184
Scanning Probe Microscopy Study of Molecular Self Assembly Behavior on Graphene Two- dimensional Material Yanlong Li Dissertation submitted to the faculty of the Virginia Polytechnic Institute and State University in partial fulfillment of the requirements for the degree of Doctor of Philosophy In Physics James R. Heflin, Chair Chenggang Tao Hans Robinson Shengfeng Cheng December 9, 2019 Blacksburg, VA Keywords: Scanning Tunneling Microscope (STM), Molecular Self Assembly, Atomic Force Microscope (AFM), Graphene, 2D materials Copyright 2019

Scanning Probe Microscopy Study of Molecular Self Assembly

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Scanning Probe Microscopy Study of Molecular Self Assembly Behavior on Graphene Two-

dimensional Material

Yanlong Li

Dissertation submitted to the faculty of the Virginia Polytechnic Institute and State University in

partial fulfillment of the requirements for the degree of

Doctor of Philosophy

In

Physics

James R. Heflin, Chair

Chenggang Tao

Hans Robinson

Shengfeng Cheng

December 9, 2019

Blacksburg, VA

Keywords: Scanning Tunneling Microscope (STM), Molecular Self Assembly, Atomic Force

Microscope (AFM), Graphene, 2D materials

Copyright 2019

Scanning Probe Microscopy Study of Molecular Self Assembly Behavior on Graphene Two-

dimensional material

Yanlong Li

Academic Abstract

Graphene, one-atom-thick planar sheet of carbon atoms densely packed in a honeycomb

crystal lattice, has grabbed appreciable attention due to its exceptional electronic, mechanical

and optical properties. Chemical functionalization schemes are needed to integrate graphene

with the different materials required for potential applications. Molecular self-assembly

behavior on graphene is a key method to investigate the mechanism of interaction between

molecules and graphene and the promising applications related to molecular devices. In this

thesis, we report the molecular self-assembly behavior of phenyl-C61-butyric acid methyl ester

(PCBM), C60, perylenetetracarboxylic dianhydride (PTCDA) and Gd3N@C80 on flat and rippled

graphene 2D material by the experimental methods of scanning tunneling microscope (STM)

and atomic force microscope (AFM) and by the theoretical method of density functional theory

(DFT). We found that molecules form ordered structures on flat graphene, while they form

disordered structure on rippled graphene. For example, PCBM forms bilayer and monolayer

structures, C60 and Gd3N@C80 form hexagonal close packed (hcp) structure on flat graphene and

PTCDA forms herringbone structure on flat graphene surface. Although C60 and Gd3N@C80 both

form hcp structure, C60 forms a highly ordered hcp structure over large areas with little defects

and Gd3N@C80 forms hcp structure only over small areas with many defects. These differences

of structure that forms on flat graphene is mainly due to the molecule-molecule interactions

and the shape of the molecules. We find that the spherical C60 molecules form a quasi-

hexagonal close packed (hcp) structure, while the planar PTCDA molecules form a disordered

herringbone structure. From DFT calculations, we found that molecules are more effected by

the morphology of rippled graphene than the molecule-molecule interaction, while the

molecule-molecule interaction plays a main role during the formation process on flat graphene.

The results of this study clearly illustrate significant differences in C60 and PTCDA molecular

packing on rippled graphene surfaces.

Scanning Probe Microscopy Study of Molecular Self Assembly Behavior on Graphene Two-

dimensional material

Yanlong Li

General Audience Abstract

As the first physical isolated two-dimensional (2D) material, graphene has attracted exceptional

scientific attention. Due to its impressive properties including high carrier density, flexibility and

transparency, graphene has numerous potential applications, such as solar cell, sensors and

electronics. 2D molecular self-assembly is an area that focuses on organization and interaction

between self-assembly behaviors of molecules on surface. Graphene is an excellent substrate

for the study of molecular self-assembly behavior, and study of molecular study is very

important for graphene due to potential applications of molecules on graphene. In this thesis,

we present investigations of the molecular self-assembly of PCBM, C60, PTCDA and Gd3N@C80

on graphene substrate.

First, we report the two types of bilayer PCBM configuration on HOPG with a step height of 1.68

nm and 1.23 nm, as well as two types of monolayer PCBM configuration with a step height of

0.7 nm and 0.88 nm, respectively. On graphene, PCBM forms one type of PCBM bilayer with a

step height of 1.37 nm and one type of PCBM monolayer with a step height of 0.87 nm. By

building and analyzing the models of PCBM bilayers and monolayers, we believe the main

differences between two configurations of PCBM bilayer and monolayer is the tilt angle

between PCBM and HOPG, which makes type I configuration the higher molecule density and

binding energy.

Secondly, we report the investigation of self-assembly behaviors of C60 and PTCDA on flat

graphene and rippled graphene by experimental scanning tunneling microscope (STM) and

theoretical density functional theory (DFT). On flat graphene, C60 forms hexagon close pack

(hcp) structure, while PTCDA forms herringbone structure. On rippled graphene, C60 forms

quasi-hcp structure while PTCDA forms disordered herringbone structure. By DFT calculation,

we study the effect of graphene curvature on spherical C60 and planar PTCDA.

Finally, we report a STM study of a monolayer of Gd3N@C80 on graphene substrate. Gd3N@C80

forms hcp structure in a small domain with a step height of 0.88 nm and lattice constant of 1.15

nm. According to our DFT calculation, for the optimal organization of Gd3N@C80 and graphene,

the gap between Gd3N@C80 and graphene is 3.3 Å and the binding energy is 0.95 eV. Besides,

the distance between Gd3N@C80 and Gd3N@C80 is 3.5 Å and the binding energy is 0.32 eV.

VI

This thesis is dedicated

To my parents Bingyan Li (李炳炎) and Hanying Zhou (周含英)

To my sister Hui Li (李蕙)

To my girlfriend Chen Song (宋晨)

VII

Acknowledgements

I would express gratitude to my research advisor, Professor James R. Heflin, for the academic

guidance and support during my PhD career. He has been my role model as a scientific

researcher. He always encouraged me to work independently and bravely different idea of

experiments. He also gave me rigorous training on experiment and knowledge in physics.

Besides, he provide many opportunities for me to work in different fields and cooperate with

different groups. Overall, he taught me how to be a good physical experimenter.

I am grateful to many faculty members in Virginia Tech, especially my committee members.

Firstly, I would like to thank Professor Chenggang Tao for guiding my research in STM field.

Professor Shengfeng Cheng, as one of my committee members, taught me to learn physics

more intuitively, without heavily depending on mathematics. Professor Hans Robinson gave me

good suggestions on scientific presentation. Professor Greg Liu help in the field of metasurface

and Ag-nanoprism.

A huge thank to Dr. Chuanhui Chen from Professor Tao group for teaching me how to operate

STM, as well as numerous other helps during our collaboration. Besides, I particular want to

thank Dr. Xiaoyang Liu from Professor Dorn group, who did the most DFT calculation in this

thesis. I also need to thank Dr. Moataz Khalifa, who trained me using AFM, and Dr. Jonathan

Metzman showing me the preparation of organic molecules solution.

VIII

Table of Contents Table of Contents ......................................................................................................................... VIII

List of Figures ............................................................................................................................... XIII

Chapter 1: Introduction .................................................................................................................. 1

1.1 2D Materials .............................................................................................................................. 2

1.1.1 Graphene .............................................................................................................................. 2

1.1.2 Other 2D Materials ................................................................................................................ 3

1.2 Molecular Self-assembly ........................................................................................................... 4

1.2.1 Two-Dimensional Self-assembly ........................................................................................... 5

1.2.2 DNA Self-assembly ................................................................................................................ 5

1.2.3 Macromolecular Assembly ................................................................................................... 6

1.2.4 Self-assembly Monolayers (SAMs) ........................................................................................ 7

1.3 Molecular Self-assembly on Graphene ..................................................................................... 8

1.4 Document Organization ........................................................................................................... 8

References .................................................................................................................................... 11

Chapter 2: Literature Review ...................................................................................................... 14

2.1 Introduction and Background ................................................................................................ 14

2.2 Graphene ................................................................................................................................ 15

IX

2.2.1 Synthesis Methods of Graphene ......................................................................................... 16

2.2.1.1 Exfoliation and Cleavage .................................................................................................. 17

2.2.1.2 Epitaxy .............................................................................................................................. 20

2.2.1.3 Chemical Vapor Deposition .............................................................................................. 22

2.2.2 Properties of Graphene ...................................................................................................... 24

2.2.2.1 Single Layer: Tight-binding Theory .................................................................................. 25

2.2.2.2 Single Layer: Properties ................................................................................................... 27

2.2.2.3 Bilayer and Trilayer Graphene ......................................................................................... 31

2.2.3 Applications of Graphene ................................................................................................... 34

2.2.3.1 Graphene Field Emission (FE) .......................................................................................... 34

2.2.3.2 Graphene Field Effect Transistors (FET) ........................................................................... 35

2.2.3.3 Graphene-based Gas and Biological Sensors .................................................................... 37

2.2.3.4 Transparent Electrode ...................................................................................................... 38

2.2.3.5 Batteries ............................................................................................................................ 40

2.3 2D Molecular Self-assembly.................................................................................................... 41

2.3.1 Metal Bonds Molecular Self-assembly ................................................................................ 43

2.3.2 Hydrogen Bonding Molecular Self-assembly ...................................................................... 45

2.3.3 Van der Waals Molecular Self-assembly ............................................................................. 47

2.3.4 Halogen‐halogen Molecular Self-assembly ........................................................................ 49

X

2.4 Molecular Self-assembly on Graphene .................................................................................. 51

2.4.1 PTCDA .................................................................................................................................. 52

2.4.2 C60 ...................................................................................................................................... 55

2.4.3 Phthalocyanines .................................................................................................................. 57

References .................................................................................................................................... 60

Chapter 3: Experimental Methods .............................................................................................. 65

3.1 Introduction to Atomic Force Microscope (AFM) ................................................................... 66

3.1.1 Working Principle: Van der Waals Force ............................................................................. 68

3.1.2 Working Modes .................................................................................................................... 71

3.1.3 Bruker Dimension Icon® AFM .............................................................................................. 75

3.1.4 The Correction of Height of AFM Measurement ................................................................. 77

3.2 Introduction to Scanning Tunneling Microscope (STM) ......................................................... 78

3.2.1 Working Principle: Tunneling Effect .................................................................................... 80

3.2.2 Working Modes .................................................................................................................... 83

3.2.3 Omicron RT® STM ................................................................................................................ 85

3.2.4 The Correction of Height of STM Measurement ................................................................. 88

3.3 Sample Preparation ................................................................................................................ 88

3.3.1 Spin Coating ......................................................................................................................... 88

XI

3.3.2 Physical Vapor Deposition ................................................................................................... 89

References .................................................................................................................................... 90

Chapter 4: Self-Assembled PCBM Bilayers on Graphene and HOPG Examined by AFM and STM

....................................................................................................................................................... 94

4.1 Introduction ............................................................................................................................ 94

4.2 Experimental Methods ........................................................................................................... 96

4.3 Results and Discussion ............................................................................................................ 97

4.3.1 PCBM Bilayer Morphology ................................................................................................... 97

4.3.2 PCBM Monolayer Morphology .......................................................................................... 100

4.3.3 Discussion........................................................................................................................... 103

4.3.4 Thermal Effects .................................................................................................................. 109

4.4 Conclusion ............................................................................................................................. 110

References .................................................................................................................................. 111

Chapter 5: Differences in Self-Assembly of Spherical C60 and Planar PTCDA on Rippled

Graphene Surfaces .................................................................................................................... 114

5.1 Introduction .......................................................................................................................... 114

5.2 Experimental and Computational Methods ......................................................................... 116

5.3 Discussions ............................................................................................................................ 117

XII

5.4 Conclusion ............................................................................................................................. 130

References .................................................................................................................................. 132

Chapter 6: Self-Assembled Gd3N@C80 Monolayer on Graphene Examined by STM ............... 135

6.1 Introduction and Background ............................................................................................... 135

6.2 Experimental and Computational Methods ......................................................................... 139

6.3 Results and Discussions ........................................................................................................ 140

6.4 Conclusion ............................................................................................................................. 149

References .................................................................................................................................. 149

Chapter 7: Conclusion and Future Work ................................................................................... 152

7.1 The Bilayer PCBM Structure Formed on Graphene and HOPG ........................................... 152

7.2 The Ordered of C60 and Disordered Structure of PTCDA Formed on Rippled Graphene ..... 153

7.3 The hcp Structure of Gd3N@C80 Formed on Graphene ........................................................ 154

7.4 Future Work .......................................................................................................................... 155

XIII

List of Figures

Chapter 2

Figure. 2.1. Graphene films. (a) Photograph (in normal white light) of a relatively large multilayer

graphene flake with thickness ∼3 nm on top of an oxidized Si wafer. (b) Atomic force microscope (AFM)

image of 2 μm by 2 μm area of this flake near its edge. Colors: dark brown, SiO2 surface; orange, 3 nm

height above the SiO2 surface. (c) AFM image of single-layer graphene. Colors: dark brown, SiO2 surface;

brown-red (central area), 0.8 nm height; yellow-brown (bottom left), 1.2 nm; orange (top left), 2.5 nm.

Notice the folded part of the film near the bottom, which exhibits a differential height of ∼0.4 nm. (d)

Scanning electron microscope image of one of our experimental devices prepared from few-layer

graphene. (e) Schematic view of the device in (d)……..…………………………………………………………………………..18

Figure. 2.2. Optical images of graphene flakes prepared by the standard exfoliation method and Dr.

Peter Sutter’s modified method. (a and b) Optical microscopy images of typical monolayer to trilayer

graphene prepared by the standard method, including a solvent wash and O2 plasma cleaning of the

substrate followed by graphene transfer. (c and d) Optical microscopy images of two graphene flakes

prepared by Dr. Peter Sutter’s modified method, with O2 plasma clean of the SiO2/Si surface, followed by

contact with graphite-loaded tape, annealing to 100 oC, cooling to room temperature and peel-off)……19

Figure 2.3. LEED patterns from graphite/SiC(0001). The sample was heated several times to successively

higher temperatures. (a) 1050 °C for 10 min. Immediately after oxide removal, showing SiC 1 × 1 pattern

at 177 eV. AES C:Si ratio 1:2. (b) 1100 °C, 3 min. The x3 × x3 reconstruction is seen at 171 eV. AES ratio

1:1.9. (c) 1250 °C, 20 min. 109 eV pattern showing diffracted beams from the 6x3 × 6x3 unit cell.

Examples of first-order SiC and graphite spots are marked. Note the surrounding hexagons of “6 × 6”

spots. AES C:Si ratio 2:1 (∼1 ML graphite). (d) 1400 °C, 8 min. 98 eV LEED pattern. AES ratio 7.5:1 (∼2.5

XIV

ML graphite). (e) STM image of a surface region of the sample described in Figure 1d. Inset: Atomically

resolved region (different sample, similar preparation). (f) UHV-SEM image of a large area of the

Ru(0001) surface after first-layer graphene growth. Inset: Carbon KLL (260.6 eV) UHV scanning Auger

microscopy image, obtained on this sample. (g) AFM image of as-grown graphene on h-BN, the scale bar

is 200 nm……………………………………………………………………………………………………………………………………………….21

Figure. 2.4. (a) SEM images of as-grown graphene films on thin (300-nm) nickel layers and thick (1-mm)

Ni foils (inset). (b) TEM images of graphene films of different thicknesses. (c) Optical image of the grown

graphene transferred from the Ni surface in panel a to another SiO2/Si substrate. (d-e) High-

magnification TEM images showing the edges of film regions consisting of one (d) and three (e)

graphene layers. The cross-sectional view is enabled by the folding of the film edge. The in-plane lattice

fringes suggest local stacking order of the graphene layers.………………………………………………………………..23

Figure. 2.5. (a) Honeycomb lattice and its Brillouin zone. Left: lattice structure of graphene, made from

two interpenetrating triangular lattices (a1 and a2 are the lattice unit vectors, and δ𝑖, i= 1, 2, 3 are the

nearest-neighbor vectors). Right: corresponding Brillouin zone. The Dirac cones are located at the K and

K′ points.37 (b) Electronic dispersion in the honeycomb lattice. Left: energy spectrum (in units of t) for

finite values of t and t′, with t= 2.7 eV and t′=−0.2t. Right: zoom in of the energy bands close to one of

the Dirac points……………………………………………………………………………………………………………………………………..26

Figure. 2.6. (a) STM image (170 nm × 170 nm) for graphene grown on Cu(111). A Moire pattern arising

due to the lattice mismatch between graphene and Cu(111) is visible that continues over the step edges.

(b) Large area STM topographic image of the rippled graphene showing well-defined linear periodic

modulation with a 0.75 nm spatial modulation frequency (Vs = 0.80 V, I = 1.0 nA). (c) STM image showing

several nanotrenches of different orientations, all exhibiting graphene nanoripples over the trenches

with the ripple crests always perpendicular to the trench edges. On the flat regions between the

XV

trenches, a Moiré pattern can be observed. (d) Atomic-resolution STM image of a nanotrench exhibiting

subnanometer graphene ripples. The magnified insets exhibit the honeycomb graphene lattice both

over the flat substrate (bottom right) and the rippled region over the trenches (top left)……….……………28

Figure 2.7. (a) Scanning electron microscopy (SEM) image of a graphene flake spanning an array of

circular holes (scale bar, 3μm). (b) Schematic illustration of nanoindentation on membranes. (c)

Photograph of a 50‐μm aperture partially covered by graphene and its bilayer. The line scan profile

shows the intensity of transmitted white light along the yellow line. Inset shows the sample design: a 20

μm thick metal support structure has apertures 20, 30, and 50 μm in diameter with graphene flakes

deposited over them. (d) Optical image of graphene flakes with one, two, three, and four layers on a

285‐nm thick SiO2‐on‐Si substrate………………………………………………………………………………………………………….30

Figure 2.8. Two-dimensional superconductivity and insulator in a graphene superlattice. (a) Schematic of

a typical twisted bilayer graphene (TBG) device and the four-probe (Vxx, Vg, I and the bias voltage Vbias)

measurement scheme. The stack consists of hexagonal boron nitride on the top and bottom, with two

graphene bilayers (G1, G2) twisted relative to each other in between. The electron density is tuned by a

metal gate beneath the bottom hexagonal boron nitride layer. (b) Fourprobe resistance Rxx = Vxx/I (Vxx

and I are defined in a) measured in two devices M1 and M2, which have twist angles of θ = 1.16° and θ =

1.05°, respectively. The inset shows an optical image of device M1, including the main ‘Hall’ bar (dark

brown), electrical contact (gold), back gate (light green) and SiO2/Si substrate (dark grey). (c) Schematic

of the TBG devices. The TBG is encapsulated in hexagonal boron nitride flakes with thicknesses of about

10–30 nm. The devices are fabricated on SiO2/Si substrates. The conductance is measured with a voltage

bias of 100 μV while varying the local bottom gate voltage Vg. ‘S’ and ‘D’ are the source and drain

contacts, respectively. (d) The band energy E of magic-angle (θ = 1.08°) TBG calculated using an ab initio

tight-binding method. The bands shown in blue are the flat bands that we study…………………..…………….33

XVI

Figure. 2.9. (a) Typical plots of the electron-emission current density (J) as a function of applied electric

field (E) for the graphene film and graphene-powder coating. (b) Corresponding F–N plots………………….35

Figure. 2.10. (a) Schematic of a graphene FET on a Si/SiO2 substrate with a heavily doped Si wafer acting

as a back gate and a gold top gate. (b) SEM micrograph showing a representative graphene top-gated

FET. The top-gate of this device is 1 µm long, with 3 µm spacing between the source–drain contacts. All

electrodes are Cr/Au. (c) Drain current (Id) as a function of source-to-drain voltage (Vsd) for Vgs-top = −0.3

V, −0.8 V, −1.3 V, −1.8 V, −2.3 V and −2.8 V (from bottom to top) for Vgs-back = 40 V. (d) Id as a function of

Vsd for Vgs-top = −0.3 V, −0.8 V, −1.3 V, −1.8 V, −2.3 V and −2.8 V (from bottom to top) for Vgs-back = −40

V……………………………………………………………………………………………………………………………………………………………36

Figure. 2.11. (a) Experimental setup for measurements performed using the graphene CO2 gas sensor.

(b) Time response of the graphene CO2 gas sensor in the presence of 100 ppm CO2, at different

temperatures ………………………………………………………………………………….…………………………………………………….38

Figure. 2.12. (a) Schematic representation of the energy level alignment (top) and the construction of

heterojunction organic solar cell fabricated with graphene as anodic electrode:

graphene/PEDOT/CuPc/C60/BCP/Al. (b) Schematic illustration of the transfer process of CVD‐graphene

onto transparent substrate. (c, d) The plots of current density vs voltage for (c) graphene and (d) ITO

devices under 100 mW cm–2 AM1.5G spectral illumination at different bending angles. Insets show the

experimental setup used in the experiments……………………………………………………………………….………………..39

Figure. 2.13. Lithium insertion/extraction properties of the GNS families. (a) charge/discharge profiles of

(a) graphite, (b) GNS, (c) GNS+CNT, and (d) GNS+C60 at a current density of 0.05 A/g. (b)

Charge/discharge cycle performance of (a) graphite, (b) GNS, (c) GNS+CNT, and (d) GNS+C60……………….41

Figure. 2.14. SAMs of thiols on gold substrate. (b) SAMs of trichlorosilanes on SiO2 substrate.……….…….42

XVII

Figure. 2.15. (a) Array of cobalt clusters following evaporation of 0.14 ML Co on the Au(111) surface at

room temperature. The deposited atoms condense in bilayer islands at the elbow sites of the chevron

reconstruction. The Co dots contain on the average ∼200 atoms representing in situ nano-reservoirs for

the formation of metal-organic complexes with co-deposited carboxylic acids (STM image size 100 × 80

nm2, I =1.3 nA, V = 20 mV). (b) Complexation reaction following deposition of 0.3 ML TPA on a Co array

(0.08 ML, corresponding to ∼120 atoms per island) on Au(111) at room temperature. A minority of

hydrogen-bonded domains (A) coexists with the dominating metal-organic compounds (B) evolving

around residual Co dots (I = 0.6 nA, V = -0.7 V). (b) Array of cobalt clusters following evaporation of 0.14

ML Co on the Au(111) surface at room temperature. The deposited atoms condense in bilayer islands at

the elbow sites of the chevron reconstruction. The Co dots contain on the average ∼200 atoms

representing in situ nano-reservoirs for the formation of metal-organic complexes with co-deposited

carboxylic acids (STM image size 100 × 80 nm2, I = 1.3 nA, V = 20 mV). (c) Fully developed rectangular

metal–organic nanogrid with a Co–TPA stoichiometry of 1:1 following annealing at 330 K. (d) The model

at the right depicts the underlying dicobalt coordination motif with both axial chelating and equatorial

bridging metal center–carboxylate bonds………….………………………………………………………………………………….44

Figure. 2.16. STM topographs of hexagonal networks from (a) trimesic acid (TMA), (b) 1,3,5-

benzenetribenzoic acid (BTB), and (c) 1,3,5-tri(4-carboxyphenylethynyl)-2,4,6-trimethylbenzene

(TCPETMB). All STM topographs are to scale and depict an area of 15 × 15 nm2 . Corresponding models

of a single supramolecular cavity bordered by six molecules are depicted below and demonstrate the

underlying building plan of these isotopological networks with dimers interconnected by 2-fold

hydrogen bonds between carboxylic groups as a structural unit……………………………………………………………46

Figure. 2.17. (a) Occupied state STM image of a 2√3 × 2√3 R 30o C60 domain on Au(111) [7.3 X 7.3 nm,

—2 V bias (sample negative)]. (b) Occupied state image of a 38 X 38 C60 domain on Au(111) (9. 1 X 9. 1

XVIII

nm, —2 V). (c) Filled states (VSB = −1.5V) and (b) empty states STM image (VSB = 1.5V) of close-packed

Sc3N@C80 on Au(111)/mica (It = 0.08 nA). The circled molecules appear bright in filled states and dark in

empty states. Vacancies always show as black holes……………………………………………………………………………..48

Figure. 2.18. The two ordered self-assembly phases of HPBI on Au(111). (a) The chemical structure of

HPBI. Carbon, grey; iodine, red; hydrogen, white. (b) Large-scale STM image of the coexisting α and β

phases. The crystalline axes of the Au substrate are labelled. (c, d) STM images of the pure α (c) and β (d)

phase domains. The white arrows in c designate the herringbone structure in the α phase; the α and β

phase unit cells are represented by red and blue rhombi, respectively. (e, f) Magnified STM images of

the α (e) and β (f) phases. HPBI molecules are labelled by the dashed white and green circles. The

dashed white triangle in e highlights the I–I trimer. The solid yellow circles in f highlight the

adatoms………………………………………………………………………………………………………………………………………………..50

Figure. 2.19. The molecular structure of (a) Phthalocyanine, (b) C60 and (c) PTCDA……………………………….52

Figure. 2.20. (a) PTCDA monolayer on bilayer graphene at T=4.7 K. The shadow-like structure

originates from the SiC interface layer below bilayer graphene. UT=1.5V, IT=3.8pA. (b) Close-up of (a) One

clearly recognizes the assembly of the PTCDA molecules. UT=1V, IT=3.8pA. (c) Monolayer coverage of

PTCDA on epitaxial graphene. (d) Molecular-resolution STM image of the PTCDA monolayer. The PTCDA

molecular structure and unit cell outline are overlaid. The monolayer continuously follows the graphene

sheet over the SiC step edge……………………………………………………………………………………………………….…………53

Figure. 2.21. (a) Calculated geometry configuration of monolayer PTCDA molecules on a

graphene/Pt(111). (b) Experimental STS spectra on one monolayer PTCDA on graphene/Pt(111). (c) Local

Density of States (LDOS) around the transport gap on PTCDA in the DFT calculations for the

PTCDA/graphene/Pt(111) system. Two peaks are clearly resolved both in the experimental and in the

XIX

theoretical spectra whose origin can be ascribed to the HOMO and LUMO of the PTCDA

molecule………………………………………………………………………….…………………………………………………………………….54

Figure. 2.22. (a) STM topographic images of the initial stages of growth of C60 molecules adsorbed on a

submonolayer of epitaxial graphene on SiC. (b) a close-up view of the blue box indicated in (a) displaying

the single vacancy of a C60 molecule and the domain boundary between molecular islands. Images (a and

b) were acquired with I = 20 pA and V = −2 V. (c) Large-area STM topography of substrate commensurate

growth of C60 molecules on G/Ru. Right part is a higher terrace of Ru(0001) surface. (Vs = 3.0 V and I =

0.05 nA) (d) Zoom-in image of the supramolecular structure. The unit cells of the underlying substrate

and molecular lattice are outlined by large and small rhombuses, respectively. (2.0 = V and 0.1 = nA). (e)

High resolution image of bimolecular and trimolecular C60 chains. Within the chains, the C60-C60

intermolecular spacing is ~1.0nm, and the interchain distance, defined as the distance between the

centers of adjacent C60 rows belonging to neighboring chains, is 1.23 nm (Vs= 1.95 V, I= 0.50nA). (f) A

line profile along the close packed orientation as marked with the dashed blue line in (e). (g) Scheme of

an individual C60 molecule preferentially trapped in the Chcp valley at RT and its corresponding STM

image as shown in (i). (h) Scheme of six C60 molecules attached to the trapped C60 as a seed for the

nucleation of monolayer C60 islands; C60-C60 cohesive energy increases. (j) RT freezing of the thermal

motions of C60 in the Chcp valleys once a C60 monolayer is formed. (k) All C60 molecules trapped in Chcp

valleys display a dumbbell shape, aligning along the <1120> directions. The bright lobes in the dumbbell-

shaped correspond to pentagons of the C60 cage at positive sample bias, which suggests C60 orients with

the 6:6 bond (the C-C bond between two carbon hexagons) facing upward, as shown in the right top of

(k)………………… ………………………………………………………………………………………………………………………………………56

Figure. 2.23. (a) High-resolution image (U = 0.1 V, I = 0.5 nA) of three distinct regions of graphene, top,

fcc, and hcp, marked by triangles and dashed and solid hexagons, respectively. (b) STM image (U = −2.0

V, I = 0.05 nA) revealing that molecules first adsorb at the fcc regions. (c) Sequences of STM images (U =

XX

−2.0 V, I = 0.05 nA) of FePc molecules with increasing coverage. (d) Details of the Kagome lattice of FePc.

A trihexagonal tiling is highlighted. The unit cell of the Kagome lattice is marked with blue lines. (e)

Structural model of the Kagome lattice showing molecular orientation disorder. (f) STM image of a close-

packed FePc molecular island in showing a square lattice (indicated by the dashed square). (g) STM

image of a close-packed CoPc molecular island in a showing a square lattice. (h) STM image of a close-

packed CuPc molecular island in a showing a square lattice. (i) STM image of a close-packed F16CoPc

molecular island showing a square lattice. (j) STM images of self-assembled ClAlPc molecular arrays of

the first layer on graphene. The ClAlPc arrays show continuous films across the Cu steps. Insets show the

magnified images (Vtip = 2.0 V, I = 75 pA). (k) The optimized configuration for the F16CuPc/graphene

[(3,4)×(4,3)] system. (l) Total DOS (thick black solid line) and projected DOS on the Pc molecule (blue

solid line), on graphene (thin black solid line), upon F16CuPc adsorption on graphene, and the DOS of the

isolated graphene (black dashed dotted line). The vertical dotted line shows the Fermi level.…..….………59

Chapter 3

Figure. 3.1. Schematic drawing shows the mechanism of imaging mode in AFM……………………………………67

Figure 3.2. Force – distance curve of a Van Der Waals force. The yellow part is the non-contact mode

region, the purple is the contact mode region and the green part is the intermittent contact (tapping)

mode region……………………………………………………………………………………………………………………………….…………70

Figure 3.3. (a) The schematic drawing of contact mode. (b) The schematic drawing of non-contact mode.

(c) The schematic drawing of tapping mode………………………………………………………………………………………….72

Figure 3.4. (a) Resonance curve of a Tapping Mode cantilever above the surface. (b) Resonance curve of

a Tapping Mode cantilever close to the surface. Note that the resonance shifts to lower frequencies and

exhibits a drop in amplitude………………………………………………………………………………………………………………….75

XXI

Figure 3.5. (a) The Dimension Icon® is an AFM system that offers a variety of nanoscale characterization

and manipulation tools. It is equipped with a closed-loop scanner offering great precision for

repositioning the tip on the sample. It has a piezo scanner based on a piezotube. (b) The AFM tip type

that is used in our experiments is NCST® from Nano world with spring constant 7.4 N/m, first longitudinal

resonance frequencies between 120 – 205 kHz. (c) A probe holder that fits on the Dimension Icon®

system. The system employs a spring loaded lever system to hold the tip in place. This holder fits directly

on the piezoelectric scanner………………………………………………………………………………………………………….………76

Figure 3.6. Left 8 images: PeakForce tapping (PFT) mode AFM height and deformation images of surface

nanobubbles on HOPG in water scanned with peak forces of F=0.24, 2.5, 9.7 and 27 nN. The line profile

(cross section) of height and deformation of surface nanobubbles with peak forces F=0.24, 2.5, 9.7 and

27 nN…….………………………………………………………………………………………………………………………………………………77

Figure. 3.7. The schematic diagram of the basic characterizing mechanism of STM………………….………..…79

Figure. 3.8. (a) Schematic illustration of the constant current mode of STM operation. (b) Schematic

illustration of constant height mode of STM operation …………………………………………….………………………..…85

Figure. 3.9. (a) The main body of the STM consists of the main chamber, the manipulator chamber and

the load lock chamber. (b) The main chamber (scanning stage) of the STM. (c) The controller of the STM.

(d) The computer of the STM…………………………………………………………………………………………………………………86

Figure. 3.10. (a) The spin coater that we used in this dissertation. (b) The detailed structure of the

homemade Knudsen cell showing the main components inside the copper shell. 1 is CF flange, 2 is

thermocouple wire, 3 is W heating filament, 4 is glass tube, 5 is ceramic piece, 6 is hollow copper rods

(A, B, C, D), 7 is supporting rods, 8 is feedthrough…………………………………………………………………………………90

XXII

Chapter 4

Figure. 4.1. AFM and STM images of PCBM films spin-coated from PCBM/Chlorobenzene solution on

graphene/Cu and HOPG substrates. (a) STM image of a graphene/Cu substrate, and the inset is the

atomic image of Moiré pattern of graphene on Cu (111) taken from the white square area in (a). (b) AFM

image of a PCBM bilayer on a graphene/Cu substrate from 0.2 mg/ml PCBM solution. (c) AFM image of a

PCBM bilayer on a graphene/Cu substrate from blue square area in (b). (d) A zoomed in AFM image of a

PCBM bilayer on a graphene/Cu substrate (e) Line profile with the height about 1.37 nm. (f) AFM image

of PCBM bilayer on HOPG substrate from 0.5 mg/ml PCBM solution. (g) Line profiles with the heights of

type I (blue line) and type II (red dashed line) indicated in Figure 4.1f. (h) AFM image of 0.15 PCBM

bilayer on HOPG from 0.1 mg/ml PCBM solution; the inset is the atomic image of HOPG taken from the

white square area. In (h), the HOPG step edges are indicated by white double arrows. (i) AFM image of a

PCBM bilayer on HOPG substrate from 1.0 mg/ml solution. (j) Line profile showing the height of type II

(red dashed line) showed in (i).……………………………………………………………………………………………………………..98

Figure. 4.2. AFM images of PCBM films spin-coated from PCBM/Chlorobenzene solution on HOPG

substrates. (a) AFM image of a PCBM bilayer on HOPG substrate from 0.15 mg/ml PCBM solution. (b)

AFM image of a PCBM bilayer on HOPG substrate from 0.5 mg/ml PCBM solution. (c) AFM image of a

PCBM bilayer on HOPG substrate from 0.75 mg/ml PCBM solution. (d) AFM image of a PCBM bilayer on

HOPG substrate from 1.0 mg/ml PCBM solution. (e) AFM image of a PCBM bilayer on HOPG substrate

from 2.0 mg/ml PCBM solution. (f) Coverage vs Concentration plotting based on AFM images…………100

Figure. 4.3. AFM images of PCBM monolayer films spin-coated from PCBM/Chlorobenzene solution on

graphene/Cu and HOPG substrates. (a) AFM image of a PCBM monolayer on graphene/Cu from 0.5

mg/ml PCBM solution after a 30 min 170 oC anneal. (b) AFM image of a PCBM monolayer on a

graphene/Cu substrate from blue square area in Fig. 2(a), and the insert is the line profile for typical

XXIII

PCBM monolayer with height of 0.87 nm. (c) AFM image of a PCBM monolayer on HOPG from 0.5 mg/ml

PCBM solution. (d) Line profile along the line marked in (c) indicates the height of 0.71 nm. (e) AFM

image of a PCBM monolayer on HOPG from 0.5 mg/ml solution. (f) Line profile along the line marked in

(e) shows the height is about 0.88 nm…………………………………………………………………………………………………102

Figure. 4.4. Two schematic model configurations of a PCBM monolayer on a graphene or HOPG

substrate: C atom in PCBM (blue), C atom in the substrate (green), O atom (red), and H atom (yellow).

(a) Top and side views of one model configuration of a PCBM monolayer with height of 0.9 nm (b) Top

and side views of another model configuration of a PCBM monolayer sample of with height of 0.7 nm.

The dashed parallelograms in (a) and (b) indicate unit cells……………………………………………………………….104

Figure. 4.5. Schematic diagrams of type I and type II configurations of a PCBM bilayer on a graphene or

HOPG substrate: C atom in PCBM (blue), C atom in the substrate (green), O atom (red), and H atom

(yellow). (a) Top and side views of type I configuration of the PCBM bilayer. In the side view, the dashed

tilted rectangle highlights a PCBM dimer cell, and the solid rectangle indicates the hydrogen binding

within a PCBM dimer, while the solid circle shows a side interaction between neighboring PCBM dimers.

(b) Top and side views of type II configuration of the PCBM bilayer. In the side view, the dashed tilted

rectangle highlights a PCBM dimer cell, while the solid rectangle indicates the hydrogen binding within a

PCBM dimer…………………………………………………………………………………………………………………………………………107

Figure. 4.6. AFM images of PCBM bilayer and size distributions of holes at different conditions. (a) AFM

image of a PCBM bilayer before annealing. (b) AFM image of a PCBM bilayer after annealing at 140 °C.

