

63

the feedback-loop adjusts the external force to compensate for the damping force,

(mω0/Q)z(t). Then, Eq. (5.1) can be rewritten as

mz(t) + kz(t) − Fts(zc + z(t)) = 0. (5.2)

This equation is normally solved with two approximations:

1) Small oscillating amplitude approximation:

If the tip oscillates with small amplitude, the tip-surface interaction forces can be

described with a first-order Taylor series expansion, Fts(zc +z) ≈ Fts(zc)+z ·∂Fts

∂s|s=zc

.

Eq. (5.2) is now written:

mz(t) +

[

k −∂Fts

∂s

s=zc

]

z(t) = Fts(zc). (5.3)

The solution for the second order differential equation is given by z(t) = A0 cos(2π(f0+

∆f)t) where A0 is the amplitude of the oscillating tip. Then the frequency shift is

approximated by

∆f ≈ −f0

2cz

∂Fts

∂s

s=zc

. (5.4)

2) Large oscillating amplitude approximation:

In most experimental conditions, frequency-modulated AFM uses a relatively large

oscillating amplitude compared to the typical tip-surface interaction distance. In

these conditions, the frequency change can be considered to be a small perturbation.

The general expression for the frequency shift based on Hamilton-Jacobi for-

malism [105, 106] is

∆f = −f0

2kA20

< Ftsz >

= −f 2

0

kA0

∫ 1

f0

0

Fts(d+ A0 −A0 cos(2πf0t)) cos(2πf0t)dt.

(5.5)

64

Substituting q = A0 cos(2πf0t) in Eq. (5.5) yields

∆f =f0

kπA20

∫ A0

−A0

Fts(d+ A0 − q)q

A20 − q2

dq. (5.6)

Integration by parts rewrites Eq. (5.6):

∆f = −f0

2k

∫ A0

−A0

kts(d+ A0 − q)2

A20π

A20 − q2dq. (5.7)

This equation is similar to Eq. (5.4), where the kts is replaced by a weighted average.

5.2.3 An efficient method for force calculations

There can be different sources causing the frequency shift in noncontact-AFM

experiments. First-principles analyses have been performed to calculate the quantum

chemical force caused by interactions between the tip and the surface. In a previous

first-principles study, Chan et al. [100] suggested an efficient method that does not

include an explicit description of the tip. Their method treats the influence of the

surface on the tip as a small perturbation, assuming that the motion of the cantilever

tip does not affect the electronic states of the surface. Based on this approximation,

the tip-surface interaction energy can be written as

Ets(r) =

|φ(r′ − r)|2Vts(r′)dr′, (5.8)

where r, Vts, and φ are the position of the tip, the potential on the tip generated

by surface atoms, and the electronic state of the tip, respectively. Since force is a

gradient of total energy, the tip-surface force can be described expanding the potential

up to the first-order expansion around tip position r:

65

Fts = −∇Ets(r)

≃ −∇Vts(r) −∇

[

∇Vts(r)

|φ(r′ − r)|2(r′ − r)dr′]

= −∇Vts(r) −∇ [∇Vts(r) · p]

= −∇Vts(r) − α∇(

|∇Vts(r)|2)

,

(5.9)

where p is the polarization of the tip, assumed to have a linear relation to ∇Vts by

α, the polarizability of the tip material. The first term is a monopole caused by the

electrons of the tip, which is canceled by the ionic potential of the tip if the tip is

neutral. In this setting, the force acting on the tip is now proportional to the dipole

interaction,

Fts(r) ∝ −∇(

|∇Vts(r)|2)

. (5.10)

The tip-surface potential is evaluated by the Hartree and the ionic potential from ab

initio calculations: Vts = Vhart + Vion. Note that the exchange-correlation potential is

not included because the tip is not treated as a quantum system.

Here, we calculate the self-consistent potential Vts(r) of the surface and eval-

uate the force using Eq. (5.10). The frequency shift of the tip is calculated with

Eq. (5.4) and (5.7) for a given height and amplitude. The nc-AFM images are gener-

ated by tracing the contour surface of constant ∆f .

The calculated tip-surface force from Eq. (5.9) can be inserted to Eq. (5.6) to

compute the frequency shift. Since Vts decays very fast in vacuum, the most relevant

integration region is only a couple of angstroms near the tip-surface turning point (d

in Fig. 5.2). Therefore, the integration region in Eq. (5.6) can be reduced from (−A0,

A0) to (−A0, −A0+2∆), with ∆ chosen to be small compared to A0.

66

5.3 Two-dimensional structures

5.3.1 GaAs(110) surface

We chose GaAs(110) surface as it is a simple surface on an extensively stud-

ied material. This surface does not reconstruct, but noticeable surface relaxation is

observed wherein the As atom moves out of the surface plane and Ga atom moves

inward. As other III-V(110) and II-VI (110) semiconducting surfaces possess a simi-

lar relaxed morphology, GaAs(110) can be considered as a prototype of these surface

systems.

5.3.1.1 Computational details

All calculations were performed with real-space pseudopotentials with the lo-

cal density approximation for the exchange-correlation functional from Ceperley and

Alder [5]. The convergence of the total energy is controlled by the grid spacing, which

was taken to be 0.35 a.u. We used Troullier-Martins pseudopotentials [7] with a par-

tial nonlinear core-correction for the Ga atom [107]. The parameters used to generate

the pseudopotentials were obtained from a previous work by Kim and Chelikowsky

for vacancies in the GaAs surface [108]. With these pseudopotentials, we obtained a

lattice parameter of 5.59 A, which agrees within 1 % of the experimental value [109].

