+ All Categories
Home > Documents > Electron Spin Resonance Study on the Magnetic Properties ...

Electron Spin Resonance Study on the Magnetic Properties ...

Date post: 06-Apr-2022
Category:
Upload: others
View: 4 times
Download: 0 times
Share this document with a friend
209
Electron Spin Resonance Study on the Magnetic Properties of Graphene and Its Derivative A Thesis Submitted to the University of Manchester for the Degree of Doctor of Philosophy in the Faculty of Science and Engineering 2019 Oka Pradipta Arjasa Putra Department of Chemistry University of Manchester
Transcript

Electron Spin Resonance Study

on the Magnetic Properties of Graphene

and Its Derivative

A Thesis Submitted to the University of Manchester for the Degree of

Doctor of Philosophy in the Faculty of Science and Engineering

2019

Oka Pradipta Arjasa Putra

Department of Chemistry

University of Manchester

P a g e | 1

Contents

Contents ................................................................................................................ 1

List of Figures and Tables ..................................................................................... 7

Figures ............................................................................................................... 7

Tables .............................................................................................................. 17

List of Abbreviations ........................................................................................... 18

Abstract ............................................................................................................... 22

Declaration .......................................................................................................... 24

Copyright Statement ........................................................................................... 25

Acknowledgements ............................................................................................. 26

1. CHAPTER ONE ......................................................................................... 27

Introduction ..................................................................................................... 27

1.0 Graphene ............................................................................................. 27

1.1 Properties of graphene......................................................................... 28

1.2 Applications of graphene .................................................................... 30

1.3 Production of graphene ....................................................................... 32

1.3.1 Anodic bonding ............................................................................... 33

1.3.2 Photo exfoliation ............................................................................. 34

1.3.3 Liquid phase exfoliation .................................................................. 34

1.3.4 Electrochemical exfoliation ............................................................ 36

1.3.5 Reduced graphene oxide ................................................................. 39

1.3.6 Thermal decomposition of SiC ....................................................... 41

P a g e | 2

1.3.7 Growth of graphene on metallic surfaces by precipitation ............. 42

1.3.8 Chemical vapour deposition ............................................................ 42

1.3.9 Molecular beam epitaxy .................................................................. 43

Analytical techniques to study graphene ........................................................ 44

1.4 Electron paramagnetic resonance spectroscopy .................................. 44

1.4.1 Electron paramagnetic resonance basic principle ........................... 44

1.4.2 State-of-the-art of EPR in graphene ................................................ 59

1.5 Raman spectroscopy ........................................................................... 67

1.5.1 Raman basic principles ................................................................... 67

1.5.2 Raman spectrum of graphene .......................................................... 69

1.6 Aims and objectives ............................................................................ 74

2. CHAPTER TWO ........................................................................................ 75

Electron Paramagnetic Resonance Study of Graphene Laminates ................. 75

2.0 Introduction ......................................................................................... 75

2.1 Sample Preparation ............................................................................. 78

2.1.1 Liquid phase exfoliation graphene laminate ................................... 78

2.1.2 Graphite ........................................................................................... 78

2.1.3 Electron paramagnetic resonance .................................................... 79

2.1.4 UV-Vis spectroscopy ....................................................................... 80

2.1.5 Atomic Force Microscope (AFM) .................................................. 80

2.1.6 Raman spectroscopy ....................................................................... 81

2.2 Results and Discussion ........................................................................ 81

P a g e | 3

2.2.1 Graphene flake characterization ...................................................... 81

2.2.2 Graphene laminate paramagnetism ................................................. 82

2.2.3 Temperature dependence of graphene laminates ............................ 85

2.2.4 EPR linewidth of graphene laminates ............................................. 87

2.2.5 Comparison to graphite ................................................................... 89

2.2.6 EPR magnetic susceptibility of graphene laminates ....................... 92

2.2.7 Relaxation times and nuclear resonances of the graphene laminates

……………………………………………………………………. 96

2.3 Conclusion .......................................................................................... 97

3. CHAPTER THREE ..................................................................................... 99

Electron Paramagnetic Resonance Study of the Electrochemical Exfoliation of

Graphite in Comparison to Graphene Laminates Produced Through

Electrochemical Exfoliation, Liquid Phase Exfoliation and Chemical

Reduction of Graphene Oxide ........................................................................ 99

3.0 Introduction ......................................................................................... 99

3.1 Sample Preparation ........................................................................... 101

3.1.1 Liquid Phase Exfoliation ............................................................... 101

3.1.2 Electrochemical Exfoliation .......................................................... 101

3.1.3 Reduced Graphene Oxide ............................................................. 102

3.1.4 Electron Paramagnetic Resonance (EPR) Spectroscopy............... 102

3.1.5 Atomic Force Microscope (AFM) ................................................ 103

3.1.6 Raman Spectroscopy ..................................................................... 103

3.2 Results and Discussion ...................................................................... 104

P a g e | 4

3.2.1 Observation of Electrochemical Exfoliated Graphite by Electron

Paramagnetic Resonance and Raman Spectroscopy .................... 104

3.2.2 Graphene flakes characterization .................................................. 109

3.2.3 Defect-induced paramagnetism ..................................................... 110

3.2.4 Temperature dependence ............................................................... 114

3.3 Conclusion ........................................................................................ 119

4. CHAPTER FOUR ..................................................................................... 121

Paramagnetic Stability and Defect Creation of Graphene Laminates Under

Controlled Conditions and Action of Laser .................................................. 121

4.0 Introduction ....................................................................................... 121

4.1 Sample Preparation ........................................................................... 123

4.1.1 Liquid phase exfoliation graphene laminate ................................. 123

4.1.2 Paramagnetic stability experiment using EPR .............................. 123

4.1.3 The aged graphene laminate experiment....................................... 123

4.1.4 In-situ defect creation experiment studied using EPR .................. 125

4.1.5 Raman Spectroscopy ..................................................................... 126

4.1.6 X-ray Photoelectron Spectroscopy (XPS) ..................................... 126

4.2 Results and Discussion ...................................................................... 126

4.2.1 Paramagnetic Stability .................................................................. 126

4.2.2 Defect creation on an aged graphene laminate sample ................. 133

4.2.3 In-situ defect creation by irradiation at 270 nm, 660 nm and 800 nm

…………………………………………………………………... 136

4.3 Conclusion ........................................................................................ 141

P a g e | 5

5. CHAPTER FIVE ....................................................................................... 142

Electron Paramagnetic Resonance Study of Fluorinated Graphene Laminate

...................................................................................................................... 142

5.0 Introduction ....................................................................................... 142

5.1 Sample Preparation ........................................................................... 144

5.1.1 Liquid phase exfoliation fluorinated graphene laminate ............... 144

5.1.2 Electron Paramagnetic Resonance spectroscopy .......................... 144

5.1.3 Fourier-Transform Infrared spectroscopy ..................................... 145

5.1.4 Raman spectroscopy ..................................................................... 145

5.2 Results and Discussion ...................................................................... 146

5.2.1 Fluorinated graphene laminate ...................................................... 146

5.2.2 Paramagnetism of fluorinated graphene laminate ......................... 150

5.2.3 Temperature-dependence of the EPR resonance ........................... 151

5.2.4 HYSCORE spectroscopy .............................................................. 156

5.3 Conclusion ........................................................................................ 160

6. CHAPTER SIX ......................................................................................... 162

Conclusions and Future Work ....................................................................... 162

6.0 Conclusions ....................................................................................... 162

6.1 Future work ....................................................................................... 165

References ......................................................................................................... 167

Appendix A ....................................................................................................... 199

Appendix B ....................................................................................................... 206

P a g e | 6

Appendix C ....................................................................................................... 207

C.1 CW EPR Code Simulation ................................................................ 207

C.2 HYSCORE Code Simulation ............................................................ 208

Word count: 43289

P a g e | 7

List of Figures and Tables

Figures

Figure 1.1 Schematic illustration showing that graphene can be rolled or stacked to form

different carbon-based nanomaterials. Taken from [5]. .......................................... 27

Figure 1.2 Schematic showing the two types of edges in graphene, the armchair edges

and the zigzag edges. Taken from [12]. .................................................................. 28

Figure 1.3 (A) Photograph of a 50 μm aperture partially covered by mono and bilayer

graphene. The line scan profile shows the intensity of the transmitted white light

along the yellow line. The inset shows a metal support structure with different sizes

of aperture. (B) Transmittance spectrum of single-layer graphene (open circles). The

red and green line is the theoretical transmittance expected for ideal Dirac electrons

and graphene, respectively. The inset shows the transmittance of white light as a

function of the number of graphene layers. Taken from [13]. ................................ 29

Figure 1.4 Methods for producing graphene. Each of them has its own advantages and

disadvantages related to graphene size, quality and application purposes. Taken from

[55]. ......................................................................................................................... 33

Figure 1.5 Schematic mechanism of anodic electrochemical exfoliation taken from [76].

................................................................................................................................. 37

Figure 1.6 High-resolution transmission electron microscopy (HRTEM) image of single-

layer rGO. Colour scheme highlighted different features. Light grey colour

represents the defect-free areas. Dark grey colour represents contaminated regions.

Blue colour represents disordered single-layer carbon network or extended

topological defects identified as remnants of the oxidation-reduction process. Red

P a g e | 8

colour represents individual adatoms or substitutions. Green colour represents

isolated topological defects. Yellow colour represents holes and their edge

reconstructions. The scale bar is 1 nm. The image is taken from [96]. ................... 41

Figure 1.7 Schematic diagram of the wet-transfer process taken from [103]. ............... 43

Figure 1.8 Energy levels of an unpaired electron spin in the applied magnetic field.

Resonant energy absorption (Equation 1.3) leads to an electron spin ‘flip’ or

transition resulting in an EPR signal. The signal can be presented in absorption

(dotted) or first derivative (solid) mode. Taken from [112]. ................................... 46

Figure 1.9 Typical anisotropic axial spectra for 𝑔𝑧 > 𝑔𝑥 = 𝑔𝑦: 1st derivative line (red)

and absorption line (blue). The Figure was made using a simulator provided in

www.eprsimulator.org [114].................................................................................... 49

Figure 1.10 Typical rhombic symmetry spectra: 1st derivative line (red) and absorption

line (blue). The Figure was made using a simulator provided in

www.eprsimulator.org [114].................................................................................... 50

Figure 1.11 Energy level diagram in a fixed magnetic field for a system with S = 1 2⁄ and

I = 1 2⁄ , in the highfield approximation, showing the electron eeeman (Ee) and

nuclear eeeman (Ne) levels, and the perturbation arising from the hyperfine

interaction (HF). The two allowed EPR transitions (solid arrows) result in the

experimentally observed resonances labelled EPR I and EPR II (shown in the inset).

Adapted from [112]. ................................................................................................ 51

Figure 1.12 The EPR spectrum of a system with S =1 2⁄ and I = 1 2⁄ . The Figure was

made using a simulator provided in www.eprsimulator.org [114]. ......................... 52

Figure 1.13 The temperature dependence of the reciprocal magnetic susceptibility. a)

Curie law behaviour of a paramagnet; b) Curie-Weiss law behaviour of a

P a g e | 9

ferromagnet; c) Curie-Weiss law behaviour of an antiferromagnet; d) behaviour of a

ferrimagnet. Figure a-b is taken from [116], Figure c-d is taken from [115]. ......... 54

Figure 1.14 The scheme of a CW EPR spectrometer employing magnetic field

modulation. The Figure is taken from [117]. .......................................................... 56

Figure 1.15 Illustration of the magnetization vector at characteristic positions in the

typical 2-pulse sequence. Adapted from [118]. ....................................................... 57

Figure 1.16 a) 2D HYSCORE spectrum where full squares ■ represent cross-peaks from

weakly coupled nuclei in the (+,+) quadrant, and full circles ● represent cross-peaks

from strongly coupled nuclei in the (-,+) quadrant. 𝑣𝐿 is the Larmor frequency for

the nucleus of interest, A is the hyperfine coupling, 𝑣𝛼(= 𝜔12) and 𝑣𝛽(= 𝜔34); b)

(+,+) quadrant for the powder HYSCORE pattern for an S = I = 1 2⁄ spin system

with an axial hyperfine tensor. The Figure is taken from [112]. ............................. 59

Figure 1.17 Temperature dependence of the EPR linewidth for mechanically exfoliated

graphene (a) [14] and LPE graphene (b) [15]. ........................................................ 62

Figure 1.18 a) Temperature dependence of the electron spin resonance (ESR) signal from

LPE graphene. b) Temperature dependence of normalized ESR susceptibility

measured after the annealing treatment showing a weaker signal which assigned to

the conducting electrons. The solid line corresponds to the Curie law. The spectrum

in the inset was recorded at 100 K with 64 accumulations. The Figure is taken from

[37]. ......................................................................................................................... 63

Figure 1.19 Temperature dependence of the linewidth from multilayer graphene. The

inset shows the temperature independence of the g-value. The Figure is taken from

[20]. ......................................................................................................................... 64

P a g e | 10

Figure 1.20 EPR spectra of (a) SGN18 graphite powder, (b) ultrasounded, (c) shear mixed,

and (d) stirred few-layer graphene. The inset shows the uniaxial g-value simulated

EPR lineshape for the stirrer prepared sample. The Figure is taken from [127]..... 65

Figure 1.21 Schematic of the Rayleigh and Raman processes. The lowest energy

vibrational state m is shown at the foot with a state one vibrational unit in energy

above it labelled n. Rayleigh scattering also occurs from higher vibrational levels

such as n. Taken from [140] .................................................................................... 68

Figure 1.22 a) Mechanically exfoliated graphene showing both monolayer and bi-layer

regions. b) Raman spectra of mono and bi-layer graphene. The top and bottom insets

represent the enlarged 2D bands of regions B and A, respectively. The Figure is

taken from [143]. ..................................................................................................... 69

Figure 1.23 Raman spectrum of defective graphene showing the main Raman features

taken with a laser excitation energy of 2.41 eV. The Figure is taken from [148]. .. 71

Figure 1.24 Typical Raman spectra of liquid-phase exfoliated graphene recorded with

514 nm laser. The Figure is taken from [62]. .......................................................... 72

Figure 2.1 Graphene flake shapes (a) and size distribution (b), analyzed by using AFM.

................................................................................................................................. 82

Figure 2.2 9.4 GHz EPR spectra of the thick graphene laminate (1.132 mg/cm2) recorded

at 295 K (a and b) and at 10 K (c and d) at two different orientations; a and c

represents 𝐻∥ ; b and d represents 𝐻⊥ . The purple line represents the overall

simulation result of the Lorentzian lineshape; the red and blue line represents

Lorentzian lineshape of narrow and broad component, respectively. The simulation

was performed by using Easyspin [176]. ................................................................ 83

P a g e | 11

Figure 2.3 The room temperature EPR linewidth of graphene laminate’s broad

component at 𝐻∥ orientation on the variation of layer thickness. ........................... 85

Figure 2.4 The EPR spectra of thin graphene laminates (0.566 mg/cm2) as a function of

temperature. The black line represents 𝐻⊥. The red line represents 𝐻∥. ................. 86

Figure 2.5 EPR linewidth of the narrow component for the thin graphene laminates (0.113

mg/cm2) on the variation of temperature. a) Red square represents 𝐻∥ . b) Black

square represents 𝐻⊥. .............................................................................................. 88

Figure 2.6 EPR linewidth of the narrow component for the thick graphene laminates

(1.132 mg/cm2) on the variation of temperature. a) Red square represents 𝐻∥ . b)

Black square represents 𝐻⊥. .................................................................................... 89

Figure 2.7 EPR spectra of a graphite flake as a function of temperature. (a) represents

𝐻⊥. (b) represents 𝐻∥. (*) marks a speculate asignment of the broad component at 70

K. ............................................................................................................................. 90

Figure 2.8 Curie-Weis behaviour of thin graphene laminates (0.113 mg/cm2) measured

from 10-75 K. Black dots represents 𝐻⊥, θ = -0.2 ± 2.5 K. Red dots represents 𝐻∥, θ

= 5.2 ± 2.5 K. Blue line represents the Curie-Weis line. ......................................... 94

Figure 2.9 Curie-Weis behaviour of thick graphene laminates (0.566 mg/cm2) measured

from 10-75 K. Black dots represents 𝐻⊥, θ = -5.4 ± 2.6 K. Red dots represents 𝐻∥, θ

= 5.9 ± 2.1 K. Blue line represents the Curie-Weis line. ......................................... 94

Figure 2.10 Curie-Weis behaviour of thick graphene laminates (1.132 mg/cm2) measured

from 10-75 K. Black dots represents 𝐻⊥, θ = -11.5 ± 2.1 K. Red dots represents 𝐻∥,

θ = 9.2 ± 1.9 K. Blue line represents the Curie-Weis line. ...................................... 95

P a g e | 12

Figure 2.11 The spin-spin relaxation time (T2) (a) and spin-lattice relaxation time (T1) (b)

of a graphene laminate over the temperature range of 10 – 70 K at the 𝐻⊥ orientation.