(c) AFM image of a PCBM bilayer after annealing at 160 °C. (d) Area distribution histogram of holes

(without PCBM area) obtained from measurements of the area of holes in AFM images of before (red

line) and after annealing at 140 °C (dark red line) and 160 °C (dark blue line)………………………………..……110

XXIV

Chapter 5

Figure 5.1. STM topographical images of planar graphene (labeled as I), 1D-rippled graphene (II) and 2D-

rippled graphene (III) on Cu. (a) Large area STM image of planar graphene (I) and 1D-rippled graphene

(II) showing the linear periodic modulation and the spatial modulation frequencies (Vs = -2.340 V, I =

0.110 nA). (b) High-resolution STM image of 1-D rippled graphene (Vs = -0.340 V, I = 1.900 nA). (c) STM

image of the 1-D rippled graphene, observed from the square region marked in (b), the schematic model

on top of the atomic image shows the ripples along zigzag direction (Vs = -0.280 V, I = 1.900 nA). (d) Line

profile perpendicular to the 1D-rippled graphene (marked as a blue line in (b)) showing the periodic

modulation. (e) STM image of graphene on two different Cu facets, planar graphene (I) and 2D-rippled

graphene (III) (Vs = -2.74 V, I = 0.045 nA). (f) High-resolution STM image of 2D-rippled graphene,

observed from the dashed square region marked in (e) (Vs = -2.600 V, I = 0.068 nA). (g) A schematic

model shows 1D-rippled graphene sheet. (h) Large area STM image of planar graphene (I) and 1D-

rippled graphene (II) (Vs = -1.850 V, I = 0.340 nA). (i) Atomic STM image showing the moiré pattern of

planar graphene, observed from the dashed squared region marked in (h) (Vs = -1.850 V, I = 0.450

nA)……………………………………………………………………………………………………………………………………………………….118

Figure 5.2. STM topography images of one single-twin rippled graphene. (a) STM topographic image of

continuous graphene forming twin wrinkles on Cu(111) (Vs = -1.06 V, I = 0.315 nA). (b) Zoomed-in image

of (a) (Vs = -0.600 V, I = 0.850 nA). (c) Line profile of the single-twin wrinkle of graphene, measured along

the blue line in b. (d)Further zoomed-in STM topographic image of single-twin wrinkle of graphene (Vs =

-0.560 V, I = 0.8750 nA). (e) Atomic STM image of one of the twin wrinkles in Figure (d) showing the

honeycomb structure of graphene (Vs = -0.560 V, I = 0.8750 nA). (f) Schematic model of one side of the

twin wrinkle……………………………………………………………….………………………………………………………………….…….120

XXV

Figure 5.3. STM images of C60 on 1D-rippled graphene (II) and on planar graphene (I). (a) Large area STM

topographic image of the C60 on 1D-rippled graphene showing well-defined linear periodic modulated

ripple (Vs = -2.00 V, I = 0.060 nA). (b) Zoomed-in STM image (measured from the dashed square of (a)) of

C60 on a long periodic graphene ripple (Vs = -2.60 V, I = 0.050 nA). (c) High-resolution image (measured

from the dashed square region of (b)) C60 on 1D-rippled graphene, shows a lattice angle α of 54.1o with a

quasi-hcp pattern (Vs = -2.60 V, I = 0.040 nA). Inset, the corresponding FFT image of (c). (d) A line profile

along the perpendicular direction of the 1D-rippled graphene marked with the blue line in (c) (top), side

view and top view showing the quasi-hcp C60 on 1D-rippled graphene (bottom). (e) Large area STM

image of the C60 on planar graphene with a well-defined moiré pattern on facet I (Vs = -2.65 V, I = 0.046

nA). (f) High-resolution STM image of C60 on planar graphene (measured from the square region of (e)),

showing a lattice angle β of 60.0o and a moiré pattern on facet I (Vs = -2.65 V, I = 0.046 nA). Inset, the

corresponding FFT image of (f)…………………………………………………………………………………………………………….121

Figure. 5.4. STM images of PTCDA on 1D-rippled graphene and on planar graphene. (a) Large area STM

image of PTCDA on 1D-rippled graphene (Vs = -2.51 V, I = 0.042 nA). (b) STM image of PTCDA on 1D-

rippled graphene showing a distorted herringbone pattern (Vs = -2.510 V, I = 0.042 nA). Inset, the

corresponding FFT image of (b). (c) Large area STM image of PTCDA on planar graphene (Vs = 1.800 V, I =

0.030 nA). Inset, the FFT image of the PTCDA herringbone structure on planar graphene. (d) Zoomed-in

STM image of PTCDA on planar graphene; a1 and a2 indicate the short and long lattice vectors of a unit

cell of the PTCDA herringbone pattern (Vs = 1.800 V, I = 0.030 nA). (e) STM images of coexistence of

substable PTCDA structure (purple curved region) and normal PTCDA herringbone structure (Vs = -2.500

V, I = 0.030 nA). (f) STM image of remaining normal PTCDA structure after the substable PTCDA was

removed by STM tip (Vs = -2.500 V, I = 0.030 nA)…………………………………………………………..…………………….124

Figure 5.5. STM images of a set of PTCDA disassembly data from the self-assembled herringbone pattern

to two sub-stable arrangements on flat graphene type I on Cu. (a-i) are typical image of the disassembly

XXVI

process. All these images were obtained under the same scanning conditions: Vs = -2.500 V, I = 0.030 nA,

and with the same size of 23 nm × 23 nm. The purple curved frames in the images show the sub-stable

arrangement……………………………………………………………………………...……………………………………………………….125

Figure 5.6. Typical fullerene orientations on graphene. The computational results suggest that (b) is the

energetically favored orientation. Each grey sphere here is a carbon atom ………………………………………..127

Figure 5.7. Computational results for C60 on 1D-rippled graphene and planar graphene showing

energetically favored orientations. (a) C60 molecule on a peak site on curved-graphene, (b) C60 molecule

with a valley site on curved-graphene and (c) C60 on planar-graphene. (d) Plot of C60-graphene distance

versus relative energy for C60 on a graphene peak (pink), C60 in a graphene valley (blue) and C60 on planar

graphene (green)…………………………………………………………………………………………………………………………………128

Figure 5.8. DFT results for adsorbed molecule/graphene interactions. (a) Energy difference of a C60

molecule on a 1D-rippled graphene surface (b) Energy difference of a PTCDA molecule on a 1D-rippled

graphene. (c) Energy curve for a PTCDA molecule rotation on 1D-rippled graphene on a peak location

(top); favored PTCDA orientations (bottom). (d) Favored PTCDA orientations at the peak site (left) and at

the valley site (right)……………………………………………………………………………………………………………………………130

Chapter 6

Figure. 6.1. (a) STM image of C60 monolayer on graphene (Vs = 1.50 V, I = 0.050 nA). (b) The right image

is the line profile of monolayer C60 (yellow dashed line in figure 6.1a). The left image is the schematic

image of C60 on graphene with a gap about 0.3 nm, according to our DFT calculation. (c) STM image of

Gd3N@C80 monolayer on graphene (Vs = -1.84 V, I = 0.140 nA). (d) The right image is the line profile of

monolayer Gd3N@C80 (red dashed line in figure 6.1c). The left image is the schematic image of

Gd3N@C80 on graphene with a gap about 0.33 nm, according to our DFT calculation…….......................142

XXVII

Figure 6.2. STM images of Gd3N@C80 on graphene. (a) Large area STM topographic image of the

Gd3N@C80 on graphene showing two domains with different orientations and many defects (Vs = -1.84

V, I = 0.248 nA). (b) Zoomed-in STM image (measured from the dashed square of (a)) of Gd3N@C80 on a

graphene showing two domains with different orientations and one defect (Vs = -1.84 V, I = 0.248 nA).

(c) High-resolution image Gd3N@C80 on graphene (Vs = -1.84 V, I = 0.240 nA). (d) Zoomed-in STM image

(measured from the dashed square of (c)) of Gd3N@C80 on a graphene showing lattice constant of an

average about 1.15 nm (Vs = -1.84 V, I = 0.240 nA)……………………………………………………………………………..143

Figure. 6.3. (a) STM image of Gd3N@C80 on graphene after annealing at 200 oC (Vs = -1.69 V, I = 0.122

nA). (b) STM image of Gd3N@C80 on graphene after annealing at 250 oC (Vs = 1.50 V, I = 0.100 nA)…….144

Figure. 6.4. Typical Gd3N@C80 orientations on graphene. The computational results suggest that

orientation 3 is the energetically favored orientation…………………………………………….……………………………146

Figure. 6.5. Computational results for Gd3N@C80 on flat graphene of different orientations, showing

energetically favored orientation 3…….....................................................................................................147

Figure. 6.6. (a) DFT results for molecule-molecule interaction of different orientation between two

Gd3N@C80 molecules. (b) The two Gd3N@C80 molecules with two metal sides facing each other with an

angle between two Gd atoms facets. (c) The two Gd3N@C80 molecules with two metal atoms facing each

other with an angle between two Gd atoms facets…...............................................................................148

1

Chapter 1

Introduction

In December 1959, Richard Feynman delivered a famous speech ‘Plenty of Room at The

Bottom’. The possibility of direct manipulation of individual atoms, which Feynman described in

this lecture, could be a much more useful chemical synthetic method than those used at the

time.1 Lately, this speech is considered to be the beginning of the birth of nanotechnology,

although Roman glassmakers were fabricating glasses containing nanosized metal to obtain

certain unique colors. Nanotechnology is the combination of science, engineering, and

technology conducted at the nanoscale and involves nano-characterization techniques, nano-

materials, self-assembly and many related areas.

Nanoscale characterization techniques are the methods that we use to explore the nanoscale

world. As traditional microscopes do not function at the scale of nanometer due to diffraction

limitations, scientists and engineers have developed a set of instruments (STM, AFM, TEM, etc.)

to image and characterize nanoscale materials. Nanoscale material involves objects such as

nanoparticles and graphene. In addition, molecular self-assembly concerns the spontaneous

formation of a complex structure (with a defined arrangement) by small components such as

molecules or nanoparticles.

In this dissertation, we present a room temperature atomic resolution scanning tunneling

microscope (STM) investigation of the self-assembly behaviors of organic semiconductor

molecules on graphene substrate. Our collaborators have used the computational density

functional theory (DFT) approach to help us understand the interaction between molecules and

2

the interaction between a molecule and graphene better. Firstly, we deposit the molecules on

graphene to create the sample. Then we scan the morphology of the sample to investigate the

self-assembly behavior. Our collaborators then used the DFT calculations and compared the

results with our AFM or STM results to provide detailed physical interpretations.

1.1 2D Materials

Nanoscience is the field of science focus on the study of objects with at least one dimension at

the nanoscale, such as nanoparticles, nanofibers and two-dimensional (2D) materials. 2D

materials as usually named as an atomic layer of a crystalline material. Since the discovery of

graphene in 2004,2 2D materials have attracted extremely high attention. In the following

fifteen years, scientists found various families of 2D materials, including metals, semiconductors

and insulators.3

1.1.1 Graphene

Graphene, an atomic thin layered graphite, was the first true 2D material to be physically

demonstrated.2 It has attracted a lot of attention due to its unique electronic structure,

ultrahigh carrier mobility, thermal conductivity, and mechanical strength.4 The modification of

graphene, for specific electronic properties or functionalities, is required for next movement

towards real device fabrication of different potential applications .5 As graphene is an intrinsic

no bandgap semiconductor, its field of applications could be extended by creating a bandgap by

3

perturbations including controlled introduction of strain, confinement to nanoribbons, or

biasing of a bilayer graphene.6-8 Covalent or noncovalent chemical functionalization could

provide natively inert graphene chemically sensitive, which is crucial for the applications, and

allowing for electron accepting/donating organic molecules to perform charge‐transfer

doping/bandgap engineering.9-10 In a number of recent papers, the electronic effects

introduced by organic molecules on graphene have been discussed.11-13

On the other hand, graphene is also an appealing test substrate to investigate the properties

and mechanism of self-assembly behavior of molecules confined to two dimensions,14 due to

the fact that it could eliminate the effect of the substrate when compared to metal substrates.

Besides, 2D molecular self-assembly is a popular research area focus on understanding the

energetics of molecular organization with the purpose of building an anticipating method of

controlled synthesis of useful 2D structures.14 As a promising useful new material, new

opportunities may be brought by graphene to control and expand the applications of confined

2D molecular structures.14

1.1.2 Other 2D Materials.

In consideration of the success in the study of graphene, the methodology and ideas learned in

these studies have been applied to other layered materials.15 Therefore, in addition to the

current limited applicability of graphene, the field of 2D materials opens up new horizons for

new variety of possibilities. Fortunately, the idea of extracting a layer of strongly covalent in-

plane bonds and inter-layer weak van der Waals like coupling from three-dimensional materials

4

is not only applicable to graphite, but also fit to other layered materials.16 Examples of 2D

inorganic nanomaterials have blossomed during the last decades. Transition metal

dichalcogenides (TMDs),3 with the formula MX2 (where M is a transition metal and X is a

chalcogen), are the main categories of 2D materials and provide a wide range of electronic

properties, from metallic or semimetallic (V,17 Nb,18 and Ta19 dichalcogenides) to insulating or

semiconducting (Ti,20 Zr,21 Mo,22 and W23 dichalcogenides). Adding molecules to a TMDs’

surface is also a very good way to tune its properties. Besides, this also serves as a platform to

investigate the mechanics behind the interactions between molecule-molecule and molecule-

substrate.

1.2 Molecular Self-assembly

Molecular self-assembly is the process in which molecules are arranged in a determined

manner without applying external forces. There are two types of self-assembly, intermolecular

self-assembly24 and intramolecular self-assembly.25 Generally, and in this dissertation,

compared to the version of intramolecular which is more commonly called self-folding, the

term molecular self-assembly refers to intermolecular self-assembly. There are several kinds of

molecular self-assembly, including two-dimensional self-assembly,26 DNA nanotechnology,27

biology macromolecular assembly,28 self-assembly monolayers (SAMs)29 and so on.

5

1.2.1 Two-dimensional Self-assembly

The spontaneous monolayer molecular assembly at the surface is often referred to as two-

dimensional self-assembly. Non-surface active molecules can be assembled into ordered

structures by the intermolecular forces.30 Early direct proofs indicates that with the

development of scanning tunneling microscopy, non-surfactant molecules can be assembled

into higher-order structure at solid surface.31 Finally, two methods for the self-assembly of 2D

structure have become popular, namely self-assembly at the solid-liquid interface and self-

assembly after ultra-high-vacuum deposition and annealing.32 Self‐assembly of molecules at a

surface depends mainly on two kinds of interactions: non‐covalent molecule‐molecule

interaction defining the relationship between neighboring molecules, and molecule‐substrate

interaction stabilizing the molecules on the surface.14 Various different intermolecular

interactions can form 2D self‐assembly. Strong hydrogen bonds (such as carboxylic dimers33)

and strong directional bonds including coordination at metal centers34 can produce porous

assembly structures. Generally, weaker bonds (such as van der Waals35 interactions and

halogen‐halogen36) result in a close‐packed structure with the maximized areal molecular

density.

1.2.2 DNA Self-assembly

DNA self-assembly is a popular research area that uses the self-assembly, bottom-up methods

to achieve nanotechnological goals.27 Over the past 30 years, DNA molecules have been used to

6

construct various nanoscale structures and devices, and prospective applications have begun to

emerge. In 1982, Dr. Nadrian Seeman wrote a paper about the field of structural DNA

nanotechnology: “It is possible to generate sequences of oligomeric nucleic acids which will

preferentially associate to form migrationally immobile junctions, rather than linear duplexes,

as they usually do.”37 Self-assembled DNA complexes with particular property is created in DNA

nanotechnology by using the unique molecular recognition properties of DNA and other nucleic

acids.27 Therefore, DNA is used as a structural material, instead of the biological informational

carrier, to make things like complex two-dimensional and three-dimensional structures (using

the tile-based method38and "DNA origami" method39-40) and 3D lattices in the shapes of

polyhedral. Scientists could use DNA tiles to assemble higher-order periodic or aperiodic

nanotubes and lattices.27 DNA origami structures are commonly organized in a limited grid.40

1.2.3 Macromolecular Self-assembly

Molecular self-assembly is an important notion in supramolecular chemistry.28 Commonly,

there are two main fields of self-assembled macromolecular, as a very extensive and well-

developed research area, i.e. self-assembly of multimolecular systems in which macromolecules

are at least one of the components, and supramolecular polymers (formed from small

molecules driven by non-covalent interactions).41 The fundamental concepts and research

methods of the latter are similar to those in supramolecular chemistry.42 For the former,

however, the main subject is self-assembly of block copolymers, while main driving force is the

cohesive interaction between the like blocks and the repulsion between unlike blocks.41 Thus,

7

this kind of research has developed in parallel for a long time, with slight influences by

supramolecular chemistry.43 Until recently, the concept and achievement of non-covalent

interactions developed by supramolecular chemistry have gradually attracted the attention of

polymer scientists for macromolecule self-assembly, which has greatly promoted its progress.41

Now, in addition to block copolymers, we can also use complementary homopolymers,44

random copolymers45 and oligomers,46 etc. as basic blocks to build regular assemblies driven by

a variety of non-covalent interactions.

1.2.4 Self-assembled Monolayers (SAMs)

Self-assembled monolayers (SAMs) were first produced by J. J. Kirkland and R. K. Iler using

microparticles in 196647-48. SAMs are organic assemblies formed by the molecular components

in solution or the gas phase adsorbed in a regular arrangement onto the solid surface or on the

liquid surface (such as mercury); the adsorbates spontaneously form crystalline (or

semicrystalline) structures.49 Nuzzo & Allara (thiols on gold) and Maoz& Sagiv (trichlorosilanes

on silicon oxide) introduced the two most popular SAMs systems in their work in early 1980s,

and brought SAMs into the popular scientific awareness. Another method that could create

self-assembled monolayers is the layer-by-layer (LBL) method,52 which is formed by alternately

depositing two oppositely charged polyions as + (PAH) and – (PAA). The fundamental concepts

and mechanisms involved in the LBL method is the electrostatic interactions between species

bearing opposite charges. It can easily create large and uniform areas of monolayers on a solid

substrate. The LBL method is widely used in optics, optoelectronics, drug delivery and

8

electrochemistry. In additon, surfactant self-assembly is another important method to produce

self-assembled monolayers.53 Surfactant molecules consist of a polar head compatible with

water and a nonpolar or hydrophobic part compatible with oil, which allows the self-assembled

monolayer of surfactant molecules.

1.3 Molecular Self-assembly on Graphene

Most research on 2D molecular self-assembly has been studied on metal surfaces, like Au54 and

Cu.55 After graphene was discovered, scientists found graphene is also a good substrate to

investigate molecular self-assembly behaviors.14 Self‐assembly of molecules on a surface

depends mainly on two kinds of interactions: molecule‐molecule interaction that define the

relationship between neighboring molecules, and molecule‐substrate interaction that stabilize

the molecules on the substrate. Compared with metal substrates, the molecule-substrate

interaction is much weaker for a graphene substrate while the molecule-molecule interaction is

same. The most common tool that we use to investigate the molecular self-assembly behaviors

is STM and the interaction between molecule-substrate and molecule-molecule can be

modeled well by DFT.

1.4 Document Organization

This chapter is followed by the literature review of Chapter 2, in which we discuss the previous

studies done in the field of molecular self-assembly behaviors on graphene substrates. The

9

previous research involved three major components, firstly the growth, properties and

application of graphene, secondly the research related to 2D molecule self-assembly on metal

substrates and finally the papers have done with molecular self-assembly behaviors on

graphene. In Chapter 3, we discuss in depth the experimental details of the methods and

equipment used in our work. The discussion covers systems and techniques of AFM, systems

and techniques of STM, and physical vapor deposition (PVD) method used in our work.

In Chapter 4, we describe the first stage of the work where we use AFM and STM to investigate

the self-assembled phenyl-C61-butyric acid methyl ester (PCBM) bilayers on graphene and highly

oriented pyrolytic graphite (HOPG). In this part of the work, we report fabrication and

characterization of PCBM bilayer structures on graphene and HOPG. Through careful control of

the PCBM solution concentration (from 0.1 mg/ml to 2 mg/ml) and the deposition conditions,

we demonstrate that PCBM molecules self-assemble into bilayer structures on graphene and

HOPG substrates. Interestingly, the PCBM bilayers are formed with two distinct heights on

HOPG, but only one unique representative height on graphene. At elevated annealing

temperatures, edge diffusion allows neighboring vacancies to merge into a more ordered

structure. This is, to the best of our knowledge, the first experimental realization of PCBM

bilayer structures on graphene. This work could provide valuable insight into fabrication of new

hybrid, ordered structures for applications to organic solar cells.

In Chapter 5, we discuss the second stage of this work where we extend the study from PCBM

to the self-assembly behavior on flat graphene to rippled graphene. We report on the

preparation of fullerene, C60 and perylenetetracarboxylic dianhydride (PTCDA) molecules

adsorbed on a rippled graphene surface. We find that the spherical C60 molecules form a quasi-

10

hexagonal close packed (hcp) structure, while the planar PTCDA molecules form a disordered

herringbone structure. These 2D layer systems have been characterized by experimental STM

imaging and computational DFT approaches. The DFT computational results exhibit interaction

energies for adsorbed molecule/rippled graphene complexes located in the 2D graphene valley

sites that are significantly larger in comparison with adsorbed idealized planar/molecule

graphene 2D complexes. In addition, we report that the adsorbed PTCDA molecules prefer

different orientations when the rippled graphene peak regions are compared to the valley

regions. This difference in orientations causes the PTCDA molecules to form a disordered

herringbone structure on the rippled graphene surface. The results of this study clearly

illustrate significant differences in C60 and PTCDA molecular packing on rippled graphene

surfaces.

In Chapter 6, we describe the final stage of the work where we extend the study to Gd3@C80 on

graphene. The self-assembly of organic semiconductor molecules on a graphene surface is a

central issue for the ultimate application in semiconductor and optoelectronic devices. In

previous studies, the packing behaviors of numerous molecules have been explored. For

example, C60 exhibits an hcp structure on a graphene surface. It has been well known that

several factors dominate the packing of molecules on graphene, such as annealing

temperature. In this study, we explore the effect of the inner cluster of a metallofullerene

molecule Gd3N@C80. The 2D layer system is characterized by experimental STM and the results

are extended by DFT based calculations. We report that the metallofullerene molecule

Gd3N@C80 shows an hcp structure on graphene surface in short range, which is similar to C60 in

long range. However, the theoretical calculations show that the orientations of the inner cluster

11

of Gd3N@C80 determine the energy level of the 2D layer system. The interactions between

Gd3N@C80 molecules is also dominated by the orientation of the inner clusters. Therefore, we

report the subtle but essential inner cluster effect and believe this effect should be considered

in future related studies.

In Chapter 7, we present a discussion of the results achieved in this dissertation and an

overview of future currently under way and possible directions in which the project could be

expanded using different molecules and 2D substrates.

References:

1. Drexler, E., There's Plenty of Room at the Bottom. 2. Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.; Zhang, Y.; Dubonos, S. V.; Grigorieva, I. V.; Firsov, A. A., Electric field effect in atomically thin carbon films. Science 2004, 306 (5696), 666-669. 3. Manzeli, S.; Ovchinnikov, D.; Pasquier, D.; Yazyev, O. V.; Kis, A., 2D transition metal dichalcogenides. Nat Rev Mater 2017, 2 (8). 4. Geim, A. K.; Novoselov, K. S., The rise of graphene. Nat Mater 2007, 6 (3), 183-191. 5. Lee, H.; Paeng, K.; Kim, I. S., A review of doping modulation in graphene. Synthetic Met 2018, 244, 36-47. 6. Dutta, S.; Pati, S. K., Novel properties of graphene nanoribbons: a review. J Mater Chem 2010, 20 (38), 8207-8223. 7. Zhao, J.; Zhang, G. Y.; Shi, D. X., Review of graphene-based strain sensors. Chinese Phys B 2013, 22 (5). 8. Ohta, T.; Bostwick, A.; Seyller, T.; Horn, K.; Rotenberg, E., Controlling the electronic structure of bilayer graphene. Science 2006, 313 (5789), 951-954. 9. Yoon, H. J.; Jun, D. H.; Yang, J. H.; Zhou, Z. X.; Yang, S. S.; Cheng, M. M. C., Carbon dioxide gas sensor using a graphene sheet. Sensor Actuat B-Chem 2011, 157 (1), 310-313. 10. Meric, I.; Han, M. Y.; Young, A. F.; Ozyilmaz, B.; Kim, P.; Shepard, K. L., Current saturation in zero-bandgap, topgated graphene field-effect transistors. Nat Nanotechnol 2008, 3 (11), 654-659. 11. Dong, X. C.; Fu, D. L.; Fang, W. J.; Shi, Y. M.; Chen, P.; Li, L. J., Doping Single-Layer Graphene with Aromatic Molecules. Small 2009, 5 (12), 1422-1426. 12. Batzill, M., The surface science of graphene: Metal interfaces, CVD synthesis, nanoribbons, chemical modifications, and defects. Surf Sci Rep 2012, 67 (3-4), 83-115. 13. Zhang, Z. X.; Huang, H. L.; Yang, X. M.; Zang, L., Tailoring Electronic Properties of Graphene by pi-pi Stacking with Aromatic Molecules. J Phys Chem Lett 2011, 2 (22), 2897-2905. 14. MacLeod, J. M.; Rosei, F., Molecular Self-Assembly on Graphene. Small 2014, 10 (6), 1038-1049.

12

15. Mas-Balleste, R.; Gomez-Navarro, C.; Gomez-Herrero, J.; Zamora, F., 2D materials: to graphene and beyond. Nanoscale 2011, 3 (1), 20-30. 16. Tang, Q.; Zhou, Z.; Chen, Z. F., Innovation and discovery of graphene-like materials via density-functional theory computations. Wires Comput Mol Sci 2015, 5 (5), 360-379. 17. Jing, Y.; Zhou, Z.; Cabrera, C. R.; Chen, Z. F., Metallic VS2 Monolayer: A Promising 2D Anode Material for Lithium Ion Batteries. J Phys Chem C 2013, 117 (48), 25409-25413. 18. Wang, X. S.; Lin, J. H.; Zhu, Y. M.; Luo, C.; Suenaga, K.; Cai, C. Z.; Xie, L. M., Chemical vapor deposition of trigonal prismatic NbS2 monolayers and 3R-polytype few-layers. Nanoscale 2017, 9 (43), 16607-16611. 19. Fu, W.; Chen, Y.; Lin, J. H.; Wang, X. W.; Zeng, Q.; Zhou, J. D.; Zheng, L.; Wang, H.; He, Y. M.; He, H. Y.; Fu, Q. D.; Suenaga, K.; Yu, T.; Liu, Z., Controlled Synthesis of Atomically Thin 1T-TaS2 for Tunable Charge Density Wave Phase Transitions. Chem Mater 2016, 28 (21), 7613-7618. 20. Park, K. H.; Choi, J.; Kim, H. J.; Oh, D. H.; Ahn, J. R.; Son, S. U., Unstable single-layered colloidal TiS2 nanodisks. Small 2008, 4 (7), 945-950. 21. Zhang, M.; Zhu, Y. M.; Wang, X. S.; Feng, Q. L.; Qiao, S. L.; Wen, W.; Chen, Y. F.; Cui, M. H.; Zhang, J.; Cai, C. Z.; Xie, L. M., Controlled Synthesis of ZrS2 Mono layer and Few Layers on Hexagonal Boron Nitride. J Am Chem Soc 2015, 137 (22), 7051-7054. 22. Lee, Y. H.; Zhang, X. Q.; Zhang, W. J.; Chang, M. T.; Lin, C. T.; Chang, K. D.; Yu, Y. C.; Wang, J. T. W.; Chang, C. S.; Li, L. J.; Lin, T. W., Synthesis of Large-Area MoS2 Atomic Layers with Chemical Vapor Deposition. Adv Mater 2012, 24 (17), 2320-2325. 23. Cong, C. X.; Shang, J. Z.; Wu, X.; Cao, B. C.; Peimyoo, N.; Qiu, C.; Sun, L. T.; Yu, T., Synthesis and Optical Properties of Large-Area Single-Crystalline 2D Semiconductor WS2 Monolayer from Chemical Vapor Deposition. Adv Opt Mater 2014, 2 (2), 131-136. 24. Huie, J. C., Guided molecular self-assembly: a review of recent efforts. Smart Mater Struct 2003, 12 (2), 264-271. 25. Schneider, J. P.; Pochan, D. J.; Ozbas, B.; Rajagopal, K.; Pakstis, L.; Kretsinger, J., Responsive hydrogels from the intramolecular folding and self-assembly of a designed peptide. J Am Chem Soc 2002, 124 (50), 15030-15037. 26. Barth, J. V.; Weckesser, J.; Trimarchi, G.; Vladimirova, M.; De Vita, A.; Cai, C. Z.; Brune, H.; Gunter, P.; Kern, K., Stereochemical effects in supramolecular self-assembly at surfaces: 1-D versus 2-D enantiomorphic ordering for PVBA and PEBA on Ag(111). J Am Chem Soc 2002, 124 (27), 7991-8000. 27. Pinheiro, A. V.; Han, D. R.; Shih, W. M.; Yan, H., Challenges and opportunities for structural DNA nanotechnology. Nat Nanotechnol 2011, 6 (12), 763-772. 28. Minton, A. P., Implications of macromolecular crowding for protein assembly. Curr Opin Struc Biol 2000, 10 (1), 34-39. 29. Ulman, A., Formation and structure of self-assembled monolayers. Chem Rev 1996, 96 (4), 1533-1554. 30. https://en.wikipedia.org/wiki/Molecular_self-assembly. 31. Foster, J. S.; Frommer, J. E., Imaging of Liquid-Crystals Using a Tunnelling Microscope. Nature 1988, 333 (6173), 542-545. 32. Rabe, J. P.; Buchholz, S., Commensurability and Mobility in 2-Dimensional Molecular-Patterns on Graphite. Science 1991, 253 (5018), 424-427. 33. Lackinger, M.; Heckl, W. M., Carboxylic Acids: Versatile Building Blocks and Mediators for Two-Dimensional Supramolecular Self-Assembly. Langmuir 2009, 25 (19), 11307-11321. 34. Clair, S.; Pons, S.; Fabris, S.; Baroni, S.; Brune, H.; Kern, K.; Barth, J. V., Monitoring two-dimensional coordination reactions: Directed assembly of Co-terephthalate nanosystems on Au(111). J Phys Chem B 2006, 110 (11), 5627-5632.

13

35. Altman, E. I.; Colton, R. J., Determination of the Orientation of C-60 Adsorbed on Au(111) and Ag(111). Phys Rev B 1993, 48 (24), 18244-18249. 36. Cheng, F.; Wu, X. J.; Hu, Z. X.; Lu, X. F.; Ding, Z. J.; Shao, Y.; Xu, H.; Ji, W.; Wu, J. S.; Loh, K. P., Two-dimensional tessellation by molecular tiles constructed from halogen-halogen and halogen-metal networks. Nat Commun 2018, 9. 37. Seeman, N. C., Nucleic-Acid Junctions and Lattices. J Theor Biol 1982, 99 (2), 237-247. 38. Park, S. H.; Yin, P.; Liu, Y.; Reif, J. H.; LaBean, T. H.; Yan, H., Programmable DNA self-assemblies for nanoscale organization of ligands and proteins. Nano Lett 2005, 5 (4), 729-733. 39. Pal, S.; Deng, Z. T.; Ding, B. Q.; Yan, H.; Liu, Y., DNA-Origami-Directed Self-Assembly of Discrete Silver-Nanoparticle Architectures. Angew Chem Int Edit 2010, 49 (15), 2700-2704. 40. Han, D. R.; Pal, S.; Nangreave, J.; Deng, Z. T.; Liu, Y.; Yan, H., DNA Origami with Complex Curvatures in Three-Dimensional Space. Science 2011, 332 (6027), 342-346. 41. Chen, G.; Jiang, M., Cyclodextrin-based inclusion complexation bridging supramolecular chemistry and macromolecular self-assembly. Chem Soc Rev 2011, 40 (5), 2254-2266. 42. Brunsveld, L.; Folmer, B. J. B.; Meijer, E. W.; Sijbesma, R. P., Supramolecular polymers. Chem Rev 2001, 101 (12), 4071-4097. 43. Zou, J.; Tao, F.; Jiang, M., Optical switching of self-assembly and disassembly of noncovalently connected amphiphiles. Langmuir 2007, 23 (26), 12791-12794. 44. Vanderkooy, A.; Taylor, M. S., Solution-Phase Self-Assembly of Complementary Halogen Bonding Polymers. J Am Chem Soc 2015, 137 (15), 5080-5086. 45. Ilhan, F.; Galow, T. H.; Gray, M.; Clavier, G.; Rotello, V. M., Giant vesicle formation through self-assembly of complementary random copolymers. J Am Chem Soc 2000, 122 (24), 5895-5896. 46. Imamura, T.; Fukushima, K., Self-assembly of metallopyridylporphyrin oligomers. Coordin Chem Rev 2000, 198, 133-156. 47. Kirkland, J. J., Porous Thin-Layer Modified Glass Bead Supports for Gas Liquid Chromatography. Anal Chem 1965, 37 (12), 1458-&. 48. Iler, R. K., Multilayers of Colloidal Particles. J Colloid Interf Sci 1966, 21 (6), 569-+. 49. Love, J. C.; Estroff, L. A.; Kriebel, J. K.; Nuzzo, R. G.; Whitesides, G. M., Self-assembled monolayers of thiolates on metals as a form of nanotechnology. Chem Rev 2005, 105 (4), 1103-1169. 50. Nuzzo, R. G.; Allara, D. L., Adsorption of Bifunctional Organic Disulfides on Gold Surfaces. J Am Chem Soc 1983, 105 (13), 4481-4483. 51. Maoz, R.; Sagiv, J., On the Formation and Structure of Self-Assembling Monolayers .1. A Comparative Atr-Wetability Study of Langmuir-Blodgett and Adsorbed Films on Flat Substrates and Glass Microbeads. J Colloid Interf Sci 1984, 100 (2), 465-496. 52. Decher, G.; Hong, J. D., Buildup of Ultrathin Multilayer Films by a Self-Assembly Process .2. Consecutive Adsorption of Anionic and Cationic Bipolar Amphiphiles and Polyelectrolytes on Charged Surfaces. Ber Bunsen Phys Chem 1991, 95 (11), 1430-1434. 53. Nagarajan, R.; Ruckenstein, E., Theory of Surfactant Self-Assembly - a Predictive Molecular Thermodynamic Approach. Langmuir 1991, 7 (12), 2934-2969. 54. Ciesielski, A.; El Garah, M.; Masiero, S.; Samori, P., Self-assembly of Natural and Unnatural Nucleobases at Surfaces and Interfaces. Small 2016, 12 (1), 83-95. 55. Li, Q.; Gao, J. Z.; Li, Y. Y.; Fuentes-Cabrera, M.; Liu, M. X.; Qiu, X. H.; Lin, H. P.; Chi, L. F.; Pan, M. H., Self-assembly directed one-step synthesis of [4]radialene on Cu(100) surfaces. Nat Commun 2018, 9.

14

Chapter 2

Literature Review

2.1 Introduction and Background

Graphene as an atomic thin layer material has attracted dramatically attention, due to its

unique electronic structure, ultrahigh carrier mobility, thermal conductivity, and mechanical

strength.1-4 Due to its intrinsic zero bandgap, graphene logic transistors possess a poor on/off

current ratios. Many attempts have been made to overcome this problem, such as bilayer

graphene manipulation by an electric field,5 chemical modification of graphene,6 generating

strain in graphene,7 and quantum confinement in graphene nanoribbons.8 None of these

methods has achieved the level of performance required for equipment applications at room

temperature. Besides these techniques, method have been proposed to create a periodic

potential modulation on graphene to open the band gap.

Molecular self-assembly is a method to create a periodic modulation on graphene. Besides,

graphene is also an excellent test substrate for the study of moleculer interactions.9 In what

follows in this chapter, we will discuss some recent papers related to molecular self-assembly

on graphene, including topics related to graphene, molecular self-assembly and molecular self-

assembly on graphene respectively. For graphene itself, we will discuss the synthesis method,10

properties11 and the applications4 of graphene. For molecular self-assembly, we will discuss

molecular self-assembly on a metal surface by metal centers,12 hydrogen bonding,13 van der

15

waals force14 and halogen-halogen interactions.15 In the last part, we will introduce the papers

related to molecular self-assembly (C60,16 PTCDA,17-18 phthalocyanines19-20) on graphene.