Our simulation cell for the GaAs(110) surface contains a 1×2 surface unit

with five atomic layers that corresponds to a slab thickness of 17.79 A, and 21 A of a

vacuum was inserted. In order to passivate the dangling bonds on the opposite side

of the surface, we generated hydrogen-like pseudopotentials with a 1s0.75 and 1s1.25

ionic configuration for Ga and As atom, respectively [110]. The ~k-points were gen-

erated by a Monkhorst-Pack scheme [111] with a 4×3 mesh. Geometry optimization

was performed until the force acting on each atom is less than 0.004 Ry/a.u. After

67

Figure 5.3: A side view (a) and a top view (b) of the relaxed GaAs(110) surface.Magenta, yellow, and blue indicate Ga, As, and H, respectively.

structural relaxation, the Ga atom moved inward and the As atom moved outward

with 0.69 A of buckling as illustrated in Fig. 5.3-(a). This value agrees very well with

the observed value of 0.69 A in the experiment [109].

To create the simulated AFM image, we set two parameters in Eq. (5.7): A0

and d. We tested the sensitivity of these parameters to the AFM images by varing the

values for A0 from 5 nm to 50 nm, but have not found significant changes in image

contrast. A0 is chosen to be similar to the typical oscillation amplitude (10 nm) used

in experiment [112]. The integration limit ∆ is selected to be 1 A for all calculations.1

5.3.1.2 Results and discussion

The parameter that has the most significant effect on the AFM image is the tip

turning point d. Figure 5.4 shows our simulation images with three different values

for d: 3 A, 4 A, and 5 A for (a), (b), and (c), respectively. In this figure, we observe

the opposite trend of how the Ga and As atoms respond. As the tip turning point

1We also tested the effect of the value for ∆ by chaning it from 0.5 A to 2.5 A, but no contrastchange was observed.

68

increases, the bright spots which correspond to the Ga atom disappear while the As

atom becomes bigger and brighter. Typically, most AFM experiments on III-V(110)

semiconducting surfaces have shown only the anion image which is similar to Fig. 5.4-

(c). This seems natural because the III-V(110) surface has a surface buckling, i.e.,

the anion is displaced to the vacuum. However, some recent experimental studies

have reported simultaneous observation of both the anion and cation sublattices. In

the case of the GaAs(110) surface, Uehara and coworkers studied image contrast by

using several frequency shift values and Si tips. [113] Among the several AFM images

they obtained, the Ga sublattice appeared in the large frequency shift AFM images,

which is shown in Fig. 5.4-(d) and (e). The larger frequency shift corresponds to the

smaller tip-surface distance and vice versa. Our simulation results are consistent with

this experiment. Also, the contrast of our simulated images is very similar to that of

the experimental images.

Previously, an ab initio AFM study on the GaAs(110) surface with several

Si-cluster tips was conducted by Ke and coworkers [114, 115]. As the tip can be

contaminated by the sample material, which is Ga and As in this case, they calculated

tip-surface energy and force with Si, Ga, and As apexes. To make a comparison

between our method and their results, we evaluated −∇(∇Vts)2 from Eq. (5.10),

which is directly related to the tip-surface force. We do not expect to calculate

absolute values for the force as the polarizability of the tip, α, is unknown. It is

unnecessary to calculate the exact force to obtain AFM images as we only need to

know the relative difference of the frequency shift at each point. Figure 5.5 presents

the calculated forces from our simulation (top panels) and Si-cluster tip simulations

(bottom panels). We chose the tip-surface distance to be 3.41 A for line A and

4.21 A for line B, and these values are calculated from the surface As plane. The

69

Figure 5.4: Simulated AFM images with respect to the tip turning point (d). ∆ isset to be 1 A for (a)-(c), and the values for d are: (a) 3 A, (b) 4 A, and (c) 5 A.The images are overlaid with the surface Ga (magenta) and As (yellow) atom. Blackand white indicate low and high frequency shift values, respectively, and the grayscale is adjusted independently. (d)-(f): Noncontact AFM images of GaAs(110) fromexperiment [113]. The frequency shift is -137 Hz, -188 Hz, -218 Hz for (d), (e), and(f).

70

Figure 5.5: Comparison of tip-surface forces. (a) A top view of the GaAs(110) surfaceand the black color indicates top layer atoms. Dashed-line A and B correspond tograph (b) and (c). Top panels in graph (b) and (c) show our results calculated fromEq. (5.10). Other three panels are previous ab initio results simulated by Si-clustertip with Si, Ga, and As apexes (Ref. [114] and [115]). The tip-surface distances are3.41 A and 4.21 A for (b) and (c), respectively.

corresponding tip-surface distances of the reference data are 3.38 A and 4.15 A for

line A and B, respectively. In this way, we can maintain a similar distance between

the tip apex and the surface As or Ga atom for each line, and can compare the closest

set of the force curve from the previous simulations. In both graphs, our results are

in good agreement with the Si apex simulation results. Our simulation and the Si

apex simulation show the strong force when the tip is close to the surface atoms. On

the other hand, Ga and As apexes detect only one surface atom: Ga apex responds

to the surface As atom while As apex responds to the Ga surface atom. This leads to

a conclusion that only the pure silicon tip would image both Ga and As sublattices at

the relatively small tip-sample distance while the larger tip-sample distance images

the As sublattice. This is the same observation that we have made in Fig. 5.4.

This result implies that our method should be suitable for simulating AFM

71

images obtained with a Si tip. Our method is based on a simple approximation: the

tip does not affect the electronic structure of the sample as expected in the limit

of a weak interaction between the tip and the surface. Our simulation may not be

applicable in a certain cases where this limit is not satisfied, e.g., when the tip and the

surface atom form a chemical bond. Since our simulations match well with previous

calculations employing a Si apex tip model [114], we believe the Si tip does not

saturate a dangling bond of either Ga or As atom. The pure Si tip does not exhibit

a strong relaxation effect with the surface Ga or As atoms if the tip-sample distance

is larger than 3 A [114]. The calculations found for shorter distances the electronic

structure of the surface is changed by the tip.