................................................................................................................................. 96

Figure 2.12 ESEEM spectrum of graphene laminate at 10 K at the 𝐻⊥ orientation. .... 97

Figure 3.1 The anode (a) and the cathode (b) after 30 seconds of the electrochemical

exfoliation process. ............................................................................................... 104

Figure 3.2 EPR spectra of the anode graphite foil before and after 30 seconds of the

electrochemical exfoliation process. The solid and dash lines represent the EPR

spectra before and after electrochemical exfoliation, respectively. The black and red

colours represent the EPR spectra at 𝐻⊥ and 𝐻∥ orientations, respectively. ......... 105

Figure 3.3 The EPR spectrum of the anode graphite foil after 30 seconds of

electrochemical exfoliation process at the 𝐻⊥ orientation (solid black line). The

green dash line represents the Lorentzian line of the broad component; the blue dash

line represents the Lorentzian line of the narrow component. The solid purple line

represents the overall simulation result. The simulation was performed by using

Easyspin [176]. ...................................................................................................... 106

Figure 3.4 Raman spectra of anode graphite foil before (black) and after (red) 30 seconds

of electrochemical exfoliation. .............................................................................. 107

Figure 3.5 EPR spectra of cathode graphite foil before and after 30 seconds of the

electrochemical exfoliation process. The solid and dash lines represent the EPR

spectra before and after electrochemical exfoliation respectively. The black and red

colours represent the EPR spectra at 𝐻⊥ and 𝐻∥ orientations, respectively. ......... 108

Figure 3.6 Raman spectra of cathode graphite foil before (black) and after 30 seconds

(red) of electrochemical exfoliation. ..................................................................... 109

P a g e | 13

Figure 3.7 The AFM images of (a) EC graphene, (b) rGO and (c) LPE graphene (taken

from Chapter 2). .................................................................................................... 110

Figure 3.8 EPR spectra at room temperature of the LPE graphene laminate (a), EC

graphene laminate (b) and rGO laminate (c). The black and red lines represent 𝐻⊥

orientation and 𝐻∥ orientation, respectively. The samples were 1.132 mg/cm2 of

graphene laminates. ............................................................................................... 111

Figure 3.9 Raman spectrum of rGO laminate (red), EC graphene laminate (blue) and LPE

graphene laminate (black). .................................................................................... 112

Figure 3.10. EPR linewidth of the EC graphene laminate on the variation of temperature.

Black dot represents 𝐻⊥. Red dot represents 𝐻∥. .................................................. 116

Figure 3.11. EPR linewidth of the rGO laminate on the variation of temperature. Black

dot represents 𝐻⊥. Red dot represents 𝐻∥. ............................................................. 116

Figure 3.12. Curie-Weis behaviour of EC graphene laminate measured from 10-75 K.

Black dot represents data at 𝐻⊥. Red dot represents data at 𝐻∥. Black line represents

the Curie-Weis line fit for 𝐻⊥, θ = -10.6 ± 2 K. Red line represents the Curie-Weis

line fit for 𝐻∥, θ = -20 ± 1.1 K. .............................................................................. 118

Figure 3.13. Curie-Weis behaviour of rGO laminate measured from 10-75 K. Black dot

represents data at 𝐻⊥. Red dot represents data at 𝐻∥. Black line represents the Curie-

Weis line fit for 𝐻⊥, θ = 7.3 ± 3.4 K. Red line represents the Curie-Weis line fit for

𝐻∥, θ = 7.4 ± 3.6 K. ............................................................................................... 119

Figure 4.1 (a) The experiment setup for the aged graphene laminate irradiation. (b) The

sample tube after 180 seconds of irradiation. (*) Marks the sample tube............. 124

P a g e | 14

Figure 4.2 EPR spectrum at room temperature of an aged graphene laminate. The black

and red signals represent the 𝐻⊥ and 𝐻∥ orientations, respectively. The blue line

represents the EPR background from the EPR tube and cavity. ........................... 128

Figure 4.3 EPR lineshape evolution at room temperature and 𝐻∥ the orientation of

graphene laminate samples stored throughout 60 days. Samples stored under normal

atmospheric conditions (a and b); samples stored under argon (c and d). Time zero

spectra are shown in black. Spectra recorded at increasing duration are lighter in

colour (black to red to yellow). The a1, b1, c1 and d1 represent the lineshape at time

zero (black) and lineshape at 60th day (bright yellow). ......................................... 129

Figure 4.4 The evolution of mean total spin concentration throughout storage time. a)

normal spin concentration vs time. b) 1/log (spin concentration) vs log time. The

blue dot represents samples stored in normal atmospheric conditions; the red dot

represents samples stored in argon. ....................................................................... 132

Figure 4.5 EPR spectra evolution of an aged graphene laminate after irradiation at 270

nm at room temperature. (a) The sample is positioned (𝐻⊥ ). (b) The sample is

positioned (𝐻∥). ..................................................................................................... 133

Figure 4.6 Raman spectrum of graphene laminate before and after ultraviolet irradiation.

............................................................................................................................... 135

Figure 4.7 Continuous-wave EPR spectrum of graphene laminate at 100 K and 𝐻∥

orientation. (a-b) 270 nm laser wavelength irradiation, (c-d) 660 nm laser

wavelength irradiation and (e-f) 800 nm laser wavelength irradiation. The black line

represents the graphene laminate before irradiation while the blue, green and red

signal represents the graphene laminate after 30 minutes of the irradiation. ........ 137

P a g e | 15

Figure 4.8 Double integration area of the EPR line at 100 K and 𝐻∥ orientation. Positive

area means that the sample gains more electron spins after the irradiation; negative

area means that the sample loses electron spins after irradiation. ......................... 139

Figure 4.9 Raman spectrum of graphene laminate before (black line) and after 30 minutes

of irradiation using a 270 nm laser (blue line), 660 nm laser (green line) and 800 nm

laser wavelengths (red line). ................................................................................. 140

Figure 5.1 FGn dispersion prepared in (a) NMP and (b) isopropanol : water (1:1). .... 146

Figure 5.2 a) FTIR spectra of FGn (orange) and FG (black). b) Four components which

correspond to stretching vibrations of C-F bonds with different local surroundings.

CF3, CF2 and CF1 annotation assigns the bonds, which have three, two and one C-

F neighbours. CFedge annotation assigns the bonds located at the graphene edges

which may be attributed to C-F2 and C-F bonds. .................................................. 148

Figure 5.3 a) Raman spectra of FGn with 325 nm laser. b) Raman spectra of LPE

graphene (black) and FGn (blue) with 514.5 nm laser. ......................................... 149

Figure 5.4 The EPR lineshape of FGn laminate at 𝐻⊥ (solid black) and 𝐻∥ (solid blue)

simulated using a single Lorentzian lineshape (dash purple and orange). The

simulation was performed using easypin [176]. ................................................... 151

Figure 5.5 The evolution of EPR linewidth on the variation of temperature. The black

rectangle represents the 𝐻⊥ orientation and the red triangle represents the 𝐻∥

orientation. ............................................................................................................ 152

Figure 5.6 The evolution of double integrated EPR intensity ( 𝜒𝐸𝑃𝑅 ) on a wide

temperature range (300 – 10 K). Black rectangle represents the 𝐻⊥ orientation and

red rectangle represents the 𝐻∥ orientation. ......................................................... 153

P a g e | 16

Figure 5.7 The Curie-Weis line fit for 𝐻⊥ orientation (solid black line) and 𝐻∥ orientation

(solid red line) at the temperature range of 100 – 10 K. The black rectangle and red

triangle represent 𝜒𝐸𝑃𝑅 at the 𝐻⊥ and 𝐻∥ orientations, respectively. .................. 154

Figure 5.8 The Curie-Weis line fit for 𝐻⊥ orientation (solid black line) and 𝐻∥ orientation

(solid red line) at the temperature range of 280 – 230 K. The black rectangle and red

triangle represent 𝜒𝐸𝑃𝑅 at the 𝐻⊥ and 𝐻∥ orientations, respectively. ................ 155

Figure 5.9 The HYSCORE 2D plot spectrum measured at 10 K and the 𝐻⊥ orientation

on FGn in frequency coordinates. a) τ = 160 ns. b) τ = 300 ns. ............................ 158

Figure 5.10 The HYSCORE 2D plot spectrum measured at 10 K and the 𝐻∥ orientation

on FGn in frequency coordinates. a) τ = 160 ns. b) τ = 300 ns. ............................ 159

Figure 5.11 The HYSCORE simulation 2D plot spectrum measuring 13C, 19F and 1H

resonances in frequency coordinates with A (13C) = 1 MHz, A (19F) = 2 MHz and A

(1H) = 0.6 MHz. a) τ = 160 ns. b) τ = 300 ns. ....................................................... 160

Figure 6.1 (A) Two layers of h-BN with B atoms are on top of the N atoms. (B) A unit

cell of the honeycomb structure of h-BN with Bravais lattice vectors. Taken from

[263]. ..................................................................................................................... 166

P a g e | 17

Tables

Table 1 CW EPR studies on graphene materials. ........................................................... 60

Table 4.1 Ultra-high vacuum XPS on graphite and graphene laminates (1.132 mg/cm2).

............................................................................................................................... 127

P a g e | 18

List of Abbreviations

⊥ Perpendicular

∥ Parallel

2D Two dimensional

A Ampere (electrical current unit)

𝐴 Hyperfine coupling constant

AFM Atomic force microscope

ATR Attenuated total reflectance

A/B the ratio describing the symmetry of an EPR line

Å Angstrom (1 Å = 10-10 m)

𝛼 Opacity

𝐵0 External magnetic field

𝑐 The speed of light (299792458 m/s)

C Curie constant

CESR Conduction electron spin resonance

cm Centimeter (1 cm = 10-2 m)

DPPH 2,2-diphenyl-1-picrylhydrazyl

CVD Chemical vapour deposition

CW Continuous wave

oC Celcius

DMA N, N-dimethylacetamide

DMEU 1,3-dimethyl-2-Imidazolidinone

Δ𝐸 Energy difference between two eeeman split states

𝑒 Electron charge

EC Electrochemical exfoliation

𝐸𝐹 Fermi level

EM Electromagnetic

EPR Electron paramagnetic resonance

ESEEM Electron spin echo envelope modulation

ESR Electron spin resonance

eV Electron volt (1 eV = 1.602 x 10-19 J)

FG Fluorinated graphite

FGn Fluorinated graphene

P a g e | 19

FID Free-induction decay

FLG Few layer graphene

FMR Ferromagnetic resonance

FTIR Fourier-transform infrared

g Gram (mass)

G Gauss (magnetic flux density/induction unit)

𝑔𝑒 g value of free electron (2.0023193043617)

GHz Gigahertz (1 GHz = 109 Hz)

GNRs Graphene nanoribbons

GO Graphene oxide

GPa Gigapascal (1 GPa = 109 Pa)

𝐻 External magnetic field

ℎ Planck’ constant (6.62607015 x 10-34 kg m2 / s)

HOPG Highly oriented pyrolitic graphite

HRTEM High resolution transmission electron microscopy

HYSCORE The hyperfine sublevel correlation

Hz Hertz (1 Hz = 1 cycle per second)

ℏ Reduced Planck’s constant (ℏ = ℎ 2𝜋⁄ )

I Nuclear spin

𝐼𝐷 The Raman intensity of D band

𝐼𝐷′ The Raman intensity of D’ band

𝐼𝐺 The Raman intensity of G band

𝐼𝑃𝑃 Peak-peak EPR intensity

ITO Indium tin oxide

J Joule (unit of energy) (1 J = 1 Nm = 1 kg m2 / s2)

K Kelvin

kg Kilogram (mass) (1 kg = 1000 g)

kHz KiloHertz (1 kHz = 1000 Hz)

kV Kilovolt (1 kV = 1000 Volt)

L Litre

LPE Liquid phase exfoliation

m Meter

mA Milliampere (1 mA = 10-3 A)

MBE Molecular beam epitaxy

mg Milligram (1 mg = 10-3 g)

P a g e | 20

MHz Megahertz (1 MHz = 106 Hz)

ml Millilitre (1 ml = 10-3 L)

mm Millimeter (1 mm = 10-3 m)

mN MilliNewton (1 mN = 10-3)

ms Millisecond (1 ms = 10-3 s)

𝑚𝑆 Spin quantum number

mT MilliTesla (1 mT = 10-3 T = 10 G)

mW MilliWatt (1 mW = 10-3 W)

𝜇 Electron magnetic moment

𝜇𝐵 Bohr magneton constant (9.274009994 x 10-24 J T-1)

µm Micrometre (1 µm = 10-6 m)

N Newton (1 N = 1 kg m/s2)

NIR Near infrared

NIT nitronyl nitroxide

nm Nanometre (1 nm = 10-9 m)

NMP 1-methyl-2-pyrrolidinone

ns Nanosecond (1 ns = 10-9 s)

OLEDs Organic light-emitting diodes

OPO Optical parametric oscillator

o-DCB Orthodichlorobenzene

Ω Ohm (electron resistance unit)

Pa Pascal (1 Pa = 1 N/m2)

PC Propylene carbonate

PDMS Polydimethylsiloxane

PMMA Polymethyl methacrylate

PVD Physical vapour deposition

rGO Reduced graphene oxide

𝑅2 Coefficient of determination

s Second (time)

S Siemens (electron conductance unit; 1 S = 1 Ω−1)

𝑆 Spin angular momentum

SDBS Sodium dodecylbenzenesulfonate

SLG Single layer graphene

SQUID Superconducting quantum interference device

S/N Signal to noise ratio

P a g e | 21

𝑡 Pulse delay

T Tesla (magnetic flux density/induction unit; 1 Tesla = 10000 G)

𝑇1 Spin-lattice relaxation time

𝑇2 Spin-spin relaxation time

𝑇𝐶 Curie temperature

𝑇𝑀 Phase memory time

𝑇𝑀(𝑁)

Nucleus phase memory time

𝑇𝑁 Néel temperature

TEM Transmission electron microscope

TPa Terapascal (1 TPa = 1012 Pa)

𝜏 Pulse delay at which the echo is detected

𝜃 Curie-Weiss constant

UHV Ultra high vacuum

UV Ultra violet

V Volt

Vis Visible

𝜈 Electromagnetic irradiation frequency

W Watt (unit of power) (1 W = 1 J/s)

Wh Watt-hour (1 Wh = 3600 Joule)

XPS X-ray photoelectron spectroscopy

𝜒 Magnetic susceptibility

P a g e | 22

Abstract

The magnetic properties of graphene are related to the presence of localized and

conduction electrons and their interplay. A variety of graphene-based materials have been

prepared and investigated using electron paramagnetic resonance (EPR) and Raman

spectroscopy in order to understand the relationship between defects and electron-

electron interaction. The graphene samples were prepared by using sonication-assisted

liquid-phase exfoliation (LPE), electrochemical exfoliation (EC), reduced graphene oxide

(rGO) and fluorinated graphene (FGn) produced from sonication-assisted LPE of

fluorinated graphite (FG). The graphene flakes produced were further characterised using

an atomic force microscope (AFM). The EPR samples analysed in the form of laminates

in order to strengthen the EPR signal.

Continuous-wave (CW) EPR experiments on the LPE graphene laminates

revealed multicomponent, anisotropic, spectra showing the presence of narrow and broad

components. A temperature-dependent study of the g value, line shape, signal intensity

and Curie-Weiss fit of the magnetic susceptibility found that the narrow component could

be attributed to localized electrons (vacancy defects) and the broad was attributed to the

interplay of electrons between graphene layers. Several different thicknesses of laminates

were prepared and further comparisons were made to graphite. It was found that an

increase of disorder could be associated with an increase in laminate thickness/graphene

stacking and further related to the interlayer electron-electron interaction of the defective

and disordered graphene.

The EPR and Raman spectroscopic analysis on the anode and cathode graphite

foils produced through electrochemical exfoliation showed the presence of defects and

expansion. The spectral analysis was consistent with the current mechanistic

understanding of electrochemically prepared graphene. The graphene laminates prepared

P a g e | 23

using electrochemical exfoliation and reduced graphene oxide showed similar spectral

characteristics and the contribution of localized and conduction electrons for each type of

graphene laminate were identified and characterized. There was evidence to suggest that

the coupled and decoupled states of localized and itinerant conduction electrons were also

influenced by defects and functionalization.

The paramagnetic stability and defects of graphene laminate samples induced by

ageing and action of a nanosecond pulsed laser irradiation were investigated. Ageing of

graphene laminates showed a reduction in the EPR intensity with time in both

atmospheric and argon atmospheres indicating passivation. Laser irradiation of the aged

sample caused an increase in the numbers of spins whereas a reduction was observed for

unaged samples. It was shown that the defects created by the laser could break the 𝑠𝑝2

carbon-carbon bonds and create new spin centres.

EPR spectroscopy of FGn revealed an isotropic line shape indicative of a

homogeneously broadened EPR resonance arising from electron-electron interactions.

The Curie-Weiss fit of the magnetic susceptibility behaviour showed two temperature

regions, which show the magnetic moments to couple both ferromagnetically and

antiferromagnetically. Hyperfine sublevel correlation (HYSCORE) spectroscopy was

able to measure the fluorine hyperfine interaction.

P a g e | 24

Declaration

I declare that no portion of this work referred to in this thesis has been submitted in

support of an application for another degree or qualification of this or any other university

of other institutes of learning.

Manchester.