2.2 Graphene

In recent years, graphene as a single-atom-thick, sp2-bonded carbon sheet that is tightly

integrated in a honeycomb lattice, and has attracted remarkable attention due to the potential

as next generation electronic device, as its excellent characteristics, high current density,

including chemical inertness, ballistic transport, optical transmittance, super hydrophobicity

and high thermal conductivity at the nanometer level.1, 4, 21, 22

Graphene was first created from graphite by using a technique called micromechanical

cleavage.1, 23 This method can easily produce high-quality graphene crystallites and further lead

to a large number of experimental activities. There are many reports about synthesis of

graphene, most of which are discussing the exfoliation of graphite or thermal epitaxy of

graphene on a SiC surface, and more recently chemical vapor deposition.24-28

Normal graphene is characterized as a no gap semiconductor or semi-metal, and its novel

electronic properties create an significant high transparency of the monolayer, with a

remarkably low white light absorption of 2.3% .29 Electrical characteristics have shown that a

surprising high electron mobility, and the experimentally reported values exceeds 15,000 cm2V−

1s− 1.2 The corresponding resistivity of the graphene sheet, which is less than the lowest

resistivity substance (sliver) known at room temperature, would be 10−6 ohm–cm.22 Graphene

16

nano ribbons (GNRs) show different electrical properties, with armchair or zigzag configuration,

armchairs are either metallic or semiconducting, while the zigzag GNRs can be metallic.22

Graphene's excellent electrical properties have attracted interest in future electronic

applications such as field emitters, ballistic transistors, integrated circuit components,

transparent conductive sensors and electrodes.4 Graphene has a high electron mobility and low

Johnson noise, making it to be used in the field effect transistor (FET) as a channel.30 Graphene

would be an excellent sensor due to the combination of its excellent electrical properties and

low noise .4 Graphene would be very efficient to detect adsorbed molecules as it is exposed to

the surroundings due to its two-dimensional structure. Graphene would be promoted as an

excellent candidate material for transparent conductive electrodes, required for prospective

applications in organic light-emitting diodes (OLEDs), liquid crystal displays, touch-screens and

organic photovoltaic cells, because of the high electrical conductivity and high optical

transparency of graphene.22

2.2.1 Synthesis Methods of Graphene

There are three basic methods, exfoliation, chemical vapor deposition and molecular beam

epitaxy, that can grow graphene samples. In 2004, graphene was first discovered by Novoselov

et al. They were the first to show repeatable graphene synthesis by exfoliation.1 Since then, the

method has been and is being developed, and broadly spread to other 2D materials, like

TMDs.23 In the same year, C. Berger et al. provided an epitaxial method to grow graphene on a

SiC surface by vacuum annealing single crystalline SiC substrates above 1100 oC.26 In 2009, A.

17

Reina et al. and K. Kim et al. published two independent papers discussing large-scale pattern

chemical vapor deposition growth of graphene films.27-28 We will discuss the advantages and

disadvantages of these methods later of this section.

2.2.1.1 Exfoliation and Cleavage

Exfoliation is the first method demonstrated to create a graphene sheet. Graphite is a stacked

layer material of a number of graphene, which are bonded by weak van der Waals force.

Therefore, if these bonds can be broken, graphene can be produced from a highly ordered

pyrolytic graphite (HOPG) sheet. These weak bonds could be broken to separate out individual

graphene sheets by exfoliation and cleavage using mechanical or chemical energy.4

Exfoliation involves the removal process of the top layers. In the investigation of Dr. Novoselov

et al., a number of 5 μm deep mesas (about area 0.4 to 4 mm2) could be made by dry etching a

commercially available HOPG sheet of 1 mm thickness in oxygen plasma.4 Then place it on the

photoresist and bake to attach the mesa to the photoresist . Then, peel off layers from the

graphite sheet by using scotch tape. It was found that the singe to few layer graphene

transferred on a Si substrate after released the thin flakes, attached to the photoresist, in

acetone. Later, the method was used to generate 2D atomic crystals of a number of other

materials, including MoS2, h-BN and so on.31 It was found that the production process of the

graphene sheets was very easy and reliable (single crystal), so it attracted direct attention of

the scientific community.32-33 With minor changes in the original process, the results show that

18

by utilizing the bonding of HOPG on the substrate and controlling exfoliation, large (about 10

μm) and flat graphene sheets can be produced.34

Figure. 2.1. Graphene films. (a) Photograph (in normal white light) of a relatively large multilayer

graphene flake with thickness ∼3 nm on top of an oxidized Si wafer. (b) Atomic force microscope (AFM)

image of 2 μm by 2 μm area of this flake near its edge. Colors: dark brown, SiO2 surface; orange, 3 nm

height above the SiO2 surface. (c) AFM image of single-layer graphene. Colors: dark brown, SiO2 surface;

brown-red (central area), 0.8 nm height; yellow-brown (bottom left), 1.2 nm; orange (top left), 2.5 nm.

Notice the folded part of the film near the bottom, which exhibits a differential height of ∼0.4 nm. (d)

Scanning electron microscope image of one of our experimental devices prepared from few-layer

graphene. (e) Schematic view of the device in (d).1 Copyright 2004 reprinted with permission from AAAS.

19

In 2015, Dr. Peter Sutter’s group reported a modification for producing monolayer and few-

layer sheets from layered materials.23 Their technique provides two procedures that

homogenize and enhance the adhesion force between the outside sheet in contact with a

substrate: Prior to exfoliation, maximization of the uniform contact area, by an additional heat

treatment, at the interface between the layered material and the substrate, and the effectively

removing of surrounding adsorbates from the substrate by oxygen plasma cleaning . For

graphene exfoliation, compared to the established exfoliation methods, these simple

processing steps can increase the transferred sheet area and the yield by more than 50 times.

Figure. 2.2. Optical images of graphene flakes prepared by the standard exfoliation method and Dr.

Peter Sutter’s modified method. (a and b) Optical microscopy images of typical monolayer to trilayer

graphene prepared by the standard method, including a solvent wash and O2 plasma cleaning of the

20

substrate followed by graphene transfer. (c and d) Optical microscopy images of two graphene flakes

prepared by Dr. Peter Sutter’s modified method, with O2 plasma clean of the SiO2/Si surface, followed by

contact with graphite-loaded tape, annealing to 100 oC, cooling to room temperature and peel-off.23

Copyright 2015 reprinted with permission from American Chemical Society.

The exfoliation method is a very simple method that can provide a high quality, single crystal

sample on many surfaces. Although the modified exfoliation could increase the area to

submillimeter size, the limited size and random shape will limit the potential applications of this

approach. Therefore, it is primarily useful for fundamental research and early stages of

prototype design.

2.2.1.2 Epitaxy

Epitaxial is defined as the directional excessive growth of thin film materials, usually refers to

the growth of single crystal thin films. The first epitaxial graphene was grown by heating a

single crystal of 6H-SiC to a temperature of 1250 to 1400 oC for a short time (1 to 20 minutes) to

thermal decompose Si on the (0001) surface plane.26 This method could effectively grow single

crystal graphene, but it is a challenge to obtain large graphene domains with uniform thickness.

Later, scientist considered the synthesis of graphene on transition metals by epitaxy. In 2008,

Dr. Sutter published a paper about graphene synthesis by thermal cycling of on Ru crystal in

UHV.24 Very sparse graphene nucleation allows the growing of true macroscopic single

crystalline domains with size exceeding 200 μm at high temperatures. Then, epitaxial graphene

21

was expanded to non-metal substrates like hexagonal boron nitride, as, in recently, hexagonal

boron nitride has become a potential substrate for graphene devices.25

Figure 2.3. (a-d) LEED patterns from graphite/SiC(0001). The sample was heated several times to

successively higher temperatures. (a) 1050 °C for 10 min. Immediately after oxide removal, showing SiC

1 × 1 pattern at 177 eV. AES C:Si ratio 1:2. (b) 1100 °C, 3 min. The x3 × x3 reconstruction is seen at 171

22

eV. AES ratio 1:1.9. (c) 1250 °C, 20 min. 109 eV pattern showing diffracted beams from the 6x3 × 6x3

unit cell. Examples of first-order SiC and graphite spots are marked. Note the surrounding hexagons of

“6 × 6” spots. AES C:Si ratio 2:1 (∼1 ML graphite). (d) 1400 °C, 8 min. 98 eV LEED pattern. AES ratio 7.5:1

(∼2.5 ML graphite).26 (e) STM image of a surface region of the sample described in Figure 1d. Inset:

Atomically resolved region (different sample, similar preparation).26 (f) UHV-SEM image of a large area

of the Ru(0001) surface after first-layer graphene growth. Inset: Carbon KLL (260.6 eV) UHV scanning

Auger microscopy image, obtained on this sample.24 (g) AFM image of as-grown graphene on h-BN, the

scale bar is 200 nm.25 Copyright 2004 reprinted with permission from American Chemical Society.

Copyright at 2008 reprinted with permission from Nature publishing group. Copyright at 2013 reprinted

with permission from Nature publishing group.

Epitaxial growth of graphene can create a relative large single crystal graphene domain. The

disadvantage is the high cost and slow growth rate of the epitaxial method.

2.2.1.3 Chemical Vapor Deposition

Many chemical methods have been developed for the synthesis of large-scale, including epitaxy

on ruthenium and silicon carbide since the discovery of the first isolated graphene made by

mechanically exfoliating graphite crystals. Although epitaxy provides high-quality multilayer

graphene samples that have strong interactions with their substrates, electrical isolation

monolayer or bilayer graphene for applications of device has not been prepared by this

method. In 2009, Dr. Hong’s group and Dr. Kong’s group provide a method for the growth of

thin layer graphene flakes by using chemical vapor deposition (CVD) on nickel to prepare the

23

large scale, high crystal quality of the graphene films with single crystal size about 20 μm.27-28 In

their CVD technology, a polycrystalline Ni film (at 900-1000 °C) is exposed to a highly diluted

hydrocarbon stream at ambient pressure. It was later proven that other transition metals, such

as Cu and Au, could also be the substrate for CVD method.35-36

Figure. 2.4. (a) SEM images of as-grown graphene films on thin (300-nm) nickel layers and thick (1-mm)

Ni foils (inset). (b) TEM images of graphene films of different thicknesses.27 (c) Optical image of the

24

grown graphene transferred from the Ni surface in panel a to another SiO2/Si substrate. (d-e) High-

magnification TEM images showing the edges of film regions consisting of one (d) and three (e)

graphene layers. The cross-sectional view is enabled by the folding of the film edge. The in-plane lattice

fringes suggest local stacking order of the graphene layers.28 Copyright at 2009 reprinted with

permission from Nature publishing group. Copyright at 2009 reprinted with permission from American

Chemistry Society.

The benefit of the CVD method is that it can create a fully covered and large scale graphene

with a relative easy method and low cost. On the other hand, it is difficult to control the

thickness of the graphene and it can’t be produced as a single crystal sample.

2.2.2 Properties of Graphene

Since graphene has been discovered, it has attracted great attention in many fields of science

due to its impressive properties. The properties of graphene derive from its single atomic

honeycomb structure. From tight-binding theory, we can derive a very unique characteristic of

graphene: a linear energy dispersion relation around the Dirac point. Besides single layer

graphene, the bilayer and trilayer graphene also shows unique properties depending on the

angle between layers.

25

2.2.2.1 Single Layer: Tight-binding Theory

To study the structure of graphene, we have to start with the lattice structure of graphene. The

structure of graphene can be seen as a triangular lattice, each unit cell has two atoms. The

lattice vectors can be written as:

a1 =𝑎

2 (3, √3), a2 =

𝑎

2 (3, −√3),

where a=1.42 Å is the carbon-carbon distance. The reciprocal lattice vectors are given by

b1 =2𝜋

3𝑎(1, √3), b2 =

2𝜋

3𝑎(1, −√3).

The tight binding theory, or tight binding approximation, is a method used to treat a complex

quantum system in solid state physics. Basically, there are there sub-assumptions: (1) consider

only interactions between the frontier atomic orbitals of nearest neighbors; (2) consider only

frontier atomic orbitals; (3) ignore the overlap integrals of separated atoms. By applying this

tight binding theory to the simplest two atom system, we obtain a simple Hamiltonian

equation:

[𝛼1 𝛽𝛽 𝛼2

] [C1

C2] = 𝐸 [

C1

C2]

For graphene, we assume the local potential: ϕ = c1𝜙1 + c2𝜙2 and the wave function: Ψ(��) =

∑ 𝑒𝑖��∙��𝜙(𝑥 − ��)�� , where �� is the lattice vector.

In this graphene system, α1 = α2 = ⟨ϕ𝑗|H|ϕ𝑗⟩, 𝛽 = ⟨ϕ1|H|ϕ2⟩,

We then get two overlap equations:

26

⟨ϕ1(R)|H|Ψ(x)⟩ = 𝐸⟨ϕ1(R)|Ψ(x)⟩ 𝑎𝑛𝑑 ⟨ϕ2(R)|H|Ψ(x)⟩ = 𝐸⟨ϕ2(R)|Ψ(x)⟩,

By applying the tight binding theory to graphene and solve these Hamiltonian equations, we get

the dispersion relation equation:

E = α ± β√3 + 2 cos(����1) + 2 cos(����2) + 2 cos(��(��1 − ��2))

At the K points, the lower band touch the upper band, where K = (±4𝜋

3√3𝑎0, 0) , (±

2𝜋

3√3𝑎0,

2𝜋

3𝑎0).

According to the dispersion relation equation, we get a massless electron at the K point.

Figure. 2.5. (a) Honeycomb lattice and its Brillouin zone. Left: lattice structure of graphene, made from

two interpenetrating triangular lattices (a1 and a2 are the lattice unit vectors, and δ𝑖, i= 1, 2, 3 are the

27

nearest-neighbor vectors). Right: corresponding Brillouin zone. The Dirac cones are located at the K and

K′ points.37 (b) Electronic dispersion in the honeycomb lattice. Left: energy spectrum (in units of t) for

finite values of t and t′, with t= 2.7 eV and t′=−0.2t. Right: zoom in of the energy bands close to one of

the Dirac points.37 Copyright at 2009 reprinted with permission from American Physical Society.

2.2.2.2 Single Layer: Properties

Graphene morphology: As graphene is a single atom layer material, the morphology is highly

dependent on the substrate. Basically, graphene forms a flat surface on a flat substrate (metal

or SiO2). In some cases, these flat graphene sheets form Moiré patterns. For example, flat

graphene forms a triangle pattern on Ru(001)38 and Cu(111)39 surfaces, while it forms a one

dimensional nanoripple on the Cu(100) surface.16 But, in 2012, Dr. László P. Biró published the

first paper about rippled graphene with a period of 1 nm and amplitude of about 0.2 nm.40 This

kind of rippled graphene is mainly due to the negative expansion coefficient (NEC) of graphene.

When graphene cools down from high temperature, the NEC of graphene makes it expaned,

while the PEC of the substrate makes the substrate shrink, which make a mismatch between

graphene and the substrate. Thus, the graphene has to form a rippled or wrinkled structure to

release the extra space of the graphene.

28

Figure. 2.6. (a) STM image (170 nm × 170 nm) for graphene grown on Cu(111). A Moire pattern arising

due to the lattice mismatch between graphene and Cu(111) is visible that continues over the step

edges.39 (b) Large area STM topographic image of the rippled graphene showing well-defined linear

periodic modulation with a 0.75 nm spatial modulation frequency (Vs = 0.80 V, I = 1.0 nA).16 (c) STM

image showing several nanotrenches of different orientations, all exhibiting graphene nanoripples over

the trenches with the ripple crests always perpendicular to the trench edges. On the flat regions

between the trenches, a Moiré pattern can be observed.40 (d) Atomic-resolution STM image of a

nanotrench exhibiting subnanometer graphene ripples. The magnified insets exhibit the honeycomb

graphene lattice both over the flat substrate (bottom right) and the rippled region over the trenches

29

(top left).40 Copyright at 2015 reprinted with permission from American Chemical Society. Copyright at

2015 reprinted with permission from Nature publishing group. Copyright at 2012 reprinted with

permission from Nature publishing group.

Electronic properties: The experimental observation that the mass of the cyclotron depends on

the square root of the electronic density of graphene is explained as the evidence of the

presence of massless Dirac quasiparticles in graphene.41 Owing to its unique band structure,

graphene exhibits novel transport effects such as ambipolar field effect and minimum

conductivity which are absent in most conventional materials, with a concentration up to 1013

cm−2 and a mobility of as high as 15 000 cm2V−1s−1 at room temperature, which is much higher

than the concentration (1010 cm-2) and mobility (1400 cm2V−1s−1) of silicon.1, 42 Mobility of

suspended graphene exceeds 200 000 cm2V−1s−1 by minimizing impurity scattering.43

Mechanical properties: The mechanical properties of monolayer graphene have been well

predicted by first principles based calculations, including the Young's modulus and fracture

strength.41 A direct measurement of mechanical properties of monolayer graphene was first

reported by Lee et al.44, by nanoindentation of suspended monolayer graphene membranes

using an atomic microscope (AFM). By force‐volume measurements in AFM, few‐layer

graphene circular membranes were also characterized.45 Recently, by the nanoindentation

method using an AFM, the inherent fracture strength and elasticity of free standing monolayer

graphene were measured (Figure 2.7a and b).46 It was reported that the Young's modulus of

30

defectless graphene is 1.0 TPa and the fracture strength is 130 GPa, compared with the Young's

modulus of 128 GPa and a fracture strength of 100 MPa of Cu.47

Figure 2.7. (a) Scanning electron microscopy (SEM) image of a graphene flake spanning an array of

circular holes (scale bar, 3μm). (b) Schematic illustration of nanoindentation on membranes.46 (c)

Photograph of a 50‐μm aperture partially covered by graphene and its bilayer. The line scan profile

shows the intensity of transmitted white light along the yellow line. Inset shows the sample design: a 20

μm thick metal support structure has apertures 20, 30, and 50 μm in diameter with graphene flakes

deposited over them.29 (d) Optical image of graphene flakes with one, two, three, and four layers on a

31

285‐nm thick SiO2‐on‐Si substrate.48 Copyright at 2008 reprinted with permission from AAAS. Copyright

at 2008 reprinted with permission from AAAS. Copyright at 2007 reprinted with permission from

American Chemical Society.

Optical properties: From the infrared to the visible range of the spectrum, the high frequency

conductivity of Dirac fermions in graphene has been expressed as a constant of πe2/2h.49-50

Then for normal incidence light, the reflectance R and transmittance T are then R = 1/4π2α2T

and T = (1 + 1/2πα)−2; the opacity is (1 − T) ≈ πα ≈ 2.3% (where α = 2πe2/hc ≈ 1/137, h is

Planck's constant, c the light speed, and e the electron charge). It is considered as the result of

the electronic and structure properties of graphene that the expression of R and T in terms of

basic constants instead of involving parameters of graphene.51 As shown in Figure 2.7c, it is

experimentally observed that graphene in the visible light range has a constant transparency

(about 97.7%), as well as the transmittance decreases proportionally with the number of

graphene layers, while the transparency of indium tin oxide (ITO, a common transparent

electrode) is around 85%.29

2.2.2.3 Bilayer and Trilayer Graphene

Monolayer graphene owns a linear energy dispersion near the Dirac points. Their bands

hybridized due to hopping results of interlayer in a fundamental modification to the low energy

band structure depending on the stacking order (AA (carbon atoms of both layers have identical

lateral coordinates) or AB (the second graphene layer is shifted relative to the first one by the

32

vector equal to the edge of the hexagon)), when two aligned graphene sheets are stacked. A

hexagonal moiré pattern consisting of alternating AB- and AA-stacked regions appears and acts

as a superlattice modulation, if there is an additional twist angle between layers. For example,

the band structure of twisted bilayer graphene can be customized to create a bandgap and

band curvatures that would not otherwise exist. In 2018, Yuan Cao et al. pulished two papers

related to twisted bilayer exist. In one paper, observation of superconductivity of bilayer

graphene with slightly different twist angles has been reported, with the maximum critical

temperature of 1.7 K.52 In the other paper, they experimentally demonstrate that the interlayer

hybridization can induce almost flat low-energy bands when the twist angle of twisted bilayer

graphene is close to the magic angle predicted by theory.53 A insulating phase at half filling of

these flat bands is led by this quenching of the quantum kinetic energy, which indicates a Mott-

like insulator in the localized flat bands.53 After the publication of these two papers, there was

intense investigation of twisted bilayer or trilayer two-dimensional materials.54

33

Figure 2.8. Two-dimensional superconductivity and insulator in a graphene superlattice. (a) Schematic of

a typical twisted bilayer graphene (TBG) device and the four-probe (Vxx, Vg, I and the bias voltage Vbias)

measurement scheme. The stack consists of hexagonal boron nitride on the top and bottom, with two

graphene bilayers (G1, G2) twisted relative to each other in between. The electron density is tuned by a

metal gate beneath the bottom hexagonal boron nitride layer.52 (b) Fourprobe resistance Rxx = Vxx/I (Vxx

and I are defined in a) measured in two devices M1 and M2, which have twist angles of θ = 1.16° and θ =

1.05°, respectively. The inset shows an optical image of device M1, including the main ‘Hall’ bar (dark

brown), electrical contact (gold), back gate (light green) and SiO2/Si substrate (dark grey).52 (c) Schematic

of the TBG devices. The TBG is encapsulated in hexagonal boron nitride flakes with thicknesses of about

10–30 nm. The devices are fabricated on SiO2/Si substrates. The conductance is measured with a voltage

bias of 100 μV while varying the local bottom gate voltage Vg. ‘S’ and ‘D’ are the source and drain

34

contacts, respectively.53 (d) The band energy E of magic-angle (θ = 1.08°) TBG calculated using an ab

initio tight-binding method. The bands shown in blue are the flat bands that we study.53 Copyright at

2018 Reprinted with permission from Nature Publishing Group. Copyright at 2018 Reprinted with

permission from Nature Publishing Group.

2.2.3 Applications of Graphene

2.2.3.1 Graphene Field Emission (FE)

The application of graphene in Field emission (FE) display is one of the potential ones. FE is

emission of electron induced by high electric field. The most common context is field emission

from a solid surface in vacuum by creating field enhancement at a sharp tip. Graphene sheets

need to be erected on the substrate to take advantage of high field enhancement. The detailed

process of a graphene field emission device can be found in Dr. Wu’s paper.55 They reported

that based on the constant FN slope in the low current region, graphene film field-

enhancement factor (β) can be determined to be around 3700, which is much higher than that

of graphene powder (about 800).55

35

Figure. 2.9. (a) Typical plots of the electron-emission current density (J) as a function of applied electric

field (E) for the graphene film and graphene-powder coating. (b) Corresponding F–N plots.55 Copyright at

2009 reprinted with permission from Wiley-VCH Publishing Group.

2.2.3.2 Graphene Field Effect Transistors (FET)

Field effect transistors (FETs) is another potential application of graphene. However, graphene

cannot be directly used for FET applications due to its zero band gap. The first observation of

characteristics of a saturating transistor in a graphene FETs was reported by Shepard et al.30 In

2008. The saturation velocity depends on the carrier concentration, which is due to the

scattering by interfacial phonons between graphene and the SiO2 substrate. Field-effect model

and single point diffusive carrier transport in the density of state explain anomalous

characteristics in the current–voltage feature. Despite low on–off current ratios, the

electrostatic modulation by an effectively coupled top gate can produce transconductance up

to 150 µSµm−1.

36

Figure. 2.10. (a) Schematic of a graphene FET on a Si/SiO2 substrate with a heavily doped Si wafer acting

as a back gate and a gold top gate. (b) SEM micrograph showing a representative graphene top-gated

FET. The top-gate of this device is 1 µm long, with 3 µm spacing between the source–drain contacts. All

electrodes are Cr/Au. (c) Drain current (Id) as a function of source-to-drain voltage (Vsd) for Vgs-top = −0.3

V, −0.8 V, −1.3 V, −1.8 V, −2.3 V and −2.8 V (from bottom to top) for Vgs-back = 40 V. (d) Id as a function of

Vsd for Vgs-top = −0.3 V, −0.8 V, −1.3 V, −1.8 V, −2.3 V and −2.8 V (from bottom to top) for Vgs-back = −40 V.30

Copyright at 2008 reprinted with permission from Nature Publishing Group.

37

2.2.3.3 Graphene-based Gas and Biological Sensors

Graphene-based sensors, including gas and biological sensors, is one of the most prospective

applications.56 The working theory of a graphene-based gas and biological sensors is based on

changes in conductivity of graphene and the molecular adsorption on the surface of graphene.

The change of conductivity of the graphene is mainly because of the absorbed gas molecules

working like acceptors or donors can create the change in electrical conductivity. In addition,

some interesting properties of graphene help increase the sensitivity to the detection of

individual atoms or molecules. First, graphene is a two-dimensional material whose entire

volume is exposed to the target analyte. Then, graphene has high conductivity and low Johnson

noise, therefore, a significant change of conductivity can be caused by a variation in carrier

concentration. What’s more, graphene has almost no crystal defects, ensuring a low level of

thermal switching noise. In the end, four-probe measurements can be made on single crystal

graphene device with low resistance ohmic electrical contacts. One of the influential papers

that demonstrate a gas sensor using a graphene sheet was published in 2011.57 In this paper, a

high performance carbon dioxide (CO2) gas sensor is reported of graphene made by exfoliation.

The graphene sensor can be operated at room temperature and under ambient conditions,

unlike other solid-state gas sensors. For a variety of concentrations of CO2 gas adsorbed on the

surface of graphene, the change in the conductance of the device was measured. With the

increasing concentration of CO2 gas, the conductivity of the graphene gas sensor increases

linearly. The advantages of this sensor are fast response time, high sensitivity, low power

consumption, and short recovery time.

38

Figure. 2.11. (a) Experimental setup for measurements performed using the graphene CO2 gas sensor.

(b) Time response of the graphene CO2 gas sensor in the presence of 100 ppm CO2, at different

temperatures.57 Copyright at 2011 reprinted with permission from Elsevier Publishing Group.

2.2.3.4 Transparent Electrode

Indium tin oxide (ITO) is broadly used in the manufacture of conductive transparent coatings for

solar cells, flat panel displays, EMI shielding, touch panels and liquid crystal displays (LCD).

However, limited supply, high cost and fragile essence of ITO limit its application in flexible

substrates, which motivate the research on highly conductive, highly transparent conductive

electrode. Graphene is considered as the most promising candidate for optoelectronic

applications in the future, including LCD displays and solar cells. Graphene’s extraordinary

chemical, mechanical and thermal stability, coupled with its atomic thin thickness and high

transparency, makes it an excellent material for transparent conductive electrode. The

application of CVD-grown graphene-based transparent electrodes for organic solar cell has

39

been reported by Zhou et al.58 They proposed a technique that takes advantage of the

continuous characteristics of graphene grown by CVD, which results in minimal surface

roughness of 0.9 nm and provides sheet resistance as low as 230Ω/□, much lower than similar

transparent stacked graphene sheets. Besides, solar cells with graphene and ITO electrodes

were made on flexible polyethylene terephthalate (PET) substrates side by side to prove a

comparable results with a power conversion efficiencies (η) of 1.18 (Figure 2.12 c) and 1.27%

(Figure 2.12 d), respectively. Besides, graphene-based solar cells have excellent capability to

operate under bending conditions as high as 138°, while the ITO devices show irreversible

failure with 60° bending.

Figure. 2.12. (a) Schematic representation of the energy level alignment (top) and the construction of

heterojunction organic solar cell fabricated with graphene as anodic electrode:

40

graphene/PEDOT/CuPc/C60/BCP/Al. (b) Schematic illustration of the transfer process of CVD‐graphene

onto transparent substrate. (c, d) The plots of current density vs voltage for (c) graphene and (d) ITO

devices under 100 mW cm–2 AM1.5G spectral illumination at different bending angles. Insets show the

experimental setup used in the experiments.58 Copyright at 2010 reprinted with permission from

American Chemical Society.

2.2.3.5 Batteries

The Lithium-ion batteries have been an important part of handheld devices due to their clean

and renewable characteristics. Because of its reversibility and reasonable specific capacity,

graphite is currently used as the anode electrode material for Li ion battery. However, new

materials with higher stability and capacity need to be studied to satisfy the growing

requirement for Li ion batteries with higher durability and energy density. Graphene has been

proposed as the most promising alternative for Lithium ion batteries among the carbonaceous

materials due to its higher surface area, chemical resistance, and electrical conductivity than

graphitic carbon. Honma et al. prepared graphene as the anode material for a lithium battery.59

In this study, by using graphene nanosheet (GNS) materials, the possibility of higher lithium

storage capacity was explored. The lithium insertion/extraction characteristics are shown in

Figure 2.13, which are graphite, GNS, GNS+carbon nano tube (CNT), and GNS+C60, respectively.

Charge/discharge curves of (a) graphite, (b) GNS, (c) GNS+CNT, and (d) GNS+C60 are shown in

Figure 2.13a. The typical insertion/extraction performance is shown that a reversible capacity of

about 320 mAh/g was obtained at a current density of 0.05 A/g with highly crystalline graphite

41

electrode materials in Profile (a) of Figure 2.13b. On the other hand, compared to graphite, the

charge/discharge profiles of GNS, GNS+CNT, and GNS+C60 show much different curves, which

indicates that lithium has different adaptability in these new carbonaceous materials. For GNS,

GNS+CNT, and GNS+C60, the reversible capacity at the same condition is 540, 730, and 784

mAh/g respectively, which is significant larger than the value of graphite (320 mAhg-1).

Figure. 2.13. Lithium insertion/extraction properties of the GNS families. (a) charge/discharge profiles of

(a) graphite, (b) GNS, (c) GNS+CNT, and (d) GNS+C60 at a current density of 0.05 A/g. (b)

Charge/discharge cycle performance of (a) graphite, (b) GNS, (c) GNS+CNT, and (d) GNS+C60.59 Copyright

at 2008 reprinted with permission from American Chemical Society.

2.3 2D Molecular Self-assembly

The idea of 2-dimensional molecular self-assembly is developed from self-assembled

monolayers (SAMs). In his 1946 paper, Dr. Zisman is often thought to originate the concept of

SAMs.60 As the free energy can be reduced by these adsorbates at the interface between the

substrate and the ambient environment, extraneous organic materials are easily adsorbed on

42

bare surfaces of metals and metal oxides.61 The stability of and the interfacial properties of

nanostructures metals and metal oxides can be significantly impacted by these adsorbates. The

reactivity of the surface atoms can be reduced because the organic material can act as an

electrically insulating film or act as an electrostatic or physical barrier against aggregation.62

SAMs are organic assemblies formed by the molecular components in solution or the gas phase

adsorbed in a regular arrangement onto the solid surface or on the liquid surface (such as

mercury); the adsorbates spontaneously form crystalline (or semicrystalline) structures.62 Nuzzo

& Allara (thiols on gold, Figure. 2.14. a)63 and Maoz& Sagiv (trichlorosilanes on silicon oxide,

Figure. 2.14. b)64 introduced the two most popular SAMs systems in their work in early 1980s,

and brought SAMs into the popular scientific awareness.

Figure. 2. 14. (a) SAMs of thiols on gold substrate. (b) SAMs of trichlorosilanes on SiO2 substrate.

As it was called, supramolecular chemistry involves how people can use noncovalent

interactions to produce higher order assemblies by studying the way molecules interact with

each other.65 Solution-based systems is the most popular research field on supramolecular

chemistry, especially in its early stages, which revealed the basic concepts in this exciting field.66

43

After the discovery of STM, it was soon used in the field of SAMs and supramolecular self-

assembly on surfaces due to its atomic scale high spatial resolution, which forms a new field

called 2D molecular self-assembly.67 In particular, STM is extremely useful to probe the

electronic properties of the surface molecule and not only the organization of molecular

surface dynamics , but also of molecules on a local scale during the self-assembly process.

2D molecular self-assembly usually means the molecules form a two-dimensional structure on a

surface, such as a metal or, more recently, graphene. Self‐assembly of molecules at a surface

depends mainly on two kinds of interactions: non‐covalent molecule‐molecule interaction

defining the relationship between neighboring molecules, and molecule‐substrate interaction

stabilizing the molecules on the surface. Many different intermolecular interactions can form

2D self‐assembly. Strong hydrogen bonds (such as carboxylic dimers) and strong directional

bonds including coordination at metal centers can produce porous assembly structures.

Generally, weaker bonds (such as van der Waals interactions and halogen‐halogen) result in a

close‐packed structure with the maximized areal molecular density. We will discuss molecular

self-assembly with these four interactions, respectively.

2.3.1 Metal Bonds Molecular Self-assembly

Metal bonds molecular self-assembly usually describe the 2D frameworks of a heterogeneous

system of metal and organic molecules by the metal-directed bonding, such as metal-

carboxylates bonding. It provided a new way to study the heterogenous system, including

44

mechanism, by STM. Johannes V.Barth et al. has reported 2D frameworks of metal and organic

molecules.12, 68 In this investigation, the formation of 2D Co-based coordination compounds

were reported in their STM observations on the reconstructed Au(111) surface (Figure 2.15).

After the deposition of terephthalic acid (TPA) molecules, the preorganized arrays of Co bilayer

islands reconstructed with the TPA molecules and formed 2D Co-based coordination

compounds. These findings indicate the widespread application of the concept of coordination

chemistry on the surface, which can be spatially limited by the use of template substrates, and

the potential to synthesize alignments that do not exist in bulk materials.

Figure. 2.15. (a) Array of cobalt clusters following evaporation of 0.14 ML Co on the Au(111) surface at

room temperature. The deposited atoms condense in bilayer islands at the elbow sites of the chevron

45

reconstruction. The Co dots contain on the average ∼200 atoms representing in situ nano-reservoirs for

the formation of metal-organic complexes with co-deposited carboxylic acids (STM image size 100 × 80

nm2, I =1.3 nA, V = 20 mV). (b) Complexation reaction following deposition of 0.3 ML TPA on a Co array

(0.08 ML, corresponding to ∼120 atoms per island) on Au(111) at room temperature. A minority of

hydrogen-bonded domains (A) coexists with the dominating metal-organic compounds (B) evolving

around residual Co dots (I = 0.6 nA, V = -0.7 V). (b) Array of cobalt clusters following evaporation of 0.14

ML Co on the Au(111) surface at room temperature. The deposited atoms condense in bilayer islands at

the elbow sites of the chevron reconstruction. The Co dots contain on the average ∼200 atoms

representing in situ nano-reservoirs for the formation of metal-organic complexes with co-deposited

carboxylic acids (STM image size 100 × 80 nm2, I = 1.3 nA, V = 20 mV). (c) Fully developed rectangular

metal–organic nanogrid with a Co–TPA stoichiometry of 1:1 following annealing at 330 K. (d) The model

at the right depicts the underlying dicobalt coordination motif with both axial chelating and equatorial

bridging metal center–carboxylate bonds.68 Copyright at 2009 reprinted with permission from Elsevier

Publishing Group.

2.3.2 Hydrogen Bonding Molecular Self-assembly

Hydrogen bonding is the main electrostatic attraction between a hydrogen (H) atom (the

hydrogen bond acceptor) and the second-row elements oxygen (O), fluorine (F), or nitrogen (N)

(the hydrogen bond donor).69 It is one of the most fundamental non-covalent forces in both

artificial systems (various soft materials, molecular self-assembly, etc.) and biological (DNA,

saccharides etc.). 2D molecular self-assembly is one of the best ways to investigate the

hydrogen bonding between molecules as we can directly ‘see’ the bonding between two atoms

by STM. In 2010, Perepichka published a review on supramolecular design with carboxylic

46

acids.13, 70 In this review, they use a carboxylic group, which is one of the most accessible and

simplest hydrogen bonding functionalities, to reveal the diversity of 2D architectures achieved

through self-assembly.70 Adjusting the (i) symmetry, (ii) valency and (iii) secondary interactions

(tailored by substituents), as well as (iv) the size of the carboxylic acid building block, they

showed control of the exact topology and the dimensionality of self-assembly. They can create

different supramolecular cycles, linear, hexagonal and other structures by this method.

Figure. 2.16. STM topographs of hexagonal networks from (a) trimesic acid (TMA), (b) 1,3,5-

benzenetribenzoic acid (BTB), and (c) 1,3,5-tri(4-carboxyphenylethynyl)-2,4,6-trimethylbenzene

(TCPETMB). All STM topographs are to scale and depict an area of 15 × 15 nm2 . Corresponding models

47

of a single supramolecular cavity bordered by six molecules are depicted below and demonstrate the

underlying building plan of these isotopological networks with dimers interconnected by 2-fold

hydrogen bonds between carboxylic groups as a structural unit.13 Copyright at 2009 reprinted with

permission from American Chemical Society.

2.3.3 Van der Waals Molecular Self-assembly

The van der Waals force is a distance-dependent force between molecules or atoms in

nanoscale physics, named after scientist Johannes Diderik van der Waals. This attraction does

not caused by chemical electronic bonds, compared to ionic or covalent bonds; it is relatively

weak, and so more impacted by disturbance.71 The Van der Waals force significantly decrease

at longer gap between interacting molecules or atoms.71 The Van der Waals force plays an

important role in many fields such as molecular self-assembly, structural biology,

nanotechnology, polymer science, condensed matter physics, and surface science. The Van Der

Waals interaction is the mechanism behind the self-assembly of the fullerene family, like C60

and Gd3@C80 (metallofulerene). The first STM images of ordered C60 were reported by Richard J.