In Fig. 5.5, we note that the position of the force peak is slightly shifted from

the actual surface atom positions. In line A, the peak is moved toward [001] direction

from Ga atom while the As atom signal moves to [001] direction in line B. These

shifts are related to the dangling bond of the unperturbed surface atom as presented

in Fig. 5.6. We plot the electron density of the surface that is close to the fermi

level. Fig. 5.6-(a) shows the fully occupied dangling bond of the As atom and 5.6-(b)

represents the empty dangling bond of the Ga atom. These position shifts were also

observed in experiment. Figure 5.6-(c) shows measured atomic structure from the

AFM experiment along the line X-X′ and Y-Y′ that correspond to the line B and

line A in Fig. 5.5, respectively. In this figure, β is expected to be 1.13 A in the ideal

situation. However, the AFM measurement determined the value between 2.4 and

2.6 A which is more than twice larger. In our simulation, the measured value is about

2.7 A.

Our method will provide a comparable AFM image as is obtained by a sili-

con tip in the region where the tip does not undergo considerable changes in their

72

Figure 5.6: (a) The dangling bonds of the surface As atom. The electron densitywithin 1 eV energy window below the Fermi level is visualized. Black and light grayrepresent Ga and As, respectively. (b) The empty dangling bonds of the surface Gaatom. (1 eV energy window above the Fermi level.) (c) Ga and As signals from AFMexperiments. (Adapted from Ref. [113]) Dashed and solid lines indicate X-X′ andY-Y′, respectively.

electronic structure. Our approach will be especially useful not only because most

commercial AFM tips are made of Si, but also the Si tip can serve as a reference for

interpreting AFM images since it supports the similar contrast mechanism for many

surfaces without regard to physical and electronic structures [116].

5.3.2 Graphene and its defect structures

Graphene has attracted much attention as a promising material for a wide

range of applications owing to its unusual electronic and mechanical properties [117,

118]. Graphene is a two-dimensional allotrope of carbon, and forms a simple honey-

comb lattice structure. The electronic structure of graphene can be greatly altered

by introducing defects or impurities [119]. Since just a small structure change can

influence the electronic properties of graphene-based materials, understanding struc-

tural changes at the atomic level is very important. Although STM has been widely

73

applied to study graphene structures [120], AFM has several advantages over STM

as mentioned in the introduction of this chapter. Even for the simple honeycomb

graphene structure, AFM experiments are difficult to interpret since the hollow site

of graphene is frequently reported rather than the carbon site [121, 122]. Moreover,

there are currently no clear nc-AFM experimental images reported for the graphene

defect structures owing to the technical difficulties in performing AFM experiment.

In this section, we present nc-AFM simulation images for graphene with and with-

out defects. Our goal is to provide AFM images that can be used to predict the

experimental ones.

5.3.2.1 Computational details

All calculations were carried out with real-space pseudopotentials with LDA

for the exchange-correlation functional from Ceperley and Alder [5]. A grid spacing

of 0.3 a.u. was employed. With this grid spacing, total energy was converged within

0.02 eV/atom. The valence electron configuration of C is 2s22p2 with 1.49 a.u. and

1.52 a.u. cutoff radii for s and p, respectively. With this pseudopotential, the C-

C bond length was optimized at 1.407 A, which is less than 1% smaller than the

experiment. For the defect-free graphene calculations, we used a 4.6×8.0×39.0 a.u.3

supercell that contains four C atoms to make a rectangular cell. The k-points were

sampled with 8×6×1 mesh by Monkhorst-Pack scheme [111]. For the defect structure

calculations, we used a larger supercell to prevent an artificial interaction between

the defect sites. A 32×32×39 a.u.3 supercell was used with a 2×2×1 k-point mesh.

Structural relaxation was also performed until the force is less than 0.004 Ry/a.u. for

each atom.

74

Figure 5.7: Simulated nc-AFM images for defect-free graphene structure. Tip-turningpoint was set to 2 A (left) and 3 A (right). Smaller d yields bright spots at carbonatom site whereas larger dts yields bright spots at hollow site.

5.3.2.2 Results and discussion

a. Graphene without defects

Figure 5.7 shows the nc-AFM images for graphene created without modeling an ex-

plicit tip. The tip-sample distance (d in Fig. 5.2) is set to 2 A and 3 A. Most of the

experimental AFM studies of graphite and carbon nanotubes have reported that the

hollow sites of the structure are visible, i.e., a hexagonal pattern is shown rather than

a honeycomb pattern. Our simulated AFM image, calculated with a larger d (3 A),

agrees with the previous experiment [122].

A few theoretical and experimental studies have reported that the AFM images

for graphene can be sensitive to the tip material and the tip-sample distance [122, 121].

A recent AFM study of the epitaxial graphene structure showed that if a metal tip

is used as a probe, hollow sites possess weaker attraction for the relatively larger d

that brings the bright spots to the hollow sites. However, for the lower tip position,

an image contrast inversion was observed, meaning that the honeycomb lattice is

brighter than the hollow site. Figure 5.8 shows how the frequency shift (∆f) changes

75

Figure 5.8: ∆f vs. tip-sample distance plot. Left and right show the results fromthe metal tip (Ir) and the CO-terminated tip, respectively. The metal tip shows theinversion of the image contrast, i.e., carbon is visible when the tip-sample is relativelyclose. Adapted from Ref. [122].

as a function of d when the two different tips are used [122]. For the Ir (metal) tip, if

the tip is positioned closer to the graphene surface, the hollow site interacts strongly

with the tip, which results in the bright spots at the carbon sites. However, for the

CO tip with oxygen atom towards the sample, bright spots always appear at the

carbon site regardless of the tip-sample distance. We note that if the tip is set to the

relatively higher position, no atomic contrast is observed in both the Ir and the CO

tip.

In order to compare our results to the experiment, we plotted ∆f , using

Eq. (5.4), as a function of tip-sample distance dts between 2.3 A and 4 A in Fig. 5.9.