Date: 25/09/2019

Signed:

Oka P. Arjasa

P a g e | 25

Copyright Statement

i. The author of this thesis (including any appendices and/or schedules to this thesis) owns

certain copyright or related rights in it (the “Copyright”) and s/he has given The

University of Manchester certain rights to use such Copyright, including for

administrative purposes.

ii. Copies of this thesis, either in full or in extracts and whether in hard or electronic copy,

may be made only in accordance with the Copyright, Designs and Patents Act 1988 (as

amended) and regulations issued under it or, where appropriate, in accordance with

licensing agreements which the University has from time to time. This page must form

part of any such copies made.

iii. The ownership of certain Copyright, patents, designs, trademarks and other intellectual

property (the “Intellectual Property”) and any reproductions of copyright works in

the thesis, for example, graphs and tables (“Reproductions”), which may be described

in this thesis, may not be owned by the author and may be owned by third parties. Such

Intellectual Property and Reproductions cannot and must not be made available for use

without the prior written permission of the owner(s) of the relevant Intellectual Property

and/or Reproductions.

iv. Further information on the conditions under which disclosure, publication and

commercialisation of this thesis, the Copyright and any Intellectual Property and/or

Reproductions described in it may take place is available in the University IP Policy

(see http://documents.manchester.ac.uk/DocuInfo.aspx?DocID=24420), in any relevant

Thesis restriction declarations deposited in the University Library, The University

Library’s regulations (see http://www.library.manchester.ac.uk/about/regulations) and

in the University’s policy on Presentation of Theses.

P a g e | 26

Acknowledgements

All the praises and thanks be to Allah (God) which made all of this possible. Also,

it would not have been possible to write this doctoral thesis without the help and support

of the kind people around me, to only some of whom it is possible to give particular

mention here.

This thesis would not have been possible without the help, support and patience

of my principal supervisor Dr Alistair J. Fielding, not to mention his advice and discussion

for which I’m extremely grateful. Also, the good advice and support of my second and

third supervisor, Prof. Cinzia Casiraghi and Prof. Eric Mcinnes, which have been

invaluable on both an academic and a personal level.

I would also like to thank all the members/staff of the EPR and graphene groups,

in particular to Bin Wang, Tong Jincheng, Dr Khaled Parvez, Dr Ana Maria Ariciu, Lidya

Nodaraki, Dr Yu Young Shin, Dr Daniele Rizzo, Adam Brookfield, Dr Guillem Brandariz

de Pedro, Lara Grangel Gutierrez, Marco earattini, Dr Daryl Mcmanus, Dr Floriana Tuna

and Prof. David Collison.

I would like to acknowledge the financial support of the Indonesian Ministry of

Research, Technology and Higher Education and as well as the Agency for the

Assessment and Application of Technology (BPPT), particularly in the award of a

postgraduate scholarship.

Lastly and most importantly, I would like to express my love and gratitude to my

mother Tri Wuryani and my father Dr Wayan Sabe Arjasa, may you both rest in peace. To

my wife Primalia Swariputri and my son eayn Lorentzian Al-Razi Arjasa, for giving me

happiness.

P a g e | 27

1. CHAPTER ONE

Introduction

1.0 Graphene

The carbon atom has two stable isotopes, 12C (98.9%, nuclear spin I = 0) and 13C

(1.1%, nuclear spin I = 1 2⁄ ) [1]. All carbon-based materials exist in different allotropes

such as graphite, diamond, carbon nanotube, graphene etc. Graphene is a single-atomic

layer made of carbon atoms with sp2 hybridization. Graphene is a two-dimensional

material, i.e. a material having a structure of a single layer of atoms. Graphene was

isolated for the first time in 2004 [2], previously it was believed to be too difficult to

isolate due to the thought that it was unstable at normal atmospheric conditions and the

lack of information on its properties [3, 4]. Graphene is considered as the basic structural

unit in sp2 carbon materials as shown in Figure 1.1 [5] i.e. graphene can be rolled or

stacked to form carbon nanotubes or graphite.

Figure 1.1 Schematic illustration showing that graphene can be rolled or stacked to form

different carbon-based nanomaterials. Taken from [5].

P a g e | 28

1.1 Properties of graphene

Graphene has a long list of outstanding properties. The most remarkable is not

only the electronic properties but also its mechanical and optical properties for several

applications. Graphene is the thinnest material (one atom thick) and has an estimated

Young’s modulus of 2.4 ± 0.4 TPa [6]. It has breaking strength of 42 N/m (on a defect-

free sheet) with a tensile strength of 130 GPa [7] and room temperature thermal

conductivity in the range (4.84 ± 0.44) x 103 to (5.3 ± 0.48) x 103 W/mK [8].

The electrons in the π-orbital of graphene behave like particles with no mass,

giving rise to extremely high charge mobility of 250.000 cm2 V-1 s-1 at room temperature,

making graphene the crystal with the highest charge mobility [9]. Graphene is a zero-gap

semiconductor where the conduction band and the valence band touch at the Dirac point

[10]. The armchair and zig-zag edges of graphene (Figure 1.2) may be related to its

magnetic properties [11].

Figure 1.2 Schematic showing the two types of edges in graphene, the armchair edges

and the zigzag edges. Taken from [12].

Graphene is also almost transparent in the visible range, Nair et al. showed that

single-layer graphene absorbs 2.3 % of the incident light, and the opacity is practically

P a g e | 29

independent of wavelength (Figure 1.3B). The absorption of light is proportional to the

number of layers. Figure 1.3A shows an image of an aperture that is partially covered by

suspended graphene to compare the opacities of different areas [13].

Figure 1.3 (A) Photograph of a 50 μm aperture partially covered by mono and bilayer

graphene. The line scan profile shows the intensity of the transmitted white light along

the yellow line. The inset shows a metal support structure with different sizes of aperture.

(B) Transmittance spectrum of single-layer graphene (open circles). The red and green

line is the theoretical transmittance expected for ideal Dirac electrons and graphene,

respectively. The inset shows the transmittance of white light as a function of the number

of graphene layers. Taken from [13].

Magnetic behaviour in graphene is associated with the interaction of the localized

and itinerant conduction electron spins [14-21]. Magnetism in graphene may arise from

impurities (i.e. adatoms) and active defects (i.e. in-plane vacancy defects / dangling bonds,

non-bonding edge defects) [22-24]. Active defects are often described as a preliminary

condition for the existence of magnetic order [25-27]. However, active defects may not

last long in normal conditions due to self-reconstruction and passivation by other

atoms/molecules [28-30]. Interestingly according to theoretical studies, edge states

P a g e | 30

consisting of nonbonding π-electrons located at the edge region [11] are often discussed

with regard to their contribution to the paramagnetic activities of graphene [31-33]. The

unpaired electron at the zig-zag edges (Figure 1.2) on graphene may spin-polarized and

could arrange parallel to each other [31-33]. The magnetic properties of different types of

graphene samples, with different properties, have been experimentally investigated [16, 19-

22, 28, 34-37]. A full report of the state-of-the-art developments is provided in Section 1.4.2.

1.2 Applications of graphene

Because of its outstanding mechanical, optical and electrical properties, graphene

can be used in several applications. In electronics, the high transmittance added with the

low sheet resistance of highly doped samples [38], allows graphene to be used as a

transparent conductive material in flexible electronics, such as in touch screen displays,

electronic paper and organic light-emitting diodes (OLEDs). Moreover, graphene has

better properties than indium tin oxide (ITO) due to its high mechanical flexibility and

chemical durability which are important characteristics for flexible electronic devices

[39]. The fracture strength of defect-free graphene is currently higher (the breaking

strength = 42 N m-1, Young’s modulus = 1 TPa) compared to many other conventional

materials [7] such as ITO (Young’s modulus = 0.116 TPa) [40], which make graphene

suitable for bendable and rollable devices [38].

In the case of transistors, graphene is gapless, so it cannot be used for digital

applications. However, for high-frequency transistor applications, graphene could be used,

but it has to compete against conventional semiconductor materials (III-V materials) [38,

41]. Graphene will probably be used when conventional semiconductor materials fail to

satisfy device requirements [38].

P a g e | 31

As discussed, graphene in principle has a wavelength-independent absorption of

2.3% [13] in the near IR-visible range. This property makes graphene suitable as a

material used in photodetectors [42]. Moreover, high carrier mobility in graphene enables

high bandwidth operation up to 640 GHz [38, 43]. Currently, the poor responsivity arising

from the very small absorption of light has hindered graphene application in

photodetectors. Photocurrent sensitivity can be increased in several ways such as coupling

the graphene with another material i.e. gold [44], using stacked monolayer graphene

separated by a thin tunnel barrier [45], or by increasing light-graphene interaction with a

waveguide [46]. Operating at a bandwidth of ~100 GHz, InGaAs and Ge are more

favourable compared to graphene in photodetector applications at the moment. Therefore,

it is predicted that graphene photodetectors will be competitive in the future providing

that the issues related to it are solved [38]. Another potential graphene application in

photonic devices is as the saturable absorber of mode-locked lasers [pulses of a laser in

extremely short duration in the order of picoseconds (10-12 s) or femtoseconds (10-15 s)]

[47]. Compared to other semiconductor saturable absorbers, the benefits of graphene as a

saturable absorber are: graphene reaches saturation at a lower intensity over a wide

spectral range [48], has ultrafast carrier relaxation times, has controllable modulation

depth [13, 38], and has high thermal conductivity [8].

In terms of energy storage applications, graphene as an active medium in solar

cells would benefit from uniform absorption over a broad spectrum [13], but on the other

hand, it would also suffer from low optical absorption [13] that it would require dopant

enhancement structures [44]. However, doped graphene as an electrode in dye-sensitized

solar cells has proved highly beneficial. Graphene electrodes, depending on the dopants,

can be used as electron/n-type [49] or hole/p-type [50] conducting mediums.

Graphene can be modified to form nanostructured materials which have relatively

high surface areas and thus have the potential to be used in energy storage applications

P a g e | 32

such as in lithium-ion batteries [51]. The high surface area allows an increase of the ion

transfer efficiency, therefore the use of graphene could reduce the amount of electrode

materials needed without reducing the power output. The high surface area of

nanostructured graphene-based materials is also applicable for sodium-ion batteries [52]

and supercapacitors [53, 54] which have shown an increase of the specific energy density

(Wh/kg). The high thermal conductivity of graphene [8] may be a benefit when it comes

to high current loads that generate heat such as in the battery system [38]. This would

help applications that require cooling (e.g. electronics).

1.3 Production of graphene

Graphene can be produced by exfoliation from graphite (top-down techniques) or

by atomic-scale growth (bottom-up techniques). It can be produced in large or laboratory

scales. Figure 1.4 shows several methods known to produce graphene [55].

Micromechanical cleavage or micromechanical exfoliation of graphene-based on

adhesive tape was the first method used to successfully isolate a single layer of graphene

[2, 55]. Micromechanical cleavage can produce μm-sizes of very high-quality graphene

suitable for fundamental studies. However, in order to use graphene in real applications,

low cost and mass scalable techniques need to be developed. Alternative methods to

micromechanical exfoliation, top-down and bottom-up methods, are discussed below.

P a g e | 33

Figure 1.4 Methods for producing graphene. Each of them has its own advantages and

disadvantages related to graphene size, quality and application purposes. Taken from [55].

1.3.1 Anodic bonding

Anodic bonding (Figure 1.4b) consists of placing graphite onto a glass substrate

and applying a high voltage (0.5-2 kV) between the graphite and the metal back contact.

A positive voltage is applied on the graphite side, and then the glass substrate is heated

(~200 oC for 10-20 minutes). A few layers of graphene stick to the glass due to

P a g e | 34

electrostatic interactions of ions in the glass and can be cleaved afterwards [55, 56].

Anodic bonding is able to produce graphene with a lateral size of 20-30 μm.

1.3.2 Photo exfoliation

The photo exfoliation technique (Figure 1.4c) uses pulsed laser irradiation

(typically 800 nm wavelength) in a vacuum or inert conditions to minimize the oxidation

of graphene [57, 58]. The irradiation results in the detachment of an entire or partial layer

of graphite. The energy density required increases with the decreasing number of

graphene layers obtained, up to ~7 layers of graphene can be obtained. This new method

is still in need of further development [55, 58].

1.3.3 Liquid phase exfoliation

Liquid phase exfoliation (LPE) (Figure 1.4d) allows exfoliation of graphite in a

liquid environment. The liquid is either water [59, 60] or an organic solvent [61, 62]. In

general, the process consists of three steps, dispersion of graphite in a liquid, exfoliation,

and purification or separation of graphene flakes from the remaining of graphite flakes.

The exfoliation process is typically done in a bath sonicator, where the formation,

growth, and collapse of bubbles or voids in liquids due to pressure fluctuations of the

sound wave will induce exfoliation [55]. The shear forces created from the collapse of

bubbles between the layers of graphite can break the π-π interaction (inter-layer

interaction) and exfoliate the layers. After the exfoliation, it is important to balance the

inter-sheet attractive forces, therefore suitable solvents need to be identified. If the

interfacial tension between the liquid and graphitic flakes is high, the adhesion (i.e. the

energy per unit area required to separate two surfaces from one interface) between them

is low, and the dispersibility of graphene flakes is poor. It was shown that solvents with

P a g e | 35

the surface tension of ~ 40 mN/m are those providing the highest amount of exfoliated

graphene [61]. Thus, solvents such as 1-methyl-2-pyrrolidinone (NMP), benzyl benzoate,

1,3-dimethyl-2-imidazolidinone (DMEU), and N, N-dimethylacetamide (DMA) are good

for exfoliation and stabilization of graphene [61]. However, all of these solvents have

some disadvantages: they are toxic, and all of them have high boiling points, making it

difficult to remove them after exfoliation. Alternatively, low boiling point solvents such

as isopropanol, acetone, ethanol, etc. can be used, but they give very low yields compared

to organic solvents. Water, a non-toxic and mild boiling point solvent with a surface

tension of 72 mN/m, is not suitable for LPE. In order to produce graphene dispersions in

water, two methods can be used:

(i) A suitable amphiphilic molecule can be used to help stabilise the graphitic flakes

in water, e.g. sodium dodecylbenzenesulfonate (SDBS) [59] and 1-pyrenesulfonic acid

[63].

(ii) The starting graphitic is oxidised, turning the material from hydrophobic to

hydrophilic because of the C-O groups formed on the surface. Graphite oxide can be

easily exfoliated in water leading to the production of graphene oxide (GO). GO shows

very different properties from graphene because of numerous oxygen species

functionalization. GO can be reduced thermally and/or chemically to produce reduced

graphene oxide (rGO) (see Section 1.3.5 for a more detailed discussion).

The liquid phase exfoliation method is an important technique because it allows

high production capacity and high concentration of graphene. The quality of graphene

produced by LPE, although it cannot compare to mechanical exfoliation methods, is

relatively good with a competitive production cost if compared to other methods capable

to produce graphene in large scale. Coleman et al. demonstrated that sonication of

graphite doesn’t cause any basal plane defects on graphene [64-68]. Their argument is

based on the observed Raman D band of thin films, prepared from vacuum filtration of

P a g e | 36

sonication-assisted LPE graphene, which is assumed due to edge defects. They suggested

if the assumption was true, then the average ID IG⁄ ratio should scale to the flake edge to

area ratio: ID IG ∝ [L−1 + w−1]⁄ , where L and w are average length and width of the

graphene flakes, respectively [64]. The prediction was in agreement with their results

indicate that the observed Raman D band was from edge defects [69]. However, several

other studies suggest that ultrasonication does introduce defects in the basal plane of

graphene. The X-ray photoelectron spectroscopy (XPS) analysis carried out by Skaltsas

et al. on NMP and orthodichlorobenzene (o-DCB) LPE graphene produced at different

tip sonication power and times showed high oxygen content present as carboxylic acid

and ether/epoxy functional groups on the graphene lattice as a result of sonication [70].

The effect of bath sonication times on defect localisation has been studied by Bracamonte

et al. which suggests that defects observed for short sonication times were mainly from

edge defects, whereas longer sonication times (> 2 hours) caused basal plane defects.

They also suggested that the observed basal plane defects are not sp3-like or vacancies or

substitutional impurities but topological defects (like pentagon-heptagon pairs) due to a

roughly constant 𝐼𝐷 𝐼𝐷′⁄ ratio of 4.5 ± 0.5 [71] and since they tend to have the lowest

formation energy [72].

1.3.4 Electrochemical exfoliation

The electrochemical exfoliation (EC) technique typically uses electric current to

trigger ion intercalation, structural expansion and exfoliation of graphite working

electrodes in a liquid electrolyte via cathodic reduction or anodic oxidation reaction. The

anodic and cathodic exfoliation are the two main strategies of the EC method and both

have their own advantages and disadvantages. Anodic exfoliation is most commonly used

because the exfoliation can be readily carried out in water using simple electrolytes (e.g.

P a g e | 37

sulphate-based) to produce graphene (1-3 layers thick) in high yield [73, 74]; whereas,

cathodic exfoliation is a slow process and in some cases requires sonication to exfoliate

the already expanded graphite. However, the graphene flakes produced from the

exfoliation of graphite cathode have a lower degree of surface oxidation (typically the

graphene produced has 2.3 wt% increase of oxygen content [75]) or functionalization,

thereby retaining the typical properties of graphene [73].

(i) Anodic exfoliation

The exfoliation for the anodic route is mostly performed in an aqueous electrolyte.

The mechanism of anodic electrochemical exfoliation using sulfate ions can be divided

into three stages [76] and the schematic illustration of the exfoliation is shown in Figure

1.5.

Figure 1.5 Schematic mechanism of anodic electrochemical exfoliation taken from [76].