Colton in 1993,14 while the first STM images of a metallofullerene was published by G.A.D.

Briggs in 2008 as the discover of metallofullerene is much later than C60.72

48

Figure. 2.17. (a) Occupied state STM image of a 2√3 × 2√3 R 30o C60 domain on Au(111) [7.3 X 7.3 nm,

—2 V bias (sample negative)]. (b) Occupied state image of a 38 X 38 C60 domain on Au(111) (9. 1 X 9. 1

nm, —2 V).14 (c) Filled states (VSB = −1.5V) and (b) empty states STM image (VSB = 1.5V) of close-packed

Sc3N@C80 on Au(111)/mica (It = 0.08 nA). The circled molecules appear bright in filled states and dark in

empty states. Vacancies always show as black holes.72 Copyright at 1993 reprinted with permission from

American Physical Society. Copyright at 2008 reprinted with permission from IOP Publishing Group.

49

2.3.4 Halogen‐halogen Molecular Self-assembly

Halogen bonding is very similar to hydrogen bonding. A similar relationship can easily be get

between hydrogen and halogen bonding. There is an electron acceptor and donor relationship

in both types of bonding. The species, which acts as the electron donor or acceptor, is the only

difference between these two bonding. In halogen bonding, a halogen atom forms a non-

covalent interaction by acting as the electron acceptor and accepting density of electron from

an electron donor.73 While in hydrogen bonding, a hydrogen atom is the electron acceptor. At

the same time, the covalent bond between halogen atom (X) or hydrogen atom (H) and other

atom (D) weakens, therefore the electron density on X or H is reduced.73 Electron density

transfer leads to a penetration of the Van der Waals volumes. Recently, a paper discussed the

self-assembly behavior of hexakis(4-iodophenyl)benzene (HPBI) through the halogen-halogen

bonding.15 In this investigation, they report a D6h symmetric molecule HPBI, is used as a basis

component to build two types of periodic self-assembly phases on Au(111). These self-assembly

phases have the different packing densities but same lattice symmetry; the change in packing

density is due to networks build by halogen bonds74-75 and halogen–Au coordination, as

confirmed by STM experiment and DFT calculations. It is worth noting that the self-similarity of

these two patterns facilitates the interweaving of the patterns to form higher order self-

assembly phases. This mechanism provides a new method for building complex 2D mosaics

through intermolecular interactions and molecular-substrate interactions.76-77

50

Figure. 2.18. The two ordered self-assembly phases of HPBI on Au(111). (a) The chemical structure of

HPBI. Carbon, grey; iodine, red; hydrogen, white. (b) Large-scale STM image of the coexisting α and β

phases. The crystalline axes of the Au substrate are labelled. (c, d) STM images of the pure α (c) and β (d)

phase domains. The white arrows in c designate the herringbone structure in the α phase; the α and β

phase unit cells are represented by red and blue rhombi, respectively. (e, f) Magnified STM images of

the α (e) and β (f) phases. HPBI molecules are labelled by the dashed white and green circles. The

dashed white triangle in e highlights the I–I trimer. The solid yellow circles in f highlight the adatoms.15

Copyright at 2018 reprinted with permission from Nature Publishing Group.

51

2.4 Molecular Self-assembly on Graphene

To date, most studied self-assemble molecules on graphene have focused on three molecules:

phthalocyanine (and its metal coordination complexes), C60 fullerenes and perylene‐3,4,9,10‐

tetracarboxylic dianhydride (PTCDA) (Figure 2.19). These molecules are well-studied

semiconductor molecules whose structure is optimal for π−π interactions with the underlying

graphene. Besides, the self-assembly behavior of these molecules on graphite or metal surface,

has been well studied. Thus, it is more useful to study these molecule on graphene, as

graphene and graphite have some similarities. What’s more, PTCDA, PC and C60 are

semiconductor molecules that are widely used in organic solar cell and graphene is an excellent

transparent electrode for organic solar cells. Therefore, studies of their behaviors on graphene

combine their adsorption and energetics studies with the electronically modified probe

introduced by their presence and is very important for future organic solar cells and graphene-

based molecule device.

52

Figure. 2.19. The molecular structure of (a) Phthalocyanine, (b) C60 and (c) PTCDA.

2.4.1 PTCDA

PTCDA is a typical n‐type organic semiconductor. Due to its potential applications in graphene-

based organic devices, it has been widely characterized on surface of graphene .9 The first

paper was published in 2008, which is about PTCDA on bilayer graphene grown on a SiC

substrate at 4.7K.78 Then in 2009, the herringbone structure of PTCDA monolayer on a

monolayer epitaxy graphene substrate at room temperature was discovered and is similar

geometrically to the structure formed on HOPG.18

53

Figure. 2.20. (a) PTCDA monolayer on bilayer graphene at T=4.7 K. The shadow-like structure

originates from the SiC interface layer below bilayer graphene. UT=1.5V, IT=3.8pA. (b) Close-up of (a) One

clearly recognizes the assembly of the PTCDA molecules. UT=1V, IT=3.8pA.78 (c) Monolayer coverage of

PTCDA on epitaxial graphene. (d) Molecular-resolution STM image of the PTCDA monolayer. The PTCDA

molecular structure and unit cell outline are overlaid. The monolayer continuously follows the graphene

sheet over the SiC step edge.18 Copyright at 2008 reprinted with permission from Wiley-VCH Publishing

Group. Copyright at 2009 reprinted with permission from Nature Publishing Group.

Recently, several theoretical calculations by density functional theory were published.13, 79-82 By

noncovalent stacking with aromatic molecules through π-π interaction, they can adjust the

54

electronic properties of graphene.82 Recent literatures report on regarding surface patterning,

surface doping, bandgap engineering, and applications in nanodevices, especially the field-

effect transistors (FETs) by different kinds of molecules (functioning as either an electron

acceptor or donor when absorbed on graphene).82

Figure. 2.21. (a) Calculated geometry configuration of monolayer PTCDA molecules on a

graphene/Pt(111). (b) Experimental STS spectra on one monolayer PTCDA on graphene/Pt(111). (c) Local

Density of States (LDOS) around the transport gap on PTCDA in the DFT calculations for the

PTCDA/graphene/Pt(111) system. Two peaks are clearly resolved both in the experimental and in the

theoretical spectra whose origin can be ascribed to the HOMO and LUMO of the PTCDA molecule.82

Copyright at 2014 reprinted with permission from American Chemical Society.

55

2.4.2 C60

STM measurements of monolayer C60 deposited onto epitaxial graphene on SiC show that the a

hexagonal close packed (hcp) layer of molecules was formed, which uniformly occupied all

regions of graphene (Figure 2.22a and b).83 In the submonolayer regime, they also measure an

energy gap on graphene around 3.5 eV for the C60 molecules. This indicate that, compared to

C60 adsorbed on metallic substrates, a pretty smaller amount of charge transfer from the

graphene to C60 and substrate induced screening.83 Additional paper reported the formation of

a hcp C60 monolayer that occupied on graphene moiré pattern deposited at 600 K and imaged

at liquid He temperature (~5K) (Figure 2. 22c and d).84 Besides, a paper was published on the

using the graphene moiré pattern to the capture C60 on homoepitaxy of graphene. Coverages of

C60 in the range 0.04–0.4 ML indicates that the spots within the moiré superlattice are

continuously occupied (Figure 2.22e-k): hcp bottom spot is first filed with a molecules, then six

closest molecules, then the fcc bottom spot is filled, then moiré top spots, and finally six other

closest molecules are arranged around the top spots.85 Furthermore, the formation of the C60

monolayer will be affected by the graphene substrate. In 2015, a quasi-1D C60 chain structure

with widths of two to three molecules formed on CVD grown graphene on Cu was reported

(Figure 2.22e and f).16 The formation of this one-dimensional chain is mainly due to the rippled

graphene showing well-defined linear periodic modulation with a 0.75nm spatial modulation

frequency.

56

Figure. 2.22. (a) STM topographic images of the initial stages of growth of C60 molecules adsorbed on a

submonolayer of epitaxial graphene on SiC. (b) a close-up view of the blue box indicated in (a) displaying

the single vacancy of a C60 molecule and the domain boundary between molecular islands. Images (a and

b) were acquired with I = 20 pA and V = −2 V.83 (c) Large-area STM topography of substrate

commensurate growth of C60 molecules on G/Ru. Right part is a higher terrace of Ru(0001) surface. (Vs =

3.0 V and I = 0.05 nA) (d) Zoom-in image of the supramolecular structure. The unit cells of the underlying

substrate and molecular lattice are outlined by large and small rhombuses, respectively. (2.0 = V and 0.1

= nA).84 (e) High resolution image of bimolecular and trimolecular C60 chains. Within the chains, the C60-

C60 intermolecular spacing is ~1.0nm, and the interchain distance, defined as the distance between the

centers of adjacent C60 rows belonging to neighboring chains, is 1.23 nm (Vs= 1.95 V, I= 0.50nA). (f) A

line profile along the close packed orientation as marked with the dashed blue line in (e).16 (g) Scheme

of an individual C60 molecule preferentially trapped in the Chcp valley at RT and its corresponding STM

image as shown in (i). (h) Scheme of six C60 molecules attached to the trapped C60 as a seed for the

57

nucleation of monolayer C60 islands; C60-C60 cohesive energy increases. (j) RT freezing of the thermal

motions of C60 in the Chcp valleys once a C60 monolayer is formed. (k) All C60 molecules trapped in Chcp

valleys display a dumbbell shape, aligning along the <1120> directions. The bright lobes in the dumbbell-

shaped correspond to pentagons of the C60 cage at positive sample bias, which suggests C60 orients with

the 6:6 bond (the C-C bond between two carbon hexagons) facing upward, as shown in the right top of

(k).85 Copyright at 2012 reprinted with permission from American Chemical Society. Copyright at 2012

reprinted with permission from American Institute of Physics. Copyright at 2015 reprinted with

permission from Nature Publishing Group. Copyright at 2012 reprinted with permission from American

Chemical Society.

2.4.3 Phthalocyanines

Phthalocyanines (shown as unsubstituted phathalocyanine, H2Pc, in Figure 2.19) can be used as

a π‐conjugated organic cell for metal atoms (FePc20, 86-87, CoPc19, F16CoPc19, NiPc, ClAlPc88) fixed

at the center of the molecule. In this style, they are called metal‐pthalocyanines (M‐Pc), and

their spin states and orbital energetics are highly influenced by the type of metal center and by

possible side groups. They lie flat on the surface of HOPG, and lateral stabilization is provided

by the Van der Waals interaction between H‐terminated M‐Pc for their close packed self‐

assembled structure. At low coverage (less than 5%), FePc molecules prefer to be deposited on

the fcc spots of the graphene moiré pattern.20 As coverage increases, they continuing occupy

more fcc regions until an ordered 2D superlattice of separated molecules is formed on the

graphene (Figure 2.23b). Once the fcc regions are all occupied, the remaining FePc tends to stay

at the hcp regions rather than the top regions (Figure 2.23c).20 At a higher coverage (75%), FePc

58

forms a Kagome lattice (consisting of the vertices and edges of the trihexagonal tiling) (Figure

2.23d) with the open networks occupying both the hcp and fcc spots of graphene, with the top

spots whole empty.86 For a fully covered monolayer, all Pc molecules create a square lattice

including FePc, CoPc, ClAlPc and other MPcs (Figure 2.23 f-j).87 In 2011, a DFT group calculated

the optimal density of states (DOS) and configuration of the adsorption system (solid black thick

line), and that projected onto the Pc molecule (solid blue line) and the graphene substrate

(solid black thin line) (Figure 2.23 k and l).89 The two characteristics above and below Fermi

Energy (EF) in the total DOS almost are caused by the contributions of LUMO and HOMO states

of F16CuPc, including, three spin-down and two spin-up orbitals (for LUMOs)and one spin-down

and two spin-up orbital (for HOMOs), respectively.89

59

Figure. 2.23. (a) High-resolution image (U = 0.1 V, I = 0.5 nA) of three distinct regions of graphene, top,

fcc, and hcp, marked by triangles and dashed and solid hexagons, respectively. (b) STM image (U = −2.0

V, I = 0.05 nA) revealing that molecules first adsorb at the fcc regions. (c) Sequences of STM images (U =

−2.0 V, I = 0.05 nA) of FePc molecules with increasing coverage.20 (d) Details of the Kagome lattice of

FePc. A trihexagonal tiling is highlighted. The unit cell of the Kagome lattice is marked with blue lines. (e)

Structural model of the Kagome lattice showing molecular orientation disorder.85 (f) STM image of a

close-packed FePc molecular island in showing a square lattice (indicated by the dashed square).80 (g)

STM image of a close-packed CoPc molecular island in a showing a square lattice. (h) STM image of a

close-packed CuPc molecular island in a showing a square lattice. (i) STM image of a close-packed

F16CoPc molecular island showing a square lattice.19 (j) STM images of self-assembled ClAlPc molecular

arrays of the first layer on graphene. The ClAlPc arrays show continuous films across the Cu steps. Insets

60

show the magnified images (Vtip = 2.0 V, I = 75 pA).88 (k) The optimized configuration for the

F16CuPc/graphene [(3,4)×(4,3)] system. (l) Total DOS (thick black solid line) and projected DOS on the Pc

molecule (blue solid line), on graphene (thin black solid line), upon F16CuPc adsorption on graphene, and

the DOS of the isolated graphene (black dashed dotted line). The vertical dotted line shows the Fermi

level.89 Copyright at 2011 reprinted with permission from American Physical Society. Copyright at 2009

reprinted with permission from American Chemical Society. Copyright at 2012 reprinted with permission

from American Chemical Society. Copyright at 2013 reprinted with permission from American Chemical

Society. Copyright at 2011 reprinted with permission from American Institute of Physics.

References:

1. Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.; Zhang, Y.; Dubonos, S. V.; Grigorieva, I. V.; Firsov, A. A., Electric field effect in atomically thin carbon films. Science 2004, 306 (5696), 666-669. 2. Saito, R.; Fujita, M.; Dresselhaus, G.; Dresselhaus, M. S., Electronic-Structure of Chiral Graphene Tubules. Appl Phys Lett 1992, 60 (18), 2204-2206. 3. Allen, M. J.; Tung, V. C.; Kaner, R. B., Honeycomb Carbon: A Review of Graphene. Chem Rev 2010, 110 (1), 132-145. 4. Choi, W.; Lahiri, I.; Seelaboyina, R.; Kang, Y. S., Synthesis of Graphene and Its Applications: A Review. Crit Rev Solid State 2010, 35 (1), 52-71. 5. Ohta, T.; Bostwick, A.; Seyller, T.; Horn, K.; Rotenberg, E., Controlling the electronic structure of bilayer graphene. Science 2006, 313 (5789), 951-954. 6. Lee, H.; Paeng, K.; Kim, I. S., A review of doping modulation in graphene. Synthetic Met 2018, 244, 36-47. 7. Zhao, J.; Zhang, G. Y.; Shi, D. X., Review of graphene-based strain sensors. Chinese Phys B 2013, 22 (5). 8. Dutta, S.; Pati, S. K., Novel properties of graphene nanoribbons: a review. J Mater Chem 2010, 20 (38), 8207-8223. 9. MacLeod, J. M.; Rosei, F., Molecular Self-Assembly on Graphene. Small 2014, 10 (6), 1038-1049. 10. Munoz, R.; Gomez-Aleixandre, C., Review of CVD Synthesis of Graphene. Chem Vapor Depos 2013, 19 (10-12), 297-322. 11. Akinwande, D.; Brennan, C. J.; Bunch, J. S.; Egberts, P.; Felts, J. R.; Gao, H. J.; Huang, R.; Kim, J. S.; Li, T.; Li, Y.; Liechti, K. M.; Lu, N. S.; Park, H. S.; Reed, E. J.; Wang, P.; Yakobson, B. I.; Zhang, T.; Zhang, Y. W.; Zhou, Y.; Zhu, Y., A review on mechanics and mechanical properties of 2D materials-Graphene and beyond. Extreme Mech Lett 2017, 13, 42-77. 12. Clair, S.; Pons, S.; Fabris, S.; Baroni, S.; Brune, H.; Kern, K.; Barth, J. V., Monitoring two-dimensional coordination reactions: Directed assembly of Co-terephthalate nanosystems on Au(111). J Phys Chem B 2006, 110 (11), 5627-5632.

61

13. Lackinger, M.; Heckl, W. M., Carboxylic Acids: Versatile Building Blocks and Mediators for Two-Dimensional Supramolecular Self-Assembly. Langmuir 2009, 25 (19), 11307-11321. 14. Altman, E. I.; Colton, R. J., Determination of the Orientation of C-60 Adsorbed on Au(111) and Ag(111). Phys Rev B 1993, 48 (24), 18244-18249. 15. Cheng, F.; Wu, X. J.; Hu, Z. X.; Lu, X. F.; Ding, Z. J.; Shao, Y.; Xu, H.; Ji, W.; Wu, J. S.; Loh, K. P., Two-dimensional tessellation by molecular tiles constructed from halogen-halogen and halogen-metal networks. Nat Commun 2018, 9. 16. Chen, C. H.; Zheng, H. S.; Mills, A.; Heflin, J. R.; Tao, C. G., Temperature Evolution of Quasi-one-dimensional C-60 Nanostructures on Rippled Graphene. Sci Rep-Uk 2015, 5. 17. Meissner, M.; Gruenewald, M.; Sojka, F.; Udhardt, C.; Forker, R.; Fritz, T., Highly ordered growth of PTCDA on epitaxial bilayer graphene. Surf Sci 2012, 606 (21-22), 1709-1715. 18. Wang, Q. H.; Hersam, M. C., Room-temperature molecular-resolution characterization of self-assembled organic monolayers on epitaxial graphene. Nat Chem 2009, 1 (3), 206-211. 19. Jarvinen, P.; Hamalainen, S. K.; Ijas, M.; Harju, A.; Liljeroth, P., Self-Assembly and Orbital Imaging of Metal Phthalocyanines on a Graphene Model Surface. J Phys Chem C 2014, 118 (24), 13320-13325. 20. Zhang, H. G.; Sun, J. T.; Low, T.; Zhang, L. Z.; Pan, Y.; Liu, Q.; Mao, J. H.; Zhou, H. T.; Guo, H. M.; Du, S. X.; Guinea, F.; Gao, H. J., Assembly of iron phthalocyanine and pentacene molecules on a graphene monolayer grown on Ru(0001). Phys Rev B 2011, 84 (24). 21. Kuzmenko, A. B.; van Heumen, E.; Carbone, F.; van der Marel, D., Universal optical conductance of graphite. Phys Rev Lett 2008, 100 (11). 22. Geim, A. K.; Novoselov, K. S., The rise of graphene. Nat Mater 2007, 6 (3), 183-191. 23. Huang, Y.; Sutter, E.; Shi, N. N.; Zheng, J. B.; Yang, T. Z.; Englund, D.; Gao, H. J.; Sutter, P., Reliable Exfoliation of Large-Area High-Quality Flakes of Graphene and Other Two-Dimensional Materials. Acs Nano 2015, 9 (11), 10612-10620. 24. Sutter, P. W.; Flege, J. I.; Sutter, E. A., Epitaxial graphene on ruthenium. Nat Mater 2008, 7 (5), 406-411. 25. Yang, W.; Chen, G. R.; Shi, Z. W.; Liu, C. C.; Zhang, L. C.; Xie, G. B.; Cheng, M.; Wang, D. M.; Yang, R.; Shi, D. X.; Watanabe, K.; Taniguchi, T.; Yao, Y. G.; Zhang, Y. B.; Zhang, G. Y., Epitaxial growth of single-domain graphene on hexagonal boron nitride. Nat Mater 2013, 12 (9), 792-797. 26. Berger, C.; Song, Z. M.; Li, T. B.; Li, X. B.; Ogbazghi, A. Y.; Feng, R.; Dai, Z. T.; Marchenkov, A. N.; Conrad, E. H.; First, P. N.; de Heer, W. A., Ultrathin epitaxial graphite: 2D electron gas properties and a route toward graphene-based nanoelectronics. J Phys Chem B 2004, 108 (52), 19912-19916. 27. Kim, K. S.; Zhao, Y.; Jang, H.; Lee, S. Y.; Kim, J. M.; Kim, K. S.; Ahn, J. H.; Kim, P.; Choi, J. Y.; Hong, B. H., Large-scale pattern growth of graphene films for stretchable transparent electrodes. Nature 2009, 457 (7230), 706-710. 28. Reina, A.; Jia, X. T.; Ho, J.; Nezich, D.; Son, H. B.; Bulovic, V.; Dresselhaus, M. S.; Kong, J., Large Area, Few-Layer Graphene Films on Arbitrary Substrates by Chemical Vapor Deposition. Nano Lett 2009, 9 (1), 30-35. 29. Nair, R. R.; Blake, P.; Grigorenko, A. N.; Novoselov, K. S.; Booth, T. J.; Stauber, T.; Peres, N. M. R.; Geim, A. K., Fine structure constant defines visual transparency of graphene. Science 2008, 320 (5881), 1308-1308. 30. Meric, I.; Han, M. Y.; Young, A. F.; Ozyilmaz, B.; Kim, P.; Shepard, K. L., Current saturation in zero-bandgap, topgated graphene field-effect transistors. Nat Nanotechnol 2008, 3 (11), 654-659. 31. Novoselov, K. S.; Jiang, D.; Schedin, F.; Booth, T. J.; Khotkevich, V. V.; Morozov, S. V.; Geim, A. K., Two-dimensional atomic crystals. P Natl Acad Sci USA 2005, 102 (30), 10451-10453. 32. Datta, S. S.; Strachan, D. R.; Khamis, S. M.; Johnson, A. T. C., Crystallographic etching of few-layer graphene. Nano Lett 2008, 8 (7), 1912-1915.

62

33. Gass, M. H.; Bangert, U.; Bleloch, A. L.; Wang, P.; Nair, R. R.; Geim, A. K., Free-standing graphene at atomic resolution. Nat Nanotechnol 2008, 3 (11), 676-681. 34. Huc, V.; Bendiab, N.; Rosman, N.; Ebbesen, T.; Delacour, C.; Bouchiat, V., Large and flat graphene flakes produced by epoxy bonding and reverse exfoliation of highly oriented pyrolytic graphite. Nanotechnology 2008, 19 (45). 35. Li, X. S.; Cai, W. W.; An, J. H.; Kim, S.; Nah, J.; Yang, D. X.; Piner, R.; Velamakanni, A.; Jung, I.; Tutuc, E.; Banerjee, S. K.; Colombo, L.; Ruoff, R. S., Large-Area Synthesis of High-Quality and Uniform Graphene Films on Copper Foils. Science 2009, 324 (5932), 1312-1314. 36. Edwards, R. S.; Coleman, K. S., Graphene Film Growth on Polycrystalline Metals. Accounts Chem Res 2013, 46 (1), 23-30. 37. Castro Neto, A. H.; Guinea, F.; Peres, N. M. R.; Novoselov, K. S.; Geim, A. K., The electronic properties of graphene. Rev Mod Phys 2009, 81 (1), 109-162. 38. Pan, Y.; Shi, D. X.; Gao, H. J., Formation of graphene on Ru(0001) surface. Chinese Phys 2007, 16 (11), 3151-3153. 39. Gottardi, S.; Muller, K.; Bignardi, L.; Moreno-Lopez, J. C.; Pham, T. A.; Ivashenko, O.; Yablonskikh, M.; Barinov, A.; Bjork, J.; Rudolf, P.; Stohr, M., Comparing Graphene Growth on Cu(111) versus Oxidized Cu(111). Nano Lett 2015, 15 (2), 917-922. 40. Tapaszto, L.; Dumitrica, T.; Kim, S. J.; Nemes-Incze, P.; Hwang, C.; Biro, L. P., Breakdown of continuum mechanics for nanometre-wavelength rippling of graphene. Nat Phys 2012, 8 (10), 739-742. 41. Zhu, Y. W.; Murali, S.; Cai, W. W.; Li, X. S.; Suk, J. W.; Potts, J. R.; Ruoff, R. S., Graphene and Graphene Oxide: Synthesis, Properties, and Applications. Adv Mater 2010, 22 (35), 3906-3924. 42. Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.; Katsnelson, M. I.; Grigorieva, I. V.; Dubonos, S. V.; Firsov, A. A., Two-dimensional gas of massless Dirac fermions in graphene. Nature 2005, 438 (7065), 197-200. 43. Bolotin, K. I.; Sikes, K. J.; Jiang, Z.; Klima, M.; Fudenberg, G.; Hone, J.; Kim, P.; Stormer, H. L., Ultrahigh electron mobility in suspended graphene. Solid State Commun 2008, 146 (9-10), 351-355. 44. Frank, I. W.; Tanenbaum, D. M.; Van der Zande, A. M.; McEuen, P. L., Mechanical properties of suspended graphene sheets. J Vac Sci Technol B 2007, 25 (6), 2558-2561. 45. Poot, M.; van der Zant, H. S. J., Nanomechanical properties of few-layer graphene membranes. Appl Phys Lett 2008, 92 (6). 46. Lee, C.; Wei, X. D.; Kysar, J. W.; Hone, J., Measurement of the elastic properties and intrinsic strength of monolayer graphene. Science 2008, 321 (5887), 385-388. 47. Gomez-Navarro, C.; Burghard, M.; Kern, K., Elastic properties of chemically derived single graphene sheets. Nano Lett 2008, 8 (7), 2045-2049. 48. Ni, Z. H.; Wang, H. M.; Kasim, J.; Fan, H. M.; Yu, T.; Wu, Y. H.; Feng, Y. P.; Shen, Z. X., Graphene thickness determination using reflection and contrast spectroscopy. Nano Lett 2007, 7 (9), 2758-2763. 49. Peres, N. M. R.; Guinea, F.; Castro Neto, A. H., Electronic properties of disordered two-dimensional carbon. Phys Rev B 2006, 73 (12). 50. Gusynin, V. P.; Sharapov, S. G.; Carbotte, J. P., Unusual microwave response of Dirac quasiparticles in graphene. Phys Rev Lett 2006, 96 (25). 51. Peres, N. M. R., The transport properties of graphene. J Phys-Condens Mat 2009, 21 (32). 52. Cao, Y.; Fatemi, V.; Fang, S.; Watanabe, K.; Taniguchi, T.; Kaxiras, E.; Jarillo-Herrero, P., Unconventional superconductivity in magic-angle graphene superlattices. Nature 2018, 556 (7699), 43-+. 53. Cao, Y.; Fatemi, V.; Demir, A.; Fang, S.; Tomarken, S. L.; Luo, J. Y.; Sanchez-Yamagishi, J. D.; Watanabe, K.; Taniguchi, T.; Kaxiras, E.; Ashoori, R. C.; Jarillo-Herrero, P., Correlated insulator behaviour at half-filling in magic-angle graphene superlattices. Nature 2018, 556 (7699), 80-+.

63

54. Chen, G. R.; Jiang, L. L.; Wu, S.; Lyu, B.; Li, H. Y.; Chittari, B. L.; Watanabe, K.; Taniguchi, T.; Shi, Z. W.; Jung, J.; Zhang, Y. B.; Wang, F., Evidence of a gate-tunable Mott insulator in a trilayer graphene moire superlattice. Nat Phys 2019, 15 (3), 237-241. 55. Wu, Z. S.; Pei, S. F.; Ren, W. C.; Tang, D. M.; Gao, L. B.; Liu, B. L.; Li, F.; Liu, C.; Cheng, H. M., Field Emission of Single-Layer Graphene Films Prepared by Electrophoretic Deposition. Adv Mater 2009, 21 (17), 1756-+. 56. Shao, Y. Y.; Wang, J.; Wu, H.; Liu, J.; Aksay, I. A.; Lin, Y. H., Graphene Based Electrochemical Sensors and Biosensors: A Review. Electroanal 2010, 22 (10), 1027-1036. 57. Yoon, H. J.; Jun, D. H.; Yang, J. H.; Zhou, Z. X.; Yang, S. S.; Cheng, M. M. C., Carbon dioxide gas sensor using a graphene sheet. Sensor Actuat B-Chem 2011, 157 (1), 310-313. 58. De Arco, L. G.; Zhang, Y.; Schlenker, C. W.; Ryu, K.; Thompson, M. E.; Zhou, C. W., Continuous, Highly Flexible, and Transparent Graphene Films by Chemical Vapor Deposition for Organic Photovoltaics. Acs Nano 2010, 4 (5), 2865-2873. 59. Yoo, E.; Kim, J.; Hosono, E.; Zhou, H.; Kudo, T.; Honma, I., Large reversible Li storage of graphene nanosheet families for use in rechargeable lithium ion batteries. Nano Lett 2008, 8 (8), 2277-2282. 60. Bigelow, W. C.; Pickett, D. L.; Zisman, W. A., Oleophobic Monolayers .1. Films Adsorbed from Solution in Non-Polar Liquids. J Coll Sci Imp U Tok 1946, 1 (6), 513-538. 61. Gast, A. P., Physical chemistry of surfaces. New York, A Wiley-Interscience Publication 1997. 62. Love, J. C.; Estroff, L. A.; Kriebel, J. K.; Nuzzo, R. G.; Whitesides, G. M., Self-assembled monolayers of thiolates on metals as a form of nanotechnology. Chem Rev 2005, 105 (4), 1103-1169. 63. Nuzzo, R. G.; Allara, D. L., Adsorption of Bifunctional Organic Disulfides on Gold Surfaces. J Am Chem Soc 1983, 105 (13), 4481-4483. 64. Maoz, R.; Sagiv, J., On the Formation and Structure of Self-Assembling Monolayers .1. A Comparative Atr-Wetability Study of Langmuir-Blodgett and Adsorbed Films on Flat Substrates and Glass Microbeads. J Colloid Interf Sci 1984, 100 (2), 465-496. 65. Lehn, J. M., Supramolecular Chemistry-Concepts and Chemistry. VCH: Wiley: Weinheim 1995. 66. Elemans, J. A. A. W.; Lei, S. B.; De Feyter, S., Molecular and Supramolecular Networks on Surfaces: From Two-Dimensional Crystal Engineering to Reactivity. Angew Chem Int Edit 2009, 48 (40), 7298-7332. 67. Binning, G.; Rohrer, H.; Gerber, C.; Weibel, E., Surface Studies by Scanning Tunneling Microscopy. Phys Rev Lett 1982, 49 (1), 57-61. 68. Barth, J. V., Fresh perspectives for surface coordination chemistry. Surf Sci 2009, 603 (10-12), 1533-1541. 69. Ivasenko, O.; Perepichka, D. F., Mastering fundamentals of supramolecular design with carboxylic acids. Common lessons from X-ray crystallography and scanning tunneling microscopy. Chem Soc Rev 2011, 40 (1), 191-206. 70. https://en.wikipedia.org/wiki?curid=13609. 71. C. Nörenberg, D. F. L., D. Cattaneo, K. Porfyrakis, A. Li Bassi, C.S. Casari, M. Passoni, J.H.G. Owen, G.A.D. Briggs, Self-assembly and electronic effects of Er3N@C80 and Sc3N@C80 on Au(111) and Ag/Si(111) surfaces. J. Phys. Conf. Ser. 2008, 100, 052080. 72. https://en.wikipedia.org/wiki/Van_der_Waals_force. 73. https://en.wikipedia.org/wiki?curid=14350687. 74. Gutzler, R.; Ivasenko, O.; Fu, C. Y.; Brusso, J. L.; Rosei, F.; Perepichka, D. F., Halogen bonds as stabilizing interactions in a chiral self-assembled molecular monolayer. Chem Commun 2011, 47 (33), 9453-9455. 75. Gutzler, R.; Fu, C. Y.; Dadvand, A.; Hua, Y.; MacLeod, J. M.; Rosei, F.; Perepichka, D. F., Halogen bonds in 2D supramolecular self-assembly of organic semiconductors. Nanoscale 2012, 4 (19), 5965-5971.

64

76. Bonifazi, D.; Mohnani, S.; Llanes-Pallas, A., Supramolecular Chemistry at Interfaces: Molecular Recognition on Nanopatterned Porous Surfaces. Chem-Eur J 2009, 15 (29), 7004-7025. 77. De Feyter, S.; De Schryver, F. C., Two-dimensional supramolecular self-assembly probed by scanning tunneling microscopy (vol 32, pg 139, 2003). Chem Soc Rev 2003, 32 (6), 393-393. 78. Lauffer, P.; Emtsev, K. V.; Graupner, R.; Seyller, T.; Ley, L., Molecular and electronic structure of PTCDA on bilayer graphene on SiC(0001) studied with scanning tunneling microscopy. Phys Status Solidi B 2008, 245 (10), 2064-2067. 79. Tian, X. Q.; Xu, J. B.; Wang, X. M., Self-Assembly of PTCDA Ultrathin Films on Graphene: Structural Phase Transition and Charge Transfer Saturation. J Phys Chem C 2010, 114 (49), 20917-20924. 80. Kozlov, S. M.; Vines, F.; Gorling, A., Bandgap Engineering of Graphene by Physisorbed Adsorbates. Adv Mater 2011, 23 (22-23), 2638-+. 81. Zhang, Z. X.; Huang, H. L.; Yang, X. M.; Zang, L., Tailoring Electronic Properties of Graphene by pi-pi Stacking with Aromatic Molecules. J Phys Chem Lett 2011, 2 (22), 2897-2905. 82. Martinez-Galera, A. J.; Nicoara, N.; Martinez, J. I.; Dappe, Y. J.; Ortega, J.; Gomez-Rodriguez, J. M., Imaging Molecular Orbitals of PTCDA on Graphene on Pt(111): Electronic Structure by STM and First-Principles Calculations. J Phys Chem C 2014, 118 (24), 12782-12788. 83. Cho, J.; Smerdon, J.; Gao, L.; Suzer, O.; Guest, J. R.; Guisinger, N. P., Structural and Electronic Decoupling of C-60 from Epitaxial Graphene on SiC. Nano Lett 2012, 12 (6), 3018-3024. 84. Li, G.; Zhou, H. T.; Pan, L. D.; Zhang, Y.; Mao, J. H.; Zou, Q.; Guo, H. M.; Wang, Y. L.; Du, S. X.; Gao, H. J., Self-assembly of C-60 monolayer on epitaxially grown, nanostructured graphene on Ru(0001) surface. Appl Phys Lett 2012, 100 (1). 85. Lu, J.; Yeo, P. S. E.; Zheng, Y.; Yang, Z. Y.; Bao, Q. L.; Gan, C. K.; Loh, K. P., Using the Graphene Moire Pattern for the Trapping of C-60 and Homoepitaxy of Graphene. Acs Nano 2012, 6 (1), 944-950. 86. Mao, J. H.; Zhang, H. G.; Jiang, Y. H.; Pan, Y.; Gao, M.; Xiao, W. D.; Gao, H. J., Tunability of Supramolecular Kagome Lattices of Magnetic Phthalocyanines Using Graphene-Based Moire Patterns as Templates. J Am Chem Soc 2009, 131 (40), 14136-+. 87. Yang, K.; Xiao, W. D.; Jiang, Y. H.; Zhang, H. G.; Liu, L. W.; Mao, J. H.; Zhou, H. T.; Du, S. X.; Gao, H. J., Molecule-Substrate Coupling between Metal Phthalocyanines and Epitaxial Graphene Grown on Ru(0001) and Pt(111). J Phys Chem C 2012, 116 (26), 14052-14056. 88. Ogawa, Y.; Niu, T. C.; Wong, S. L.; Tsuji, M.; Wee, A. T. S.; Chen, W.; Ago, H., Self-Assembly of Polar Phthalocyanine Molecules on Graphene Grown by Chemical Vapor Deposition. J Phys Chem C 2013, 117 (42), 21849-21855. 89. Ren, J.; Meng, S.; Wang, Y. L.; Ma, X. C.; Xue, Q. K.; Kaxiras, E., Properties of copper (fluoro-)phthalocyanine layers deposited on epitaxial graphene. J Chem Phys 2011, 134 (19).

65

Chapter 3

Experimental methods

For the purpose of characterizing the morphology of molecules, atoms and hybrid structures at

the nanoscale, a tool with super-high resolution is absolutely essential to observe and record

the nanostructure. As we have already introduced in Chapter 1, all these hybrid structures are

typically at atomic scale, so this experimental tool should have the resolution of at least a

nanometer. Second of all, the structure should not be distorted or damaged during the

measuring process. Scanning tunneling microscopy (STM) and atomic force microscopy (AFM)

are ideally suited due to their capability of atomic resolution and weak interaction compared to

other measurement tools. An ultrahigh vacuum (UHV) STM1® STM from Omicron

Nanotechnology is one of our main instrument used in this thesis, and all the measurements

were done under a base pressure of 1 × 10-10 Torr. A Dimension Icon® AFM from Bruker is the

other equipment that we use during the experiments.