This figure shows a certain point, approximately at 2.6 A, that inverts the contrast as

can be seen in Fig. 5.7-(a). For larger dts, however, no significant change is observed

for both the carbon and the hollow site. This is because of the fast decaying behavior

of Vts as we used LDA and did not include van der Waals interactions. This should

not be relevant if we consider only the atomic resolution image. If the tip is far from

76

2 2.5 3 3.5 4 4.5d

ts(Å)

∆f (

arb.

uni

t) HollowCarbon

Figure 5.9: Frequency shift with respect to the tip-graphene distance obtained byEq. (5.10) and Eq. (5.4)

the sample, in which the vdW interaction is dominant, no atomic resolution AFM

images are available.

This comparison implies that our simulation results are similar to the exper-

iment performed by a metal tip. This is consistent with the results that we present

in the next section about small molecules. Based on our observations on defect-free

graphene, we expect that our method predicts a similar contrast pattern obtained by

a metal tip for other graphene structures.

b. Single vacancy and Stone-Wales defect

To our knowledge, there is no atomic resolution AFM experiment currently available

for the graphene defects. Here, we present predictions for the nc-AFM images for a

few simple defect structures, which are frequently found in graphene [119].

A single vacancy can be found in any material. The atomic structure and

their AFM images are shown in Fig. 5.10-(a). There are two different types of single

vacancy structures with D3h and Cs symmetry. Cs symmetry can be found when

the reconstruction occurs around the vacancy [123]. In our simulation, we did not

77

observe the reconstructed configuration after the structural relaxation. The AFM

images were simulated with dts at 3 A, and we maintained this tip-sample distance

for all of the defect simulations. In Fig. 5.10-(a), the brightest spot is observed at the

vacancy site that has a pattern with a 3-fold symmetry.

Figure 5.10-(b) shows the Stone-Wales (SW) defect structure and the AFM

image. The SW defect is the simplest defect type that can occur by rearranging the

carbon atoms. As the hollow sites are imaged when we set a larger dts for graphene,

one might expect both pentagonal and heptagonal hollow sites of the SW defect to

appear as a bright spot in AFM. However, our simulation shows that the brightest

spot is located at the pentagonal hollow site rather than the heptagonal hollow site,

which is rather dark. This contrast difference between the pentagonal and the hep-

tagonal hollow sites should be useful in identifying the orientation of the SW defect

configuration.

Although both the single vacancy and the SW defect AFM images show the

bright spots at the defect hollow sites, the underlying physics may be different. In

order to explain our results, we composed a contour plot for the electron density at

1.5 A, 2.5 A, and 3.5 A distances from the graphene surface, and they are shown in

Fig. 5.11. Because of the conjugated π bond electrons, the electron density plot looks

like the honeycomb. For the single vacancy, however, higher electron density around

the single vacancy site is observed at 2.5 A and 3.5 A heights. For the single vacancy

structure with the D3h symmetry, previous DFT calculations showed the existence

of localized π orbitals lying on the single vacancy site [123], which looks very similar

to Fig. 5.11-(c) [124]. For the SW defect, the electron density is higher around the

heptagonal rings as shown in Fig. 5.11-(f).

In the previous section, we observed that the high electron density (π bond

78

of the carbon backbone) causes the bright spot when the tip is at the lower position.

Once the tip is positioned farther from the sample, hollow sites become less attractive

than carbon sites as the Pauli repulsion is no longer dominant at the carbon sites.

The same explanation can be applied for the AFM images of the defect structure.

Even though dts is set to 3 A in both cases, for the single vacancy case, the tip experi-

ences less attractive interaction owing to the high electron density. This explains the

observation of the bright spot at the vacancy site. For the SW defect, on the other

hand, the electron density is not found to be higher at any of the hollow sites. In this

case, the AFM image shows bright spots in which the electron density is lower than

other regions. This bright spot with low density is the hollow site. This explains why

our image shows the brightest spots at the pentagonal hollow site.

5.4 Small molecules

Remarkable advances in the AFM technique have enabled us to see very de-

tailed atomic structure of small molecules. Not only the bond order differences, but

also intermolecular hydrogen bonds have been detected by this technique [125, 126,

127]. The exact mechanism of how this technique, specifically with the CO func-

tionalized tip, achieves such a high resolution is still controversial. To address this

question, we performed AFM simulations for a few planar molecules. In particular,

we introduced the CO molecule as a model for the tip to compare two different AFM

simulation methods.

5.4.1 Computational details

All calculations we performed are based on the real-space pseudopotential

method. We utilized both the local density approximation (LDA) by Ceperley and

79

Figure 5.10: nc-AFM simulation results for two graphene defect structures. The tip-sample turning point is set to 3 A based on our results from the previous section.Yellow dots indicate the carbon atoms around the defects.

80

Figure 5.11: Electron density plots for the single vacancy (a)-(c), and the Stone-Walesdefect (d)-(f). From the left column to the right column, isosurfaces are taken fromthe graphene surface at 1.5 A, 2.5 A, and 3.5 A distances from the graphene sheet.

81

Alder [5], and the generalized gradient approximation (GGA) by Perdew-Burke-

Ernzerhof (PBE) [86] for the exchange correlation functional. For the GGA func-

tional, we added van der Waals force by following Tkatchenko-Scheffler scheme [128]

that does not require any preliminary determined parameters.2 Pseudopotentials

were generated with the same valence electron configuration and the cutoff radii as

we reported in the previous sections. Grid spacing was chosen to be 0.3 a.u. and all

structures were relaxed until the maximum force is less than 0.004 Ry/a.u. For the

tip model, we used only a CO molecule for a model for the CO-functionalized tip.

Several theoretical AFM studies have used the CO molecule attached to the metal

clusters such as Cu2CO. Our test for the choice of the tip finds no noticeable change

of the contrast in the AFM image between the Cu2CO and the CO tip models.

To simulate AFM images for the small molecules, we used a different equation

to calculate frequency shift (∆f). Most high resolution AFM experiments take the

measurement with a tip that oscillates with a very small amplitude [129, 126, 130].

To make our simulations more similar to them, we used Eq. 5.4, which is based on

the small amplitude oscillating motion of the tip.