The first stage consists of a reduction of water at the cathode due to bias voltage,

creating hydroxyl ions [OH−] that act as a strong nucleophile in the electrolyte and

initially attack graphite at the edge sites and grain boundaries. The second stage consists

of oxidation at the edge sites and grain boundaries and then leads to depolarization and

expansion of the graphite layers, thereby facilitating the intercalation of sulfate ions

P a g e | 38

[SO42−] within the graphitic layers. During this stage, water molecules may co-intercalate

with the [SO42−] anions. The third stage consists of reduction of [SO4

2−] anions and self-

oxidation of water to produce gaseous species such as SO2, O2, and others, as evidenced

by the vigorous gas evolution during the electrochemical process [77, 78]. These gaseous

species can exert large forces on the graphite layers, which are sufficient to separate

weakly bonded graphite layers from one another [79].

The sulfate ion is suitable for intercalation and exfoliation of graphite because: a)

the ionic size of sulfate ion (0.46 nm) is close to the graphite interlayer spacing (0.335

nm) [80]; b) the reduction of sulfate ion and the oxidation of water lead to the formation

of gaseous species i.e. SO2, O2, and H2, which could promote the exfoliation of graphene

sheets [81]. Ion intercalation in acidic electrolytes e.g. H2SO4 is so fast that it occurs

simultaneously with exfoliation at all graphite edges and the interplay between [H+] and

[SO42−] results in excess oxidation of graphene at low pH [74]. These problems can be

overcome by the use of melamine additives in sulfuric acid. Chen et al. demonstrate the

use of various melamine additives in aqueous acids (H2SO4 in deionized water) [82]. The

interplay between melamine and the basal plane of graphene was thought to facilitate

exfoliation and provides in-situ protection of the graphene flake surface against further

oxidation resulting in graphene with high C/O ratio (26.2), good uniformity (over 80 %

are less than 3 layers), and low defect density (𝐼𝐷 𝐼𝐺⁄ < 0.45) [82]. Another solution is to

use inorganic salts such as ammonium sulfate, sodium sulfate, and potassium sulfate

which have been previously investigated [76]. Among them, ammonium sulfate

represents the best performance which shows a good C/O ratio of 17.2 with over 80 %

thin graphene produced (consisting of 1-3 layers) and low defect density (𝐼𝐷 𝐼𝐺⁄ = 0.25)

[76].

P a g e | 39

Non-aqueous electrolyte such as organic solution and ionic liquid (IL) could also

be used as the electrolyte. However, cost, safety and efficiency problems restricted the

development of the non-aqueous electrolyte [83].

(ii) Cathodic exfoliation

Unlike the anions intercalation in the anode exfoliation process, the kinetics of

cation intercalation in the cathode exfoliation process is slow [73]. Huang et al. tried to

accelerate the kinetics of the intercalation by using molten LiOH at 600 oC but this was

still not enough to achieve complete exfoliation of the graphite and sonication steps were

required in order to achieve reasonable yields of graphene (80 wt%) [84]. Although the

ion size of [Li+] is small (0.146 nm in diameter) [85] which should be favourable for the

intercalation, however, due to the slow kinetics of the intercalation, [Li+] alone could not

effectively expand the graphite cathode. In principle, incorporation of a [Li+] metal

complex i.e. [Li+]/propylene carbonate (PC) could intercalate into the interlayer space of

cathodic graphite effectively and cause an expansion of the graphite interlayer. However,

subsequent ultrasonication is still needed to achieve a 70 wt% yield of few-layer graphene

and the potentials required are in excess of -15 V [86]. Yang et al. demonstrate a cathodic

intercalation route that could directly exfoliate the graphite cathode by using N-butyl,

methylpyrrolidinium bis(trifluoromethylsulfonyl)imide. However, the potentials required

were around -30 V [87].

1.3.5 Reduced graphene oxide

One of the most popular methods to produce graphene is by oxidising the graphite

to generate graphite oxide which could easily be exfoliated by ultrasonication in various

solvents to produce graphene oxide (GO). The most commonly used method for oxidising

P a g e | 40

graphite is the Hummers’ method, which includes strong oxidising agents like potassium

permanganate, nitric acid and sulfuric acid [88]. The oxidation results in epoxy, hydroxyl

and carbonyl group functionalization of the graphene. The functionalization increases the

graphite interlayer spacing and after subsequent sonication, the hydrophilic nature of the

functional groups facilitates excellent stability of GO in water [89]. GO is negatively

charged due to the ionisation of the functional groups which provides electrostatic

repulsion and increases GO stability in water, alcohols and certain organic solvents [90].

Reduced graphene oxide can be produced by reduction via chemical or thermal

methods. The chemical approach uses reducing agents such as hydrazine (N2H4) [91, 92],

sodium borohydride (NaBH4) [93, 94] and hydrogen iodide (HI) [95] etc. During the

reduction, the brown coloured GO turns black and precipitates in the solution. The

reduction process, however, could not remove the oxygen-containing species completely

and could not restore carbon sp2 hybridisation. Therefore leaving carbon sp3 hybridized

and vacancies [96] (Figure 1.6).

Heat treatment at high temperature could increase the efficiency of the restoration

process with dramatic reduction of surface defects and residual oxygen. Several studies

have demonstrated the restoration of sp2 carbon lattice by applying high temperature (≥

1500 oC) i.e. graphitisation. The obtained graphene showed improved electronic

properties with electron mobility ~1000 cm2/Vs (the precursor rGO had a value of 130

cm2/Vs) [97] and electrical conductivity of 577000 S/m [98].

P a g e | 41

Figure 1.6 High-resolution transmission electron microscopy (HRTEM) image of single-

layer rGO. Colour scheme highlighted different features. Light grey colour represents the

defect-free areas. Dark grey colour represents contaminated regions. Blue colour

represents disordered single-layer carbon network or extended topological defects

identified as remnants of the oxidation-reduction process. Red colour represents

individual adatoms or substitutions. Green colour represents isolated topological defects.

Yellow colour represents holes and their edge reconstructions. The scale bar is 1 nm. The

image is taken from [96].

1.3.6 Thermal decomposition of SiC

Thermal decomposition of SiC/growth of graphene on SiC technique (Figure

1.4e) uses SiC as the carbon source. The method is performed by annealing the SiC under

ultra-high vacuum. The SiC decomposes above 1000 oC, the carbon graphitizes due to

P a g e | 42

evaporation of Si. The problem that arises in this method is that the graphene grows on

SiC which has a different atomic structure compared to graphene. The mismatch of the

substrate is thought to lead to defects on the graphene produced [55].

1.3.7 Growth of graphene on metallic surfaces by precipitation

The growth of graphene on metallic surfaces by precipitation (Figure 1.4f) refers

to techniques allowing precipitation of carbon atoms on metal surfaces. The techniques

used are flash evaporation, physical vapour deposition (PVD), CVD, spin coating etc [55].

The carbon source is in the form of solid, liquid or gas. Carbon precipitation is affected

by the amount of pressure, temperature, annealing time, cooling rate and metal thickness

[99].

1.3.8 Chemical vapour deposition

Chemical vapour deposition (CVD) (Figure 1.4g) allows the growth of

polycrystalline graphene by depositing a mixture of hydrocarbon gas on a metal plate at

high temperature. The technique is proven to be promising and has been able to produce

square metres of graphene [39]. However, it does require a transfer/removal process as

the most cost-effective graphene produced so far is grown on inexpensive metals such as

copper, nickel and cobalt [99], which may not be a suitable substrate for many

applications. The high cost of production and the difficulty to control the grain size have

hindered the development of this technique.

The wet-transfer method is often used to transfer graphene from a metallic

substrate to other substrates. A well-known wet-transfer method uses polymethyl

methacrylate (PMMA) [100] or polydimethylsiloxane (PDMS) [101] to coat the graphene

P a g e | 43

on the metal substrate followed by etching the metal substrate in an etchant (e.g. iron(III)

chloride, FeCl3) (Figure 1.7). After successful transfer of the graphene onto the target

substrate, the polymer can be dissolved by using acetone. However, a small residue of the

polymer remains on the graphene resulting in p-doped graphene in some samples [102].

Figure 1.7 Schematic diagram of the wet-transfer process taken from [103].

1.3.9 Molecular beam epitaxy

Molecular beam epitaxy (MBE) (Figure1.4h) is an epitaxial growth technique

based on the interaction of species adsorbed from molecular beams of thermal energy on

a heated crystalline substrate under ultra-high vacuum (UHV) conditions. The UHV

conditions are required to minimize the incorporation of contaminants at the growth

surface and to prevent such contamination. It is also required to use high-purity materials

as source materials [104].

Amongst the various insulating substrates available, SiC is often used because of

its graphene-like crystalline structure. Nevertheless, with the graphene lattice parameter

of 0.246 nm (2.46 Å, the second neighbour distance) and the corresponding parameter for

SiC is 0.307 nm (3.07 Å, projected in the [0001] plane), the lattice mismatch between the

graphene and SiC is below 0.3 % [105]. The drawback with SiC is that the bulk material

is expensive and the graphene produced contains numerous defective regions [106, 107].

P a g e | 44

Interest in using hexagonal boron nitride (h-BN) as the insulating substrate

appeared recently because it was shown that the transport properties of graphene were

better [108]. The lattice parameter of h-BN is 0.25 nm (2.5 Å, the second neighbour

distance) which makes the h-BN almost similar to graphene [109]. The main problem

with h-BN is the low and heterogeneous nucleation of graphene [110, 111].

Analytical techniques to study graphene

Many different analytical techniques have been used to study graphene including

Raman spectroscopy, atomic force microscope (AFM), electron paramagnetic resonance

(EPR) spectroscopy etc. The following sections describe the basic principles of EPR

followed by examples of its use in graphene research. This will then be complemented by

a section on the fundamentals of Raman spectroscopy and its use to study graphene.

1.4 Electron paramagnetic resonance spectroscopy

1.4.1 Electron paramagnetic resonance basic principle

Electron paramagnetic resonance (EPR) spectroscopy also known as electron spin

resonance (ESR) spectroscopy is a spectroscopic technique used to characterise

substances or molecules that have unpaired electrons [112]. The samples that are analysed

can be in the form of fluid or solid. The most common EPR experiment consists of

applying a continuous-wave (CW) of electromagnetic radiation and sweeping the

magnetic field on the sample. The first EPR spectrum was observed in 1944 by a Russian

physicist, E.K. eavoisky [113].

P a g e | 45

An electron has spin angular momentum 𝑆 and spin quantum number 𝑚𝑠 . The

magnetic moment of an electron 𝜇 is proportional to the spin angular momentum 𝑆

𝜇 = −𝑔𝑒𝜇𝐵𝑚𝑠 Equation 1.1

with 𝜇𝐵 = 𝑒ℏ2𝑚𝑒

⁄ is Bohr magneton (9.274009994 x 10-24 J T-1), ℏ = ℎ2𝜋⁄ , ℎ is

Planck's constant (6.62607015 x 10-34 J s) and g is the g value (Equation 1.1). The exact

g value for a free electron of 𝑔𝑒 = 2.0023193043617 is derived from quantum

electrodynamics. The negative sign means that the magnetic momentum of the electron

is collinear but antiparallel to the spin itself [112].

In the presence of an external magnetic field 𝐵0 , eeeman splitting occurs

depending on the electron magnetic quantum number and the strength of the magnetic

field, as shown by Equation 1.2.

𝐸 = ±1

2𝑔𝜇𝐵𝐵0 Equation 1.2

Electromagnetic irradiation with a frequency 𝜈 that matches the energetic

difference ∆𝐸 will result in absorption, as shown in Figure 1.8. The g value of the

absorption can be calculated by using Equations 1.3 and 1.4, with 𝜈 in GHz and 𝐵0 in

Gauss.

∆𝐸 = ℎ𝜈 = 𝑔𝜇𝐵𝐵0 Equation 1.3

𝑔 = 714.5 𝜈 𝐵0⁄ Equation 1.4

P a g e | 46

Figure 1.8 Energy levels of an unpaired electron spin in the applied magnetic field.

Resonant energy absorption (Equation 1.3) leads to an electron spin ‘flip’ or transition

resulting in an EPR signal. The signal can be presented in absorption (dotted) or first

derivative (solid) mode. Taken from [112].

EPR uses this electromagnetic absorption principle to detect molecules or atoms

that have unpaired electrons by detecting the changes in the electromagnetic resonance

frequency. The resonance detection can be conducted in two ways; either the

electromagnetic frequency is held constant and the magnetic field is swept or the applied

electromagnetic frequency is varied while the magnetic field is kept constant. EPR

spectroscopy uses the former case because it is easier to vary the magnetic field than to

change the frequency. Field modulation is used to increase the sensitivity of the detection.

The resultant of the modulated signal is its first derivative as shown in Figure 1.8.

P a g e | 47

(i) Relaxation

During the EPR transition, an electron in a lower energy state (spin-down) will

absorb the electromagnetic (EM) radiation and move into a higher energy state (spin-up).

To maintain the net energy between two spin energy states, an electron in the higher

energy state will release a phonon ℎ𝜈 to move into a lower energy state. The transition

from a higher into a lower energy state is called relaxation. The rate of the relaxation

process is expressed as relaxation time. Saturation occurs if the relaxation rate is too slow.

These phenomena can be observed when the absorption does not increase, or line

broadening starts to happen.

There are two types of relaxation process; spin-lattice relaxation and spin-spin

relaxation. In spin-lattice relaxation, the energy is released within the lattice as phonons

(vibrational, rotational and translational energy). The spin-lattice relaxation is

characterised by an exponential decay of energy as a function of time. The exponential

time constant or the spin-lattice relaxation time is denoted as 𝑇1 . In the spin-spin

relaxation, the energy exchange between the spins occurs without transfer of energy to

the lattice. The time constant or spin-spin relaxation time is known as 𝑇2. Both spin-lattice

and spin-spin relaxation contribute to the EPR linewidth (Equation 1.5):

Δ𝐼𝑝𝑝 ∝1

𝑇1+

1

𝑇2 Equation 1.5

The linewidth, in general, depends mainly on spin-spin interactions (𝑇1 > 𝑇2). 𝑇2

increases if the spin concentration is decreased which causes the spin-spin distance in the

system to be larger. 𝑇1 is inversely proportional to the absolute temperature (𝑇1 ∝ T−n)

with n depending on the relaxation mechanism. Thus, cooling the sample increases 𝑇1

and may lead to detectable resonances.

P a g e | 48

(ii) The g value

The g value of the free electron 𝑔𝑒 = 2.002319 is a fundamental constant. In real

materials, 𝑔 ≠ 𝑔𝑒 because of the orbital angular momentum contribution to the magnetic

moment. This is often discussed as the g shift (Δ𝑔, Equation 1.6). The mixing of spin

angular momentum and orbital angular momentum is called spin-orbit coupling (SOC,

Equation 1.7).

Δ𝑔 = 𝑔 − 𝑔𝑒 Equation 1.6

𝑔 = 𝑔𝑒 −𝑛𝜆𝑎2

Δ𝐸 Equation 1.7

where 𝜆 is the SOC constant (larger for heavier element), 𝑎2 is the covalency parameter

(≤1), and n is the quantum mechanical coefficient.

Organic free-radicals usually have large Δ𝐸 and small 𝜆 . Inconsequent,

hydrocarbon radicals usually have g values ranging from 2.002 – 2.003, N/O based

radicals have g values between 2.003 – 2.006, and S-based radicals give g-values of 2.007

– 2.010. Transition metal ions with small Δ𝐸 and large 𝜆 can have large g shifts. In the

case of transition metals, g values can be < 𝑔𝑒 when the SOC occurs at an empty orbital.

(iii) Lineshape

The g value can be affected by the orientation of the molecule in the magnetic

field because orbitals are oriented in the molecule. In other words, g values can be

anisotropic. In the fluid form, all of this anisotropy is averaged out. However, in the solid

form, the g value can change as the sample is rotated in different directions.

P a g e | 49

Each molecule has a unique axis system called the principal axis system. The g

values measured along the axis are called the principal g values and denoted as 𝑔𝑥, 𝑔𝑦

and 𝑔𝑧. The anisotropic interactions lead to powder spectra (Figures 1.9-1.10).

In powder spectra, the anisotropic resonance can appear as isotropic, axial or

rhombic symmetry. Typical axial and rhombic symmetry spectra are presented in Figures

1.9 and 1.10, respectively. The average g-value is known as 𝑔𝑖𝑠𝑜 can be written as shown

in Equation 1.8.

𝑔𝑖𝑠𝑜 =𝑔𝑥+𝑔𝑦+𝑔𝑧

3 Equation 1.8

Figure 1.9 Typical anisotropic axial spectra for 𝑔𝑧 > 𝑔𝑥 = 𝑔𝑦: 1st derivative line (red)

and absorption line (blue). The Figure was made using a simulator provided in

www.eprsimulator.org [114].

P a g e | 50

Figure 1.10 Typical rhombic symmetry spectra: 1st derivative line (red) and absorption

line (blue). The Figure was made using a simulator provided in www.eprsimulator.org

[114].

(iv) Hyperfine Interaction

The hyperfine interaction also known as the hyperfine coupling is the interaction

between the electron and the nuclei. The nuclei of the atoms in a molecule or complex

often have magnetic moments, which produce a local magnetic field at the electron. There

are two selection rules in EPR that allow the interaction between electron spin and nuclear

spin. The selection rules are shown in Equation 1.9:

|∆𝑚𝑠| = 1 𝑎𝑛𝑑 ∆𝑚𝑙 = 0 Equation 1.9

P a g e | 51

Figure 1.11 Energy level diagram in a fixed magnetic field for a system with S = 12⁄ and

I = 12⁄ , in the highfield approximation, showing the electron eeeman (Ee) and nuclear

eeeman (Ne) levels, and the perturbation arising from the hyperfine interaction (HF).