In this chapter, our main tools – STM and AFM will be introduced, including a brief history

development and the fundamental physical mechanism of them. Then we will show the main

designs for the Omicron STM and Bruker ‘Dimension Icon’ AFM in our lab. Finally, we will

explain the other important parts (sample preparation) employed in our experiments, including

the spin coating method and physical vapor deposition method.

66

3.1 Introduction to Atomic Force Microscope (AFM)

After the invention of STM in 1981, the AFM was invented by G. Binning and Ch. Gerber in

1986.1 Since the emergence of commercially available equipment by the end of that 1980s,

AFM has been an important tool for materials, biological and surface research.2-4 AFM exists in

the core facilities of most research universities and in many single researcher laboratories, and

is a standard feature of large corporation research laboratories.5 The AFM is an imaging device

to see the morphology of a surface in three-dimensions (3D) that relies on the atomic force

(Van der Waals forces) between a sharp tip, which has an average diameter of a few

nanometers, and the scanned sample. AFM can image most materials (soft or hard, natural or

synthetic, including biological materials such as biomolecules and cells) regardless of

conductivity or opaqueness. The sample is not only operated in air, but also in liquid

environments, as well as under vacuum. It owns good resolution on the order of 10-9 m, 1000

times higher than the optical microscope. Additionally, overall the scanning efficiency of the

AFM is much higher than STM even though their scanning speed is comparable, because of the

convenience of loading a new sample and AFM tips compared to the time-consuming process

of transferring samples or installing tips during the STM measurement. Therefore, it is a good

supplement for STM during our lab experiments, especially in certain cases, such as a polymer

or insulator surface.

67

Figure. 3.1. Schematic drawing shows the mechanism of imaging mode in AFM.

How exactly does AFM determine the height of a surface? As shown in Figure 3.1, when we

start scanning a sample, the AFM automatically approaches a sharp tip to contact the surface,

which mounted at the end of the cantilever, while measuring the vertical displacement.

However, this contact can be very subtle; in other word, the metaphor can be understood

literally. In most AFM designs (Figure 3.1), the tip is connected to a cantilever that bends under

the effect of force. Its behavior is that of a tip attached to a spring; a cantilever bent downward

or upward is that of an extended or compressed spring. The bending is usually measured by

reflecting the laser beam off the cantilever and reflecting it onto the photodiode, the output of

68

the photodiode measures the position of the laser spot. The vertical movement of the tip is

used to measured this cantilever bending, while the horizontal movement of the laser spot is

used to measure the lateral force that torque the tip. A vertical tip will be usually handled

during the measurement with subnanometer resolution on the order of hundreds of

nanometers. The deflection of the laser will be transformed to electrical information and will be

input to a feedback loop in the controller electronics. The atomic force will keep at a constant

value between tip and sample by a Z direction piezoelectric through this feedback loop. The Z

information of the Z piezo will be recorded as the Z information of the sample, as well as the X

and Y information. In the simplest way, the tip will be brought into contact with a surface,

started to move or scan laterally, and the vertical movement of the tip is measured as the

cantilever bents up and down to measure the surface height, while the tip is sliding on the

surface. By doing so, a surface topography image can established on a 2D grid of locations

across the surface: height versus X and Y, using a pseudocolor key.

3.1.1 Working Principle: Van der Waals Force

Typically, AFM system are operated in three modes: (i) contact mode, (ii), non-contact mode

and (iii) tapping mode. The non-contact mode is utilized to probe atomic forces of a sample by

moving the cantilever slightly off the surface of the sample and having the cantilever near its

resonance frequency or near its natural vibration frequency.6 In non-contact mode,

topographical information of the sample can be extracted by mounting the cantilever on a

piezoelectric element and measuring the deviation from its resonance frequency due to

69

attraction between tip and sample.7 While, the image of the sample is acquired in contact

mode contributing to monitor the interaction forces with the target sample with the cantilever

tip remaining in contact.8 By oscillating the cantilever tip at or near its intrinsic resonance

frequency while reducing the time of impact by cantilever tip on sample to the shortest, the

qualities of both the non-contact and contact modes is combined in tapping mode .9-10

In order to perform nanocharacterization, we need an atomic force or atomic phenomenon

that works in the nanoscale range. The mechanism of imaging in AFM is due to the Van der

Waals interaction between the tip end and sample. The van der Waals force is a distance-

dependent force between molecules or atoms in nanoscale physics, named after scientist

Johannes Diderik van der Waals, and can be described as shown in Figure 3.2 (the force-

displacement curve). These attractive forces, unlike covalent or ionic bonds, are not caused by a

chemical electronic bond; they are relatively weak, and so more impacted by disturbance. The

Van der Waals force dramatically decrease at longer distance between interacting molecules or

atoms. The attraction increases until the molecules are so close that the repulsion appear due

to electrostatic interaction between dipoles or multipoles. As the intermolecular distance

decreasing, this repulsion between the atoms progressively weaken the attraction. When the

distance between the molecules is a few angstroms, the interaction force becomes zero and

then becomes completely repulsive as the molecules are in contact.

70

Figure 3.2. Force – distance curve of a Van Der Waals force. The yellow part is the non-contact mode

region, the purple is the contact mode region and the green part is the intermittent contact (tapping)

mode region.

For the three modes, the working region is different. The non-contact mode is in the purely

repulsive region (yellow region in Figure 3.2). The contact mode is in the attractive region

(purple region in Figure 3.2). The tapping mode is alternating between the attractive and

repulsive region (green region in Figure 3.2). From Figure 3.2, we can see that the contact

mode is nearest to the surface with the largest interaction, the non-contact mode is the

farthest from the surface with the lowest interaction, while the tapping mode is the

intermittent between the contact mode region and the non-contact mode region.

71

3.1.2 Working Modes

There are three basic modes in AFM, contact, non-contact and tapping mode.

Contact mode: Contact mode is the mode that used for AFM when it invented. In this imaging

mode, the AFM tip is the closest to a surface among these three modes (the order of

Angstrom), owing to the repulsive predominant force with an average of 10-9 N. This mode is

called contact mode because the atoms in top of the AFM tip and the atoms in surface of the

sample come into such close place that overlapping in electronic orbitals exists. AFM tips used

for contact mode imaging are usually characterized by low spring constants such that the net

force the tip exerts on the surface being imaged is less than the interatomic forces between the

atoms in the surface. As we operate the AFM in contact mode, the deflection of the cantilever

need to be measured and compared to the desired value of deflection in a DC feedback

amplifier. A voltage is applied through the feedback amplifier to the piezoelectric to lower or

raise the cantilever to restore the desired deflection. The voltage that applied to the

piezoelectric is a measure of the characteristic height of the sample surface. Then the AFM

starts to scan. During the scanning process, the AFM can measure 256 points height

information and deformation information each line forward and backward 256 lines with a

settled scanning rate. The computer will display this information will be displayed as a function

of the X and Y position of the sample and save as an image. As the gradient and the amount of

the repulsive force of contact mode is largest among these three modes, the interaction

between tip and sample is very strong. Therefore, it is more suitable for a hard surface without

danger of scratching the sample.

72

Figure 3.3. (a) The schematic drawing of contact mode. (b) The schematic drawing of non-contact mode.

(c) The schematic drawing of tapping mode.11

Non-contact mode: Since the invention of AFM, although contact mode cannot achieve atomic

resolution stably, it has successfully obtained a number of excellent results at nanoscales level

or even the atomic level through simple contact measurements. In 1994 F.J. Giessibl

successfully obtained atomic resolution AFM images of the Si(111)-(77) surface by a non-

contact AFM by utilizing a frequency modulation detection method under a weak tip-sample

attraction at room temperature in UHV.12 Soon the non-contact mode successfully

accomplished atomically resolved images on various surfaces.13-15 Non-contact mode depends

73

on oscillating the tip near its resonant frequency. In this mode, the separation between tip and

surface of the sample is larger than it is in contact mode, while smaller than tapping mode. The

tip – sample separation is this mode is between tens to hundreds of Angstroms. The oscillation

amplitudes are a bit smaller than this. In this mode, the smallest separation between the tip

and the surface is larger than the smallest separation for both contact and tapping modes. A

strong interaction can be prevented between the tip and the surface by keeping the distance

relatively large between them. In this imaging mode, the system simultaneously monitors the

amplitude as well as the frequency of oscillation of the tip. Any changes in either the amplitude

or the frequency are used through the feedback system to control the piezoelectric scanner in

the Z direction as to minimize these changes and maintain the original amplitude and frequency

of oscillation. This keeps the separation between tip and surface constant, and the profile of

how the tip was moved to keep this constant separation yields the height profile of the sample.

Tapping mode: The technique of tapping mode is an important improvement in AFM. It is

another mode of imaging that depends on oscillating the tip near its resonant frequency, which

firstly was used in biological surfaces field in 1993.16 In tapping mode, the AFM tip is oscillated at

or near its resonant frequency at an optimum amplitude. The tips used for this mode of imaging

are characterized with higher spring constant to prevent them from oscillating out of control

with large amplitudes. The spring constants for tips used for tapping mode imaging range

between 0.1 Nm-1 to 200 Nm-1. The amplitudes of the oscillations are usually in the range of

hundreds of Angstroms. As the tip scans the surface while it is oscillating, the changes in the

surface topography will affect the amplitude of the oscillation. Monitoring these changes and

74

using this data through the feedback loop to extend or contract the piezoelectric scanner to

minimize these changes produces a topographic and phase map of the surface of the sample.

When the parameters are set properly for this imaging mode, the forces applied on the tip from

the surface can be minimized compared to contact mode. The working principle of tapping

mode is resonating the tip from contact to non-contact mode, which reduces the interaction

between tip and surface. The resonating of tip weakens the lateral force and strengthen the

penetrating force between tip and surface. Another advantage of tapping mode is phase

images. Phase images provide nanoscale information about surface, which often not provided

by other AFM modes. During scanning, phase imaging goes beyond simple topographical

mapping by mapping the phase of the cantilever oscillation to detect information about the

properties of adhesion, composition, viscoelasticity, friction and so on. Applications include

identifying contaminants, distinguishing area of high and low surface adhesion or hardness, and

mapping different components in composite materials. Therefore, tapping mode is suited for

soft surfaces or polymer samples, especially for the study of phase separation in polymers.

75

Figure. 3.4. (a) Resonance curve of a TappingMode cantilever above the surface. (b) Resonance curve of

a TappingMode cantilever close to the surface. Note that the resonance shifts to lower frequencies and

exhibits a drop in amplitude.17

3.1.3 Bruker Dimension Icon® AFM

Figure 3.5 (a) shows the main body of the Dimension Icon® AFM system we used for the

majority of our work in this dissertation. The scanner in this system consisted of controller,

main body, e-box and computer. The scanning part is the direct part related with scanning and

consists with three main part, scanning head, stage and camera. The maximization of XY range

of the scanner used in this system is 90 μm X 90 μm, while the maximum Z range for this system

76

is 13.5 μm. This system is capable of a variety of imaging modes and probing techniques.

Among the imaging modes that can be performed on this system are contact, tapping, non-

contact, lateral force, electric force microscopy, magnetic force microscopy and liquid modes.

The scanning part is on a pneumatic vibration insulation table operated with compressed air to

minimize the vibrational noise from the floor. The system also has an optional vibration

insulation hood. Several types of AFM tip holders exist to accommodate the various needs for

imaging different samples. One of the main types of probe holders that are widely used are

probe holders that is functional in air or gas (Figure 3.5 (c)). This probe holder is capable of

performing both contact and tapping mode imaging. Figure 3.5 (b) is the type of tip that we

mainly used in this dissertation with a spring constant of 7.4 Nm-1 and first longitudinal

resonating frequencies between 120 – 205 kHz.

Figure 3.5. (a) The Dimension Icon® is an AFM system that offers a variety of nanoscale characterization

and manipulation tools. It is equipped with a closed-loop scanner offering great precision for

77

repositioning the tip on the sample. It has a piezo scanner based on a piezotube.18 (b) The AFM tip type

that is used in our experiments is NCST® from Nano world with spring constant 7.4 N/m, first longitudinal

resonance frequencies between 120 – 205 kHz. (c) A probe holder that fits on the Dimension Icon®

system. The system employs a spring loaded lever system to hold the tip in place. This holder fits directly

on the piezoelectric scanner.

3.1.4 The correction of height of AFM measurement

During the operation of an AFM, the Z range will be highly affected by the force between tip

and surface. The effect of tip is the highest in contact mode, while for tapping mode, the effect

is weaker. In 2013 Wiktoria Walczyk published a paper describing how the force of the AFM

would affect the height and deformation of the surface of nanobubbles on an HOPG surface.19

Depending on the force of the tip, the height changes from 8 nm to 24 nm.

Figure 3.6. Left 8 images: PeakForce tapping (PFT) mode AFM height and deformation images of surface

nanobubbles on HOPG in water scanned with peak forces of F=0.24, 2.5, 9.7 and 27 nN. The line profile

78

(cross section) of height and deformation of surface nanobubbles with peak forces F=0.24, 2.5, 9.7 and

27 nN.19

3.2 Introduction to Scanning Tunneling Microscope (STM)

Scanning Tunneling Microscope (STM) was invented by G. Binnig, H. Rohrer at the IBM Zurich

Research Lab in 1981 by observing vacuum electron tunneling between a sharp tungsten tip

and a platinum sample.20-21 They won the Nobel Prize of 1986 in Physics for their invention of

STM. Since then, STM experienced revolutionary development broadening our perception

about atomic scale processes and structure and become one of the most advanced atomic

structure probing instruments in the world. With increasing availability of low-temperature

STM, local electronic properties can be investigated with unprecedented space and energy

resolution which opens the vista to completely new applications. The sharp tip can be

considered to be a powerful local probe, allowing one to measure physical properties of

materials on a small scale by using a variety of different spectroscopic methods. A variety of

STMs with different sizes in different environments have been designed over a wide range of

pressures and temperatures.22-27

The general principle of operation of a STM is surprising simple. Figure 3.7 shows the

fundamental schematic diagram of STM. From the figure 3.7, we can see separation of the scale

of a few Angstrom unit between end of the sharp conducting tip and the scanned sample. A

tunneling current will form in this gap as the quantum tunneling effect, after a bias voltage

applied between the sharp metal tip and a conductive sample (metal or semiconductor). The

79

tunneling current can be used by the controller to locally probe the physical properties at the

surface and adjust the height of the tip to control the separation between the tip and sample

surface. As the tunneling current is strongly (exponentially) dependent on this gap, the distance

control based on tunneling current is very sensitive to small variations in the gap between the

two electrodes, as we will discuss later (section 3.2.1). We can track the surface profile of the

tip by scanning the tip over the surface while keeping the tunneling current constant through a

feedback loop, which will keep at constant distance from the sample surface. We can obtain a

3D image z(x, y) of the surface by recording the vertical z information of the tip as a function of

the x-y plane. The software will store the height, voltage and current information of 256×256

points on the region chosen to form an image.

Figure. 3.7 The schematic diagram of the basic characterizing mechanism of STM.

80

3.2.1 Working Principle: Tunneling Effect

The phenomenon that enables STM is the tunneling effect. The tunneling effect is a quantum

mechanical phenomenon in which subatomic particles pass through a classically forbidden

potential barrier. The laws of classical mechanics cannot explain quantum tunneling because

the energy required to pass the barrier is greater than the barrier.

The most important part for theoretical description of STM is how to calculate the tunneling

current. The actual tunneling current is pretty complicated due to different kinds of samples,

but we can start with the simplest model: both tip and sample are metals, separated by a tiny

vacuum gap.

According to the tunneling mechanism, the simplest equation is:

𝑑𝐼

𝑑𝑉|𝑉 ~ 𝐿𝐷𝑂𝑆𝑠𝑎𝑚𝑝𝑙𝑒(𝐸𝐹 + 𝑒𝑉) 3.1

In this equation, V is the voltage applied on the sample, I is the total tunneling current,

𝐿𝐷𝑂𝑆𝑠𝑎𝑚𝑝𝑙𝑒 are the local density of state (LDOS) for the sample and Ef is the Fermi energy.

From this equation, we see that the tunneling current conductance is in proportion to LODS of

the surface.

In 1961, Bardeen proposed the time-dependent perturbation theory that is used most widely

now.28 There are three presumptions in this theory: there is overlap in the density of state

between the tip and the sample because of the very small gap; this kind of the weak overlap

does not influence each density of states; and the wave functions exponentially drop to zero in

the gap area.

81

After using the time-dependent perturbation theory, we can get:

I𝑡→𝑠 = −2𝑒 ∙2𝜋

ℏ ∙ |𝑀𝑡𝑠|2 ∙ 𝐿𝐷𝑂𝑆𝑡𝑖𝑝(𝐸)f(E) ∙ (𝐿𝐷𝑂𝑆𝑠𝑎𝑚𝑝𝑙𝑒(𝐸 + 𝑒𝑉)[1 − 𝑓(𝐸 + 𝑒𝑉)]) 3.2

In this equation, the factor 2 is due to the spin degeneracy of the electron, -e represents the

classical negative electron charge, 2𝜋/ℏ is the factor from the time-dependent perturbation

theory, |𝑀𝑡𝑠| is the tunneling matrix influenced by the separation distance between the tip29

and the sample and f(E)=1/(1 + eE/k𝐵𝑇) is the famous Fermi Distribution.

Eq. 3.2 is the dominant tunneling current between the tip and sample, but there are also

electrons tunneling from the sample to the tip:

I𝑠→𝑡 = −2𝑒 ∙ 2𝜋

ℏ |𝑀𝑠𝑡|2 ∙ 𝐿𝐷𝑂𝑆𝑠𝑎𝑚𝑝𝑙𝑒(𝐸 + 𝑒𝑉)𝑓(E + eV) ∙ (𝐿𝐷𝑂𝑆𝑡𝑖𝑝(𝐸)[1 − 𝑓(𝐸)]) 3.3

These two tunneling currents are in opposite directions, so the total current should be the

integral over all energies E after the subtraction of Eq. 3.3 from Eq. 3.2. Then we get the total

tunneling current:

I𝑡𝑜𝑡𝑎𝑙 = −2𝑒 ∙ 2𝜋

ℏ ∙ ∫ |𝑀|2

−𝐸𝐹(𝑠𝑎𝑚𝑝𝑙𝑒)∙ 𝐿𝐷𝑂𝑆𝑡𝑖𝑝(𝐸) ∙ LDOS𝑠𝑎𝑚𝑝𝑙𝑒(E + eV) ∙ {𝑓(E)[1 −

𝑓(𝐸 + 𝑒𝑉)] − [1 − 𝑓(𝐸)]𝑓(𝑒 + 𝑒𝑉) ∙ 𝑑𝐸 3.4

If we assume the temperature (T) is 0 K, then 𝑓(E ≤ E𝐹) = 1, 𝑓(E > E𝐹) = 0, so we can

simplify Eq. 3.4 to:

I𝑡𝑜𝑡𝑎𝑙 ≈ −2𝑒 ∙ 2𝜋

ℏ ∙ ∫ |𝑀|2

0

−𝐸𝐹(𝑠𝑎𝑚𝑝𝑙𝑒)∙ 𝐿𝐷𝑂𝑆𝑡𝑖𝑝(𝐸) ∙ LDOS𝑠𝑎𝑚𝑝𝑙𝑒(E + eV) ∙ 𝑑𝐸 3.5

82

For the simplest situation, the vacuum barrier is a square barrier with a smaller bias compared

to the sum of the work function of both the tip and sample, the tunneling matrix |𝑀| can

written as Eq. 3.6 under the WKB approximation:30-31

|𝑀|2 = 𝑒−2𝑠√

2𝑚

ℏ(

𝜙𝑠+𝜙𝑡2

) 3.6

where 𝜙𝑠 𝑎𝑛𝑑 𝜙𝑡 represent the work functions of the tip and sample, respectively. Finally, we

get the total tunneling current:

I𝑡𝑜𝑡𝑎𝑙 = −2𝑒 ∙ 2𝜋

ℏ ∙ 𝑒

−2𝑠√2𝑚

ℏ(

𝜙𝑠+𝜙𝑡2

)∫ 𝐿𝐷𝑂𝑆𝑡𝑖𝑝(𝐸) ∙ LDOS𝑠𝑎𝑚𝑝𝑙𝑒(E + eV) ∙ 𝑑𝐸

0

−𝐸𝐹(𝑠𝑎𝑚𝑝𝑙𝑒) 3.7

J. Tersoff, and D.R. Hamann proposed an s-wave-model to simplify the DOS of the tip to a

constant, so we can modify the tunneling current to:

I𝑡𝑜𝑡𝑎𝑙 = −2𝑒 ∙ 2𝜋

ℏ ∙ 𝑒

−2𝑠√2𝑚

ℏ(

𝜙𝑠+𝜙𝑡2

)∙ 𝐿𝐷𝑂𝑆𝑡𝑖𝑝(0) ∫ LDOS𝑠𝑎𝑚𝑝𝑙𝑒(E + eV) ∙ 𝑑𝐸

0

−𝑒𝑉 3.8

Therefore, we can differentiate Eq. 3.8 to get the Eq. 3.1:

𝑑𝐼𝑡𝑜𝑡𝑎𝑙

𝑑𝑉|𝑉 ~ 𝐿𝐷𝑂𝑆𝑠𝑎𝑚𝑝𝑙𝑒(𝐸𝐹 + 𝑒𝑉) 3.1

So, we verify that the tunneling current conductance is just in proportion to the LDOS of the

surface.

83

3.2.2 Working Modes

We now focus on two primary and popular modes of STM: the constant current mode and the

constant height mode.

Constant current mode: Constant current mode is the first and most commonly used mode of

STM operation as shown in Figure.3.8a. The height of the tip is tuned by the feedback loop

during scanning so that the voltage and current remain constant when the STM tip moves

across the sample surface. The lateral position (x, y) is controlled by applying the voltage Vx and

Vy to the corresponding x- and y-piezoelectric drives while the vertical position (z) of the tip is

determined by applying a voltage Vz in the z-piezoelectric drive. Therefore, as long as the

sensitivities of the three orthogonal piezoelectric drives are known, the recorded signal Vz(Vx,

Vy) can be translated to the topography z(x, y). In the constant current mode, the z range is

wide; it is fit for most cases, especially for large scale scanning or a rough surface.

The principle of constant current mode sounds rather simple. However, the interpretation of

the obtained contour map z(x, y) is not at all trivial. In generally, it is not only the topologies of

the surface but also the local density of state of the surface that determines z(x, y), although

the contour map z(x, y) is often called the topography image of the sample surface. The contour

map can reflect a constant current surface based on the experimental process. To interpret the

contour map properly, we must consider the contributions of specific sample and tip properties

to the tunneling current.

84

Constant height mode: A notable disadvantage of the constant current mode is the limited

response time of the feedback loop, which greatly limits data acquisition time and the scanning

speed. In principle, the scanning speed of the tip is faster than the response of the feedback.32-

33 Consequently, the feedback can not respond to the higher frequency components

modulating the tunneling current due to it has to maintain an desired tunneling current.

Alternatively, you can switch off the feedback completely. If the response to the preamplifier is

still faster than the scanning speed, the tunneling current will be exponential depend on the

spacing between the tip and surface which will reflect the morphology at the atomic scale.

The tip will be scanned at a constant z height in the constant height mode, while the tunneling

current I is measured as a function of the position. The voltage applied on the tip or sample is

constant, and the z-piezoelectric feedback system is turned off. In this case, a surface bump will

be reflected in a higher tunneling current. As the height is constant and the separation between

tip and sample is of the sub-nanometer level, the scanning area has to be very flat. The current

mainly reflects the local density of states. This mode is especially fit for small area scanning.

This constant height mode can also used to obtain STM images at video speed, which provide

the opportunity to investigate dynamic atomic level procedure at surfaces, such as surface

diffusion.

85

Figure. 3.8. (a) Schematic illustration of the constant current mode of STM operation. (b) Schematic

illustration of constant height mode of STM operation.

3.2.3 Omicron RT® STM

STM Overview: The STM system that we used in this dissertation is the RT STM1® from

Omicron. This STM system consists of with 4 components, the main body, the pumping system,

the controller and the computer (software and UI). In this section we mainly discuss the main

body (Figure 3.9 a). The main body was three chambers, the main chamber, manipulator

chamber and load lock chamber. The main chamber is for scanning the sample, including a tip,

a scanner, a sample stage and a coarse runner, which are suspended by springs (Figure 3.9 b).

86

The manipulator chamber is the chamber where we can store and prepare the sample for

scanning. The load lock chamber is an important transfer chamber, which is used to connect

between the manipulator chamber and outside. We can also deposit the molecules on the

samples in the load lock chamber. There are two tools, which are used to transfer the sample,

the pincer grip wobble stick to transfer between the main chamber and the manipulator

chamber and a magnetic linear drive to transfer between the load lock chamber and the

manipulator chamber.

Figure. 3.9 (a) The main body of the STM consists of the main chamber, the manipulator chamber and

the load lock chamber. (b) The main chamber (scanning stage) of the STM. (c) The controller of the STM.

(d) The computer of the STM.

The STM is kept under ultra-high vacuum, with a base pressure in the range of 10-10 torr. There

are two main purposes to maintain the STM in such high vacuum. Firstly, the tip and samples

can remain clean for a longer time in UHV, which is crucial for getting high quality images. The

87

scanning gap between the tip and sample is only a few Angstroms, so a cleaner chamber has

less chance to contaminate the sample, in return. Secondly, the mechanism of the STM is that

the tunneling current go through the tip and the sample, and being in air will make the

tunneling situation more complicated, decreasing the stability of the measurement.

The STM vacuum system has four pumps, the mechanical pump, turbo pump, titanium

sublimation pump (tsp) and ion pump. The mechanical pump is a classical roughing pump or

backing pump. It is use to rough pump the system to 0.01 torr and also a backing pump to

support the turbo pump. The turbo pump is the secondary pump which provide a greater

probability of gas moving towards the outlet by transferring momentum from the rotating

blades to the molecules.34 The highest turbo vacuum system could reach as high as 10-7 torr.

After the vacuum reach as the 10-7 level, we can engage the ion pump. The ion pump is the

pump that can reach as high as 10-11 (our STM can reach 10-10 torr) by sputtering the ionized gas

to a metal gutter. After keeping the ion pump on, we can turn off the turbo pump and

mechanical pump. When the vacuum level in the main chamber is not good, we can turn on the

TSP to improve the vacuum. A high current (typically around 40 Amps, for us) is periodically

passed through TSP, which consists of a titanium filament. A thin layer of clean titanium can be

formed on the surrounding chamber walls as this current causes the filament to reach the

sublimation temperature of titanium.35 The residual gas components, colliding with the

chamber wall, tend to react and to form a solid, stable product due to the reactivity of clean

titanium in the chamber.35 Thus, the gas pressure is reduced in the chamber.

88

3.2.4 The correction of height of STM measurement

During STM measurement, the height difference typically appears a shift, similar as the AFM

measurement. According to quantum mechanics, the tunneling current is inverse to the gap

between the tip and sample (height) and proportional to the local density of states (LDOS). For

two different materials having different LDOS, in order to obtain same tunneling current, the

gap would be different. For example, for the C60 domain on graphene, the height difference

between the C60 domain and graphene is about 0.7 nm during the STM measurement. In

theory, the size of C60 is about 0.7 nm and the gap between C60 and graphene is about 0.3 nm,

so the height difference should be 1.0 nm. There is a difference 0.3 nm between STM

measurement and theoretical calculation.

3.3 Sample preparation

A key aspect of this work is the preparation of the samples. The methods that we use to create

samples with nanoscale control are the spin coating method and the physical vapor deposition

method.

3.3.1 Spin Coating

Spin coating is a widely used method for the deposition of uniform thin films onto a substrate,

especially useful for deposition of organic materials. Spin coating is widely used in organic solar

cells, polymer phase separation research and so on. Generally, a small amount of the target

89

material is dropped onto the substrate, which is either spinning at low speed or not spinning at

all, and then it is rotated at a higher speed to spread the target material by centrifugal force. In

chapter 4, we use spin coating to create monolayer or bilayer samples for AFM testing. During

the preparation, we spin coat a PCBM/Chlorobenzene solution on HOPG and graphene

substrates with a rate of 1000 rpm. The concentration of the solution that we used varied from

0.15 mg/ml to 2.0 mg/ml. Figure 3.10a is the spin coater that we used in this dissertation.

3.3.2 Physical Vapor Deposition

Physical vapor deposition (PVD) is one of the methods to create nanometer scale thin films. The

PVD method is realized by creating vapor by a physical method and then deposit the vapor on

the substrate under UHV condition. In this dissertation, the PVD method is used to create a

monolayer of C60 or PTCDA molecules in the load lock chamber. Figure 3.10 b is the home made

Knudesen cell that we used for the PVD method. The molecular powder was loaded into the

homemade Knudsen cell and mounted in the load lock of the STM system. The molecular

sources were degassed at around 20 K below the deposition temperature under a vacuum of

about 1 x 10-6 torr prior to deposition. Then molecules were deposited onto graphene at a

deposition rate around 0.5 monolayer/min with a background pressure below 1.8×10-8 torr.

Thus, we could get monolayer or sub-monolayer samples by controlling the deposition time.

90

Figure. 3.10 (a) The spin coater that we used in this dissertation. (b) The detailed structure of the

homemade Knudsen cell showing the main components inside the copper shell. 1 is CF flange, 2 is

thermocouple wire, 3 is W heating filament, 4 is glass tube, 5 is ceramic piece, 6 is hollow copper rods

(A, B, C, D), 7 is supporting rods, 8 is feedthrough.36

References:

1. Binnig, G.; Quate, C. F.; Gerber, C., Atomic Force Microscope. Phys Rev Lett 1986, 56 (9), 930-933. 2. Variola, F., Atomic force microscopy in biomaterials surface science. Phys Chem Chem Phys 2015, 17 (5), 2950-2959. 3. Horber, J. K. H.; Miles, M. J., Scanning probe evolution in biology. Science 2003, 302 (5647), 1002-1005. 4. Marshall, G. W.; Balooch, M.; Gallagher, R. R.; Gansky, S. A.; Marshall, S. J., Mechanical properties of the dentinoenamel junction: AFM studies of nanohardness, elastic modulus, and fracture. J Biomed Mater Res 2001, 54 (1), 87-95. 5. Haugstad, G., Atomic force microscopy: understanding basic modes and advanced applications. John Wiley & Sons, Inc 2012. 6. Jalili, N.; Laxminarayana, K., A review of atomic force microscopy imaging systems: application to molecular metrology and biological sciences. Mechatronics 2004, 14 (8), 907-945.

91

7. Basso, M. G., L.; Dahleh, M.; Mezic, I., Numerical analysis of complex dynamics in atomic force microscopes. Proc IEEE Conf Control Appl. Trieste, Italy 1998, 1026-1030. 8. Fung, R. F.; Huang, S. C., Dynamic modeling and vibration analysis of the atomic force microscope. J Vib Acoust 2001, 123 (4), 502-509. 9. Salapaka, M. C. D., Stability and sensitivity analysis of periodic orbits in tapping mode atomic force microscopy. Proc Conf Decision Control. Tampa FL 1998, 2047-2052. 10. Sebastian, A. S., M.; Chen, D.; Cleveland, J., Harmonic analysis based modeling of tapping-mode AFM. Proc Am Control Conf. San Diego CA 1999, 232-236. 11. http://slideplayer.com/slide/9702681/31/images/9/Modes+of+operation.+There+are+3+ modes+of+AFM+operation+Contact+mode.jpg. 12. Giessibl, F. J.; Trafas, B. M., Piezoresistive Cantilevers Utilized for Scanning Tunneling and Scanning Force Microscope in Ultrahigh-Vacuum. Rev Sci Instrum 1994, 65 (6), 1923-1929. 13. Albrecht, T. R.; Grutter, P.; Horne, D.; Rugar, D., Frequency-Modulation Detection Using High-Q Cantilevers for Enhanced Force Microscope Sensitivity. J Appl Phys 1991, 69 (2), 668-673. 14. Sugawara, Y.; Minobe, T.; Orisaka, S.; Uchihashi, T.; Tsukamoto, T.; Morita, S., Non-contact AFM images measured on Si(111)root 3 x root 3-Ag and Ag(111) surfaces. Surf Interface Anal 1999, 27 (5-6), 456-461. 15. Orisaka, S.; Minobe, T.; Uchihashi, T.; Sugawara, Y.; Morita, S., The atomic resolution imaging of metallic Ag(111) surface by noncontact atomic force microscope. Appl Surf Sci 1999, 140 (3-4), 243-246. 16. Zhong, Q.; Inniss, D.; Kjoller, K.; Elings, V. B., Fractured Polymer Silica Fiber Surface Studied by Tapping Mode Atomic-Force Microscopy. Surf Sci 1993, 290 (1-2), L688-L692. 17. https://www.azonano.com/article.aspx?ArticleID=3010. 18. https://www.bruker.com/products/surface-and-dimensional-analysis/atomic-force-microscopes/dimension-icon/overview.html. 19. Walczyk, W.; Schon, P. M.; Schonherr, H., The effect of PeakForce tapping mode AFM imaging on the apparent shape of surface nanobubbles. J Phys-Condens Mat 2013, 25 (18). 20. Binning, G.; Rohrer, H.; Gerber, C.; Weibel, E., Surface Studies by Scanning Tunneling Microscopy. Phys Rev Lett 1982, 49 (1), 57-61. 21. Binnig, G.; Rohrer, H.; Gerber, C.; Weibel, E., Tunneling through a Controllable Vacuum Gap. Appl Phys Lett 1982, 40 (2), 178-180. 22. Besenbacher, F.; Laegsgaard, E.; Mortensen, K.; Nielsen, U.; Stensgaard, I., Compact, High-Stability, Thimble-Size Scanning Tunneling Microscope. Rev Sci Instrum 1988, 59 (7), 1035-1038. 23. Mcintyre, B. J.; Salmeron, M.; Somorjai, G. A., A Variable Pressure Temperature Scanning Tunneling Microscope for Surface Science and Catalysis Studies. Rev Sci Instrum 1993, 64 (3), 687-691. 24. Besenbacher, F., Scanning tunnelling microscopy studies of metal surfaces. Rep Prog Phys 1996, 59 (12), 1737-1802. 25. Ferris, J. H.; Kushmerick, J. G.; Johnson, J. A.; Youngquist, M. G. Y.; Kessinger, R. B.; Kingsbury, H. F.; Weiss, P. S., Design, operation, and housing of an ultrastable, low temperature, ultrahigh vacuum scanning tunneling microscope. Rev Sci Instrum 1998, 69 (7), 2691-2695. 26. Jensen, J. A.; Rider, K. B.; Chen, Y.; Salmeron, M.; Somorjai, G. A., High pressure, high temperature scanning tunneling microscopy. J Vac Sci Technol B 1999, 17 (3), 1080-1084. 27. Kugler, M.; Renner, C.; Fischer, O.; Mikheev, V.; Batey, G., A He-3 refrigerated scanning tunneling microscope in high magnetic fields and ultrahigh vacuum. Rev Sci Instrum 2000, 71 (3), 1475-1478. 28. Bardeen, J., Tunnelling from a Many-Particle Point of View. Phys Rev Lett 1961, 6 (2), 57-&. 29. Chenggang, T., LUCTUATIONS ON METAL SURFACES AND MOLECULE/METAL INTERFACES PhD thesis. University of Maryland 2007. 30. Chen, C. J., Introduction to scanning tunneling microscopy. Oxford University Press 1993.