5.4.2 Results and discussion

The choice of molecules was made from the recent nc-AFM experiment found

in the literature [127, 126]. Our first case is an 8-hydroxyquinoline (8-hq) molecule

(Fig. 5.12-(a)). This molecule was deposited on the Cu(111) surface either as an

individual molecule or randomly assembled aggregates [127]. One significant feature

of Zhang et al.’s work [127] is capturing the intermolecular hydrogen bond. Although

2This scheme was implemented in PARSEC by Ido Azuri in Prof. Leeor Kronik’s group atWeizmann Institute of Science.

82

this is a phenomenal observation, the individual molecule itself also possesses an

interesting feature in their AFM image. Figure 5.12-(a) shows the relaxed structure of

the 8-hq molecule. Fig. 5.12-(b) shows the AFM experiment adapted from Ref. [127].

One noticeable difference between the relaxed structure and measured AFM image

appears around the nitrogen atom (indicated as a blue circle in Fig. 5.12-(b)). The

relaxed structure shows that the N atom is slightly displaced toward to the O atom,

but the change remains relatively small. In the experimental AFM image, we make

two observations: the -OH group is almost invisible and the right side of the molecule

is vertically stretched. To investigate if the adsorbed 8-hq molecule goes through a

substantial structural relaxation, such as a distorsion, due to the Cu substrate, we

simulated AFM images with the 8-hq molecule without substrate, and the tip is set to

3.4 A above from the molecule using structure shown in (a). Our simulation images

are shown in Fig. 5.12-(d) which (d) was created by calculating force using Eq. 5.10,

and Fig. 5.12-(e), which was obtained by using the CO molecule as a tip model. In

this case, the frequency shift is calculated by using the equation∂2IE∂z2

where IE is

an interaction energy calculated with IE = E(8-hq+CO)-[(E(8-hq)+E(CO)] [131].

Both simulation methods resulted in the slightly stretched bright spots around

the N atom (toward to the H atom in the -OH group), and did also not image the

-OH group. In the previous section, we explained that the electron density is relevant

to the AFM image. The electron density at 3.4 A is shown in Fig 5.12-(c). This figure

confirms that the electron density contour plot shows a similar contrast to the AFM

images obtained from both of our simulation results.

The slightly elongated and shortened bond lengths have been reported in other

nc-AFM experiments as well. Gross and coworkers [126] reported that the bond or-

der, which is closely related to the bond length, of organic molecules such as C60 can

83

Figure 5.12: (a) 8-hydroxyquinoline molecule. (b) AFM experiment from Ref. [127].(c) Electron density contour plot at 3.4 A above from the molecule. (d) SimulatedAFM image without the explicit model for the tip. Tip height is set to 3.4 A. (e)Simulated frequency shift map by using CO tip. Tip height is set to 3.4 A.

84

be discriminated by nc-AFM. We chose one of the planar molecules from Ref. [126],

Dibenzo(cd,n)naphtho(3,2,1,8-para)perylene (DBNP), which is shown in Fig. 5.13-

(a). We performed the AFM simulations with two different methods, with and with-

out the AFM tip. Fig. 5.13-(d) and (e) present the AFM images simulated by without

the tip and with the CO tip, respectively. Even though both simulation methods de-

livered similar image contrast for the 8-hq molecule, in this DBNP case, we were

not able to observe the C-C bond in Fig. 5.13-(d) without the tip model. Interest-

ingly enough, this image shows a similar contrast obtained by the Xe tip experiment

(Fig. 5.13-(b)) [129]. Using the Xe tip, the C-C bonds are hardly visible unless it is

on the edge of the molecule. In fact, these results can be understood in the same

way as illustrated in graphene. In the graphene simulation, our method showed a

similar character to the Ir tip (metal tip) rather than the CO tip. Hence, we reaffirm

that our method that excludes an explicit model for the tip would bring comparable

results that we can expect from the metal tip.

5.5 Summary

In this chapter, we performed noncontact atomic force microscopy simulations.

We introduced an efficient method that can greatly reduce the computational costs

and eliminate the uncertainty of the tip morphology. Using this method, only one

fully self-consistent calculation is required to compose nc-AFM images. We tested our

method on various systems, from a semiconducting surface to hydrocarbon molecules.

We showed that our method can be applied to predict AFM images for the systems

in which the tip does not form a chemical bond with the substrate. For the carbon-

based systems such as graphene and organic molecules, our method is found to be

similar to the nc-AFM image obtained by a metal tip.

85

Figure 5.13: (a) Dibenzo(cd,n)naphtho(3,2,1,8-para)perylene molecule. (b)-(c) AFMexperiment from Ref. [129]. CO tip provides much higher resolution for C-C bondthan the Xe tip. (d) Simulated AFM image without the explicit model for the tip.Tip height is set to 3.4 A. (e) Simulated frequency shift map by using the CO tip.Tip height is set to 3.4 A. (f) Electron density contour plot at 3.4 A above from themolecule.

86

Bibliography

[1] P. Hohenberg and W. Kohn. Phys. Rev., 136:B864, 1964.

[2] W. Kohn and L. J. Sham. Phys. Rev., 140:A1133, 1965.

[3] K. Burke. J. Chem. Phys., 136:150901, 2012.

[4] R. M. Martin. Electronic Structure: Basic Theory and Practical Methods.

CAMBRIDGE, 2004.

[5] D. M. Ceperley and B. J. Alder. Phys. Rev. Lett., 45:566, 1980.

[6] J. P. Perdew and A. Zunger. Phys. Rev. B, 23:5048, 1981.

[7] N. Troullier and J. L. Martins. Phys. Rev. B, 43:1993, 1991.

[8] J. R. Chelikowsky. J. Phys. D: Appl. Phys., 33:R33–R50, 2000.

[9] B. Fornberg. Mathematics of compuatation, 51:699, 1988.

[10] J. R. Chelikowsky, N. Troullier, and Y. Saad. Phys. Rev. Lett., 72:1240, 1994.