The two allowed EPR transitions (solid arrows) result in the experimentally observed

resonances labelled EPR I and EPR II (shown in the inset). Adapted from [112].

The energies E1, E2, E3 and E4 are derived from equation 1.10 with E1 (𝑚𝐼 = − 12⁄ )

and E2 (𝑚𝐼 = 12⁄ ) related to 𝑚𝑠 = − 1

2⁄ and E3 (𝑚𝐼 = − 12⁄ ) and E4 (𝑚𝐼 = 1

2⁄ ) related to

𝑚𝑠 = 12⁄ . The equation 1.10 describe the general equation for hyperfine interaction:

𝐸 = 𝑔𝜇𝐵𝐵𝑚𝑠 − 𝑔𝑁𝜇𝑁𝐵𝑚𝐼 + 𝐴𝑚𝑠𝑚𝐼 Equation 1.10

with 𝑔𝑁 is the nuclear g value, 𝜇𝑁 is the nuclear magneton (5.0508 x 10-27 J T-1) and A is

the hyperfine coupling constant.

Figure 1.12 shows a system with S =12⁄ and I = 1 2⁄ (i.e. hydrogen, 1H), and the

result is the appearance of two EPR resonances corresponding to a hyperfine with a

coupling constant a. If there is a second nucleus with I = 12⁄ , each of the signals is further

split into a pair resulting in four signals. For n number of nuclei with I = 1 2⁄ , there are 2𝑛

P a g e | 52

EPR signals. The number of hyperfine lines follows the general rule 2𝑛𝐼 + 1, where n is

the number of nuclei and I is the nuclear spin.

Figure 1.12 The EPR spectrum of a system with S =12⁄ and I = 1 2⁄ . The Figure was

made using a simulator provided in www.eprsimulator.org [114].

(v) Temperature dependence of magnetic susceptibility

The temperature dependence of magnetic susceptibility (𝜒) is proportional to the

temperature dependence of the double integral of the EPR signal (𝜒𝐸𝑃𝑅). Therefore, the

magnetic susceptibility behaviour of a material can be determined by fitting the

temperature dependence of the EPR signal intensity with the Curie, Curie-Weiss or Pauli

laws. However, Pauli paramagnetism, characterized by a nearly temperature-independent

susceptibility, is mainly observed on materials having conductive electrons i.e. metals and

the susceptibility is proportional to the density of conduction electron states at the Fermi

level (EF) [115].

Materials with paramagnetic behaviour are known to have weak interacting

permanent magnetic moments. As a consequence of negligible exchange interactions

between moments, a true paramagnetic material shows no sign of magnetic ordering down

to the lowest temperatures. The magnetic susceptibility of a paramagnetic material is

P a g e | 53

positive and strongly dependent on the temperature and follows the Curie law (Equation

1.11 and Figure 1.13a):

𝜒 = 𝐶T⁄ Equation 1.11

where C is the Curie constant and T is the absolute temperature. The Curie law

dependence in the sample is indicative of the presence of isolated paramagnetic ions,

radicals or atoms in the material. Diamagnetism is the opposite of paramagnetism since

it has negative magnetic susceptibility due to a moving electron charge in a manner

described by Lenz's law of electromagnetism [115]. Materials with diamagnetism

characteristics do not have unpaired electrons and therefore cannot be analysed by using

EPR spectroscopy.

Ferromagnetism is when the magnetic moments align parallel to each other. The

ferromagnetic susceptibility of a material diverges at the Curie temperature (TC). In

principle, at temperatures below TC, a spontaneous magnetization (magnetization in zero

magnetic fields) within one domain occurs. The spontaneous magnetization increases

with decreasing temperature. A block of a ferromagnet contains a number of domains

whose spontaneous magnetizations compensate mutually so that the total bulk

magnetization is zero. At temperatures above TC, the magnetic moments become

disordered as in a paramagnet due to the fact that the thermal energy is greater than the

magnetic interactions. The transition is reversible. At temperatures above TC, the

magnetic susceptibility follows the Curie-Weiss law (Equation 1.12 and Figure 1.13b):

𝜒 = 𝐶

T−𝜃 Equation 1.12

where the Curie-Weiss constant, θ, is positive, has the dimensions of temperature and has

a value usually close to TC.

P a g e | 54

Figure 1.13 The temperature dependence of the reciprocal magnetic susceptibility. a)

Curie law behaviour of a paramagnet; b) Curie-Weiss law behaviour of a ferromagnet; c)

Curie-Weiss law behaviour of an antiferromagnet; d) behaviour of a ferrimagnet. Figure

a-b is taken from [116], Figure c-d is taken from [115].

The magnetic moments can align in an antiparallel fashion due to energetically

favourable conditions and this is called antiferromagnetism. Above a temperature called

the Néel temperature (TN), the arrangement of the magnetic moments becomes disordered

and behaves as a paramagnet. Above the Néel temperature, the magnetic susceptibility

follows Curie-Weiss law (Figure 1.13c and Equation 1.11) with the constant θ is negative.

At temperatures below TN, the susceptibility decreases (inverse susceptibility increases).

Antiferromagnets exhibit zero spontaneous magnetization due to pairs of magnetic

moments coupled in an antiparallel manner and mutually compensate.

Ferrimagnetism happens when the antiparallel coupling of magnetic moments do

not mutually compensate. As a consequence, a net spontaneous magnetization is observed

a) b)

c) d)

P a g e | 55

at temperatures below the ordering temperature TC. At temperatures well above TC, the

magnetic susceptibility behaviour is paramagnetic and follows the Curie-Weiss law,

usually with a negative value of θ. At a certain temperature interval above TC, the

temperature dependence of the inverse magnetic susceptibility forms a curve as shown in

Figure 1.13d.

(vi) Instrumentation

The main parts of the continuous-wave (CW) EPR instrument are the microwave

bridge, the cavity, the magnet and the console for signal processing. Figure 1.14 shows

the scheme of a CW EPR spectrometer without the console. The electromagnetic radiation

source and the detector are placed in a box called the microwave bridge. The cavity is a

metal box and its function is not only as a sample holder but also helps to amplify weak

signals from the sample. The magnetic field is modulated at high frequency (100 kHz).

As a result of the field modulation and phase-sensitive detection, the spectrum is recorded

as the first derivative of the absorption (Figure 1.8). The sensitivity of measurement

increases at higher frequencies, but the sample volume decreases and the instrument

becomes more difficult to use.

Phase-sensitive detection with magnetic field modulation can increase the

sensitivity by several orders of magnitude. However, care should be taken when choosing

the appropriate modulation amplitude, frequency and time constants. Using too large a

modulation amplitude (larger than the linewidth of the EPR signal) will make the detected

EPR signal broaden and distorted. Although higher modulation amplitudes cause

broadening of the signal, the integrated intensity of the signal continues to increase

linearly with modulation amplitude. In this case, high modulation amplitude could be

applied if the main goal of the experiment is spin quantitation. Time constants can affect

P a g e | 56

the noise in the spectrum. The increasing time constant can suppress the noise level. If

the time constant is too long for the scan rate of the magnetic field, the signal may be

distorted or even missed. A slower scan rate must be used if the user wants to use a long

time constant to suppress the noise further.

Figure 1.14 The scheme of a CW EPR spectrometer employing magnetic field

modulation. The Figure is taken from [117].

(vii) Pulsed EPR

Pulsed EPR uses a range of microwave frequencies for set periods of time defined by

pulse sequences. The pulse sequences can be divided into the following steps (Figure

1.15): a) At equilibrium, the average magnetic moment of a group spins of a sample will

be parallel to the magnetic field. b) A 90-degree pulse at resonant frequency radiation is

applied, and spins respond in bulk by tipping into the X-Y plane. c) Due to local

inhomogeneities of magnetic field (variation in the magnetic field at different parts of the

sample and at constant time), as the net magnetic moment precesses, some spins slow

down due to lower local field strength while some speed up due to higher field strength

P a g e | 57

and thus make the signal decay [free-induction decay (FID)]. d) A 180-degree pulse is

then applied, and the magnetic moment flips 180 degrees, the slower spins now lead ahead

of the main magnetic moment and the fast ones trail behind (rephasing spins). e)

Progressively, the fast and the slow magnetic moments eventually catch-up with one

another resulting in a refocusing spin or “spin echo”. The intensity of the echo is affected

by the time between the two pulses (tau) and can be monitored by integration in pulsed

EPR experiments. Therefore, the spin-echo decay can be observed by recording the

changes in the size of the echo for different values of tau resulting in an exponential decay

diagram, which describes the spin-spin relaxation time (T2).

Figure 1.15 Illustration of the magnetization vector at characteristic positions in the

typical 2-pulse sequence. Adapted from [118].

(viii) Inversion recovery

Inversion recovery with echo detection can be used to measure spin-lattice

relaxation time ( 𝑇1 ). Three-pulse inversion recovery uses a 𝜋 − 𝑡 − 𝜋 2 −⁄ 𝜏 − 𝜋

sequence to generate an echo at tau after the last pulse. A π pulse inverts the magnetization

and the echo integration is detected as the delay pulse (t) is increased. In order to extract

the signal from noise, the echo needs to be averaged by repeat scans.

P a g e | 58

(ix) Electron spin echo envelope modulation

Electron spin echo envelope modulation (ESEEM) can be used to detect weak

hyperfine couplings. Two-pulse ESEEM uses a simple 𝜋 2⁄ − 𝜏 − 𝜋 pulse sequence to

generate a primary echo detected at time τ after the second pulse. The issues with the two-

pulse ESEEM are: a) unresolved Fourier transformation of the low-frequency modulation

time trace due to the typically short phase memory time (TM) which leads to unresolved

spectra. b) Spectrometer dead-time (𝜏𝑑, ~100 ns at X-band frequency) which leads to

distortions or artefacts in the resulting frequency-domain spectrum [112]. Some of the

limitations of two-pulse ESEEM can be overcome by three-pulse ESEEM which uses

𝜋 2⁄ − 𝜏 − 𝜋 2 − 𝑡 − 𝜋 2⁄⁄ as a pulse sequence. The three-pulse ESEEM sequence

generates a stimulated echo observed at time τ after the third pulse. The experimental time

trace of an electron coupled to a single nuclear spin equation includes TM and the nuclear

phase memory time (TM(N)), which is longer than TM, leading to narrow lines in the

frequency-domain spectrum and therefore increases the spectral resolution. However, the

three-pulse ESEEM is still affected by blind-spot behaviour [112].

(x) Hyperfine sublevel correlation

The hyperfine sublevel correlation (HYSCORE) spectroscopy is a four-pulse

microwave sequence in which a mixing π pulse is inserted between the second and the

third 𝜋 2⁄ pulse of the three-pulse ESEEM experiment; the total sequence

becomes 𝜋 2⁄ − 𝜏 − 𝜋 2⁄ − 𝑡1 − 𝜋 − 𝑡2 − 𝜋 2⁄ . The two pulse delays 𝑡1 and 𝑡2 are

varied independently to produce a two-dimensional (2D) time delay array. The Fourier

transformation of the modulated time decay data for both 𝑡1 and 𝑡2 is presented in a 2D

frequency-domain spectrum with 𝑣1and 𝑣2 as axes. In HYSCORE, the frequencies from

weakly-coupled nuclei appear as cross-peaks in the (+,+) quadrant, whereas strongly-

P a g e | 59

coupled nuclei are observed in the (-,+) quadrant. For disordered systems with broad

ESEEM features, the correlation peaks broaden into ridges as illustrated in Figure 1.16.

Figure 1.16 a) 2D HYSCORE spectrum where full squares ■ represent cross-peaks from

weakly coupled nuclei in the (+,+) quadrant, and full circles ● represent cross-peaks from

strongly coupled nuclei in the (-,+) quadrant. 𝑣𝐿 is the Larmor frequency for the nucleus

of interest, A is the hyperfine coupling, 𝑣𝛼(= 𝜔12) and 𝑣𝛽(= 𝜔34); b) (+,+) quadrant for

the powder HYSCORE pattern for an S = I = 1 2⁄ spin system with an axial hyperfine

tensor. The Figure is taken from [112].

1.4.2 State-of-the-art of EPR in graphene

A selection of published papers on CW EPR of graphene materials has been

summarized in Table 1. Although some results have been presented, the field is still open

for discussion as different types of graphene have been investigated (LPE, CVD grown

etc.) with very little information on the type of graphene analysed (thickness, lateral size,

defects etc.). Therefore, a more detailed study, with well-characterized samples, is

necessary. EPR spectroscopy from a single layer of graphene may give a very weak signal

and limited information as the conduction electron spins which affected by the Fermi

level (EF) position would not be resolved. The EPR spectroscopy from a stack of graphene

P a g e | 60

sample is stronger and may be affected by the conduction electron spins due to interlayer

interaction affecting the EF position causing a more complicated EPR spectroscopy.

Year Type of Graphene EPR result References

2009

Mechanically exfoliated graphene

on Kapton scotch-tape;

Dominated by multilayer

graphene

g value 2.0034; single

Lorentzian lineshape; Curie

behaviour; linewidth broadening

at low temperature

[14]

2011

Commercial graphene deposited

on amorphous SiO2 under

vacuum; Produced by substrate-

free gas-phase synthesis

g value 2.00245 ± 0.00005;

single isotropic Lorentzian

lineshape; Curie-Weiss

behaviour (antiferromagnetic)

[15]

2012 Synthesized graphene nanoribbon

g value 2.0032; isotropic non-

Lorentzian lineshape; Curie

behaviour

[36]

2012

Synthesized graphene

nanoribbons; Split (GNRs) and

oxidative unzipped

ribbon(CCGNRs)

g value 2.0025 (GNRs) and

2.0032 (CCGNRs);

ferromagnetic (GNRs); the

number of spin GNRs is 2

orders of magnitude higher than

CCGNRs

[16]

2013

Commercial graphene deposited

on SiO2

g value 2.00245 ± 0.00005;

antiferromagnetic; linewidth

decreased linearly with

temperature

[18]

2014 Synthesized graphene multilayer

g value 2.0044; single isotropic

Lorentzian lineshape; linewidth

broadening at low temperature;

[20]

2014 CVD graphene deposited on SiO2

g value ~2.002; a mixture of

Gaussian and Lorentzian

lineshape with small

anisotropic; antiferromagnetic

[19]

2016 CVD graphene in a transistor

device

g value ~2.0033-2.0036 (with

applied voltage); a mixture of

Gaussian and Lorentzian; Pauli

paramagnetism

[119]

2016 Annealed reduced graphene

oxides

g value ~2.004 (annealed at

1000 oC); Isotropic single EPR

line;

[120]

2017 Reduced graphene oxides Three Lorentzian components

with g value 2.0000-2.0031 [121]

Table 1 CW EPR studies on graphene materials.

P a g e | 61

From a theoretical point of view, the origin of magnetism in graphene may come

from vacancies defects [122], zig-zag edge states of graphene [28, 123], absorption of

adatoms [124, 125]. In regard to zig-zag edges, electron spins would be strongly coupled

in parallel with each other through strong ferromagnetic interactions. However,

Kunstmann et al. revealed that magnetic edge states might not exist in real systems and

showed that there are at least three very natural mechanisms which dramatically reduce

the effect of edge states or even totally eliminate them. The three mechanisms are edge

reconstruction, edge passivation, and edge closure [28].

In the case of graphene exfoliation, the exfoliation method can significantly affect

the composition of the samples. As consequences, each sample can have different EPR

behaviour. A temperature-dependent study on mechanically exfoliated graphene revealed

that the g-value is slightly decreased below 70 K, and the linewidth begins to broaden

below 70 K (Figure 1.17a). The former is due to strong coupling of the defects to the

conduction electrons at a lower temperature, while the latter is due to the motional

broadening of the dipolar linewidth of localized spins. The graphene sample was reported

to consist of mono-layer and ultra-thin graphite [14].

a)

P a g e | 62

Figure 1.17 Temperature dependence of the EPR linewidth for mechanically exfoliated

graphene (a) [14] and LPE graphene (b) [15].

A study on LPE graphene has shown different temperature-dependent behaviour.

Figure 1.17b shows the linewidth narrowed as the temperature decreases. The authors did

not discuss further the linewidth changes and they postulated that the source of localised

states in their sample was due to vacancies created from vacuum treatment [15].

In another study on LPE graphene, the authors reported that they were able to

observe an EPR signal after the sample was put inside a high vacuum for 60 hours. The

resonance situated at ~210 mT was thought to be a ferromagnetic resonance (FMR) line.

The resonance centre for the FMR signal was shifted to a lower value at low temperature,

as shown in Figure 1.18, due to the presence of the local magnetic field [37].

b)

P a g e | 63

Figure 1.18 a) Temperature dependence of the electron spin resonance (ESR) signal from

LPE graphene. b) Temperature dependence of normalized ESR susceptibility measured

after the annealing treatment showing a weaker signal which assigned to the conducting

electrons. The solid line corresponds to the Curie law. The spectrum in the inset was

recorded at 100 K with 64 accumulations. The Figure is taken from [37].

Tadyszak et al. assigned the signal shown in Figure 1.18b to itinerant spins on the

basis of the Pauli type temperature dependence of the ESR susceptibility. Further, the

weak temperature dependence was thought due to the interactions with the localized states

on the zigzag edges [37].