92

31. Wiesendanger, R., Contributions of Scanning Probe Microscopy and Spectroscopy to the Investigation and Fabrication of Nanometer-Scale Structures. J Vac Sci Technol B 1994, 12 (2), 515-529. 32. Bryant, A.; Smith, D. P. E.; Quate, C. F., Imaging in Real-Time with the Tunneling Microscope. Appl Phys Lett 1986, 48 (13), 832-834. 33. Bryant, A.; Smith, D. P. E.; Binnig, G.; Harrison, W. A.; Quate, C. F., Anomalous Distance Dependence in Scanning Tunneling Microscopy. Appl Phys Lett 1986, 49 (15), 936-938. 1. Binnig, G.; Quate, C. F.; Gerber, C., Atomic Force Microscope. Phys Rev Lett 1986, 56 (9), 930-933. 2. Variola, F., Atomic force microscopy in biomaterials surface science. Phys Chem Chem Phys 2015, 17 (5), 2950-2959. 3. Horber, J. K. H.; Miles, M. J., Scanning probe evolution in biology. Science 2003, 302 (5647), 1002-1005. 4. Marshall, G. W.; Balooch, M.; Gallagher, R. R.; Gansky, S. A.; Marshall, S. J., Mechanical properties of the dentinoenamel junction: AFM studies of nanohardness, elastic modulus, and fracture. J Biomed Mater Res 2001, 54 (1), 87-95. 5. Haugstad, G., Atomic force microscopy: understanding basic modes and advanced applications. John Wiley & Sons, Inc 2012. 6. Jalili, N.; Laxminarayana, K., A review of atomic force microscopy imaging systems: application to molecular metrology and biological sciences. Mechatronics 2004, 14 (8), 907-945. 7. Basso, M. G., L.; Dahleh, M.; Mezic, I., Numerical analysis of complex dynamics in atomic force microscopes. Proc IEEE Conf Control Appl. Trieste, Italy 1998, 1026-1030. 8. Fung, R. F.; Huang, S. C., Dynamic modeling and vibration analysis of the atomic force microscope. J Vib Acoust 2001, 123 (4), 502-509. 9. Salapaka, M. C. D., Stability and sensitivity analysis of periodic orbits in tapping mode atomic force microscopy. Proc Conf Decision Control. Tampa FL 1998, 2047-2052. 10. Sebastian, A. S., M.; Chen, D.; Cleveland, J., Harmonic analysis based modeling of tapping-mode AFM. Proc Am Control Conf. San Diego CA 1999, 232-236. 11. http://slideplayer.com/slide/9702681/31/images/9/Modes+of+operation.+There+are+3+ modes+of+AFM+operation+Contact+mode.jpg 12. Giessibl, F. J.; Trafas, B. M., Piezoresistive Cantilevers Utilized for Scanning Tunneling and Scanning Force Microscope in Ultrahigh-Vacuum. Rev Sci Instrum 1994, 65 (6), 1923-1929. 13. Albrecht, T. R.; Grutter, P.; Horne, D.; Rugar, D., Frequency-Modulation Detection Using High-Q Cantilevers for Enhanced Force Microscope Sensitivity. J Appl Phys 1991, 69 (2), 668-673. 14. Sugawara, Y.; Minobe, T.; Orisaka, S.; Uchihashi, T.; Tsukamoto, T.; Morita, S., Non-contact AFM images measured on Si(111)root 3 x root 3-Ag and Ag(111) surfaces. Surf Interface Anal 1999, 27 (5-6), 456-461. 15. Orisaka, S.; Minobe, T.; Uchihashi, T.; Sugawara, Y.; Morita, S., The atomic resolution imaging of metallic Ag(111) surface by noncontact atomic force microscope. Appl Surf Sci 1999, 140 (3-4), 243-246. 16. Zhong, Q.; Inniss, D.; Kjoller, K.; Elings, V. B., Fractured Polymer Silica Fiber Surface Studied by Tapping Mode Atomic-Force Microscopy. Surf Sci 1993, 290 (1-2), L688-L692. 17. https://www.azonano.com/article.aspx?ArticleID=3010. 18. https://www.bruker.com/products/surface-and-dimensional-analysis/atomic-force-microscopes/dimension-icon/overview.html. 19. Walczyk, W.; Schon, P. M.; Schonherr, H., The effect of PeakForce tapping mode AFM imaging on the apparent shape of surface nanobubbles. J Phys-Condens Mat 2013, 25 (18). 20. Binning, G.; Rohrer, H.; Gerber, C.; Weibel, E., Surface Studies by Scanning Tunneling Microscopy. Phys Rev Lett 1982, 49 (1), 57-61.

93

21. Binnig, G.; Rohrer, H.; Gerber, C.; Weibel, E., Tunneling through a Controllable Vacuum Gap. Appl Phys Lett 1982, 40 (2), 178-180. 22. Besenbacher, F.; Laegsgaard, E.; Mortensen, K.; Nielsen, U.; Stensgaard, I., Compact, High-Stability, Thimble-Size Scanning Tunneling Microscope. Rev Sci Instrum 1988, 59 (7), 1035-1038. 23. Mcintyre, B. J.; Salmeron, M.; Somorjai, G. A., A Variable Pressure Temperature Scanning Tunneling Microscope for Surface Science and Catalysis Studies. Rev Sci Instrum 1993, 64 (3), 687-691. 24. Besenbacher, F., Scanning tunnelling microscopy studies of metal surfaces. Rep Prog Phys 1996, 59 (12), 1737-1802. 25. Ferris, J. H.; Kushmerick, J. G.; Johnson, J. A.; Youngquist, M. G. Y.; Kessinger, R. B.; Kingsbury, H. F.; Weiss, P. S., Design, operation, and housing of an ultrastable, low temperature, ultrahigh vacuum scanning tunneling microscope. Rev Sci Instrum 1998, 69 (7), 2691-2695. 26. Jensen, J. A.; Rider, K. B.; Chen, Y.; Salmeron, M.; Somorjai, G. A., High pressure, high temperature scanning tunneling microscopy. J Vac Sci Technol B 1999, 17 (3), 1080-1084. 27. Kugler, M.; Renner, C.; Fischer, O.; Mikheev, V.; Batey, G., A He-3 refrigerated scanning tunneling microscope in high magnetic fields and ultrahigh vacuum. Rev Sci Instrum 2000, 71 (3), 1475-1478. 28. Bardeen, J., Tunnelling from a Many-Particle Point of View. Phys Rev Lett 1961, 6 (2), 57-&. 29. Chenggang, T., LUCTUATIONS ON METAL SURFACES AND MOLECULE/METAL INTERFACES PhD thesis. University of Maryland 2007. 30. Chen, C. J., Introduction to scanning tunneling microscopy. Oxford University Press 1993. 31. Wiesendanger, R., Contributions of Scanning Probe Microscopy and Spectroscopy to the Investigation and Fabrication of Nanometer-Scale Structures. J Vac Sci Technol B 1994, 12 (2), 515-529. 32. Bryant, A.; Smith, D. P. E.; Quate, C. F., Imaging in Real-Time with the Tunneling Microscope. Appl Phys Lett 1986, 48 (13), 832-834. 33. Bryant, A.; Smith, D. P. E.; Binnig, G.; Harrison, W. A.; Quate, C. F., Anomalous Distance Dependence in Scanning Tunneling Microscopy. Appl Phys Lett 1986, 49 (15), 936-938. 34. https://vacaero.com/information-resources/vacuum-pump-technology-education-and-training/1039-an-introduction-to-vacuum-pumps.html. 35. https://en.wikipedia.org/wiki/Titanium_sublimation_pump#cite_note-vacgen-1. 36. Chuanhui Y. , A. M., Husong Z. , Yanlong L. , Chenggang T., Preparation and Characterization of C60/Graphene Hybrid Nanostructures. Journal of Visualized Experiments 2018, 135 (e57257).

94 This chapter closely follows the publication: Li, Y.; Chen, C.; Burton, J.; Park, K.; Heflin, J.; Tao, C., Self-Assembled PCBM Bilayers on Graphene and HOPG Examined by AFM and STM, Nanotechnology 29, 185703 (2018)

Chapter 4

Self-Assembled PCBM Bilayers on Graphene and HOPG

Examined by AFM and STM

The majority of this chapter is from a manuscript published in Institute of Physics (IOP)

nanotechnology, with slight modifications.1 The experimental part of this chapter is done by

Yanlong Li with the help of Chuanhui Chen. The discussion part is due to the effort of Yanlong Li,

Chuanhui Chen and John Burton.

4.1 Introduction

In the past several decades, organic solar cells have attracted tremendous scientific and

industrial interest because their power conversion efficiency has dramatically increased and

reached 17 % to date. 2-5 In addition, organic solar cells have potential advantages compared to

traditional solar cells in flexibility of chemical modification as well as low-cost mass production.6

Typically, an organic solar cell generates electric current through photon-induced electron

transfer that separates electrons from holes.7 the behavior of a solar cell depends on the

materials serving as electron donor and electron acceptor, respectively. As light enters a solar

cell, the photons induce electrons to transfer from the excited state of the donor to the lowest

unoccupied molecular orbital (LUMO) of the acceptor. Subsequently, the separated electrons

and holes reach the cathode and anode, respectively, delivering a direct current to an outer

95

circuit.8 The power conversion efficiency of a solar cell depends on various properties including

electron affinity of electron acceptor.

The overall performance of organic solar cells hinges on material properties of an active layer,

which is composed of a variety of donors (e.g., poly[2-methoxy-5-(3′,7′-dimethyloctyloxy)-1,4-

phenylenevinylene] (MDMO-PPV), poly(3-hexylthiophene-2,5-diyl) (P3HT)5, 9-12 ) and acceptors

(e.g., phenyl-C61-butyric acid methyl ester (PCBM)13-14). There have been intensive previous

investigations on the various factors that impact the efficiency of organic solar cell devices, such

as solvent/thermal annealing, weight ratio of donor and acceptor, thickness of the active layer,

etc.15-18 In the past several years, utilization of emerging two-dimensional (2D) materials such as

graphene for energy-related applications has attracted major research efforts.19-21 Notably,

graphene is a promising candidate for a transparent electrode material in solar cells.22-25 To

design efficient organic/2D material hybrid solar cells, it is crucial to understand the

morphology of the donor/acceptor nanostructures on 2D materials. Although morphology of

donor/acceptor nanostructures has been well characterized on bulk substrates, such as metals

or ITO,16, 26-38 similar studies on 2D materials are still lacking.

In this work, we present the self-assembled structure of PCBM, a promising acceptor material

for organic solar cells, deposited on graphene and HOPG. We discover novel bilayer

nanostructures of PCBM on graphene and HOPG, and investigate how thermal annealing tunes

the morphology of the PCBM bilayer, by using AFM and STM. Interestingly, PCBM bilayers are

formed with two typical heights on HOPG, but only one on graphene. At different annealing

temperatures, edge diffusion causes neighboring vacancies to emerge into a more ordered

structure. This first experimental realization of PCBM bilayer structures on graphene may pave

96

a way to fabricate hybrid structures of organic donor/acceptor molecules and graphene for

applications in organic solar cells.

4.2 Experimental Methods

PCBM was purchased from NanoC Inc. (Purity: 99.5%). A solution of PCBM was prepared by

stirring PCBM powder in chlorobenzene and then the solution was set on a hot plate at about

70 °C for 24 hours. The film samples were prepared by spin-coating the solution onto freshly

cleaved HOPG substrates (SPI-1 grade, purchased from SPI supplies) or graphene on Cu foils

synthesized by CVD. AFM measurements were carried out on a Dimension Icon (Bruker

Corporation) instrument in a dark environment. Monolithic silicon cantilevers (NCST, NANO

WORLD) with a spring constant of 7.4 N/m, first longitudinal resonance frequencies between

120 – 205 kHz, and nominal tip radius of 8 nm were employed in soft tapping mode.

Simultaneous height and phase images were acquired and reproduced across multiple samples.

STM characterizations were carried out in an ultra-high vacuum (UHV) scanning tunneling

microscope system (Omicron STM) with a base pressure of low 10-9 Torr. The STM tip was a

chemically etched tungsten tip.

97

4.3 Results and Discussion

4.3.1 PCBM Bilayer Morphology

We first investigated the self-assembled structure of PCBM deposited on a graphene/Cu

substrate. The main facet of Cu underneath monolayer graphene is (111) oriented, which was

determined by typical Moiré patterns of graphene (inset of Figure 4.1a). Typical AFM

topography images of the PCBM bilayer on graphene (Figure 4.1b, c and d), showed a random

distribution of PCBM islands, similar to previous SEM results,39 in contrast to highly ordered hcp

or “double row” structures such as have been previously reported.40-41 These islands are

identified as PCBM bilayers with a measured height of ~1.37 nm (Figure 4.1d and e, blue lines),

which is close to twice the diameter of PCBM molecules (~0.7 nm). The height of this domain

differs significantly from the height of chlorobenzene residue and the height of a PCBM

monolayer,40 and so we can exclude the possibility that the connected islands are due to

solvent or PCBM monolayers.

In order to characterize the large-scale morphology of such a PCBM domain, we also deposited

PCBM bilayer nanostructures on a HOPG substrate, which provides much larger atomically flat

terraces in comparison with the flat facets in the case of the graphene/Cu substrate. The lowest

coverage (~0.15) bilayer films on HOPG were spin-coated from 0.1 mg/ml PCBM in

chlorobenzene solution (Figure 4.1h). Atomically resolved STM images of the area outside the

film domain verified that it was bare HOPG substrate (inset of Figure 4.1h). The PCBM bilayer

film showed continuity when crossing step edges on the HOPG substrate (Figure 4.1h, white

double arrows). The overall differences between Figure 4.1b (long stripes) to 4.1f, 4.1h, and 4.1i

(irregular islands) are due to the differences in the underlying graphene/Cu to HOPG substrate.

98

Figure. 4.1. AFM and STM images of PCBM films spin-coated from PCBM/Chlorobenzene solution on

graphene/Cu and HOPG substrates. (a) STM image of a graphene/Cu substrate, and the inset is the

atomic image of Moiré pattern of graphene on Cu (111) taken from the white square area in (a). (b) AFM

image of a PCBM bilayer on a graphene/Cu substrate from 0.2 mg/ml PCBM solution. (c) AFM image of a

PCBM bilayer on a graphene/Cu substrate from blue square area in (b). (d) A zoomed in AFM image of a

PCBM bilayer on a graphene/Cu substrate (e) Line profile with the height about 1.37 nm. (f) AFM image

of PCBM bilayer on HOPG substrate from 0.5 mg/ml PCBM solution. (g) Line profiles with the heights of

type I (blue line) and type II (red dashed line) indicated in Figure 4.1f. (h) AFM image of 0.15 PCBM

bilayer on HOPG from 0.1 mg/ml PCBM solution; the inset is the atomic image of HOPG taken from the

white square area. In (h), the HOPG step edges are indicated by white double arrows. (i) AFM image of a

PCBM bilayer on HOPG substrate from 1.0 mg/ml solution. (j) Line profile showing the height of type II

(red dashed line) showed in (i).

99

Strikingly, we found two typical heights in the PCBM bilayer deposited on the HOPG substrate:

1.64 ± 0.09 nm (Type I, blue line in Figure 4.1f), and 1.23 ± 0.03 nm (Type II, red lines in Figure

4.1f, i). In some films, one type of PCBM bilayer dominated the sample (Figure 4.1i), while in

others the two types coexisted in a sample (Figure 4.1f). Our observations of PCBM

monolayers, which will be discussed in the following section, showed a very different height,

and hence we exclude the possibility that the Type II structure originates from a monolayer.

Combining the above observations together, we conjecture that the height difference reflects

two distinct types of the PCBM dimer-HOPG substrate interaction with different arrangement

angles leading to different heights.

Note that these novel PCBM bilayers were observed on graphene and HOPG surfaces rather

than monolayers with hexagonal close-packed structures30, 40 or double row structures on Au

surfaces.40-41 The typical height of a PCBM monolayer is about 0.7 nm, and the height of a

PCBM bilayer is 1.64 nm. It is most likely that the observed PCBM bilayer structure consists of

PCBM dimers standing up on the graphene surface with some tilt angle due to the weak

interaction between PCBM dimers and the graphene surface. The tilt angle between PCBM

dimers and the graphene surface is about 40.6o. In a previous theoretical study, Bredas’ group

found a similar structure where pentacene molecules on a gold surface are tilted with an angle

of about 37.7°.42

Spin-coating from higher concentrations of PCBM resulted in higher coverage of the substrate

by PCBM bilayers (Figure 4.2.a-e), but the morphology of the bilayer remained as irregular

networks with randomly distributed holes. Obvious, with the increasing the concentrations of

PCBM, the coverage almost increases linearly until to a fully coverage (Figure 4.2.f).

100

Figure. 4.2 AFM images of PCBM films spin-coated from PCBM/chlorobenzene solution on HOPG

substrates. (a) AFM image of a PCBM bilayer on HOPG substrate from 0.15 mg/ml PCBM solution. (b)

AFM image of a PCBM bilayer on HOPG substrate from 0.5 mg/ml PCBM solution. (c) AFM image of a

PCBM bilayer on HOPG substrate from 0.75 mg/ml PCBM solution. (d) AFM image of a PCBM bilayer on

HOPG substrate from 1.0 mg/ml PCBM solution. (e) AFM image of a PCBM bilayer on HOPG substrate

from 2.0 mg/ml PCBM solution. (f) Coverage vs Concentration plotting based on AFM images.

4.3.2 PCBM Monolayer Morphology

Next, we examine self-assembled PCBM monolayers deposited on graphene/Cu and HOPG. On

graphene, a monolayer was formed after a post annealing for 30 minutes at 170 oC (Figure 4.3a

and 4.3b). On HOPG, the monolayer sample is more favorable as the stabilization time of the

101

solution increases. The overall morphology resembles the irregular network observed in the

bilayer structure (Figure 4.3a and 4.3b). The Figure indicates the morphology of a PCBM

monolayer on the graphene/Cu substrate, while the inseted line profile shows the height is

about 0.87 nm, which is in the range of monolayer height. In order to further examine

monolayers, additional PCBM monolayer samples were studied by spin coating 0.5 mg/ml

PCBM/chlorobenzene solution on HOPG under the same deposition conditions as the bilayer

samples. The height of a monolayer has two typical values on the HOPG substrate: one is

around 0.71 nm (Figure 4.3c, d); the other one is about 0.88 nm (Figure 4.3e, f). These two

values are relatively close to each other, and both of them are comparable to those reported in

previous literature.40-41 Thus, it is suggested that both of these two heights originate from

PCBM monolayers with different orientations. Compared to the mixed structures that can be

observed in the bilayers, these two typical heights always appear in different samples, which

may be due to a higher energy barrier (1 kcal/mol - about 200 oC thermal energy) between the

two PCBM monolayer orientations39. Note that our PCBM monolayer sample is filled with

randomly shaped monolayer terraces, in contrast to elbow nucleating structure found

previously.41

102

Figure. 4.3. AFM images of PCBM monolayer films spin-coated from PCBM/chlorobenzene solution on

graphene/Cu and HOPG substrates. (a) AFM image of a PCBM monolayer on graphene/Cu from 0.5

mg/ml PCBM solution after a 30 min 170 oC anneal. (b) AFM image of a PCBM monolayer on a

graphene/Cu substrate from blue square area in (a), and the inset is the line profile for typical PCBM

monolayer with height of 0.87 nm. (c) AFM image of a PCBM monolayer on HOPG from 0.5 mg/ml PCBM

solution. (d) Line profile along the line marked in (c) indicates the height of 0.71 nm. (e) AFM image of a

PCBM monolayer on HOPG from 0.5 mg/ml solution. (f) Line profile along the line marked in (e) shows

the height is about 0.88 nm.

103

4.3.3 Discussion

To understand the observed bilayer and monolayer structures, we constract molecular models

based on van der Waals interactions between neighboring fullerene moieties, hydrogen-

bonding between tail functional groups, and the interactions between molecules and

substrates. We first discuss the model for the monolayer structure. We propose that the two

typical monolayer heights of ~0.9 nm and ~0.7 nm observed in the PCBM monolayers reflect

different orientations of the PCBM relative to the substrate. The thicker monolayer (~0.9 nm)

corresponds to a vertical configuration of PCBM, with its tail perpendicular to the HOPG surface

(Figure 4.4a), whereas the thinner monolayer (~0.7 nm) corresponds to a horizontal

configuration, with PCBM tails parallel to the HOPG surface and interacting in pairs (Figure

4.4b). The monolayer height of 0.7 nm is consistent with the previous measurement of the size

of C60 by Robey’s group40, while the monolayer height of 0.9 nm has never been reported.

Now we turn to the small differences between the modeled and measured heights of a PCBM

monolayer in both of the proposed configurations. These differences arise from the fact that

the tapping mode AFM tends to underestimate the height of sample surface features, and that

the underestimate amount depends on the stiffness of the measured areas.43-44 For the

horizontal configuration, the actual monolayer height should be ~1 nm, because the

equilibrium van der Waals gap between C60 and the HOPG substrate is calculated to be 0.25~0.3

nm wide (Figure 4.4).40 But due to pressure exerted by the AFM tip, the monolayer height was

measured at ~0.7 nm. In fact, this measured value is comparable to a previous AFM tapping

mode image of a C60 shuttlecocks monolayer on HOPG.43 For the vertical configuration, the

monolayer height that the model suggests is 1.45 nm, marking an even larger difference from

104

the measured height of 0.9 nm. This large difference is likely due to softness of the monolayer

caused by the PCBM tail groups – these tails are likely much softer than the stiff C60 cages in the

horizontal configuration.

Figure. 4.4. Two schematic model configurations of a PCBM monolayer on a graphene or HOPG

substrate: C atom in PCBM (blue), C atom in the substrate (green), O atom (red), and H atom (yellow).

(a) Top and side views of one model configuration of a PCBM monolayer with height of 0.9 nm (b) Top

105

and side views of another model configuration of a PCBM monolayer sample of with height of 0.7 nm.

The dashed parallelograms in (a) and (b) indicate unit cells.

Our analysis also suggests that a higher coverage of the monolayer favors the vertical

configuration. This is reasonable because this configuration can accommodate a much higher

area density (1.15 molecule/nm2) of PCBM than that of the horizontal configuration (0.73

molecule/nm2). Using the energy values of neighboring fullerene moieties, the hydrogen-

bonding, and the interactions between molecules and substrates from previous experimental

and theoretical investigations,40-41, 45 we calculated the formation energy of these two

configurations and the results show an energy barrier of about 1 kcal/mol exists between the

two configurations, which inhibits a thermally-induced transition between the configurations at

room temperature.

We now discuss the model for the PCBM bilayer. In most cases, we observed PCBM bilayers

rather than monolayers on graphene and HOPG. As described before, there are two typical

heights in PCBM bilayers on graphene and HOPG: 1.64 nm (modeled in Figure 4.5a), and 1.23

nm (modeled in Figure 4.5b). It was previouly reported that on a gold substrate, PCBM dimers

are formed with a twin chain structure in low density, but with a double row structure in high

density,39-40, 45 because the affinity between C60 cages (0.28 eV)46-48 is higher than the hydrogen

bonding between the PCBM tail groups (0.114 eV).46 However, in this work, PCBM molecules

were deposited by spin coating rather than physical vapor deposition. It is known that steric

hindrance of sidechain-substituted PCBM molecules would forbid a possibility of a C60-to-C60

coordinated structure in the region defined by their first solvation shells.49 Considering

106

hydrogen bonding and dipole-dipole interactions, the energy of a side-to-side dimer (0.114 eV)

is higher than side-to-C60 dimer (0.001 eV).46 Hence, we suggest that the PCBM molecules form

side-to-side dimers in the chlorobenzene solvent before the spin coating. After the spin coating,

the side-to-side PCBM dimers are distributed on the HOPG, forming the PCBM bilayers (type I

and type II in Figure 4.5). For the thicker bilayer, we propose that the PCBM dimers form a

‘double hcp’ structure: one C60 cell of the dimer lies in the lower layer, the other in the upper

layer, such that neighboring PCBM dimers interact sideways (Type I, Figure 4.5a). For the

thinner bilayer, the PCBM monomers in each dimer are also located in the lower and upper

layers (Type II, Figure 4.5b), but without strong sideways interactions between PCBM dimers.

The two typical PCBM bilayer heights indicates the different tilt angles between a PCBM dimer

and the substrate in the two configurations. A tilt angle of 49.4o corresponds to the 1.64 nm

bilayer, while a tilt angle of 60.8o the 1.23 nm bilayer, where the tilt angle is defined to be an

angle between an axis connecting the centers of two C60 in a given dimer and the direction

normal to the substrate. Similarly to the PCBM monolayer case, the measured heights of a

PCBM bilayer are ~ 0.3 nm smaller than those predicted by the model in both configurations.

These differences can be similarly explained by the compression caused by the AFM tips.

107

Figure. 4.5. Schematic diagrams of type I and type II configurations of a PCBM bilayer on a graphene or

HOPG substrate: C atom in PCBM (blue), C atom in the substrate (green), O atom (red), and H atom

(yellow). (a) Top and side views of type I configuration of the PCBM bilayer. In the side view, the dashed

tilted rectangle highlights a PCBM dimer cell, and the solid rectangle indicates the hydrogen binding

within a PCBM dimer, while the solid circle shows a side interaction between neighboring PCBM dimers.

(b) Top and side views of type II configuration of the PCBM bilayer. In the side view, the dashed tilted

rectangle highlights a PCBM dimer cell, while the solid rectangle indicates the hydrogen binding within a

PCBM dimer.

108

We further estimate the energy per unit area for the different monolayer and bilayer

configurations in order to deduce their relative stability. For Type I bilayer, the binding energy

of one upper layer cell of PCBM is 2.52 eV, which arises from the sum of the binding energy of

the 9 nearest C60 molecules with 0.28 eV each; for the lower layer PCBM cells, the binding

energies are 3.5 eV each, which arises from the sum of the binding energy of the 9 nearest C60

molecules and of the binding energy between C60 and the HOPG (0.98 eV).45, 50 Combining the

above information with hydrogen binding energies of the neighboring tail groups, we find that

the total energy for each Type I PCBM dimer is 6.13 eV. For Type II bilayer, the upper-layer

PCBM cell only has a binding energy of 1.12 eV (4 nearest C60), while the lower-layer PCBM cell

on HOPG has a binding energy of 2.1 eV, considering the binding energy of one C60 and HOPG

and the binding energy of the nearest C60 molecules. The binding energy of each PCBM dimer is

3.33 eV. Combining the above information with the hydrogen bonding of the tail groups, we

find that for the horizontal configuration the binding energy one type II PCBM dimer is 4.31 eV.

Considering the dimer concentration density of 1.15 dimer/nm2, we find that the energy

density of Type I PCBM bilayer is 7.05 eV/nm2, which is higher than that of Type II bilayer by

1.35 eV/nm2 and that of the horizontal monolayer by 1.57 eV/nm2). As concentration density

increases, the PCBM dimers are compressed to form more vertical configurations in order to

accommodate more PCBM dimers on the HOPG surface.

109

4.3.4 Thermal Effects

We now investigate annealing effects on the morphology of the PCBM bilayer nanostructures.

In order to quantify the large-scale morphology changes of PCBM domains, we focused on the

PCBM bilayer nanostructures on HOPG substrates, which provide large atomically flat terraces.

For this experiment, we annealed the samples at 140 °C and 160 °C for 10 minutes and

measured the samples by AFM immediately after annealing. Figure 4.6a shows the topography

before annealing and Figure 4.6b and 4.6c are the PCBM bilayer after 140 °C and 160 °C

annealing, respectively. To quantify the morphology changes, we performed size distribution

analysis for the holes (i.e., bare HOPG area without PCBM) as shown in Figure 4.6d. by using the

standard nanoparticle size distribution analysis method.51 For data analysis, we use the method

introduced in ‘On optimal and data-based histograms’.52 We find a right shift of the distribution

peak after annealing at 140 °C and 160 °C. The peaks are located at 1490 nm2 before annealing,

1884 nm2 for annealing at 140 °C, and 2291 nm2 for annealing at 160 °C, respectively. The result

indicates that the smaller holes are merged together.

110

Figure. 4.6. AFM images of PCBM bilayer and size distributions of holes at different conditions. (a) AFM

image of a PCBM bilayer before annealing. (b) AFM image of a PCBM bilayer after annealing at 140 °C.

(c) AFM image of a PCBM bilayer after annealing at 160 °C. (d) Area distribution histogram of holes

(without PCBM area) obtained from measurements of the area of holes in AFM images of before (red

line) and after annealing at 140 °C (dark red line) and 160 °C (dark blue line).

4.4 Conclusion

In summary, we demonstrated the self-assembly of PCBM bilayer nanostructures on graphene

and HOPG, by using AFM and STM, and analyzed the observed morphology by comparison to

111

molecular models. The PCBM bilayer revealed two distinct configurations on HOPG with

different heights, and only one configuration on graphene. Post thermal annealing can induce

merging of the bilayer nanostructures. Our results will shed light on improvement of the energy

efficiency in solar cells containing graphene and organic molecules, by increasing the donor-

acceptor interface area.

References:

1. Li, Y. L.; Chen, C. H.; Burton, J.; Park, K.; Heflin, J. R.; Tao, C. G., Self-assembled PCBM bilayers on graphene and HOPG examined by AFM and STM. Nanotechnology 2018, 29 (18). 2. Meng, L. X.; Zhang, Y. M.; Wan, X. J.; Li, C. X.; Zhang, X.; Wang, Y. B.; Ke, X.; Xiao, Z.; Ding, L. M.; Xia, R. X.; Yip, H. L.; Cao, Y.; Chen, Y. S., Organic and solution-processed tandem solar cells with 17.3% efficiency. Science 2018, 361 (6407), 1094-+. 3. Zhao, W. C.; Li, S. S.; Yao, H. F.; Zhang, S. Q.; Zhang, Y.; Yang, B.; Hou, J. H., Molecular Optimization Enables over 13% Efficiency in Organic Solar Cells. J Am Chem Soc 2017, 139 (21), 7148-7151. 4. Xiao, Z.; Jia, X.; Ding, L. M., Ternary organic solar cells offer 14% power conversion efficiency. Sci Bull 2017, 62 (23), 1562-1564. 5. Cui, Y.; Yao, H. F.; Gao, B. W.; Qin, Y. P.; Zhang, S. Q.; Yang, B.; He, C.; Xu, B. W.; Hou, J. H., Fine-Tuned Photoactive and Interconnection Layers for Achieving over 13% Efficiency in a Fullerene-Free Tandem Organic Solar Cell. J Am Chem Soc 2017, 139 (21), 7302-7309. 6. Hoppe, H.; Sariciftci, N. S., Organic solar cells: An overview. J Mater Res 2004, 19 (7), 1924-1945. 7. Sariciftci, N. S.; Smilowitz, L.; Heeger, A. J.; Wudl, F., Photoinduced Electron-Transfer from a Conducting Polymer to Buckminsterfullerene. Science 1992, 258 (5087), 1474-1476. 8. Bagher, A. M., Introduction to organic solar cells Sustain. Energy 2014, 2, 85–90. 9. Gunes, S.; Neugebauer, H.; Sariciftci, N. S., Conjugated polymer-based organic solar cells. Chem Rev 2007, 107 (4), 1324-1338. 10. Scharber, M. C.; Wuhlbacher, D.; Koppe, M.; Denk, P.; Waldauf, C.; Heeger, A. J.; Brabec, C. L., Design rules for donors in bulk-heterojunction solar cells - Towards 10 % energy-conversion efficiency. Adv Mater 2006, 18 (6), 789-+. 11. Wienk, M. M.; Kroon, J. M.; Verhees, W. J. H.; Knol, J.; Hummelen, J. C.; van Hal, P. A.; Janssen, R. A. J., Efficient methano[70]fullerene/MDMO-PPV bulk heterojunction photovoltaic cells. Angew Chem Int Edit 2003, 42 (29), 3371-3375. 12. Liang, Y. Y.; Xu, Z.; Xia, J. B.; Tsai, S. T.; Wu, Y.; Li, G.; Ray, C.; Yu, L. P., For the Bright Future-Bulk Heterojunction Polymer Solar Cells with Power Conversion Efficiency of 7.4%. Adv Mater 2010, 22 (20), E135-+. 13. Padinger, F.; Rittberger, R. S.; Sariciftci, N. S., Effects of postproduction treatment on plastic solar cells. Adv Funct Mater 2003, 13 (1), 85-88.

112

14. Muhlbacher, D.; Scharber, M.; Morana, M.; Zhu, Z. G.; Waller, D.; Gaudiana, R.; Brabec, C., High photovoltaic performance of a low-bandgap polymer. Adv Mater 2006, 18 (21), 2884-+. 15. Li, G.; Shrotriya, V.; Yao, Y.; Yang, Y., Investigation of annealing effects and film thickness dependence of polymer solar cells based on poly(3-hexylthiophene). J Appl Phys 2005, 98 (4). 16. Ma, W. L.; Yang, C. Y.; Gong, X.; Lee, K.; Heeger, A. J., Thermally stable, efficient polymer solar cells with nanoscale control of the interpenetrating network morphology. Adv Funct Mater 2005, 15 (10), 1617-1622. 17. Wei, G. D.; Wang, S. Y.; Sun, K.; Thompson, M. E.; Forrest, S. R., Solvent-Annealed Crystalline Squaraine: PC70BM (1:6) Solar Cells. Adv Energy Mater 2011, 1 (2), 184-187. 18. Dang, M. T.; Hirsch, L.; Wantz, G., P3HT:PCBM, Best Seller in Polymer Photovoltaic Research. Adv Mater 2011, 23 (31), 3597-3602. 19. Bonaccorso, F.; Colombo, L.; Yu, G. H.; Stoller, M.; Tozzini, V.; Ferrari, A. C.; Ruoff, R. S.; Pellegrini, V., Graphene, related two-dimensional crystals, and hybrid systems for energy conversion and storage. Science 2015, 347 (6217). 20. Quesnel, E.; Roux, F.; Emieux, F.; Faucherand, P.; Kymakis, E.; Volonakis, G.; Giustino, F.; Martin-Garcia, B.; Moreels, I.; Gursel, S. A.; Yurtcan, A. B.; Di Noto, V.; Talyzin, A.; Baburin, I.; Tranca, D.; Seifert, G.; Crema, L.; Speranza, G.; Tozzini, V.; Bondavalli, P.; Pognon, G.; Botas, C.; Carriazo, D.; Singh, G.; Rojo, T.; Kim, G.; Yu, W. J.; Grey, C. P.; Pellegrini, V., Graphene-based technologies for energy applications, challenges and perspectives. 2d Mater 2015, 2 (3). 21. Zhang, Y. Y.; Hu, J. P.; Bernevig, B. A.; Wang, X. R.; Xie, X. C.; Liu, W. M., Localization and the Kosterlitz-Thouless Transition in Disordered Graphene. Phys Rev Lett 2009, 102 (10). 22. Hong, W. J.; Xu, Y. X.; Lu, G. W.; Li, C.; Shi, G. Q., Transparent graphene/PEDOT-PSS composite films as counter electrodes of dye-sensitized solar cells. Electrochem Commun 2008, 10 (10), 1555-1558. 23. Lee, Y. Y.; Tu, K. H.; Yu, C. C.; Li, S. S.; Hwang, J. Y.; Lin, C. C.; Chen, K. H.; Chen, L. C.; Chen, H. L.; Chen, C. W., Top Laminated Graphene Electrode in a Semitransparent Polymer Solar Cell by Simultaneous Thermal Annealing/Releasing Method. Acs Nano 2011, 5 (8), 6564-6570. 24. Wang, X.; Zhi, L. J.; Mullen, K., Transparent, conductive graphene electrodes for dye-sensitized solar cells. Nano Lett 2008, 8 (1), 323-327. 25. Wu, J. B.; Becerril, H. A.; Bao, Z. N.; Liu, Z. F.; Chen, Y. S.; Peumans, P., Organic solar cells with solution-processed graphene transparent electrodes. Appl Phys Lett 2008, 92 (26). 26. Takeuchi, O.; Takeuchi, N.; Ochiai, T.; Kato, H.; Yoshida, S.; Shigekawa, H., Microscopic description of the current-voltage characteristics of a bulk-heterojunction organic solar cell under illumination. Appl Phys Express 2014, 7 (2). 27. Chen, D. A.; Nakahara, A.; Wei, D. G.; Nordlund, D.; Russell, T. P., P3HT/PCBM Bulk Heterojunction Organic Photovoltaics: Correlating Efficiency and Morphology. Nano Lett 2011, 11 (2), 561-567. 28. Vanlaeke, P.; Swinnen, A.; Haeldermans, I.; Vanhoyland, G.; Aernouts, T.; Cheyns, D.; Deibel, C.; D'Haen, J.; Heremans, P.; Poortmans, J.; Manca, J. V., P3HT/PCBM bulk heterojunction solar cells: Relation between morphology and electro-optical characteristics. Sol Energ Mat Sol C 2006, 90 (14), 2150-2158. 29. Peet, J.; Kim, J. Y.; Coates, N. E.; Ma, W. L.; Moses, D.; Heeger, A. J.; Bazan, G. C., Efficiency enhancement in low-bandgap polymer solar cells by processing with alkane dithiols. Nat Mater 2007, 6 (7), 497-500. 30. Otero, R.; Ecija, D.; Fernandez, G.; Gallego, J. M.; Sanchez, L.; Martin, N.; Miranda, R., An organic donor/acceptor lateral superlattice at the nanoscale. Nano Lett 2007, 7 (9), 2602-2607. 31. Masui, A.; Sakaue, H.; Takahagi, T.; Suzuki, H., Intermixing behaviors of PCBM with CuPc on Au(111) surface. Chem Phys Lett 2016, 661, 215-218.