[11] Y. Saad, J. R. Chelikowsky, and S. M. Shontz. SIAM Review, 52:3, 2010.

[12] Y. Zhou, Y. Saad, M. L. Tiago, and J. R. Chelikowsky. Phys. Rev. B,

74:066704, 2006.

[13] K. H. Khoo, M. Kim, G. Schofield, and J. R. Chelikowsky. Phys. Rev. B,

82:064201, 2010.

87

[14] Y. Zhou, Y. Saad, M. L. Tiago, and J. R. Chelikowsky. J. Comp. Phys.,

219:172, 2006.

[15] Y. Zhou, Y. Saad, M. L. Tiago, and J. R. Chelikowsky. Phys. Rev. E,

74:066704, 2006.

[16] A. Polman, D. C. Jacobson, D. J. Eaglesham, R. C. Kistler, and J. M. Poate.

J. Appl. Phys., 70:3778, 1991.

[17] J. Shieh, W. C. Yu, J. Y. Huang, C. K. Wang, B. T. Dai, H. Y. Jhan, C. W.

Hsu, H. C. Kuo, F. L. Yang, and C. L. Pan. Appl. Phys. Lett., 94:241108,

2009.

[18] Y. N. Xu and W. Y. Ching. Phys. Rev. B, 44:11048, 1991.

[19] F. Mauri, A. Pasquarello, B. G. Pfrommer, Y. G. Yoon, and S. G. Louie. Phys.

Rev. B, 62:R4786, 2000.

[20] S. von Alfthan, A. Kuronen, and K. Kaski. Phys. Rev. B, 68:073203, 2003.

[21] J. P. Rino, I. Ebbsjo, R. Kalia, A. Nakano, and P. Vashishta. Phys. Rev. B,

47:3053, 1993.

[22] T. Uchino, M. Takahashi, and T. Yoko. Phys. Rev. B, 62:2983, 2000.

[23] J. Sarnthein, A. Pasquarello, and R. Car. Phys. Rev. B, 52:12690, 1995.

[24] J. Sarnthein, A. Pasquarello, and R. Car. Phys. Rev. Lett., 74:4682, 1995.

[25] B. B. Karki, D. Bhattarai, and L. Stixrude. Phys. Rev. B, 76:104205, 2007.

[26] M. M. G. Alemany and J. R. Chelikowsky. Phys. Rev. B, 68:054206, 2003.

88

[27] J. R. Chelikowsky, J. J. Derby, V. V. Godlevsky, M. Jain, and J. Y. Raty.

Journal of Physics: Condensed Matter, 13:R817, 2001.

[28] K. Vollmayr, W. Kob, and K. Binder. Phys. Rev. B, 54:15808, 1996.

[29] K. H. Khoo, T.-L. Chan, M. Kim, and J. R. Chelikowsky. Phys. Rev. B,

84:214203, 2011.

[30] L. Kronik, A. Makmal, M. L. Tiago, M. M. G. Alemany, M. Jain, X. Huang,

Y. Saad, and J. R. Chelikowsky. phys. stat. sol., 243:1063, 2006.

[31] N. Binggeli, J. L. Martins, and J. R. Chelikowsky. Phys. Rev. Lett., 68:2956,

1992.

[32] M. M. G. Alemany, R. C. Longo, L. J. Gallego, D. J. Gonzalez, L. E. Gonzalez,

M. L. Tiago, and J. R. Chelikowsky. Phys. Rev. B, 76:214203, 2007.

[33] J. C. Tully, G. H. Gilmer, and M. Shugard. J. Chem. Phys., 71:1630, 1989.

[34] L. Martin-Samos, Y. Limoge, J.-P. Crocombette, G. Roma, N. Richard, E. Anglada,

and E. Artacho. Phys. Rev. B, 71:014116, 2005.

[35] A. Carre, J. Horbach, S. Ispas, and W Kob. Europhysics Letter, 82:17001,

2008.

[36] A. Einstein. Annalen der Physik, 17:549, 1905.

[37] J. R. Silva, D. G. Pinatti, C. E. Anderson, and M. L. Rudee. Philosophical

Magazine, 31:712, 1975.

89

[38] S. Susman, K. J. Volin, D. L. Price, M. Grimsditch, J. P. Rino, R. K. Kalia,

P. Vashishta, G. Gwanmesia, Y. Wang, and R. C. Liebermann. Phys. Rev. B,

43:1194, 1991.

[39] P. A. V. Johnson, A. C. Wright, and R. N. Sinclair. Journal of Non-Crystalline

Solids, 58:109, 1983.

[40] E. Dupree and R. F. Pettifer. Nature, 308:523, 1984.

[41] P. G. Coombs, J. F. D. Natale, P. J. Hood, E. K. McElfresh, R. S. Wortman,

and J. F. Schackelford. Philos. Mag. Lett., 51:L39, 1985.

[42] D. Ceresoli, M. Bernasconi, S. Iarlori, M. Parrinello, and E. Tosatti. Phys.

Rev. Lett., 84:3887, 2000.

[43] D. R. Hamann. Phys. Rev. B, 55:14784, 1997.

[44] T. Uchino, Y. Kitagawa, and T. Yoko. Phys. Rev. B, 61:234–240, 2000.

[45] G. Boureau and S. Carniato. Solid State Comm., 98:485–487, 1996.

[46] G. Pacchioni and G. Ierano. Phys. Rev. B, 56:7304, 1997.

[47] N. Capron, S. Carniato, A. Lagraa, G. Boureau, and A. Pasturel. J. Chem.

Phys., 112:9543, 2000.

[48] R. D. Bendale and L. L. Hench. Surface Science, 338:322, 1995.

[49] B. A. Morrow and I. A. Cody. J. Phys. Chem., 80:1995, 1976.

[50] T. A. Michalske and B. C. Bunker. J. Appl. Phys., 56:2686, 1984.