Itinerant and localised electrons below 50 K on multilayer graphene have been

detected by using multi-frequency (9.4 - 420 GHz) EPR [20, 126]. At 315 GHz and below

40 K the single Lorentzian line started to diminish into a powder spectrum with small g-

a)

b)

P a g e | 64

value anisotropy. The anisotropic powder spectrum is formed completely at 2 K with

𝑔𝑥𝑥 = 2.00441, 𝑔𝑦𝑦 = 2.00452 and 𝑔𝑧𝑧 = 2.00431. The isotropic conduction electron

spin, resonant with 𝑔𝐶𝐸𝑆𝑅 = 2.00434, was also observed [126]. The temperature

dependence of the line-width exhibited an anomaly below 70 K (Figure 1.19). They

attributed the anomaly to temperature-induced decoupling of the localised and conduction

electrons [20].

Figure 1.19 Temperature dependence of the linewidth from multilayer graphene. The

inset shows the temperature independence of the g-value. The Figure is taken from [20].

In the terms of EPR lineshape, Tampieri et al. (2014) indicated that the electron

interaction of graphene and graphite samples can be categorized into three types of

situations: (a) non-interacting localized electrons, like in radicals, with small hyperfine

(electron-nuclear) interactions, exhibiting a Gaussian lineshape, (b) electrons localized or

delocalized in narrow regions, with strong electron-electron interaction, exhibiting a

Lorentzian lineshape, and (c) mobile electrons in conductive particles with dimensions

larger than the microwave penetration depth (skin depth), exhibiting a Dysonian lineshape

and normally having low intensity [21].

P a g e | 65

Uniaxial g-value anisotropy lineshape was observed from graphene samples

prepared by three different exfoliation routes [127]. The exfoliation routes were

sonicating, shear mixing and stirring (Figure 1.20).

Figure 1.20 EPR spectra of (a) SGN18 graphite powder, (b) ultrasounded, (c) shear mixed,

and (d) stirred few-layer graphene. The inset shows the uniaxial g-value simulated EPR

lineshape for the stirrer prepared sample. The Figure is taken from [127].

Figure 1.20 shows graphene samples which contain two EPR resonances with two

g-values that come from the two crystallite directions (𝐵0 ∥ 𝑐 − 𝑎𝑥𝑖𝑠 𝑎𝑛𝑑 𝐵0 ⊥ 𝑐 −

P a g e | 66

𝑎𝑥𝑖𝑠). The c-axis is perpendicular to the graphene sheets. Their graphene samples are

dominated by few-layer graphene with the g-value between the free electron and the

graphite powder. The SGN18 graphite powder displays a broad line of 12.2 mT and g-

value at 2.0148. Ultrasounded few-layer graphene gave a Lorentzian line of 1.1 mT

linewidth at g = 2.0059, while the shear mixed presented a 1.4 mT linewidth at g = 2.0082.

The stirred sample has a uniaxial anisotropic signal with a width of 1.2 mT at g = 2.0094.

The broad line components for the SGN18 graphite sample was thought due to conduction

electrons present in graphite. The narrow anisotropic lines are associated with defects and

dangling bonds in all cases [127].

Graphite, however, is different than graphene. Graphite, either natural graphite or

synthetic graphite such as highly oriented pyrolytic graphite (HOPG), has a paramagnetic

signal and the EPR lineshape may depend on its crystal size [127-131]. Natural graphite

usually has relatively large crystals and Dysonian lineshapes. The Dysonian lineshape

associated with conduction electron typically appears for crystals with a size of more than

3 μm (skin depth) [129, 132, 133]. Nano graphite, however, showed different EPR spectra.

The typical Dysonian line was not present in the sample. The spectrum was a single and

narrow EPR line that could be simulated with a single Lorentzian lineshape with g-value

at 2.0035 [128].

P a g e | 67

1.5 Raman spectroscopy

1.5.1 Raman basic principles

The Raman effect was discovered in 1928 by Sir Chandrasekhra Venkata Raman

[134]. Light is an electromagnetic wave quantised as a photon, with energy ∆𝐸 (Equation

1.12).

Δ𝐸 = ℎ𝜈 Equation 1.13

with ℎ is the Planck constant (6.626 𝑥 10−34 𝐽𝑠) and 𝜈 is the wave frequency.

Matter interacts with the electromagnetic wave in different ways, depending on

the wave frequency or the wavelength. X-ray light will be diffracted by atomic lattice;

light in the UV-Visible spectrum will cause electron excitation - molecules with

conjugated π systems tends to fluorescence under UV [135-137]; light in the IR-Visible

region will cause excitations of vibrational and rotational levels of molecules and crystals

[138, 139]; microwaves will cause rotational levels of gaseous molecules [139].

Incident light is scattered elastically (re-scattered with the same frequency as the

incident light) from atoms of the material. The light interacts with the molecule and

distorts (polarize) the cloud of electrons around the nuclei to form a short-lived state

called a virtual state. This state is not stable and the photon is quickly re-radiated with the

same frequency as the incident radiation while the cloud of electrons drops back down to

the original state [140]. This effect is known as Rayleigh scattering.

However, a small portion of the incident light is also scattered inelastically (re-

scattered with a different frequency than the incident light). This happens when the

nuclear motion is induced during the scattering process causing the energy transfer either

from the incident photon to the molecule or from the molecule to the scattered photon

P a g e | 68

[140]. The energy of the scattered photon is different from the incident photon by one

vibrational unit. This effect is called Raman scattering. Raman scattering is a weak effect

as it happens once in 106 - 108 incident photons. Figure 1.21 shows the basic processes

which occur for one vibration.

Figure 1.21 Schematic of the Rayleigh and Raman processes. The lowest energy

vibrational state m is shown at the foot with a state one vibrational unit in energy above

it labelled n. Rayleigh scattering also occurs from higher vibrational levels such as n.

Taken from [140]

At room temperature, most molecules, but not all, are present in the lowest energy

vibrational level. The Raman scattering process from the ground vibrational state m leads

to absorption of energy by the molecule and its promotion to the higher energy excited

vibrational state n. This is called Stokes scattering. However, due to thermal energy, some

molecules may be present initially in an excited state as represented by n in Figure 1.21.

Scattering from these states to the ground state m is called anti-Stokes scattering and

involves the transfer of energy from the molecule to the scattered photon. This process is

less probable than the Stokes process. The relative intensity of the anti-Stoke process is

P a g e | 69

very weak and depend on the Boltzman population states distribution [140]. At room

temperature, the number of molecules expected to be in an excited vibrational state other

than really low energy states will be small. Thus, a typical Raman spectrum only shows

the Stoke lines. In general, anti-Stokes scattering will become weaker the higher the

energy of the vibration, due to the decreasing population of the excited vibrational states.

However, anti-Stokes scattering will increase relative to Stokes scattering as the

temperature rises [140]. In this report, we will only show the Stokes side of the Raman

spectrum.

1.5.2 Raman spectrum of graphene

Raman spectroscopy is the most used non–destructive method for the

characterization of graphene [62, 141, 142]. The spectrum of graphene shows two main

peaks, called the G and 2D peaks, in the region of 1200-3000 cm-1, Figure 1.22.

Figure 1.22 a) Mechanically exfoliated graphene showing both monolayer and bi-layer

regions. b) Raman spectra of mono and bi-layer graphene. The top and bottom insets

P a g e | 70

represent the enlarged 2D bands of regions B and A, respectively. The Figure is taken

from [143].

The G band which lies around ~1580 cm-1 corresponds to the bond stretching of

all pairs of sp2 atoms. The 2D band (~2700 cm-1) is the second-order harmonic (overtone

or higher-order Raman processes) of the D band. The D band (usually lies around ~1350

cm-1) is not visible in the Raman spectrum of defect-free graphene because the D band is

defects-activated. However, in the case of a second-order mode, no defects are required

for its activation. The 2D band is always present, even if the D peak is not visible because

it is an overtone and therefore always satisfies momentum conservation. The 2D peak is

very important because its shape allows identification of graphene. In graphene, the 2D

peak is a single and narrow peak, in contrast to few-layers and graphite, where the 2D

peak has a more complex lineshape [144] as shown in Figure 1.22.

If graphene contains defects, then additional peaks are observed in the Raman

spectrum as shown in Figure 1.23. The D band, at around ~1350 cm-1, is due to the

breathing modes of sp2 atoms and requires a defect for its activation [145, 146]. The D

peak intensity has been related to the amount of disorder [145]. Another defect-activated

peak is the D’ peak, seen at around ~1620 cm-1. Defects are also responsible for the

appearance of the combination mode at around ~2950 cm-1 [147]. This mode is the

combination of the D and D’ phonons and requires a defect for its activation.

P a g e | 71

Figure 1.23 Raman spectrum of defective graphene showing the main Raman features

taken with a laser excitation energy of 2.41 eV. The Figure is taken from [148].

The ratio between D and D’ bands is very sensitive to the type of defect according

to previous studies [142, 149, 150]. A ratio 𝐼𝐷 𝐼𝐷′⁄ is maximum ( ≅ 13) for defects

associated with sp3 hybridization, it decreases for vacancy-like defects (≅ 7), reaches a

minimum for boundary-like defects in graphite (≅ 3.5) [150].

In the case of LPE graphene, due to the processing involved in the production of

this type of graphene, the Raman spectrum is different from the typical Raman spectrum

shown in Figure 1.22 (where graphene was produced by mechanical exfoliation). The

typical Raman spectrum of LPE graphene is shown in Figure 1.24.

P a g e | 72

Figure 1.24 Typical Raman spectra of liquid-phase exfoliated graphene recorded with

514 nm laser. The Figure is taken from [62].

The Raman spectrum of LPE graphene shows a D peak, which is activated by the

edges of the graphene sheets having the size of the sheets comparable to that of the laser

spot [141]. Furthermore, the Raman spectrum of LPE graphene shows a more complex

2D lineshape [63, 151] because of the LPE processing [62]. Some of the sheets in LPE

graphene will re-stack in random stacking and result in complex 2D peak lineshapes as

observed in a twisted bilayer [152, 153].

A simple method to estimate the amount of graphene in a graphene dispersion

produced by LPE was proposed [62, 151, 154, 155]. The protocol is based on a statistical

analysis of the 2D peak lineshape. In general, the single-layer graphene (SLG), few-layer

graphene (FLG, restacked or retaining AB stacking) and graphitic material (> 7 layers

with AB stacking) can be distinguished by evaluating the coefficient of determination R2

of Lorentzian lineshape fitting. Graphite is easily identified by its two peak shape. The

single-layer can be identified by the narrow symmetrical 2D lineshape with 𝑅2 ≥ 0.987.

P a g e | 73

The restacked FLG is identified when the 2D shape shows a single asymmetric peak with

0.987 > 𝑅2 ≥ 0.985. The 2D shape of FLG (retaining AB stacking) can be distinguished

by the single 2D peak with 𝑅2 < 0.985. The spectra obtained are fitted with a Lorentzian

line to determine 𝐼𝐷

𝐼𝐺⁄ and 𝑅2 (coefficient of determination). Previous results showed

that the qualitative Raman analysis obtained following this protocol was in agreement

with the results obtained by transmission electron microscopy (TEM) [151, 154, 155].

The procedure requires measurement of 20-50 isolated flakes, drop cast on a silicon

substrate. Graphite residuals are excluded from the analysis. The flake must look like a

small and transparent (almost invisible) individual dot.

P a g e | 74

1.6 Aims and objectives

The thesis aim is to investigate the paramagnetism of graphene and its derivative

(i.e. fluorinated graphene) by using EPR spectroscopy. The application of EPR

spectroscopy in graphene characterization is less known compared to Raman

spectroscopy. There are only a few published papers about EPR on graphene (Section

1.4.2). Furthermore, most of the EPR studies have been conducted on poorly

characterized graphene. However, it is of vital importance to correlate the properties of

the materials with the EPR signal in order to get insights about the magnetic properties of

graphene which are crucial for realising its many proposed applications. The objectives

of this project are the following:

1) To study the temperature dependence of the EPR spectrum of well-

characterized graphene and its derivatives

2) To study the EPR signal decay over time of graphene samples.

3) To observe how defects in graphene samples affect the EPR spectrum.

The graphene samples will be made from several production methods (i.e. liquid-

phase exfoliation, electrochemical exfoliation, and graphene made from reduced

graphene oxide), which will allow comparison.

P a g e | 75

2. CHAPTER TWO

Electron Paramagnetic Resonance Study of Graphene

Laminates

2.0 Introduction

Since the first successful isolation of graphene in 2004 at the University of

Manchester [2] many properties have been identified and it is expected to be exploited in

a wide range of applications [5, 38, 156]. Studies on the electron spin that can generate a

magnetic moment in graphene attract a lot of attention due to the possibility of using

graphene in advanced and niche applications [24, 157-159]. Moreover, graphene also

shows high electron mobility [9] and shows some intrinsic spin-orbit interaction and

hyperfine interaction of the electron spins with carbon nuclei [160] that make it attractive

for spintronic devices.

Magnetic properties in graphene are often associated with the interaction of the

localized and itinerant conduction electron spins [14-21]. Paramagnetism in graphene

may arise from active defects such as in-plane vacancy defects/dangling bonds and non-

bonding edge defects [22-24]. Interestingly according to theoretical studies, edge states

consisting of nonbonding π-electrons located at the edge region [11] are often discussed

due to their contribution to the paramagnetic activities of graphene [31-33]. Most of the

studies presented focused on the interaction of magnetic moments within a single layer

of graphene, while only a few discussed the interaction of magnetic moments within

multilayer stacked graphene [161-163]. Other theoretical studies show that magnetic

states existing in the active edge defects might not survive in normal conditions due to

P a g e | 76

self-reconstruction and passivation by other atoms/molecules [28-30]. An active defect

can be introduced by removing a carbon atom via irradiation [19, 164], sonication [64,

65, 70, 71] or by attaching a stable radical [165]. The latter method is appealing because

the method allows for controlling the defect setting but was also inefficient as the method

used a bottom-up approach.

The sonication method also known as liquid-phase exfoliation (LPE) is a well-

known method to produce high-quality graphene in large quantity. However, to identify

the area of the defect induced by sonication remains a challenge [64, 65, 70, 71]. Active

defects are often described as a preliminary condition for the existence of magnetic order

[25-27], although the recent results on the existence of magnetic ordering in graphene

[35, 166-168] raise some doubt because the observed magnetization of graphene was

considered small (0.1-1 emu/g) [22, 35, 168]. Recently, Slota et al. [165] have

demonstrated a method to inject spin density into the edge states of stable graphene

nanoribbons (GNRs) by attachment of nitronyl nitroxide (NIT) radicals. The results

indicated the presence of delocalized spin states at the GNRs edges. The delocalized spins

at the GNRs edges were clearly visible at high-frequency electron paramagnetic

resonance (EPR) bands [Q (~34 GHz) and W (~94 GHz)], while at X (~9.8 GHz) band,

the edge spin signal overlapped with the NIT radicals signal.

A point of reference in the EPR study of graphitic materials is the seminal work

of Wagoner [169] in a study of perfect single crystals of graphite. The EPR line shape

was of the Dysonian form which is characteristic of the presence of conduction electron

spins in metals. The peak heights A/B ratio was 3.0 where A was the distance of the peak

from the baseline in the positive intensity direction and B was the distance of the peak

from the baseline in the negative intensity direction. It was found that the g shift (g factor

minus the free-electron value) and the linewidth are strongly anisotropic. When the

magnetic field was parallel to the graphene planes the g value had a minimum value and

P a g e | 77

was 2.0026; a value slightly higher than the free electron value due to small spin-orbit

coupling with carbon atoms. A strong g shift was measured when the field was

perpendicular to the planes giving a value of 2.05 at room temperature. This was most

often defined in the literature as g = 2.05 and g = 2.0026, the c-axis of the crystal being

perpendicular to the carbon planes. The value anisotropy increased with decreasing

temperature with g = 2.127 at 77 K while g remained constant. The magnitude of the

anisotropy of g depends strongly on temperature and on the position of the Fermi level

with respect to the band edge. An increase of temperature was thought to shift the

population of states from close to the band edge to those further away which have smaller

g shifts [169]. The full theoretical distribution of these observations is complex [170] and

a good summary can be found in the work of Beuneu et al [171].

The magnetic properties of different types of graphene samples, often with very

different properties, have been experimentally investigated [16, 19-22, 28, 34-37], making

it difficult to compare the results. The EPR spectroscopy of graphitic materials has shown

large variations in g values [19, 20, 121], line widths [14, 15, 18, 20] and intensities [37,

127] which was not surprising as the resonance phenomena was very complex depending

on the mobility of the charge carriers and its interplay with the spin-lattice relaxation

time. The dimensions, interlayer interactions, defects, and disorder were all important

factors in governing the electron mobility in each crystal and thereby influencing the EPR

signal.

In this study, graphene flakes were produced using liquid-phase exfoliation [61]

and made into laminates. Defects induced paramagnetism on the flakes were assumed to

be a mixture of edge defects [64, 65] and basal plane defects [70, 71]. The linewidth and

magnetic susceptibility were analyzed in the range 10-295 K by using continuous-wave

EPR (CW EPR) at two sample orientations with respect to the magnetic field.