113

32. Gallego, J. M.; Ecija, D.; Martin, N.; Otero, R.; Miranda, R., An STM study of molecular exchange processes in organic thin film growth. Chem Commun 2014, 50 (69), 9954-9957. 33. Shao, Q.; Tskipuri, L.; Reutt-Robey, J. E., Vertical Phase Separation in Bilayer [6,6]-Phenyl-C-61-butyric Acid Methyl Ester:Zinc Phthalocyanine Films. J Phys Chem C 2014, 118 (32), 18612-18617. 34. Zheng, L. D.; Liu, J. G.; Han, Y. C., Polymer-regulated epitaxial crystallization of methanofullerene on mica. Phys Chem Chem Phys 2013, 15 (4), 1208-1215. 35. Hoppe, H.; Sariciftci, N. S., Morphology of polymer/fullerene bulk heterojunction solar cells. J Mater Chem 2006, 16 (1), 45-61. 36. Jin, Z. W.; Wang, J. Z., PIN architecture for ultrasensitive organic thin film photoconductors. Sci Rep-Uk 2014, 4. 37. Pratihar, P.; Ghosh, S.; Stepanenko, V.; Patwardhan, S.; Grozema, F. C.; Siebbeles, L. D. A.; Wurthner, F., Self-assembly and semiconductivity of an oligothiophene supergelator. Beilstein J Org Chem 2010, 6, 1070-1078. 38. Ji, A. C.; Xie, X. C.; Liu, W. M., Quantum magnetic dynamics of polarized light in arrays of microcavities. Phys Rev Lett 2007, 99 (18). 39. Yang, X. N.; van Duren, J. K. J.; Rispens, M. T.; Hummelen, J. C.; Janssen, R. A. J.; Michels, M. A. J.; Loos, J., Crystalline organization of a methanofullerene as used for plastic solar-cell applications. Adv Mater 2004, 16 (9-10), 802-+. 40. Tskipuri, L.; Shao, Q.; Reutt-Robey, J., Molecular Ordering in PCBM-Au(111) Interface Formation. J Phys Chem C 2012, 116 (41), 21874-21879. 41. Ecija, D.; Otero, R.; Sanchez, L.; Gallego, J. M.; Wang, Y.; Alcami, M.; Martin, F.; Martin, N.; Miranda, R., Crossover site-selectivity in the adsorption of the fullerene derivative PCBM on Au(111). Angew Chem Int Edit 2007, 46 (41), 7874-7877. 42. Li, H.; Duan, Y. Q.; Coropceanu, V.; Bredas, J. L., Electronic structure of the pentacene-gold interface: A density-functional theory study. Org Electron 2009, 10 (8), 1571-1578. 43. Huebener, K.; Scheloske, M.; Hauschild, J.; Harneit, W.; Zehl, G.; Fiechter, S., AFM investigation of the formation of one-dimensional structures of C(60) shuttlecocks on HOPG. Phys Status Solidi B 2006, 243 (13), 2990-2994. 44. Walczyk, W.; Schon, P. M.; Schonherr, H., The effect of PeakForce tapping mode AFM imaging on the apparent shape of surface nanobubbles. J Phys-Condens Mat 2013, 25 (18). 45. Girard, C.; Lambin, P.; Dereux, A.; Lucas, A. A., Van-Der-Waals Attraction between 2 C60 Fullerene Molecules and Physical Adsorption of C60 on Graphite and Other Substrates. Phys Rev B 1994, 49 (16), 11425-11432. 46. Wang, Y.; Alcami, M.; Martin, F., Understanding the supramolecular self-assembly of the fullerene derivative PCBM on gold surfaces. Chemphyschem 2008, 9 (7), 1030-1035. 47. Girifalco, L. A.; Hodak, M.; Lee, R. S., Carbon nanotubes, buckyballs, ropes, and a universal graphitic potential. Phys Rev B 2000, 62 (19), 13104-13110. 48. Branz, W.; Malinowski, N.; Enders, A.; Martin, T. P., Structural transition in (C-60)(n) clusters. Phys Rev B 2002, 66 (9). 49. Wang, C. I.; Hua, C. C., Solubility of C-60 and PCBM in Organic Solvents. J Phys Chem B 2015, 119 (45), 14496-14504. 50. Gravil, P. A.; Devel, M.; Lambin, P.; Bouju, X.; Girard, C.; Lucas, A. A., Adsorption of C-60 molecules. Phys Rev B 1996, 53 (3), 1622-1629. 51. Igathinathane, C.; Pordesimo, L. O.; Columbus, E. P.; Batchelor, W. D.; Methuku, S. R., Shape identification and particles size distribution from basic shape parameters using ImageJ. Comput Electron Agr 2008, 63 (2), 168-182. 52. Scott, D. W., Optimal and Data-Based Histograms. Biometrika 1979, 66 (3), 605-610.

114 This chapter closely follows the publication: Li, Y.; Liu, X.; Chen, C.; Duchamp, J.; Huang, R.; Chung, T.; Young, M.; Chalal, T.; Chen, Y.P.; Heflin, J.R.; Dorn, H.; Tao, C., Differences in Self-Assembly of Spherical C60 and Planar PTCDA on Rippled Graphene Surfaces, Carbon 145, 549 (2019)

Chapter 5

Differences in Self-Assembly of Spherical C60 and Planar

PTCDA on Rippled Graphene Surfaces

The majority of this chapter is from a manuscript published in Carbon, with slight modifications.

The experimental part of this chapter is done by Yanlong Li with the help of Chuanhui Chen. The

DFT calculation part is due to the work of Xiaoyang Liu.

5.1 Introduction

Graphene is a unique two-dimensional (2D) material that exhibits fascinating physical and

chemical properties and has a wide range of applications.1-3 For instance, thanks to its single-

atom thickness and flexibility, graphene is an excellent candidate for flexible electronics and gas

sensors.4-9 To optimize the applications of graphene and other 2D materials, it is essential to

investigate how curvature affects and tunes their properties. It has been reported that

graphene on rough substrates (e.g. SiO2) or suspended exhibits nonplanar aberrations.10-11

Furthermore, rippling the graphene to induce a curved surface would introduce variability into

the properties of graphene and changes interactions with adsorbed molecules, which has not

been experimentally examined.

Significant research efforts have recently been devoted to investigate the adsorption and

desorption of various molecules on planar graphene and other 2D materials, such as fabricating

115

and tuning molecule/graphene hybrid structures.12-16 Among the organic species, C60 and

perylenetetracarboxylic dianhydride (PTCDA) have attracted a huge amount of research

interest partially because they are key components, as effective electron acceptors, in

photovoltaic cells.12, 16-24 In the past two decades, the power conversion efficiency of organic

solar cells has rapidly increased, currently beyond 17%.25 To further improve the efficiency of

organic cells, it is necessary to understand the interactions between the organic species and

other building blocks like graphene, which is an excellent material for transparent electrodes in

solar cells.26-27 Development and study of hybrid nanostructures based on rippled graphene,

C60/rippled graphene, and PTCDA/rippled graphene could provide significant insights for

improving the efficiency of organic solar cells.

Previous experimental and computational studies have found that C60 and PTCDA on a planar

graphene surface form a hexagonal close packed (hcp) structure and a herringbone structure,

respectively.12-13, 16-19, 28-29 The major interaction present in the C60-planar graphene system is a

π-π stacking interaction.30 π-π stacking interactions are common in parallel aromatic systems,

have distances ranging from 3.0 to 4.0 Å and are mainly based on van der Waals forces.31-32 The

PTCDA-graphene system also contains π-π stacking interactions, but the dominant interaction

that leads to a herringbone pattern is intermolecular hydrogen bonding.17 We here report

significant experimental and computational differences of spherical C60 and planar PTCDA self-

assembled structures on rippled graphene surfaces. The inherent ability to tune the

interactions between rippled graphene and structurally different molecules will undoubtedly

open the door to interesting properties and potential applications of curved 2D materials, such

as flexible sensors.33-36

116

5.2 Experimental and Computational Methods

Experimental: All STM measurements were carried out in an ultra-high vacuum (UHV) scanning

tunneling microscope system (Omicron RT-STM). Before C60 deposition, the graphene was

grown using chemical vapor deposition (CVD) onto Cu foil37 and annealed for 12 hours at 673K

in a preparation chamber with a base pressure of 1 x 10-10 torr. PTCDA powder (TCI AMERICA,

99.0% purity) was loaded into the homemade Knudsen cell and mounted in the load lock side A

of the STM system. C60 powder (MER Corporation, 99.5% purity) was loaded into the

homemade Knudsen cell and mounted in the load lock side B of the STM system. The C60 and

PTCDA sources were degassed to 1 x 10-6 torr prior to deposition. C60 and PTCDA molecules

were then simultaneously deposited onto a same graphene at a deposition rate of ~ 0.5

monolayer/min with the background pressure below 1.8×10-8 torr. During the deposition

process, the substrate was kept at 413 K. The sample was subsequently annealed at 423 K for

one hour in the preparation chamber of the STM system with a base pressure of 1.0 x 10-10 torr.

All of the STM measurements were performed at room temperature with a base pressure of 1.9

× 10-10 torr. The STM used a chemically etched tungsten tip.

Computational: My collaborator Xiaoyang Liu does this computational part. Density functional

theory (DFT) based calculations are used to obtain further understanding of the self-assembled

systems. A model containing an adsorbed molecule and a curved graphene surface is used to

simulate the attachment of C60 and PTCDA on the rippled graphene surface. The structures of

C60 and PTCDA are fully optimized at the B3LYP level with def2-SVP basis set as provided in

ORCA 3.0.3. The curved graphene is constructed based on experiment observation and then is

optimized with constraints to maintain the bending angles and size. The energies of the

117

combined system are estimated based on single point calculations.38-40 DFT based approaches

with D3-correction are used to address the intermolecular interactions between the C60/PTCDA

molecules and the graphene.41-42 The distance between C60 and the curved graphene surface

are changed consistently and the system energy for each distance is calculated. The energy of

the C60- curved graphene complex is sensitive to the orientation of the C60 molecules on

graphene. To solve the orientation-related challenge, we employed a detailed minimum

potential search on representative orientations.43-44 Comparative calculations with molecules

adsorbed on planar graphene are also reported.

5.3 Discussions

The self-assembled structure of molecules adsorbed on graphene relates to the geometry of

graphene underneath. We identified three different patterns of graphene: planar graphene

labeled as I (Figures 5.1h and i)), one-dimensional (1D)-rippled graphene labeled as II (Figures

5.1a-d), 2D-rippled graphene labeled as III (Figures 5.1e and f). Planar graphene usually forms a

moiré pattern on Cu substrate. Figure 5.1i shows a moiré pattern with a hexagonal moiré super

lattice with periodicity of 2.0 nm. Besides the planar graphene areas, we also observed 1D-

rippled graphene (Figure 5.1a-c). Figure 5.1a shows the coexistence of planar graphene (I) and

1D-rippled graphene (II). Figure 5.1d, a line profile of the blue line in Figure 5.1b, shows a

periodicity of ~5 nm with amplitude of 0.23 nm. Typically, the peaks of the rippled graphene

have heights that range from 0.2 nm to 0.4 nm and periodicities that range from 3 nm to 10

nm. High resolution STM images (Figure 5.1c and Figure 5.2d and e) reveal a honeycomb lattice

118

of graphene on the top of each ripple. For 2D-rippled graphene (Figures 5.1e and f), the ripples

are along two perpendicular directions with the peak height and periodicity similar to 1D-

rippled graphene. Considering the heights of the ripples are much higher than the surface

smoothness of the underneath Cu(111), Cu(100) or Cu(110) that is in the range of less than 0.1

nm,34, 45 the line profiles of the ripples (Figure 5.2 b, c) indicate that the graphene in these areas

is quasi-suspended over the Cu substrate.

Figure 5.1. STM topographical images of planar graphene (labeled as I), 1D-rippled graphene (II) and 2D-

rippled graphene (III) on Cu. (a) Large area STM image of planar graphene (I) and 1D-rippled graphene

(II) showing the linear periodic modulation and the spatial modulation frequencies (Vs = -2.340 V, I =

119

0.110 nA). (b) High-resolution STM image of 1-D rippled graphene (Vs = -0.340 V, I = 1.900 nA). (c) STM

image of the 1-D rippled graphene, observed from the square region marked in (b), the schematic model

on top of the atomic image shows the ripples along zigzag direction (Vs = -0.280 V, I = 1.900 nA). (d) Line

profile perpendicular to the 1D-rippled graphene (marked as a blue line in (b)) showing the periodic

modulation. (e) STM image of graphene on two different Cu facets, planar graphene (I) and 2D-rippled

graphene (III) (Vs = -2.74 V, I = 0.045 nA). (f) High-resolution STM image of 2D-rippled graphene,

observed from the dashed square region marked in (e) (Vs = -2.600 V, I = 0.068 nA). (g) A schematic

model shows 1D-rippled graphene sheet. (h) Large area STM image of planar graphene (I) and 1D-

rippled graphene (II) (Vs = -1.850 V, I = 0.340 nA). (i) Atomic STM image showing the moiré pattern of

planar graphene, observed from the dashed squared region marked in (h) (Vs = -1.850 V, I = 0.450 nA).

The formation of rippled graphene is mainly due to the negative thermal expansion coefficient

of graphene.34-35, 45 The rippled graphene always emerges near the boundary of graphene. The

planar graphene appears on the Cu (111) facet and the 1D and 2D rippled graphene appear on

the Cu (100) and Cu (110) facets. Our measurements lead us to believe the rippled graphene is

caused by the negative thermal expansion coefficient of graphene and the interaction between

graphene and different Cu facets. As the sample is cooled from the annealing temperature, the

graphene expands as the Cu surfaces contract. The excess graphene on Cu surfaces leads to

graphene ripples. In the middle domain, the excess graphene diffuses towards planarity, while

near the boundaries the spatial constraints cause rippled graphene to form. Since graphene has

a stronger interaction with the Cu (111) facet, graphene preferentially forms a moiré pattern on

Cu (111) instead of a rippled pattern.

120

Figure 5.2. STM topography images of one single-twin rippled graphene. (a) STM topographic image of

continuous graphene forming twin wrinkles on Cu(111) (Vs = -1.06 V, I = 0.315 nA). (b) Zoomed-in image

of (a) (Vs = -0.600 V, I = 0.850 nA). (c) Line profile of the single-twin wrinkle of graphene, measured along

the blue line in b. (d)Further zoomed-in STM topographic image of single-twin wrinkle of graphene (Vs =

-0.560 V, I = 0.8750 nA). (e) Atomic STM image of one of the twin wrinkles in Figure (d) showing the

honeycomb structure of graphene (Vs = -0.560 V, I = 0.8750 nA). (f) Schematic model of one side of the

twin wrinkle.

We deposited C60 molecules on both rippled and planar graphene with low coverage, typically

less than 10%. On planar graphene, C60 molecules self-assemble to an hcp arrangement (Figure

5.3e) similar to previous studies.12, 16, 29 Figure 5.3f is the zoomed image of the area marked in

Figure 5.3e showing a moiré pattern. This moiré pattern originates from the structure of

121

graphene and the Cu (111) surface as shown in Figure 5.1i. For C60 on rippled graphene, the C60

self-assembly is more complicated than that on the planar case. C60 molecules form a quasi-hcp

structure as shown in Figures 5.3a-c at various scales. The quasi-hcp structure formed on

rippled graphene has a different angle from that formed on the planar area. For example, the

angle shown in Figure 5.3c is 54.1o instead of 60.0o shown in Figure 5.3f. The angle difference

between C60 on rippled graphene and planar graphene is due to the geometric curvature of the

rippled graphene. When compared to C60 adsorbed on planar graphene, the hcp structure on

rippled graphene is distorted by the curvature of the surface. The difference is also reflected in

the corresponding Fast Fourier Transform (FFT) images (the insets in Figures 5.3c and e).

Figure 5.3. STM images of C60 on 1D-rippled graphene (II) and on planar graphene (I). (a) Large area STM

topographic image of the C60 on 1D-rippled graphene showing well-defined linear periodic modulated

122

ripple (Vs = -2.00 V, I = 0.060 nA). (b) Zoomed-in STM image (measured from the dashed square of (a)) of

C60 on a long periodic graphene ripple (Vs = -2.60 V, I = 0.050 nA). (c) High-resolution image (measured

from the dashed square region of (b)) C60 on 1D-rippled graphene, shows a lattice angle α of 54.1o with a

quasi-hcp pattern (Vs = -2.60 V, I = 0.040 nA). Inset, the corresponding FFT image of (c). (d) A line profile

along the perpendicular direction of the 1D-rippled graphene marked with the blue line in (c) (top), side

view and top view showing the quasi-hcp C60 on 1D-rippled graphene (bottom). (e) Large area STM

image of the C60 on planar graphene with a well-defined moiré pattern on facet I (Vs = -2.65 V, I = 0.046

nA). (f) High-resolution STM image of C60 on planar graphene (measured from the square region of (e)),

showing a lattice angle β of 60.0o and a moiré pattern on facet I (Vs = -2.65 V, I = 0.046 nA). Inset, the

corresponding FFT image of (f).

The difference between the C60 structure formed on planar graphene and rippled graphene is

primarily due to the differences of van der Waals forces for peaks and valleys. Based on our DFT

calculations, C60 will initially deposit in the valleys of rippled graphene. As additional C60 is

deposited covering the peaks, the quasi-hcp structure is formed. DFT calculations identify the

binding energy of a C60-graphene valley site to be 0.34 eV more than the binding energy of a

C60-graphene peak site. The high-resolution STM image suggests that adsorbed C60 is not

continuously deposited in the valleys of the 1D-rippled graphene. The C60 molecules are not

fully revealed in the STM images because the STM tip is not sharp enough to measure into the

narrow valley regions.

123

In order to better understand the interactions between adsorbed molecules and rippled

graphene, we investigated PTCDA on 1D-rippled and planar graphene substrates. Similar to C60

on graphene, the coverage of PTCDA on rippled and planar areas is low, typically less than 8%.

The most common arrangement for PTCDA molecules on planar graphene is a herringbone

structure (Figures 5.4c and d). The high resolution STM image (Figure 5.4d) reveals a

herringbone arrangement with a1 = 1.3 nm, a2 = 1.96 nm, and ɣ = 90o, consistent with previous

reports.13, 17-18, 28 The inset in Figure 5.4c is the FFT image of the herringbone structure obtained

from an ordered area shown as the right part of Figures 5.5c. On 1D-rippled graphene, the

herringbone structure of the adsorbed PTCDA molecules is influenced by the graphene

curvature. In Figure 5.4b, we see there are a few PTCDA herringbone structures at the top right

corner, while other regions show PTCDA molecules forming a distorted herringbone pattern.

The FFT images of PTCDA on rippled and planar graphene (the insets in Figures 5.4b and 5.4c)

also show the difference.

124

Figure. 5.4. STM images of PTCDA on 1D-rippled graphene and on planar graphene. (a) Large area STM

image of PTCDA on 1D-rippled graphene (Vs = -2.51 V, I = 0.042 nA). (b) STM image of PTCDA on 1D-

rippled graphene showing a distorted herringbone pattern (Vs = -2.510 V, I = 0.042 nA). Inset, the

corresponding FFT image of (b). (c) Large area STM image of PTCDA on planar graphene (Vs = 1.800 V, I =

0.030 nA). Inset, the FFT image of the PTCDA herringbone structure on planar graphene. (d) Zoomed-in

STM image of PTCDA on planar graphene; a1 and a2 indicate the short and long lattice vectors of a unit

cell of the PTCDA herringbone pattern (Vs = 1.800 V, I = 0.030 nA). (e) STM images of coexistence of

substable PTCDA structure (purple curved region) and normal PTCDA herringbone structure (Vs = -2.500

V, I = 0.030 nA). (f) STM image of remaining normal PTCDA structure after the substable PTCDA was

removed by STM tip (Vs = -2.500 V, I = 0.030 nA).

Our results (Figures 5.4e, f and Figure 5.5) further show that PTCDA sub monolayer is very easy

to disassemble, due to the weak interaction between the PTCDA molecule and graphene on

125

copper substrate. During disassembly, we found a very interesting substable PTCDA structure

(purple curved region shown in Figures 5.4e and Figure 5.5) with a rectangular lattice. This

substable structure is formed by the interaction between PTCDA molecules and STM tip. After

growing to a critical size, the substable PTCDA would be moved away by the tip (Figures 5.5g

and h).

Figure 5.5. STM images of a set of PTCDA disassembly data from the self-assembled herringbone pattern

to two sub-stable arrangements on flat graphene type I on Cu. (a-i) are typical image of the disassembly

process. All these images were obtained under the same scanning conditions: Vs = -2.500 V, I = 0.030 nA,

and with the same size of 23 nm × 23 nm. The purple curved frames in the images show the sub-stable

arrangement.

When compared to planar graphene, molecules adsorbed onto 1D-rippled graphene show

different patterns. To obtain further understanding of the formation mechanism Xiaoyang Liu

126

have calculated the magnitude of the adsorbed molecule/curved graphene interaction. We

employed computational approaches based on DFT with van der Waals dispersion corrections

to explore the interactions between C60 molecules and the curved graphene surface. As

previously noted,46 the potential energy of the C60-graphene complex is sensitive to C60

molecular orientation on the graphene surface. Previous work established that the energy

minima of different C60 orientations are similar and are in the range of rotation energy

barriers.43 Inspired by previous studies, we inspect typical orientations (Figure 5.6). A detailed

examination of typical C60 orientations is employed to investigate the effects of orientations

and to find the most stable configuration. It has been confirmed that the offset face-to-face

alignment (Figure 5.6. b) is energetically favored and shows an ~1 kcal/mol lower energy than

other orientations.47-48 As shown in Figure 5.7 (a,b), there are two archetype locations on

curved graphene surface for arranging C60 molecules, the peak and the valley. C60 molecules

located on the peak area may be modeled with C60 on a convex aromatic surface and C60

molecules located in the valley area may be modeled with C60 on a concave aromatic surface. A

previous computational study revealed that C60 molecules on a concave aromatic surface have

larger intermolecular interactions which provide greater stability.43 As illustrated in Figure

(5.7d), a C60 molecule located on a graphene peak has a relative interaction energy of -0.92 eV

while the interaction energy for C60 in a graphene valley is -1.26 eV. The relative interaction

energy for C60 and planar graphene falls in between with a value of -1.07 eV. Calculations

reveal an optimized C60 molecule – curved graphene distance of 3.1 Å. The energy curves shown

in Figure 5.7d show a significant energy difference between C60 molecules located on a peak

and those located in a valley. Figure 5.8 shows results of DFT calculations for adsorbed

127

molecule-graphene interactions. The curved graphene structure increases the stability for C60

molecules in valley regions. The additional stability leads to the self-assembled quasi-hcp

configurations of C60 molecules on the 1D-rippled graphene surface.

Figure 5.6. Typical fullerene orientations on graphene. The computational results suggest that (b) is the

energetically favored orientation. Each grey sphere here is a carbon atom.

128

Figure 5.7. Computational results for C60 on 1D-rippled graphene and planar graphene showing

energetically favored orientations. (a) C60 molecule on a peak site on curved-graphene, (b) C60 molecule

with a valley site on curved-graphene and (c) C60 on planar-graphene. (d) Plot of C60-graphene distance

versus relative energy for C60 on a graphene peak (pink), C60 in a graphene valley (blue) and C60 on planar

graphene (green).

In contrast to the hcp pattern for C60 molecules on a planar graphene surface, it has been

shown that PTCDA molecules form herringbone structure on a planar graphene surface.

129

However, on a 1D-rippled graphene surface, we observed a disordered herringbone pattern for

the PTCDA molecules. The disordered structure implies that the curved graphene surface has a

significant effect on the self-assembly of the PTCDA molecules. Xiaoyang utilized DFT-based

calculations to obtain further understanding on the structures of a PTCDA molecule on a 1D-

curved graphene surface. Our calculations show the relative energies between PTCDA and

curved graphene on a peak and in a valley are -2.17 eV and -2.61 eV, respectively. This may be

compared to a relative energy of -2.40 eV for PTCDA on planar graphene. Calculations reveal

that PTCDA molecules prefer different orientations on a graphene peak and in a graphene

valley as shown in Figure 5.8a, b. A PTCDA molecule in a valley aligns symmetrically while the

lowest relative energy for a molecule on a graphene peak makes a 30˚ angle with the ridge

(Figure 5.8c). A possible explanation for the angle is the electronegative oxygens minimizing

contact with the graphene while maximizing π-π stacking interactions. The difference in

preferred orientations for PTCDA molecules on peaks and in valleys show that curved graphene

may be used to help to regulate the orientation of molecules. Compared to the C60/curved

graphene interaction, the PTCDA/curved graphene interactions are slightly stronger due to the

larger contact area between PTCDA molecules and the curved graphene. In both cases,

adsorbed molecules in valley sites show a stronger interaction than the same molecule

adsorbed on planar graphene.

130

Figure 5.8. DFT results for adsorbed molecule/graphene interactions. (a) Energy difference of a C60

molecule on a 1D-rippled graphene surface (b) Energy difference of a PTCDA molecule on a 1D-rippled

graphene. (c) Energy curve for a PTCDA molecule rotation on 1D-rippled graphene on a peak location

(top). (d) Favored PTCDA orientations at the peak site (left) and at the valley site (right).

5.4 Conclusion

In this chapter, we have demonstrated the self-assembly of C60 and PTCDA molecules on rippled

graphene with characterization using both experimental STM and DFT calculations. The

adsorbed molecules on 1D-rippled graphene systems reveal distortions when compared with

the analogous planar graphene system. Specifically, the nearly spherical Ih-symmetrical C60

molecules form a quasi-hexagonal close packed structure, while the planar PTCDA molecules

form a disordered herringbone structure on the rippled graphene surface. The change in the

131

monolayer packing pattern of C60 and PTCDA molecules on a curved graphene surface is due to

competition between the adsorbate-graphene interaction and the intermolecular adsorbate

interactions. Because of the nearly spherical C60 molecules, the rippled graphene surface

exhibits only diminished effects on the π-π intermolecular interactions. In contrast, the planar

PTCDA molecules have different sides for intermolecular interactions, namely, sp2 hybridized on

opposite two sides and significantly more electronegative oxygen anhydride moieties on the

other two sides. This leads to the well characterized herringbone structure that is assembled

mainly based on the relative weak PTCDA hydrogen bonds, C-H…O, with a strength estimated

as 0.1 eV. In this case, the PTCDA-graphene interaction, which have energies at the peak and

valley sites of 2.17 and 2.61 eV, respectively, are far more important than the PTCDA

intermolecular interaction. These results are also consistent with the tendency for dissembling

the PTCDA submonolayer vide supra. Furthermore, the DFT computational results demonstrate

significant increases in π-π interactions for both the adsorbed PTCDA and C60/rippled graphene

complexes located in the 2D graphene valley sites in comparison with adsorbed more idealized

molecule/planar graphene 2D complexes. In addition, we find that the adsorbed planar PTCDA

molecules prefer different orientations when the rippled graphene peak regions are compared

to the valley regions. These fundamental experimental and computational results are important

for understanding any potential application of structurally diverse molecules adsorbed on

graphene and/or rippled graphene surfaces.

132

References:

1. Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.; Katsnelson, M. I.; Grigorieva, I. V.; Dubonos, S. V.; Firsov, A. A., Two-dimensional gas of massless Dirac fermions in graphene. Nature 2005, 438 (7065), 197-200. 2. Zhang, Y. B.; Tan, Y. W.; Stormer, H. L.; Kim, P., Experimental observation of the quantum Hall effect and Berry's phase in graphene. Nature 2005, 438 (7065), 201-204. 3. Castro Neto, A. H.; Guinea, F.; Peres, N. M. R.; Novoselov, K. S.; Geim, A. K., The electronic properties of graphene. Rev Mod Phys 2009, 81 (1), 109-162. 4. Georgiou, T.; Jalil, R.; Belle, B. D.; Britnell, L.; Gorbachev, R. V.; Morozov, S. V.; Kim, Y. J.; Gholinia, A.; Haigh, S. J.; Makarovsky, O.; Eaves, L.; Ponomarenko, L. A.; Geim, A. K.; Novoselov, K. S.; Mishchenko, A., Vertical field-effect transistor based on graphene-WS2 heterostructures for flexible and transparent electronics. Nat Nanotechnol 2013, 8 (2), 100-103. 5. Kim, Y. H.; Kim, S. J.; Kim, Y. J.; Shim, Y. S.; Kim, S. Y.; Hong, B. H.; Jang, H. W., Self-Activated Transparent All-Graphene Gas Sensor with Endurance to Humidity and Mechanical Bending. Acs Nano 2015, 9 (10), 10453-10460. 6. Sentef, M. A.; Claassen, M.; Kemper, A. F.; Moritz, B.; Oka, T.; Freericks, J. K.; Devereaux, T. P., Theory of Floquet band formation and local pseudospin textures in pump-probe photoemission of graphene. Nat Commun 2015, 6, 7047(1-8). 7. Geim, A. K., Graphene: Status and Prospects. Science 2009, 324 (5934), 1530-1534. 8. Zurutuza, A.; Marinelli, C., Challenges and opportunities in graphene commercialization. Nat Nanotechnol 2014, 9 (10), 730-734. 9. Avouris, P.; Xia, F. N., Graphene applications in electronics and photonics. Mrs Bull 2012, 37 (12), 1225-1234. 10. Bao, W. Z.; Miao, F.; Chen, Z.; Zhang, H.; Jang, W. Y.; Dames, C.; Lau, C. N., Controlled ripple texturing of suspended graphene and ultrathin graphite membranes. Nat Nanotechnol 2009, 4 (9), 562-566. 11. Ishigami, M.; Chen, J. H.; Cullen, W. G.; Fuhrer, M. S.; Williams, E. D., Atomic structure of graphene on SiO2. Nano Lett 2007, 7 (6), 1643-1648. 12. Li, G.; Zhou, H. T.; Pan, L. D.; Zhang, Y.; Mao, J. H.; Zou, Q.; Guo, H. M.; Wang, Y. L.; Du, S. X.; Gao, H. J., Self-assembly of C-60 monolayer on epitaxially grown, nanostructured graphene on Ru(0001) surface. Appl Phys Lett 2012, 100 (1), 013304(1-4). 13. Wang, Q. H.; Hersam, M. C., Room-temperature molecular-resolution characterization of self-assembled organic monolayers on epitaxial graphene. Nat Chem 2009, 1 (3), 206-211. 14. MacLeod, J. M.; Rosei, F., Molecular Self-Assembly on Graphene. Small 2014, 10 (6), 1038-1049. 15. Jarvinen, P.; Hamalainen, S. K.; Ijas, M.; Harju, A.; Liljeroth, P., Self-Assembly and Orbital Imaging of Metal Phthalocyanines on a Graphene Model Surface. J Phys Chem C 2014, 118 (24), 13320-13325. 16. Chen, C. H.; Zheng, H. S.; Mills, A.; Heflin, J. R.; Tao, C. G., Temperature Evolution of Quasi-one-dimensional C-60 Nanostructures on Rippled Graphene. Sci Rep-Uk 2015, 5, 14336(1-7). 17. Martinez-Galera, A. J.; Nicoara, N.; Martinez, J. I.; Dappe, Y. J.; Ortega, J.; Gomez-Rodriguez, J. M., Imaging Molecular Orbitals of PTCDA on Graphene on Pt(111): Electronic Structure by STM and First-Principles Calculations. J Phys Chem C 2014, 118 (24), 12782-12788. 18. Emery, J. D.; Wang, Q. H.; Zarrouati, M.; Fenter, P.; Hersam, M. C.; Bedzyk, M. J., Structural analysis of PTCDA monolayers on epitaxial graphene with ultra-high vacuum scanning tunneling microscopy and high-resolution X-ray reflectivity. Surf Sci 2011, 605 (17-18), 1685-1693. 19. Lu, J.; Yeo, P. S. E.; Zheng, Y.; Yang, Z. Y.; Bao, Q. L.; Gan, C. K.; Loh, K. P., Using the Graphene Moire Pattern for the Trapping of C-60 and Homoepitaxy of Graphene. Acs Nano 2012, 6 (1), 944-950.

133

20. Sariciftci, N. S.; Smilowitz, L.; Heeger, A. J.; Wudl, F., Photoinduced Electron-Transfer from a Conducting Polymer to Buckminsterfullerene. Science 1992, 258 (5087), 1474-1476. 21. Morita, S.; Zakhidov, A. A.; Yoshino, K., Doping Effect of Buckminsterfullerene in Conducting Polymer - Change of Absorption-Spectrum and Quenching of Luminescence. Solid State Commun 1992, 82 (4), 249-252. 22. Smilowitz, L.; Sariciftci, N. S.; Wu, R.; Gettinger, C.; Heeger, A. J.; Wudl, F., Photoexcitation Spectroscopy of Conducting-Polymer-C(60) Composites - Photoinduced Electron-Transfer. Phys Rev B 1993, 47 (20), 13835-13842. 23. Peumans, P.; Yakimov, A.; Forrest, S. R., Small molecular weight organic thin-film photodetectors and solar cells. J Appl Phys 2003, 93 (7), 3693-3723. 24. Rand, B. P.; Burk, D. P.; Forrest, S. R., Offset energies at organic semiconductor heterojunctions and their influence on the open-circuit voltage of thin-film solar cells. Phys Rev B 2007, 75 (11), 115327(1-11). 25. Meng, L.; Zhang, Y.; Wan, X.; Li, C.; Zhang, X.; Wang, Y.; Ke, X.; Xiao, Z.; Ding, L.; Xia, R.; Yip, H.-L.; Cao, Y.; Chen, Y., Organic and solution-processed tandem solar cells with 17.3% efficiency. Science 2018, 361 (6407), 1094-1098. 26. Wang, X.; Zhi, L. J.; Mullen, K., Transparent, conductive graphene electrodes for dye-sensitized solar cells. Nano Lett 2008, 8 (1), 323-327. 27. Wu, J. B.; Becerril, H. A.; Bao, Z. N.; Liu, Z. F.; Chen, Y. S.; Peumans, P., Organic solar cells with solution-processed graphene transparent electrodes. Appl Phys Lett 2008, 92 (26), 263302(1-3). 28. Huang, H.; Chen, S.; Gao, X. Y.; Chen, W.; Wee, A. T. S., Structural and Electronic Properties of PTCDA Thin Films on Epitaxial Graphene. Acs Nano 2009, 3 (11), 3431-3436. 29. Cho, J.; Smerdon, J.; Gao, L.; Suzer, O.; Guest, J. R.; Guisinger, N. P., Structural and Electronic Decoupling of C-60 from Epitaxial Graphene on SiC. Nano Lett 2012, 12 (6), 3018-3024. 30. Reveles, J. U.; Karle, N. N.; Baruah, T.; Zope, R. R., Electronic and Structural Properties of C-60 and Sc3N@C-80 Supported on Graphene Nanoflakes. J Phys Chem C 2016, 120 (45), 26083-26092. 31. Mishra, B. K.; Arey, J. S.; Sathyamurthy, N., Stacking and Spreading Interaction in N-Heteroaromatic Systems. J Phys Chem A 2010, 114 (36), 9606-9616. 32. Gung, B. W.; Wekesa, F.; Barnes, C. L., Stacking interactions between nitrogen-containing six-membered heterocyclic aromatic rings and substituted benzene: Studies in solution and in the solid state. J Org Chem 2008, 73 (5), 1803-1808. 33. Tapaszto, L.; Dumitrica, T.; Kim, S. J.; Nemes-Incze, P.; Hwang, C.; Biro, L. P., Breakdown of continuum mechanics for nanometre-wavelength rippling of graphene. Nat Phys 2012, 8 (10), 739-742. 34. Tian, J. F.; Cao, H. L.; Wu, W.; Yu, Q. K.; Guisinger, N. P.; Chen, Y. P., Graphene Induced Surface Reconstruction of Cu. Nano Lett 2012, 12 (8), 3893-3899. 35. Bai, K. K.; Zhou, Y.; Zheng, H.; Meng, L.; Peng, H. L.; Liu, Z. F.; Nie, J. C.; He, L., Creating One-Dimensional Nanoscale Periodic Ripples in a Continuous Mosaic Graphene Monolayer. Phys Rev Lett 2014, 113 (8), 086102(5). 36. Deng, S. K.; Berry, V., Wrinkled, rippled and crumpled graphene: an overview of formation mechanism, electronic properties, and applications. Mater Today 2016, 19 (4), 197-212. 37. Yu, Q. K.; Jauregui, L. A.; Wu, W.; Colby, R.; Tian, J. F.; Su, Z. H.; Cao, H. L.; Liu, Z. H.; Pandey, D.; Wei, D. G.; Chung, T. F.; Peng, P.; Guisinger, N. P.; Stach, E. A.; Bao, J. M.; Pei, S. S.; Chen, Y. P., Control and characterization of individual grains and grain boundaries in graphene grown by chemical vapour deposition. Nat Mater 2011, 10 (6), 443-449. 38. Kruse, H.; Grimme, S., A geometrical correction for the inter- and intra-molecular basis set superposition error in Hartree-Fock and density functional theory calculations for large systems. J Chem Phys 2012, 136 (15), 154101(16).