[51] J. M. Carpenter and D. L. Price. Phys. Rev. Lett., 54:441, 1985.

90

[52] A. Pasquarello and R. Car. Phys. Rev. Lett., 80:5145, 1998.

[53] B. Fischer, R. A. Pollak, T. H. DiStefano, and W. D. Grobman. Phys. Rev.

B, 15:3139, 1977.

[54] J. R. Chelikowsky and M. Schuter. Phys. Rev. B, 15:4020, 1977.

[55] R. B. Laughlin, J. D. Joannopoulos, and D. J. Chadi. Phys. Rev. B, 20:5228,

1979.

[56] S. T. Pantelides and W. A. Harrison. Phys. Rev. B, 13:2667, 1976.

[57] V. I. Klimov, S. A. Ivanov, J. Nanda, M. Achermann, I. Bezel, J. A. McGuire,

and A. Piryatinski. Nature, 447:441, 2007.

[58] X. Michalet, F. F. Pinaud, L. A. Bentolila, J. M. Tsay, S. Doose, J. J. Li,

G. Sun-daresan, A. M. Wu, S. S. Gambhir, and S. Weiss. Science, 307:538,

2005.

[59] I. Gur, N. A. Fromer, M. L. Geier, and A. P. Alivisatos. Science, 310:462, 2005.

[60] T. Takagahara and K. Takeda. Phys. Rev. B, 46:15578, 1992.

[61] A. Stadler. Materials, 5:661, 2012.

[62] R. G. Gordon. MRS bulletin, August2000:52, 2000.

[63] M. K. Singh, M. C. Mathpal, and A. Agarwal. Chem. Phys. Lett., 536:87,

2012.

[64] J. Suffner, P. Agoston, J. K., and H. Hahn. J. Nanopart. Res., 12:2579, 2010.

91

[65] D. G. Stroppa, L. A. Montoro, A. Beltran, T. G. Conti, R. O. da Silva, J. An-

dres, E. Longo, E. R. Leite, and A. J. Ramirez. J. Am. Chem. Soc., 131:14544,

2009.

[66] X. Xu, J. Zhuang, and X. Wang. J. Am. Chem. Soc., 130:12527, 2008.

[67] C.-T. Liu, W.-H. Lee, and T.-L. Shih. J. Nanotechnology, 2012:710908, 2012.

[68] H.-X. Deng, S.-S. Li, and J. Li. J. Phys. Chem. C, 114:4841, 2010.

[69] D. V. Melnikov and J. R. Chelikowsky. Phys. Rev. Lett., 92:046802, 2004.

[70] H. Kwak. Size-dependent properties of impurity doped semiconductor nanos-

tructures. PhD thesis, The University of Minnesota, 2009.

[71] P. D. Borges, L. M. R. Scolfaro, H. W. Leite Alves, and E. F. da Silva Jr.

Theor. Chem. Acc., 126:39, 2010.

[72] E. J. H. Lee, C. Ribeiro, T. R. Giraldi, E. Longo, and E. R. Leite. Appl. Phys.

Lett., 84:1745, 2004.

[73] M. Leslie and M. J. Gillan. J. Phys. C: Solid State Phys., 18:973, 1985.

[74] E. J. H. Lee, C. Ribeiro, T. R. Giraldi, E. Longo, E. R. Leite, and J. A. Varela.

Appl. Phys. Lett., 84:1745, 2004.

[75] G. M. Dalpian and J. R. Chelikowsky. Phys. Rev. Lett., 96:226802, 2006.

[76] T.-L. Chan, M. L. Tiago, E. Kaxiras, and J. R. Chelikowsky. Nano Lett., 8:596,

2008.

[77] S. Lany and A. Zunger. Phys. Rev. B, 78:235104, 2008.

92

[78] O. I. Velikokhatnyi and P. N. Kumta. Physica B, 406:471, 2011.

[79] U. zum Felde, M. Haase, and H. Weller. J. Phys. Chem. B, 104:9388, 2000.

[80] D. G. Stroppa, L. A. Montoro, A. Beltran, T. G. Conti, R. O. da Silva,

J. Andres, E. R. Leite, and A. J. Ramirez. Chem. Eur. J., 17:11515, 2011.

[81] N. S. Bobbitt, M. Kim, N. Marom, N. Sai, and J. R. Chelikowsky. (unpub-

lished).

[82] K. Nakataa and A. Fujishima. Journal of Photochemistry and Photobiology C:

Photochemistry Reviews, 13:169, 2012.

[83] H.-J. Zhai L.-S. Wang. Chem. Phys. Lett., 500:185, 2010.

[84] H.-J. Zhai and L.-S. Wang. J. Am. Chem. Soc., 129:3022, 2007.

[85] R. Gehrke and K. Reuter. Phys. Rev. B, 79:085412, 2009.

[86] J. P. Perdew, K. Burke, and M. Ernzerhof. Phys. Rev. Lett., 77:3865, 1996.

[87] V. Blum, R. Gehrke, F. Hanke, P. Havu, V. Havu, X. Ren, K. Reuter, and

M. Scheffler. Computer Physics Communications, 180:2175, 2009.

[88] N. Marom, M. Kim, and J. R. Chelikowsky. Phys. Rev. Lett., 108:106801,

2012.

[89] L. Hedin. Phys. Rev., 139:A796–A823, 1965.

[90] M. S. Hybertsen and S. G. Louie. Phys. Rev. B, 34:5390–5413, 1986.

[91] C. Adamo and V. Barone. J. Chem. Phys., 110:6158, 1999.

93

[92] L. Kronik, R. Fromherz, E. Ko, G. Gantefor, and J. R. Chelikowsky. Nature

Materials, 1:49, 2002.

[93] G. Binnig, H. Rohrer, Ch. Gerber, and E. Weibel. Phys. Rev. Lett., 49:57,

1982.

[94] G. Binnig, C. F. Quate, and Ch. Gerber. Phys. Rev. Lett., 56:930, 1986.