P a g e | 78

2.1 Sample Preparation

2.1.1 Liquid phase exfoliation graphene laminate

Graphene dispersions were prepared by following a liquid-phase exfoliation

method reported in previous work [61] with modifications. In detail, 3 mg/ml of graphite

(Graphexel Ltd.) was added into 5 ml of N-methyl-2-pyrrolidone (NMP) (Sigma-

Aldrich). The mixtures were bubbled with nitrogen for 1 minute and then sonicated for 6

days in a bath sonicator (Hilsonic, 40 Hz and 600 W). The graphene dispersion (see

Appendix A, Figure A1a) was obtained after centrifugation at 4000 rpm (1180 g) for 60

minutes to remove the unexfoliated flakes. The obtained supernatant liquids (graphene

dispersion) were put inside a sealed glass bottle and then stored in the fridge for later use.

The graphene laminates (Figure A1b) were prepared by filtering the graphene dispersion

using a durapore membrane (from Merck Millipore) which is EPR silent (Figure A10).

Acetone (Sigma-Aldrich) was added into the graphene dispersion before filtrating to

flocculate the graphene flakes. Three different graphene laminates were prepared by

filtering a certain amount of graphene dispersion onto a filter membrane. The filtration

was repeated until the solution was clear. Each sample was prepared on a filter membrane

with an area of around ~ 1.77 cm2 (diameter ~ 1.5 cm) (Figure A1b). The three graphene

laminates were 0.113 mg/cm2, 0.566 mg/cm2 and 1.132 mg/cm2. The increasing amounts

of graphene deposited onto the filter membrane caused the thickness of the graphene

laminates to increase.

2.1.2 Graphite

A graphite flake (Graphexel Ltd.) was attached to the inside of the EPR tube by

using scotch tape. The scotch tape was EPR silent (Appendix Figure A10). The sample

P a g e | 79

was then cut into 2 x 250 mm2 (W x L) pieces. The graphite plane was parallel with the

scotch tape plane.

2.1.3 Electron paramagnetic resonance

The laminates samples for EPR of each thickness type were prepared by cutting

the membranes to ± 2 mm wide, and stacked into ± 11 layers and put into Suprasil EPR

tubes. Around eleven layers were needed to strengthen the signal, so that the EPR

lineshape could easily be observed at room temperature because one layer of graphene

laminate of the same concentration gave a significantly weaker signal (Figure A4). The

spin concentration of the graphene laminate was calculated by using 2,2-diphenyl-1-

picrylhydrazyl (DPPH) as a standard. DPPH is known to have a g value of 2.0036. All

EPR measurements were taken using 2 mW, 1 G modulation amplitude, 10 scans, 40.96

ms time constant and conversion time, under non-saturating conditions. During the

experiments, the graphene laminates were rotated to produce two different orientations.

The z-axis of the samples was positioned 90o (𝐻⊥) and 0o (𝐻∥) with respect to the magnetic

field (H) (see Figure A2). The EPR measurements were performed on a Bruker EMX X

band (~9.4 GHz) spectrometer equipped with a Bruker cryostat and an Oxford

Instruments Cryospares temperature controller.

Pulsed EPR was performed on a Bruker pulsed ELEXSYS E580 (9.7 GHz)

spectrometer equipped with a cryostat and an Oxford Instruments Cryospares temperature

controller. The measurement was carried out at the 𝐻⊥ orientation using ~3480 G centre

field, ~9.7 GHz frequency with pulse lengths of 16 ns for π/2 and 32 ns for π and pulse

delays of t = 300 and τ = 180 ns. The ESEEM experiment used a step size of 16 ns. The

equation to fit the inversion recovery traces to extract T1 was:

𝑦 = 𝑦0 + 𝐴1 ∗ exp (− 𝑥 𝑡1) + 𝐴2⁄ ∗ exp (− 𝑥 𝑡2⁄ ) Equation 2.1

P a g e | 80

𝜏1 = 𝑡1 ∗ 𝑙𝑛(2) Equation 2.2

𝜏2 = 𝑡2 ∗ 𝑙𝑛(2) Equation 2.3

where the fast component was attributed to spectral diffusion (any process that takes spins

resonance [172]) and ignored. The equation to extract T2 value was:

𝑦 = 𝐴1 ∗ exp(− 𝑥 𝑡⁄ ) + 𝐴2 ∗ 𝑥 + 𝐴3 Equation 2.4

𝜏 = 𝑡 ∗ ln (2) Equation 2.5

2.1.4 UV-Vis spectroscopy

The concentration of graphene dispersion was determined by using a Perkin-

Elmer l-900 UV-Vis-NIR spectrometer. The UV spectrum of graphene dispersion tends

to be featureless in the visible-IR region. A Beer-Lambert equation with an absorption

coefficient of 2460 L g−1m−1 was used to calculate the concentration at 660 nm

absorption [61].

2.1.5 Atomic Force Microscopy (AFM)

The average graphene flakes size and shape were analyzed by using a Bruker

Multimode 8 atomic force microscopy (AFM). AFM samples were prepared through

drop-casting the diluted graphene dispersion onto the surface of silicon dioxide. 100

flakes were selected to determine the size distribution. The AFM results were supplied

by Tong Jincheng, School of Chemistry, University of Manchester.

P a g e | 81

2.1.6 Raman spectroscopy

A Renishaw Invia Raman spectrometer was used to determine the quality of the

graphene dispersion. The sample for Raman measurements was prepared by drop-casting

the diluted graphene dispersion onto the surface of silicon dioxide. The measurements

were taken by using a 514.5 nm laser excitation with 1 mW laser power, 100X NA0.85

objective lens and 2400 grooves/mm grating.

2.2 Results and Discussion

2.2.1 Graphene flake characterization

The average graphene dispersion concentration was found to be around 0.8 mg/ml.

The graphene flakes were analyzed by Raman spectroscopy using the procedure described

in previous work [62, 154]. Representative Raman spectra are found in Figure A3. Our

qualitative Raman analysis, based on the shape of 2D peak [62] shows that the dispersion

contained ~40 % single-layer graphene, 50 % few-layer graphene (restacked or retaining

AB stacking) and 3 % graphitic material (>10 layers with AB stacking), confirming high

exfoliation efficiency.

The flakes size and shape shown by the AFM are presented in Figure 2.1. The

average flake size is 58 ± 18 nm with random shapes. The topographic height of the flakes

was in the range of 0.7 – 1.6 nm, which is consistent with the existence of thin layers [2,

173, 174], as found by the qualitative Raman analysis.

P a g e | 82

(a) (b)

Figure 2.1 Graphene flake shapes (a) and size distribution (b), analyzed by using AFM.

2.2.2 Graphene laminate paramagnetism

The graphene laminates had a spin concentration of 2.9333 x 1018 spins/g (

1.5864 x 1017 spins/g) measured at room temperature. The spin concentration was

determined by comparing the double integration of the EPR signal intensity of the

graphene laminates with a known DPPH standard at room temperature [175]. The

measured concentration was two orders of magnitude greater than a previously reported

sample prepared using mechanically exfoliated graphene [14] and the difference is

probably related to the graphene production process. In this study, the relatively long

sonication time may result in a smaller flakes size as well as an increase in the number of

paramagnetic active sites.

20 40 60 80 100 120 1400

10

20

30

40

50

60

Count

Size / nm

P a g e | 83

340 344 348 340 344 348In

tens

ity (a

.u.)

Magnetic Field (mT)

a) b)

340 344 348 352 340 344 348 352

Inte

nsity

(a.u

.)

Magnetic Field (mT)

c) d)

Figure 2.2 9.4 GHz EPR spectra of the thick graphene laminate (1.132 mg/cm2) recorded

at 295 K (a and b) and at 10 K (c and d) at two different orientations; a and c represents

𝐻∥; b and d represents 𝐻⊥. The purple line represents the overall simulation result of the

Lorentzian lineshape; the red and blue line represents Lorentzian lineshape of narrow and

broad component, respectively. The simulation was performed by using Easyspin [176].

The EPR lineshape of the graphene laminates at room temperature was best fitted

using Lorentzian lineshapes (Figure 2.2) indicative of homogeneous contributions to the

linewidth and has been attributed to a strong electron-electron interaction, localized or

delocalized [21]. A composite linewidth was observed for the 𝐻⊥ orientation at room

temperature, which could be simulated with overlapping narrow and broad components.

The A/B ratios were both ~1 for the narrow and broad components at both orientations.

The 𝐻∥ orientation at room temperature also showed two separate EPR signals which

represent a narrow and broad component individually. The intensity of the narrow

a

)

b

)

c

)

d

)

P a g e | 84

component was stronger at 𝐻⊥ orientation. The linewidth narrowed at 10 K for both

orientations and could be simulated with one component. The changes in the spectra were

further understood by the performance of a more detailed temperature-dependent study

as described in Section 2.2.3.

The linewidth of the narrow component at room temperature and 𝐻∥ orientation

was found to increase from 0.45 ± 0.02 mT (0.113 mg/cm2) to 0.53 ± 0.06 mT (1.132

mg/cm2). The linewidth of the narrow component for the 0.566 mg/cm2 samples was

found to be 0.55 ± 0.18 mT. The change appears to reach a limit and this is probably due

to passivation (see Chapter 4 for further discussion on lineshape evolution due to

passivation) and we choose not to interpret the results in detail. Nevertheless, the

linewidth on thin graphene laminate (0.113 mg/cm2) was narrower compared to the thick

graphene laminate (1.132 mg/cm2). This may arise from an interplay of electrons between

layers. The narrow component is affected by the interaction between localized and

conduction electrons between layers which depend on the degree of disorder (i.e. defects

and stacking disorder). Thus, the disorder in thick graphene laminate may disrupt the

conduction electrons mobility and broaden the linewidth.

The linewidth of the graphene laminate’s broad component was found to increase

as the thickness of the layer increases (Figure 2.3). A linewidth of 2.515 ± 0.211 mT was

found for the 0.113 mg/cm2 samples, while the 0.566 mg/cm2 and 1.132 mg/cm2 samples

displayed linewidths of 2.855 ± 0.267 mT and 4.072 ± 0.295 mT, respectively. Graphite

with an increase of disorder also showed similar broadening [177]. In the current case, an

increase of disorder could be associated with an increase in laminate thickness/graphene

stacking. This, again, could be related to the interlayer electron-electron interaction of the

defective and disordered graphene stacking, which could lead to a change in energy bands

and Fermi level position [169, 171, 178].

P a g e | 85

0.0 0.2 0.4 0.6 0.8 1.0 1.2

2.5

3.0

3.5

4.0

4.5

Lin

ew

idth

(m

T)

Graphene laminates (mg/cm2)

Figure 2.3 The room temperature EPR linewidth of graphene laminate’s broad

component at 𝐻∥ orientation on the variation of layer thickness.

2.2.3 Temperature dependence of graphene laminates

The narrow component is considered the main spectral component as it is always

present at a significant intensity. The narrow component had a g value of 2.0031 ± 0.0002.

The deviation of g value from the free-electron value is determined by the contribution

of orbital angular momentum in the electronic states of spins giving rise to resonance.

The g value of the narrow component was temperature-independent and was also in the

range where the mixture of carbon sp2 and sp3-hybridized states has been frequently

observed [128, 179]. This indicates that the EPR signals are originating from carbon-

centred spin states.

The changing of the EPR lineshape as a function of temperature for the 0.566

mg/cm2 sample is shown in Figure 2.4 and for the other samples in Figures A5-A6. At

room temperature, a broad component was observed at both orientations. As the

temperature decreased, the broad component’s intensity weakened and shifted downfield

P a g e | 86

away from the narrow component and eventually became unobservable around 50 – 75

K. The shifting of the broad component for 𝐻∥ orientation from room temperature to 75

K was in the range of 0.8 - 2.6 mT. At 𝐻∥ and room temperature, the broad component

had a g value of 2.0151 ± 0.0006 for the 0.113 mg/cm2 sample, 2.0172 ± 0.0053 for the

0.566 mg/cm2 sample and 2.0183 ± 0.0031 for the 1.132 mg/cm2 sample.

338 340 342 344 346 348 350 352

10 K

20 K

30 K

40 K

50 K

75 K

100 K

150 K

200 K

250 K

295 K

Inte

nsity (

a.u

.)

Magnetic Field (mT)

Figure 2.4 The EPR spectra of thin graphene laminates (0.566 mg/cm2) as a function of

temperature. The black line represents 𝐻⊥. The red line represents 𝐻∥.

The g value of the broad component and its temperature dependence appeared to

be affected by the thickness of the graphene laminates/stacking of graphene planes. This

characteristic has also been found in graphite [177] with an increasing amount of disorder

in graphite causing an increase of the g value as the temperature is lowered. The

temperature dependence of the g value for graphite and nanographites has also been

P a g e | 87

reported to be affected by the mobility of electrons between graphene planes; the g value

increased as the temperature decreases [129, 169].

The intensity of the broad component decreased proportionally with temperature

and disappeared at around 75 - 100 K. To the best of our knowledge, this has not been

observed before in any graphene or graphite experiments. In addition, the temperature

limit at which the broad components started to disappear was found to get lower as the

graphene laminates got thicker. For example, for the thin graphene laminate (0.113

mg/cm2) the broad component disappeared at ~100 K and for the medium thickness

laminate (1.132 mg/cm2) the broad component disappeared at ~ 75 K. Although the broad

component has never been observed and discussed in previous graphene literature, similar

behaviour has been observed in the case of nanographite [129, 180]. They attributed the

broad component to be as the result of itinerant conduction electron spins, while the

narrow component was representative of localized electron spins [129, 180]. However,

the appearance of broad components was absent in other nano graphite samples [128].

2.2.4 EPR linewidth of graphene laminates

The interaction between localized and itinerant conduction electron spins can be

observed through changes in the EPR linewidth. The coupling of the localized and

itinerant conduction electron spins can cause the average EPR linewidth to narrow as the

temperature decreases. The narrowing behaviour has been reported on nano graphite [128,

181] and graphene [20, 119]. The EPR linewidth of graphitic materials has also been

found to broaden as the temperature decreases [14, 20, 128, 181]. The broadening

behaviour was thought due to decoupling of the localized and conduction electrons [20].

The narrow component of graphene laminate is affected by coupled and decoupled states

of localized and conduction electrons. Figures 2.5 and 2.6 show plots of the EPR

P a g e | 88

linewidth of the narrow component against temperature for different graphene laminates

thickness and orientation. Error bars arise from the average of three samples. The

linewidth from 100 K – 10 K for 0.113 mg/cm2 samples for both sample orientations

showed only a slight narrowing while the linewidths for 0.566 mg/cm2 samples practically

remains unchanged (Figures 2.5 and A7). However, the low-temperature EPR linewidth

broadening can be observed in the thicker graphene laminate (1.132 mg/cm2) samples at

the 𝐻∥ orientation (Figure 2.6). Interestingly, this result suggests that the coupling and

decoupling of localized and itinerant conduction electron spins observed at 𝐻∥ orientation

might involve out-of-plane inter-graphene layers interactions. The out-of-plane interlayer

magnetic interactions arising from the disorder are much stronger in the thick graphene

laminates (high number of graphene layers stacked) with high localized spin

concentration leading to a change in the energy bands and shifting of the Fermi level [171,

177]. In addition, the linewidth may also be affected by spin-lattice relaxation [181].

0 50 100 150 200 250 300

0.32

0.36

0.40

0.44

0.48

0.52

Lin

ew

idth

(m

T)

Temperature (K)

a)

0 50 100 150 200 250 300

0.4

0.8

1.2

1.6

Lin

ew

idth

(m

T)

Temperature (K)

b)

Figure 2.5 EPR linewidth of the narrow component for the thin graphene laminates

(0.113 mg/cm2) on the variation of temperature. a) Red square represents 𝐻∥. b) Black

square represents 𝐻⊥.

P a g e | 89

0 50 100 150 200 250 3000.35

0.40

0.45

0.50

0.55

0.60

0.65

Lin

ew

idth

(m

T)

Temperature (K)

a)

0 50 100 150 200 250 300

0.4

0.8

1.2

1.6

2.0

Lin

ew

idth

(m

T)

Temperature (K)

b)

Figure 2.6 EPR linewidth of the narrow component for the thick graphene laminates

(1.132 mg/cm2) on the variation of temperature. a) Red square represents 𝐻∥. b) Black

square represents 𝐻⊥.

The linewidths at the 𝐻⊥ orientation from 300 – 150 K was affected by the

presence of the broad component overlapping with the narrow component due to the

anisotropic behaviour of the broad component. Thus, the EPR peak-peak linewidth was

difficult to measure. Thus, we choose not to interpret the data from 300 – 150 K.

In contrast to the temperature dependence of the linewidth for the narrow

component, the linewidth of the broad component for all samples at 𝐻∥ was seen to

broaden as the temperature decreases. This was observed as the intensity of the broad

component decreased with decreasing temperature. The broadening of the linewidth and

the decreasing of the intensity was possibly due to the limitation of electron mobility as

the temperature decreases and the displacement of Fermi level position [129, 169].

2.2.5 Comparison to graphite

The comparison was made to a single flake of graphite in order to help understand

the lineshape of the graphene laminates and the influence of itinerant conduction

electrons. Figure 2.7 displays the EPR spectra of graphite recorded at different

P a g e | 90

temperatures, showing the influence of conduction electrons. The graphite flake showed

Dysonian lineshapes at room temperature with an A B⁄ (∥) ratio of 2.77 and A B⁄ () ratio

of 1.75. Similarities and differences at equivalent temperatures were found in comparison

to the laminates:

i) The resonances at the 𝐻∥ orientation shifted to a lower magnetic field and the intensity

was reduced and finally disappeared.

ii) The resonance along the 𝐻∥ orientation also came into resonance at a much lower

magnetic field.