134

39. Schafer, A.; Horn, H.; Ahlrichs, R., Fully Optimized Contracted Gaussian-Basis Sets for Atoms Li to Kr. J Chem Phys 1992, 97 (4), 2571-2577. 40. Schafer, A.; Huber, C.; Ahlrichs, R., Fully Optimized Contracted Gaussian-Basis Sets of Triple Zeta Valence Quality for Atoms Li to Kr. J Chem Phys 1994, 100 (8), 5829-5835. 41. Grimme, S., Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J Comput Chem 2006, 27 (15), 1787-1799. 42. Grimme, S.; Ehrlich, S.; Goerigk, L., Effect of the Damping Function in Dispersion Corrected Density Functional Theory. J Comput Chem 2011, 32 (7), 1456-1465. 43. Cabaleiro-Lago, E. M.; Rodriguez-Otero, J.; Carrazana-Garcia, J. A., A theoretical study of complexes between fullerenes and concave receptors with interest in photovoltaics. Phys Chem Chem Phys 2017, 19 (39), 26787-26798. 44. Gonzalez-Veloso, I.; Rodriguez-Otero, J.; Cabaleiro-Lago, E. M., Carbon-nanorings ([10]CPP and [6]CPPA) as fullerene (C-60 and C-70) receptors: a comprehensive dispersion-corrected DFT study. Phys Chem Chem Phys 2016, 18 (46), 31670-31679. 45. Gao, L.; Guest, J. R.; Guisinger, N. P., Epitaxial Graphene on Cu(111). Nano Lett 2010, 10 (9), 3512-3516. 46. Ojeda-Aristizabal, C.; Santos, E. J. G.; Onishi, S.; Yan, A. M.; Rasool, H. I.; Kahn, S.; Lv, Y. C.; Latzke, D. W.; Velasco, J.; Crommie, M. F.; Sorensen, M.; Gotlieb, K.; Lin, C. Y.; Watanabe, K.; Taniguchi, T.; Lanzara, A.; Zettl, A., Molecular Arrangement and Charge Transfer in C-60/Graphene Heterostructures. Acs Nano 2017, 11 (5), 4686-4693. 47. Headen, T. F.; Howard, C. A.; Skipper, N. T.; Wilkinson, M. A.; Bowron, D. T.; Soper, A. K., Structure of pi-pi Interactions in Aromatic Liquids. J Am Chem Soc 2010, 132 (16), 5735-5742. 48. Feng, C.; Lin, C. S.; Fan, W.; Zhang, R. Q.; Van Hove, M. A., Stacking of polycyclic aromatic hydrocarbons as prototype for graphene multilayers, studied using density functional theory augmented with a dispersion term. J Chem Phys 2009, 131 (19), 194702(8).

135

Chapter6

Self-Assembled Gd3N@C80 Monolayer on Graphene

Examined by STM

The experimental part of this chapter is done by Yanlong Li. The DFT calculation part is due to

the work of Xiaoyang Liu. The majority of this chapter is based on a submitted manuscript.

6.1 Introduction and Background

Since their discovery back to the 1980s, fullerenes have attracted extensive attention due to

their extraordinary electronic and structural properties.1 The hollow structure of a fullerene is

an ideal container to encapsulate clusters or molecules inside the cage.2 In 1991, the Smally

group reported the first synthesized and isolated metallofullerene, La@C82,3 and a large class of

metallofullerenes with Sc, Y and most of the lanthanoid elements clusters encaged have been

discovered.4 After that, metallofullerenes were studied extensively and it has been well

established that metallofullerenes have potential in various applications, such as biomedical

diagnostics and therapeutics5-6 and solar cells.7 However, the synthetic yield of

metallofullerenes is generally at a low level, which is a major difficulty that limits further

investigations and applications of metallofullerene materials. In 1999, the Dorn group first

reported the synthesis of the first trimetallic nitride (TNT) endohedral metallofullerene (EMF),

Sc3N@C80, and recently, a large family of TNT EMFs have been discovered and utilized in

136

research and in industry.8-10 For example, Gd3N@C80 has been applied as a magnetic resonance

imaging contest agent and a large increase in performance has been reported.11 Currently, a

non-toxic derivative of Gd3N@C80 is synthesized to selectively detect a tumor and has the

potential to be used in future MRI diagnosis.6 In addition, the electronic properties have been

studied and metallofullerenes have been employed in fabricating solar cells. In 2006, the

Akasaka group reported their studies on the reverse ground electron transfers between La@C82

and an organic donor, N, N, N’, N’-tetramethyl-p-phen-ylenediamine (TMPD) in solution.12 In

the Dorn group’s previous studies, they observed the electron transfer between various

metallofullerenes and organic donors.7, 13 The facile electron transfer property of

metallofullerenes make them good candidates for the electron transfer layer material in solar

cells. In addition, the hydrophobic carbon cage prevents possible performance loss because of

moisture. Recently, Wang et al. reported their studies in a novel perovskite solar cells utilizing

Sc3N@C80 as the electron transfer layer material to achieve a high energy conversion efficiency

and an extraordinary stability at the same time.7 Although the novel electronic properties have

been widely employed for various devices and achieve much performance improvements, the

fundamental mechanism of the electron transfer is still open and further investigations

regarding the electron transfer are needed.

Single layer graphene, which provides a large scale aromatic surface, is an ideal example to

study the adsorption behaviors. It has been recognized that C60 shows a hexagonal close packed

structure on the graphene surface. In 2015, Chen et al. arranged C60 molecules on a rippled

graphene surface and observed organized quasi-1D C60 chain structures with a two or three

molecules width.14 The effects of the ripple region on the adsorption of molecules on a

137

graphene surface is quite interesting since ripple areas extensively exist, which may shift the

adsorption performance away from our expectation. It has been well established that the

spherical fullerene molecules are only slightly affected by the rippled area and can still conserve

the hexagonal close packed structure, however, other planar molecules will lose the order in

rippled regions.15 In a seminal study, direct imaging of endohedral metallofullerenes, Sc2@C84-

D2d and Sc2@C84-D2, on a silicon surface has been reported, and a quasi-close pack structures

were formed on the surface with no Sc atoms directly detected, which presented the evidence

that the Sc metals are encapsulated inside the cage.16 A theoretical study of the complexity of

the metalofullerene Sc3N@C80 on a graphene flake surface was reported and the adsorptions of

fullerenes and metallofullerenes were found to vary due to the inside cluster.17 Computational

results also indicate electron transfers and the interactions between the metallofullerene

molecules and the graphene surface. Experimental investigations of metallofullerenes on a

graphene surface are limited and the mechanism of the adsorption of metallofullerenes

remains unknown.

Several factors dominate the packing of organic molecules on the graphene surface. In previous

studies, it has been shown that molecules of different shapes prefer different packing

structures to minimize the energies of the 2D layer system.15, 18-20 For example, a planar

perylenetetracarboxylicdianhydride (PTCDA) molecule forms a herring bone structure on a

planar graphene surface.21 While, a spherical C60 molecule prefers a hexagonal close packed

(hcp) structure.21-22 In our previous study (Chapter 5), we found that the nonplanar aberrations

of the graphene surface also affect the packing styles of molecules.15 It has been shown that

rippled areas on the graphene surface have distinct effects on molecules of different sizes and

138

shapes. For examples, the valley sites of the rippled graphene surface have stronger

interactions with C60 than the peak sites. However, the rippled area breaks the ordered packing

structure of PTCDA. Several other factors also impact the interactions between organic

molecules and the graphene surface and between neighbor organic molecules.

Metallofullerenes provide readily available examples to study the subtle effect of inner clusters.

The inner cluster of a metallofullerene transfers electrons on the outer cage and slightly

changes the shape of the cage.9, 24 The electron distribution of the outer cage is also influenced

by the inner cluster. In 2018, Dr. Warner published a paper about the Gd atoms studied by

deposition Gd3N@C80 on graphene by using TEM.25 At room temperature, the orientation of 3

Gd atoms are random in the C80 cage. While, with the temperature increasing to above 500 oC,

the Gd3N@C80 molecules would break down. Therefore, the orientations of the inner cluster

may vary the interactions and then slightly alter the packing style.

Here, we examine Gd3N@C80 molecules on the chemical vapor deposition (CVD) graphene

surface. The metallofullerenes are deposited on the graphene surface following a similar

procedure as depositing C60 molecules. The complexities have been treated with various

annealing temperatures to achieve organized structures and make it convenient for STM

observations. In this study, we investigate the interactions between metallofullerenes and the

graphene surface focusing on the effects of the inner metal clusters. The results suggest that

Gd3N@C80 forms hexagonal close packed (hcp) structure, which is the same as for C60

molecules, although the C80-Ih cage shifts from a perfect sphere to an ellipsoid shape. The

experimental observations are then augmented by DFT based computations to understand the

structures. Various orientations of the inner cluster are examined and the energy favored

139

orientation is reported. The interactions between two Gd3N@C80 molecules are also explored.

The results are compared with previously reported C60-graphene 2D layer system.

6.2 Experimental and Computational Methods

Experimental: All STM measurements were carried out in an ultra-high vacuum (UHV) scanning

tunneling microscope system (Omicron RT-STM). Before Gd3N@C80 deposition, the graphene

was grown using chemical vapor deposition (CVD) onto Cu foil and annealed for 12 hours at

673K in a preparation chamber with a base pressure of 4.0 x 10-10 torr. Gd3N@C80 powder was

loaded into the homemade Knudsen cell and mounted in the load lock of the STM system. The

Gd3N@C80 source was degassed to 1 x 10-7 torr prior to deposition. Gd3N@C80 molecules were

then deposited onto graphene at a deposition rate of ~ 0.1 monolayer/min with the

background pressure below 1.8×10-8 torr. During the deposition process, the substrate was kept

at 353 K. The sample was subsequently annealed at 423 K for two hours in the preparation

chamber of the STM system with a base pressure of 4.0 × 10-10 torr. All of the STM

measurements were performed at room temperature with a base pressure of 4.0 × 10-10 torr.

The STM used a chemically etched tungsten tip.

Density functional theory (DFT) based calculations (done by Xiaoyang Liu) are used to obtain

further understanding of the self-assembled systems. A model containing an adsorbed

molecule and a planar graphene surface is used to simulate the attachment of

metallofullerenes (Gd3N@C80) on the graphene surface. The structures of all molecules are fully

optimized at the B3LYP level with def2-SVP basis set as provided in ORCA 3.0.3. The graphene is

140

represented as a large scale graphene flake to simplify the computations. The energies of

combined system are estimated based on single point calculation. DFT based approaches with

D3-correction are used to address the intermolecular interactions between the Gd3N@C80

molecules and the graphene. The distance between metallofullerenes and the graphene surface

are changed consistently and system energy for each distance is calculated. The energy of the

metallofullerene/graphene complex is sensitive to the orientation of the metallofullerene

molecules on graphene. To solve the orientation-related challenge, Xiaoyang employed a

detailed minimum potential search on representative orientations. Comparative calculations

with molecules adsorbed on planar graphene are also reported.

6.3 Results and Discussion

In a seminal study, it has been demonstrated that graphene grown on a metal surface by

chemical vapor deposition (CVD) is an ideal platform for the arrangement of numerous

molecules. Graphene shows strong ability for the adsorption and desorption of various

molecules, especially those with a large π system. Significant efforts have been devoted to

investigate the arrangement of molecules on the graphene surface.14, 15 In a previous study, a

quasi-one-dimensional (quasi-1D) C60 nanostructure on rippled graphene surface has been

reported,14 which provides a unique class for fabricating C60/graphene hybrid structures. The

absorption ability and the arrangement of molecules on the graphene surface are dominated by

the structures of the organic molecules as well as the graphene form. We previously reported

that rippled areas on graphene surface affect the arrange of organic molecules (Chapter 5).15

141

The arrangement of organic molecules on rippled graphene are determined by the structures as

well as the folding of graphene. There are also subtle factors affecting the packing styles and

the interactions between organic molecules. Since the interaction between an organic molecule

and a graphene surface is controlled by π-π stacking interactions, the electron distribution and

the electron density of the organic molecule and graphene change the interactions and

therefore, change the packing styles.

At a pressure below 2.0 × 10-8 mbar, we deposited Gd3N@C80 molecules on the flat graphene

surface. Then, in the large scale STM image (Figure 6.1c), we observed atomically flat facets

with sizes that range from 50 nm to 200 nm, which confirms the successful monolayer

deposition of Gd3N@C80 on the graphene surface. The yellow dashed line shows that the step

height of this domain is about 0.88 nm. When compared to the step height of C60 about 0.67

nm (Figure 6.1a), the step height of Gd3N@C80 is 0.2 nm larger, which is larger than the

difference between the size of C60 and Gd3N@C80. The difference here is mainly due to the

difference in the local density of states between Gd3N@C80 and C60, which affects the apparent

height by STM.

142

Figure. 6.1. (a) STM image of C60 monolayer on graphene (Vs = 1.50 V, I = 0.050 nA). (b) The left image is

the line profile of monolayer C60 (yellow dashed line in Figure 6.1a). The right image is the schematic

image of C60 on graphene with a gap about 0.3 nm, according to our DFT calculation. (c) STM image of

Gd3N@C80 monolayer on graphene (Vs = -1.84 V, I = 0.140 nA). (d) The left image is the line profile of

monolayer Gd3N@C80 (red dashed line in Figure 6.1c). The right image is the schematic image of

Gd3N@C80 on graphene with a gap about 0.33 nm, according to our DFT calculation.

The STM atomic resolution images (Figure 6.2a) represent the hcp structure of Gd3N@C80

molecules and the closer zoom (Figure 6.2b) illustrates more details about the arrangement

pattern that this hcp structure only forms over a small area with many defects. The two

domains (Figure 6.2b) shows different orientations with one defect in the middle. Figure 6.2d

shows that the lattice distance of Gd3N@C80 is about 1.15 nm

143

Figure 6.2. STM images of Gd3N@C80 on graphene. (a) Large area STM topographic image of the

Gd3N@C80 on graphene showing two domains with different orientations and many defects (Vs = -1.84

V, I = 0.248 nA). (b) Zoomed-in STM image (measured from the dashed square of (a)) of Gd3N@C80 on

graphene showing two domains with different orientations and one defect (Vs = -1.84 V, I = 0.248 nA).

(c) High-resolution image Gd3N@C80 on graphene (Vs = -1.84 V, I = 0.240 nA). (d) Zoomed-in STM image

(measured from the dashed square of (c)) of Gd3N@C80 on a graphene showing lattice constant with an

average about 1.15 nm (Vs = -1.84 V, I = 0.240 nA).

In this study, we also inspect the temperature evolution of Gd3N@C80 nanostructures and alter

the annealing temperatures to investigate the corresponding detailed arrangements (Figure

144

6.3). After annealing at 150 oC for 2 hours, Gd3N@C80 forms an hcp structure with relatively less

defects. While, with the annealing temperature increasing, Gd3N@C80 molecules tend to

aggregate together. After annealing at 200 oC for 2 hours, the Gd3N@C80 molecules still forms

the hcp structure, while some nanobubbles form on the monolayer. With higher annealing

temperature (250 oC) for 2 hours, Gd3N@C80 molecules aggregate together without forming any

ordered structure.

Figure. 6.3. (a) STM image of Gd3N@C80 on graphene after annealing at 200 oC (Vs = -1.69 V, I = 0.122

nA). (b) STM image of Gd3N@C80 on graphene after annealing at 250 oC (Vs = 1.50 V, I = 0.100 nA).

As illustrated in previous and the current studies, the highly ordered arrangements of C60 shows

relatively strong long range intermolecular interactions, while the lower ordered structure of

Gd3N@C80 indicates a relatively weak long range intermolecular interactions. It has been long

recognized that two π systems have noncovalent interactions and the π-π stacking interactions

occur in biological molecules, such as DNA, RNA and proteins, and play an essential role in

determining their 3D structures.26 It has been reported that the stacking interaction not only

145

exists between two planar surfaces but also between curved-planar and curved-curved

surfaces.15 Thus, Xiaoyang employ theoretical calculations based on DFT with dispersion

correction augmenting the experimental results to explore the interactions between Gd3N@C80

and graphene and to understand the scattering pattern at different annealing temperature. It

has been well recognized that the minima energy is relevant to the orientation of Gd3N@C80

and it is too cumbersome and time consuming to traverse all possible orientations with

different orientations of Gd3N@C80 on graphene (Figure 6.4). A seminal study shows that a

partial search of typical structures rather than a full search is able to obtain the most stable

structures.27-28 We also confirm that the minima are quite similar for different orientations of

Gd3N@C80. In Figure 6.5, the line profiles obtained by STM experiment shows the energy

favored orientation is Orientation 3 and the distance between Gd3N@C80 and the graphene

surface in Orientation 3 is 3.3 Ȧngstrom. The energy differences between different orientations

is very small, being less than 0.09 eV, while the energy differences between two nearest

orientations is 0.02 eV which is comparable to room temperature thermal energy (0.025 eV).

This calculation explains why the Gd3N@C80 molecules show random orientations when

deposited on graphene, as the energy gap between two nearest orientations are comparable to

the thermal energy at room temperature, which also support our STM results that Gd3N@C80

molecules form an hcp structure over small distances with many defects. Firstly, the geometry

of Gd3N@C80 molecule is more like an out of sphere shape than a perfect sphere shape, as the

four inner atoms are too big and expand the C80 cage. What’s more, the energy differences

between different orientations is very small and comparable to thermal energy. These two

146

factors make it difficult for Gd3N@C80 molecules to form perfect ordered hcp structure over

long distances.

Figure. 6.4. Typical Gd3N@C80 orientations on graphene. The computational results suggest that

Orientation 3 is the energetically favored orientation.

As illustrated by the energy curves in Figure 6.5, the energy of the system decreases first as the

distance between a Gd3N@C80 molecule and the graphene surface increases. After reaching a

minimum, the energy increases again. This is due to the combined effect of affinity and repulse,

and shows that the interactions between a Gd3N@C80 molecule and the graphene is dominated

by π-π stacking. The computed distance is 3.4 – 3.5 Ȧngstroms, which is not consistent with the

147

distance measured by STM experiments as a result of differences between the local density of

state of graphene and Gd3N@C80 molecules.

Figure. 6.5. Computational results for Gd3N@C80 on flat graphene of different orientations, showing

energetically favored orientation 3.

Figure 6.6a shows π-π interaction between two Gd3N@C80 molecules with two orientations,

which is two molecules with the two metal sides facing each other with an angle between two

Gd atoms facets and two metal atoms facing each other with an angle between two Gd atoms

facets respectively. The lowest energy (Orientation II) is 0.32 eV with the distance about 3.5

148

Ȧngstroms between molecules. The energy difference between the two orientations is only

0.02 eV, which is comparable to the thermal energy. According to our STM results, the lattice

constant of the Gd3N@C80 hcp structure is 1.15 nm, subtracting the size of the molecule of 0.8

nm, the gap between two molecules is about 0.35 nm which is highly consistent with our DFT

calculation.

Figure. 6.6. (a) DFT results for the molecule-molecule interaction of two different orientations between

two Gd3N@C80 molecules. (b) The two Gd3N@C80 molecules with two metal sides facing each other with

an angle between two Gd atoms facets. (c) The two Gd3N@C80 molecules with two metal atoms facing

each other with an angle between two Gd atoms facets.

149

6.4 Conclusion

In summary, we observed an hcp structure of Gd3N@C80 molecules on flat graphene with

characterization by both experimental STM and DFT calculations. Compared to the perfect

structure of C60 over long range, Gd3N@C80 molecules form an hcp structure over a small range

with many defects. The lattice constant that we measured is 1.15 nm and the monolayer step

height is about 0.87 nm. In addition, we studies the temperature evolution of a Gd3N@C80

monolayer on graphene. With the annealing temperature increasing, the Gd3N@C80 molecules

eventually aggregate together and break the ordered structure. Meanwhile, we compared DFT

calculations of the different orientations of a Gd3N@C80 molecule on graphene. We found the

energy favored orientation is Orientation 3 with 3.5 Ȧngstroms distance between molecule and

graphene and a binding energy of 0.95 eV. The energy difference between all orientations is

smaller than 0.09 eV. Furthermore, we calculated the binding energy between two Gd3N@C80

molecules with two orientations. The lowest energy (Orientation II) is 0.32 eV with a distance of

about 3.5 Ȧngstroms between molecules. The energy difference between two orientations is as

small as 0.02 eV. Our DFT calculation is highly agreed with our STM measurement.

References:

1. Kroto, H. W.; Heath, J. R.; Obrien, S. C.; Curl, R. F.; Smalley, R. E., C-60 - Buckminsterfullerene. Nature 1985, 318 (6042), 162-163. 2. Popov, A. A.; Yang, S. F.; Dunsch, L., Endohedral Fullerenes. Chem Rev 2013, 113 (8), 5989-6113. 3. Laasonen, K.; Andreoni, W.; Parrinello, M., Structural and Electronic-Properties of La-at-the-Cost-of-C82. Science 1992, 258 (5090), 1916-1918. 4. Zhang, J. Y.; Stevenson, S.; Dorn, H. C., Trimetallic Nitride Template Endohedral Metallofullerenes: Discovery, Structural Characterization, Reactivity, and Applications. Accounts Chem Res 2013, 46 (7), 1548-1557.

150

5. Zhang, J. Y.; Ye, Y. Q.; Chen, Y.; Pregot, C.; Li, T. H.; Balasubramaniam, S.; Hobart, D. B.; Zhang, Y. F.; Wi, S.; Davis, R. M.; Madsen, L. A.; Morris, J. R.; LaConte, S. M.; Yee, G. T.; Dorn, H. C., Gd3N@C-84(OH)(x): A New Egg-Shaped Metallofullerene Magnetic Resonance Imaging Contrast Agent. J Am Chem Soc 2014, 136 (6), 2630-2636. 6. Li, T. H.; Murphy, S.; Kiselev, B.; Bakshi, K. S.; Zhang, J. Y.; Eltahir, A.; Zhang, Y. F.; Chen, Y.; Zhu, J.; Davis, R. M.; Madsen, L. A.; Morris, J. R.; Karolyi, D. R.; LaConte, S. M.; Sheng, Z.; Dorn, H. C., A New Interleukin-13 Amino-Coated Gadolinium Metallofullerene Nanoparticle for Targeted MRI Detection of Glioblastoma Tumor Cells. J Am Chem Soc 2015, 137 (24), 7881-7888. 7. Wang, K.; Liu, X. Y.; Huang, R.; Wu, C. C.; Yang, D.; Hu, X. W.; Jiang, X. F.; Duchamp, J. C.; Dorn, H.; Priya, S., Nonionic Sc3N@C-80 Dopant for Efficient and Stable Halide Perovskite Photovoltaics. Acs Energy Lett 2019, 4 (8), 1852-1861. 8. Stevenson, S.; Rice, G.; Glass, T.; Harich, K.; Cromer, F.; Jordan, M. R.; Craft, J.; Hadju, E.; Bible, R.; Olmstead, M. M.; Maitra, K.; Fisher, A. J.; Balch, A. L.; Dorn, H. C., Small-bandgap endohedral metallofullerenes in high yield and purity (vol 401, pg 55, 1999). Nature 1999, 402 (6764), 898-898. 9. Lu, J.; Sabirianov, R. F.; Mei, W. N.; Gao, Y.; Duan, C. G.; Zeng, X. C., Structural and magnetic properties of Gd3N@C-80. J Phys Chem B 2006, 110 (47), 23637-23640. 10. Popov, A. A.; Zhang, L.; Dunsch, L., A Pseudoatom in a Cage: Trimetallofullerene Y-3@C-80 Mimics Y3N@C-80 with Nitrogen Substituted by a Pseudoatom. Acs Nano 2010, 4 (2), 795-802. 11. Nafradi, B.; Antal, A.; Pasztor, A.; Forro, L.; Kiss, L. F.; Feher, T.; Kovats, E.; Pekker, S.; Janossy, A., Molecular and Spin Dynamics in the Paramagnetic Endohedral Fullerene Gd3N@C-80. J Phys Chem Lett 2012, 3 (22), 3291-3296. 12. Tsuchiya, T.; Sato, K.; Kurihara, H.; Wakahara, T.; Maeda, Y.; Akasaka, T.; Ohkubo, K.; Fukuzumi, S.; Kato, T.; Nagase, S., Spin-site exchange system constructed from endohedral metallofullerenes and organic donors. J Am Chem Soc 2006, 128 (45), 14418-14419. 13. Cai, T.; Slebodnick, C.; Xu, L.; Harich, K.; Glass, T. E.; Chancellor, C.; Fettinger, J. C.; Olmstead, M. M.; Balch, A. L.; Gibson, H. W.; Dorn, H. C., A pirouette on a metallofullerene sphere: Interconversion of isomers of N-tritylpyrrolidino I-h Sc3N@C-80. J Am Chem Soc 2006, 128 (19), 6486-6492. 14. Chen, C. H.; Zheng, H. S.; Mills, A.; Heflin, J. R.; Tao, C. G., Temperature Evolution of Quasi-one-dimensional C-60 Nanostructures on Rippled Graphene. Sci Rep-Uk 2015, 5. 15. Li, Y. L.; Liu, X. Y.; Chen, C. H.; Duchamp, J.; Huang, R.; Chung, T. F.; Young, M.; Chalal, T.; Chen, Y. P.; Heflin, J. R.; Dorn, H. C.; Tao, C. G., Differences in self-assembly of spherical C-60 and planar PTCDA on rippled graphene surfaces. Carbon 2019, 145, 549-555. 16. Shinohara, H.; Hayashi, N.; Sato, H.; Saito, Y.; Wang, X. D.; Hashizume, T.; Sakurai, T., Direct Stm Imaging of Spherical Endohedral Sc-2-at-C-84 Fullerenes. J Phys Chem-Us 1993, 97 (51), 13438-13440. 17. Reveles, J. U.; Karle, N. N.; Baruah, T.; Zope, R. R., Electronic and Structural Properties of C-60 and Sc3N@C-80 Supported on Graphene Nanoflakes. J Phys Chem C 2016, 120 (45), 26083-26092. 18. Tian, X. Q.; Xu, J. B.; Wang, X. M., Self-Assembly of PTCDA Ultrathin Films on Graphene: Structural Phase Transition and Charge Transfer Saturation. J Phys Chem C 2010, 114 (49), 20917-20924. 19. Li, G.; Zhou, H. T.; Pan, L. D.; Zhang, Y.; Mao, J. H.; Zou, Q.; Guo, H. M.; Wang, Y. L.; Du, S. X.; Gao, H. J., Self-assembly of C-60 monolayer on epitaxially grown, nanostructured graphene on Ru(0001) surface. Appl Phys Lett 2012, 100 (1). 20. Ren, J.; Meng, S.; Wang, Y. L.; Ma, X. C.; Xue, Q. K.; Kaxiras, E., Properties of copper (fluoro-)phthalocyanine layers deposited on epitaxial graphene. J Chem Phys 2011, 134 (19). 21. Wang, Q. H.; Hersam, M. C., Room-temperature molecular-resolution characterization of self-assembled organic monolayers on epitaxial graphene. Nat Chem 2009, 1 (3), 206-211. 22. Lu, J.; Yeo, P. S. E.; Zheng, Y.; Yang, Z. Y.; Bao, Q. L.; Gan, C. K.; Loh, K. P., Using the Graphene Moire Pattern for the Trapping of C-60 and Homoepitaxy of Graphene. Acs Nano 2012, 6 (1), 944-950.

151

23. Cho, J.; Smerdon, J.; Gao, L.; Suzer, O.; Guest, J. R.; Guisinger, N. P., Structural and Electronic Decoupling of C-60 from Epitaxial Graphene on SiC. Nano Lett 2012, 12 (6), 3018-3024. 24. Chen, L.; Carpenter, E. E.; Hellberg, C. S.; Dorn, H. C.; Shultz, M.; Wernsdorfer, W.; Chiorescu, I., Spin transition in Gd3N@C-80, detected by low-temperature on-chip SQUID technique. J Appl Phys 2011, 109 (7). 25. Sinha, S.; Sheng, Y. W.; Griffiths, I.; Young, N. P.; Zhou, S.; Kirkland, A. I.; Porfyrakis, K.; Warner, J. H., In Situ Atomic-Level Studies of Gd Atom Release and Migration on Graphene from a Metallofullerene Precursor. Acs Nano 2018, 12 (10), 10439-10451. 26. Hunter, C. A.; Sanders, J. K. M., The Nature of Pi-Pi Interactions. J Am Chem Soc 1990, 112 (14), 5525-5534. 27. Cabaleiro-Lago, E. M.; Rodriguez-Otero, J.; Carrazana-Garcia, J. A., A theoretical study of complexes between fullerenes and concave receptors with interest in photovoltaics. Phys Chem Chem Phys 2017, 19 (39), 26787-26798. 28. Gonzalez-Veloso, I.; Rodriguez-Otero, J.; Cabaleiro-Lago, E. M., Carbon-nanorings ([10]CPP and [6]CPPA) as fullerene (C-60 and C-70) receptors: a comprehensive dispersion-corrected DFT study. Phys Chem Chem Phys 2016, 18 (46), 31670-31679.

152

Chapter 7

Conclusion and Future Work

The experiments discussed in this dissertation have focused on the molecular self-assembly

behavior of PCBM, PTCDA, C60 and Gd3N@C80 molecules on planar graphene and rippled

graphene by experimental STM and AFM methods and theoretical DFT calculations. The main

motivation in this dissertation is to investigate the mechanism behind molecular self-assembly

behavior to aid in the development of the future graphene applications. Here, we present a

summary of the conclusions and the potential future work for this area of research.

7.1 The Bilayer PCBM Structure Formed on Graphene and HOPG

In the first part of our investigation, we demonstrated the self-assembly behavior of PCBM

bilayer nanostructures on HOPG and graphene, by using STM and AFM, and analyzed the

observed structure by comparison to molecular models. Through careful control of the PCBM

solution concentration (from 0.1 mg/ml to 2 mg/ml) and the deposition conditions, we

demonstrate that PCBM molecules self-assemble into bilayer structures with increasing

coverage on graphene and HOPG substrates. Interestingly, the PCBM bilayer revealed two

distinct configurations on HOPG with two different heights (1.64 nm and 1.23 nm respectively),

and only one configuration (1.37 nm) on graphene. We also found two monolayer

configurations on HOPG with two different heights (0.71 nm and 0.88 nm respectively) and one

153

configuration of 0.87 nm on graphene. We believe the two configurations of a bilayer are due

to the different tiled angle by contracting the model of molecular dimer. While, for the

monolayer, the two configurations are more due to the standing up or laying down position of

PCBM molecules. Post thermal annealing could induce the merging of bilayer nanostructures

due to the edge diffusion. This is, to the best of our knowledge, the first experimental

realization of PCBM bilayer structures on graphene. Our results will shed light on improvement

of the energy efficiency in solar cells containing graphene and organic molecules, by increasing

the donor-acceptor interface area.

7.2 The Ordered of C60 and Disordered Structure of PTCDA Formed on Rippled Graphene

In the second part, we have found the self-assembly of PTCDA and C60 molecules on rippled

graphene with by using both STM and DFT calculations. The adsorbed molecules on 1D-rippled

graphene systems show distortions when compared with analogous planar graphene system.

Specifically, the nearly spherical C60 molecules form a quasi-hexagonal close packed structure,

while the planar PTCDA molecules form a disordered herringbone structure on the rippled

graphene surface. The change in the packing pattern of C60 and PTCDA molecules on a curved

graphene surface is due to competition between the intermolecular adsorbate interactions and

the adsorbate-graphene interaction. Because of the nearly spherical C60 molecules, the rippled

graphene surface exhibits only depreciated effects on the π-π intermolecular interactions. In

opposition, the planar PTCDA molecules receive the effect that is more significant by the

rippled graphene. In addition, the DFT computational results demonstrate significant increases

154

in π-π interactions for both the adsorbed PTCDA and C60/rippled graphene complexes located

in the 2D graphene valley sites compared with adsorbed more idealized molecule/planar

graphene 2D complexes. Furthermore, we find that the adsorbed planar PTCDA molecules

prefer different orientations when on the rippled graphene peak regions compared to the

valley regions. These fundamental experimental and computational results are important for

understanding any potential application of structurally diverse molecules adsorbed on

graphene and/or rippled graphene surfaces.

7.3 The hcp Structure of Gd3N@C80 Formed on Graphene

In last part, we reported hcp structure of Gd3N@C80 molecules on flat graphene with

characterization using both experimental STM and DFT calculations. Comparable to the perfect

structure of C60 in long range, Gd3N@C80 molecules forms hcp structure in small range with

many defects. The lattice constant that we measured is 1.15 nm and the monolayer step height

is about 0.87 nm. Besides, we did the temperature evolution of Gd3N@C80 monolayer on

graphene. With the annealing temperature increasing, the Gd3N@C80 molecules would

aggregate together breaking the ordered structure. Meanwhile, we did the DFT calculation

about the different orientations of Gd3N@C80 molecule on graphene. We found the energy

favorite orientation is orientation 6 with 3.5 angstroms distance between molecule and

graphene and binding energy of 0.95 eV. The energy difference between all orientations is

smaller than 0.09 eV, while the energy differences between two nearest orientations is 0.02 eV

which is compared to room temperature thermal energy (0.025 eV). That means the

155

orientation of Gd3N@C80 molecule could change easily in room temperature. What’s more, we

calculated the binding energy between two Gd3N@C80 molecules with two orientations. The

lowest energy (orientation 2) is 0.32 eV with the distance about 3.5 angstroms between

molecules. The energy difference between two orientations is as small as 0.2 eV. Our DFT

calculation is highly agreed with our STM measurement. This is, to the best of our knowledge,

the first experimental realization of Gd3N@C80 molecules monolayer structures on graphene.

These fundamental experimental and computational results are important for understanding

any potential application of Gd3N@C80 molecules adsorbed on graphene.

7.4 Future Work

The immediate next step in this project is the investigation of molecular self-assembly behavior

of the mixture of organic semiconductor donor and acceptor molecules (such as C60 and

DTS(PTTh2)2 and other combinations) on graphene substrate. This is of interest as donor

acceptor combinations are essential for efficient organic photovoltaics (OPVs). We are hoping

to see the phase separation of donor and acceptor molecules on graphene substrate. OPVs

require proximity of the donors and acceptors for charge transfer and bicontinuous networks to

enable charge transfer to the substrate. Besides, we could better understand the interaction

and mechanism between donor and acceptor molecules by combining scanning tunneling

microscope (STM) and scanning tunneling spectroscopy (STS) to study the molecules at both

sides of the interfere between donor and acceptor domains, especially at low temperature. This

156

could help us better understand the mechanism behind organic solar cell for the goal of

increasing the efficiency.

In addition, it would be useable to examine other members of the fullerene familiy including

metallofullerenes. For example, we could study magnetic phenomena and charge transfer

effect of Gd3N@C80 molecule by using a magnetic tip. We could also study the electronic

properties and morphology of larger fullerenes, like C90, C100 and C120. C120 is the largest

fullerene molecule and was an interesting structure. A C120 molecule is an armchair carbon

nanotube with two C60 hemisphere at the two ends. The electronic properties of C120 are very

interesting as the armchair carbon nanotube is metallic while the C60 is semiconductor.

The most important future direction of the project is the molecular self-assembly on graphene

established by other forces. For now, the molecular self-assembly behavior on graphene is all

based on hydrogen bonding. For example, we could study the molecular self-assembly behavior

on graphene induced by molecules and metal atoms. Besides, the halogen-halogen molecule

self-assembly behavior on graphene is also a good topic for investigation by depositing

hexakis(4-iodophenyl)benzene (HPBI) molecules on graphene.

157

List of Publications

The work of this thesis is partial based on the following publications:

1. Li, Y.; Chen, C.; Burton, J.; Park, K.; Heflin, J.; Tao, C., Self-Assembled PCBM Bilayers on

Graphene and HOPG Examined by AFM and STM, Nanotechnology 29, 185703 (2018)

2. Li, Y.; Liu, X.; Chen, C.; Duchamp, J.; Huang, R.; Chung, T.; Young, M.; Chalal, T.; Chen, Y.P.;

Heflin, J.R.; Dorn, H.; Tao, C., Differences in Self-Assembly of Spherical C60 and Planar PTCDA on

Rippled Graphene Surfaces, Carbon 145, 549 (2019)