[95] W. R. Bowen and N. Hilal, editors. Atomic Force Microscopy in Process Engi-

neering. Elsevier, 2009.

[96] J. B. Pethica and W. C. Oliver. Phys. Scri. T, 19:61, 1987.

[97] F. Ohnesorge and G. Binnig. Science, 260:1451, 1993.

[98] Y. Martin, C. C. Williams, and H. K. Wickramasinghe. J. Appl. Phys.,

61:4723, 1987.

[99] T. R. Albrecht, P. Trutter, D. Horne, and D. Rugar. J. Appl. Phys., 69:668,

1991.

[100] T.-L. Chan, C. Z. Wang, K. M. Ho, and J. R. Chelikowsky. Phys. Rev. Lett.,

102:176101, May 2009.

[101] S. Morita, R. Wiesendanger, and E. Meyer, editors. Noncontact Atomic Force

Microscopy. Springer, 2002.

[102] G. H. Enevoldsen. PhD thesis, University of Aarhus, 2007.

[103] W. Melitz, J. Shen, A. C. Kummel, and S. Lee. Surface Science Reports, 66:1,

2011.

94

[104] R. Garcia and R. Perez. Surface Science Reports, 47:197–301, 2002.

[105] F. J. Giessibl. Phys. Rev. B, 56:16010, 1997.

[106] F. J. Giessibl. Appl. Phys. Lett., 78:123, 2001.

[107] S. G. Louie, S. Froyen, and M. L. Cohen. Phys. Rev. B, 26:1738, 1982.

[108] H. Kim and J. R. Chelikowsky. Surf. Sci., 409:435, 1998.

[109] J. L. A. Alves, J. Hebenstreit, and M. Scheffler. Phys. Rev. B, 44:6188, 1991.

[110] T. Ohno and K. Shiraishi. Phys. Rev. B, 42:11194, 1990.

[111] H. J. Monkhorst and J. D. Pack. Phys. Rev. B, 13:5188, 1976.

[112] A. Schwarz, W. Allers, U. D. Schwarz, and R. Wiesendanger. Phys. Rev. B,

61:2837, 2000.

[113] N. Uehara, H. Hosoi, K. Sueoka, and K. Mukasa. Nanotechnology, 15:S97,

2004.

[114] S. H. Ke, T. Uda, R. Perez, I. Stich, and K. Terakura. Phys. Rev. B, 60:11631–

11638, 1999.

[115] S. H. Ke, T. Uda, I. Stich, and K. Terakura. Phys. Rev. B, 63:245323, 2001.

[116] A. S. Foster, A. Y. Gal, J. M. Airaksinen, O. H. Pakarinen, Y. J. Lee, J. D.

Gale, A. L. Shluger, and R. M. Nieminen. Phys. Rev. B, 68:195420, 2003.

[117] A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, and A. K.

Geim. Rev. Mod. Phys., 81:109, 2009.

95

[118] S. Das Sarma, S. Adam, E. H. Hwang, and E. Rossi. Rev. Mod. Phys., 83:407,

2011.

[119] F. Banhart, J. Kotakoski, and A. V. Krasheninnikov. ACS NANO, 5:26, 2011.

[120] M. J. Allen, V. C. Tung, and R. B. Kaner. Chem. Rev., 110:132, 2010.

[121] M. Ondracek, P. Pou, V. Rozsıval, C. Gonzalez, P. Jelınek, and R. Perez. Phys.

Rev. Lett., 106:176101, 2011.

[122] M. P. Boneschanscher, J. van der Lit, Z. Sun, I. Swart, P. Liljeroth, and D. Van-

maekelbergh. ACS NANO, 6:10216, 2012.

[123] H. Amara, S. Latil, V. Meunier, Ph. Lambin, and J.-C. Charlier. Phys. Rev.

B, 76:115423, 2007.

[124] H. Amara, S. Latil, V. Meunier, Ph. Lambin, and J.-C. Charlier. Phys. Rev.

B, 76:115423, 2007.

[125] D. G. de Oteyza, P. Gorman, Y. C. Chen, S. Wickenburg, A. Riss, D. J. Mow-

bray, G. Etkin, Z. Pedramrazi, H. Z. Tsai, A. Rubio, M. F. Crommie, and F. R.

Fischer. Science, 340:1434, 2013.

[126] L. Gross, F. Mohn, N. Moll, B. Schuler, A. Criado, E. Guitian, D. Pena,

A. Gourdon, and G. Meyer. Science, 337:1326, 2012.

[127] J. Zhang, P. Chen, B. Yuan, W. Ji, Z. Cheng, and X. Qiu. Science, 342:611,

2013.

[128] A. Tkatchenko and M. Scheffler. Phys. Rev. Lett., 102:073005, 2009.

96

[129] F. Mohn, B. Schuler, L. Gross, and G. Meyer. Appl. Phys. Lett., 102:073109,

2013.

[130] L. Gross, F. Mohn, N. Moll, P. Liljeroth, and G. Meyer. Science, 325:1110,

2009.

[131] C.-S. Guo, M. A. Van Hove, R.-Q. Zhang, and C. Minot. Langmuir, 26:16271,

2010.

[132] C. Barth, A. S. Foster, C. R. Henry, and A. L. Shluger. Advanced Materials,

23:477, 2011.

[133] J. K. Wassei and R. B. Kaner. Materials Today, 13:52, 2010.

97

Vita

Minjung Kim was born in Seoul, South Korea. After graduating from Seoul

Arts High School, she entered Seoul National University where she received Bachelor

of Science degree in Chemical and Biological Engineering with Physics minor. After

completing her undergraduate study, she entered graduate school at The University

of Texas at Austin in August 2008. She joined Professor James R. Chelikowsky’s

group in October 2008 and continued her work until May 2014.

Permanent email: [email protected]

This dissertation was typeset with LATEX† by Minjung Kim.

†LATEX is a document preparation system developed by Leslie Lamport as a special version ofDonald Knuth’s TEX Program.

98