Figure 2.7 EPR spectra of a graphite flake as a function of temperature. (a) represents

𝐻⊥. (b) represents 𝐻∥. (*) marks a speculate asignment of the broad component at 70 K.

P a g e | 91

The g value of the resonances at the 𝐻∥ orientation for the graphite flake was

2.0419 ± 0.0046 while the laminates have g values in the range of 0.0236-0.0268 less than

the graphite (see Section 2.2.3). The magnitude of the g value of graphite at 𝐻∥ orientation

has been reported to depend strongly on temperature and on the position of the Fermi

level with respect to the band edge [169]. Further, the large g-shift in graphite along the

𝐻∥ orientation has been explained due to the change in the energy bands (degeneracy) of

graphite at the zone edge. The g value anisotropy of the graphite is affected by the

stacking of graphene planes and the mobility of electrons between planes [129, 177].

Accordingly, the graphene laminates were prepared by using vacuum filtration and

according to a previous investigation, they consist of numerous graphene flakes which

are randomly stacked [182]. The loss of AB stacking in the graphene laminate may be

explained due to the change in the degeneracy of the energy bands at the zone edge and

increase of electron mobility [183] causing the decrease of the g value [129].

The linewidth of graphite’s resonance along the 𝐻∥ orientation (Figure A8) was

narrower (∆Hpp = 1.29 mT, at room temperature) compared to the laminates (see Section

2.2.4, Figure 2.6). The alignment of the AB Bernal system in the graphite may be

responsible for the narrower EPR linewidth [177]. Chehab et al demonstrated that an

increase of stacking disorder in graphite broadened the linewidth of graphite observed for

𝐻∥ [177]. The linewidth of graphite’s resonance at 𝐻⊥ orientation and at room

temperature was 0.54 mT (Figure A8).

Studying further Figure 2.7, it clearly shows that the EPR intensity of the graphite

flake at both orientations decreased proportionally with temperature whereas the g value

is constant for 𝐻⊥orientation and increases for 𝐻∥ orientation. Figure 2.7b shows that at

𝐻∥, the resonance can be seen to disappear at ~70 K. The spectra clearly show the change

and movement of the signal. This is similar to the observation in the laminates albeit at a

P a g e | 92

different temperature range. To the best of our knowledge, this observation has never

been reported before. Previously, Matsubara et al [170] was able to observe the g value

of highly oriented pyrolytic graphite (HOPG), synthetic graphite, down to ~5 K at 𝐻∥.

The g value increased as the temperature decreased and exhibited a peak at 20 K and then

decreased. The EPR intensity was shown to decrease parallel with the temperature down

to 20 K and then increase. The decrease of g value and the increase of EPR intensity after

20 K was thought to arise due to the contribution of localized spins which become

dominant at low temperatures in accordance with the Curie law. However, in our graphite

flake sample, although the EPR intensity of the resonance at 𝐻∥ was reduced and even

completely unobservable below 70 K, we did not observe a narrow component associated

with localized spins. This could be because the localized spin concentration in our

graphite was probably too small and may be suppressed by the π character of the Fermi-

energy states and therefore the narrow component was unobservable. Interestingly, the

intensity behaviour measured by double integration of our graphite flake at 𝐻⊥ shows

similar behaviour as observed by Matsubara et al. at 𝐻∥ such that the intensity decreases

and reaches a minimum at around 30 K and then increases (Figure A9).

2.2.6 EPR magnetic susceptibility of graphene laminates

The EPR magnetic susceptibility measurements from 10-70 K of graphene

laminates showed anisotropic Curie-Weis behaviour (Figures 2.8-2.10). The equation

used to fit the Curie-Weis behaviour was:

𝜒𝐸𝑃𝑅 =𝐶

T−𝜃 Equation 2.6

where C is the Curie constant, T is the temperature of the sample and θ is the critical

temperature. Both antiferromagnetism and ferromagnetism were observed in the

P a g e | 93

graphene laminate samples as judged by fitting of the Curie-Weis equation to the inverse

susceptibility (using double integration of the whole spectrum) below 75 K. Above 75 K

where the broad component exists, the Pauli contribution tends to be the dominant state

due to the EPR line intensity behaviour being dictated by electron-electron interactions

[181].

The thin graphene laminates (0.113 mg/cm2) showed negligible

antiferromagnetism with θ = -0.2 ± 2.5 K close to zero at 𝐻⊥ (Figure 2.8).

Ferromagnetism was observed at 𝐻∥ with θ = 5.2 ± 2.5 K. As the graphene laminate got

thicker, the Curie-Weis temperature changed. The 0.566 mg/cm2 graphene laminate

sample gave θ = -5.4 ± 2.6 K and 5.9 ± 2.1 K, while the 1.132 mg/cm2 graphene laminate

sample displayed θ = -11.5 ± 2.1 K and 9.2 ± 1.9 K for 𝐻⊥ and 𝐻∥ , respectively (Figures

2.9 and 2.10).

The ferromagnetic interaction coexisting with the antiferromagnetic interaction in

graphene has been reported previously in a study which points out that flake size and the

number of layers in the graphene sample can influence the magnetic properties [184]. In

the present work, since the size of the flakes of the three graphene laminates was assumed

to be similar due to the same production method, it is likely that the thickness/the number

of layers stacked should give the most influence to differences in the magnetic properties.

Indeed, Figures 2.8-2.10 clearly reveal that the occurrence of anti/ferromagnetism and it

can be observed clearly as the laminates thickness increases. The Curie-Weis temperature

moved away from 0 K as the laminates thickness increases. As the graphene laminate

thickness increases, the stacking disorder increases. This variable could also impart the

change of magnetism [27, 161].

P a g e | 94

0 10 20 30 40 50 60 70 80

1

2

EP

R-1

(10

-5)

Temperature (K)

1

2

3

(10

-5)

Figure 2.8 Curie-Weis behaviour of thin graphene laminates (0.113 mg/cm2) measured

from 10-75 K. Black dots represents 𝐻⊥, θ = -0.2 ± 2.5 K. Red dots represents 𝐻∥, θ = 5.2

± 2.5 K. Blue line represents the Curie-Weis line.

1

2

3

4

5

0 10 20 30 40 50 60 70 80

0.5

1.0

1.5

(10

-5)

E

PR

-1(1

0-5

)

Temperature (K)

Figure 2.9 Curie-Weis behaviour of thick graphene laminates (0.566 mg/cm2) measured

from 10-75 K. Black dots represents 𝐻⊥, θ = -5.4 ± 2.6 K. Red dots represents 𝐻∥, θ = 5.9

± 2.1 K. Blue line represents the Curie-Weis line.

P a g e | 95

0 10 20 30 40 50 60 70 80

2

4

6

8

1

2

3

E

PR

-1(1

0-6

)

Temperature (K)

(10

-5)

Figure 2.10 Curie-Weis behaviour of thick graphene laminates (1.132 mg/cm2) measured

from 10-75 K. Black dots represents 𝐻⊥, θ = -11.5 ± 2.1 K. Red dots represents 𝐻∥, θ =

9.2 ± 1.9 K. Blue line represents the Curie-Weis line.

A study on graphene magnetism reveals that the alignment of the spins in

graphene might depend on the distribution of the spin population within graphene

sublattices and stacking disorder [162]. In addition, the antiferromagnetic behaviour in

graphene has been discussed previously [15, 19]. The exchange interaction between

localized states and conduction electrons might be responsible for such behaviour.

Another theoretical study on stacked graphene, realizing the energy difference between

different stack sequences, predicted that the net magnetic ordering was due to lower total

energy causing stable conditions to allow magnetic moment alignment [163]. Figures 2.8-

2.10 showed that the ferromagnetism in the graphene laminate could be observed at 𝐻∥,

while the antiferromagnetism appeared at 𝐻⊥ . The results suggest energetically

favourable orientation-dependence of the interlayer interactions of localized spins and

indirectly conduction electron spins. According to theoretical studies, the spins at zig-zag

edges (edge states) are strongly polarized and are coupled through ferromagnetic

P a g e | 96

interactions [16, 37]. However, the discussion of the ferromagnetic source has to be

limited because of the existence of zig-zag edges within the graphene laminates cannot

be quantitatively verified. The percentage of the zig-zag edges within the samples can be

determined statistically by observing several numbers of individual flakes. However, the

method is time-consuming and would require high-resolution AFM or Transmission

Electron Microscopy (TEM).

2.2.7 Relaxation times and nuclear resonances of the graphene laminates

Pulsed EPR experiments were used to measure relaxation rates of the graphene

laminates (Figures 2.11). The spin-spin relaxation time (T2) and spin-lattice relaxation

time (T1) of the graphene laminates were observed over a temperature range of 10 – 70 K

(Figure 2.11) which showed a weak temperature dependence. The T2 was practically

unchanged and fluctuated between 0.9 – 1.2 µs over the temperature measurement.

Similarly, the T1 values showed little variation with a weak trend for decreasing values at

a lower temperature. For example, T1 varied from ~20 µs at 10 K to ~15 µs at 70 K. This

is consistent with relaxation playing only a minor role in the changes of line width.

1.0 1.2 1.4 1.6 1.8

5

6

7

Lo

g(1

/T2 s

-1)

Log(T (K))

a)

1.0 1.2 1.4 1.6 1.8

4

5

6

Lo

g(1

/T1 s

-1)

Log(T (K))

b)

Figure 2.11 The spin-spin relaxation time (T2) (a) and spin-lattice relaxation time (T1) (b)

of a graphene laminate over the temperature range of 10 – 70 K at the 𝐻⊥ orientation.

P a g e | 97

Electron spin echo envelope modulation (ESEEM) experiments were performed

to analyse the presence of nuclear-electron spin coupling. Figure 2.12 shows the

identification of nuclear magnetic resonances arising from 13C and possibly for 14N. The

signal to noise (S/N) ratio was low to confirm the 14N resonance, nevertheless, the

detected resonance was possibly from the residual NMP solvent. The magnetic coupling

can be a source of relaxation.

0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30

0

1

2

3

4

ES

EE

M A

mp

l. x

10

7(a

.u.)

Frequency (MHz)

14N

13C

Figure 2.12 ESEEM spectrum of graphene laminate at 10 K at the 𝐻⊥ orientation.

2.3 Conclusion

The experiments have shown that localized and itinerant conduction electron spins

exist in graphene laminates and their interactions are affected by spin concentration,

temperature, the order of stacking and the magnetic field orientation. Broad and narrow

components were observed in the CW EPR spectra. The broad component of graphene

laminate was affected by the interlayer coupling of localized and itinerant conduction

electron spins. The g value and linewidth of the broad component was affected by electron

P a g e | 98

mobility, the laminate’s thickness, disorder, temperature and orientation in the external

magnetic field. The most likely origin of the broad component is from interlayer

interactions of localized and itinerant conduction electrons. The narrow component was

attributed to localized electrons (vacancy defects). The g value of the narrow component

was assigned to carbon centred spins. The linewidth of the narrow component was

affected by the interlayer interaction between localized and itinerant conduction electrons

as well as spin-lattice relaxation. This was particularly observed through the increase of

linewidth at low temperature for the thick graphene laminate revealing the decoupled and

coupled states of the localized-itinerant conduction electron spins with high localized spin

concentration for the narrow component and 𝐻∥ orientation. The ferromagnetic and

antiferromagnetic interactions present in the graphene laminates may be related to

energetically favourable conditions due to interlayer interaction of localized and

indirectly conduction electron spins within the magnetic field orientation.

P a g e | 99

3. CHAPTER THREE

Electron Paramagnetic Resonance Study of the

Electrochemical Exfoliation of Graphite in Comparison

to Graphene Laminates Produced Through

Electrochemical Exfoliation, Liquid Phase Exfoliation

and Chemical Reduction of Graphene Oxide

3.0 Introduction

Low-cost mass production of solution-processable and high-quality graphene

remains a major challenge. Several mass production methods have been developed [185].

Chemical vapour deposition (CVD) has been able to produce high-quality graphene films

with large-area [186]. However, production variables such as high-temperature, a

sacrificial metal catalyst, and the multistep transfer onto the desired substrates could

become obstacles to the cost-effective mass-production scale of CVD graphene. Chemical

or thermal reduction of graphene oxide (GO) is a well-known and cost-effective method

for the mass production of graphene but suffers from the use of toxic reducing agents

such as hydrazine or sodium borohydride. Moreover, the reduced graphene oxide (rGO)

produced was found to only partially restore the electronic properties of graphene [187].

Recently, the use of environmentally friendly reducing agents has become an appealing

topic for the generation of rGO [188-190]. Liquid phase exfoliation (LPE) is another

attractive efficient method to produce graphene. Exfoliation can be achieved by

sonicating graphite either in solvents or in surfactant/stabilizer-aqueous media. The

graphene dispersion obtained was a mixture of single and multilayer graphene. Cascade

P a g e | 100

centrifugation could maximize the selection to obtain a single layer graphene enriched

dispersion [191].

Electrochemical exfoliation of graphite is a fast and easy top-down method to

produce graphene in large quantities [192-196]. The exfoliation follows either anodic

oxidation or cathodic reduction. Anodic exfoliation is the most commonly used method

due to the relatively low-cost electrolyte solution and fast intercalation process compared

to cathodic exfoliation. The intercalation at the cathode is usually only able to expand the

graphite; sonication was needed to fully exfoliate the graphite [84, 197-199]. However,

the anodic exfoliation approach was not without problems. The graphene produced

through anodic exfoliation method was partially oxidised or functionalised by other

species involved in the process [73]. The anodic exfoliation mechanism has been

explained in Section 1.3.4 and can be summarized as the following: 1) the positively

charge graphite foil anode attracts anions (negative ions), 2) oxidation of graphene,

intercalation of graphene layers by negative ions, and the formation of gaseous species,

3) expansion and exfoliation of graphene layers [76].

Previously, Krivenko et al [200] demonstrated the use of electron paramagnetic

resonance (EPR) spectroscopy, infrared spectroscopy (IR), X-ray photoelectron

spectroscopy (XPS) and scanning electron microscopy (SEM) to analyse the graphene

powder produced by electrochemical exfoliation of graphite rods. The results suggested

that the powder produced through anode exfoliation had a significantly higher number of

defects compared to the powder obtained in the cathode. The present report further

highlights the use of EPR spectroscopy to study the electrochemical exfoliation of

graphite foil. In addition, this Chapter studies the magnetic properties of electrochemical

exfoliated (EC) graphene laminate with comparison to rGO and LPE laminates. The EPR

spectra generated from the laminates were able to separate and characterize the

contribution of the localized and conduction electrons within the samples.

P a g e | 101

3.1 Sample Preparation

3.1.1 Liquid Phase Exfoliation

Graphene dispersions and laminates were prepared and characterised using the

methods described in Chapter 2, Section 2.1.1.

3.1.2 Electrochemical Exfoliation

The graphene solution was prepared as followed. The graphite foil (Sigma-

Aldrich) was cut into ± 1.5 (W) x 3 (L) cm2 strips and was used as the electrodes. A 0.5

M solution of K2SO4 (Sigma-Aldrich) in water was used as the electrolyte. The

exfoliation was carried out by applying 10 V at room temperature for 5 minutes. After

exfoliation, the material was collected and washed with deionized water via vacuum

filtration. The filtered material was then dispersed in NMP (100 mL) using ultrasonication

for 30 minutes. The graphene solution was supplied by Khaled Parvez, School of

Chemistry, University of Manchester.

The graphene solution was sonicated for 6 days. The sample was then centrifuged

at 3500 rpm for 20 min. The electrochemical exfoliated (EC) graphene laminate was

prepared by filtering the graphene dispersion using a Durapore membrane. Acetone was

added into the graphene dispersion before filtrating to flocculate the graphene flakes. The

filtration was repeated until the solution was clear. The sample was prepared on a filter

membrane with an area of around ~ 1.77 cm2 (diameter ~ 1.5 cm). The graphene

laminate had 1.132 mg/cm2 of graphene.

To allow the study of the electrochemical exfoliation process, the sample was

prepared using the following procedure. The graphite foil (Sigma-Aldrich) was cut into ±

2.5 (W) x 100 (L) mm2 strips and was used as the electrodes. A 0.5 M solution of K2SO4

P a g e | 102

(Sigma-Aldrich) in water was used as the electrolyte. The exfoliation carried out by

applying 10 V at room temperature for 30 seconds. The electrodes were then washed by

using deionised water followed by an acetone wash and then dried at room temperature.

Thereafter, the electrodes were analysed immediately by using EPR and Raman

Spectroscopy.

3.1.3 Reduced Graphene Oxide

The reduced graphene oxide (rGO) was prepared by using the chemical reduction

process in a pressurized vessel. A mixture of hydrazine monohydrate (0.02 mol) (Sigma-

Aldrich) and graphene oxide (GO) suspension (2 mg/mL, 30 mL) (Sigma-Aldrich)

solution was sonicated for half-an-hour. Then the mixture was transferred to a Teflon-

coated autoclave and heated for 12 h at 180 °C. The product was washed with acetone

and water via filtration. Finally, the powder was obtained by freeze-drying the sample.

The obtained rGO powder was dissolved in isopropanol/water mixture via sonication for

1 h. The 1.132 mg/cm2 of rGO laminate was prepared by vacuum filtration using

Durapore membranes. The rGO laminates were supplied by Bin Wang, School of

Chemistry, University of Manchester.

3.1.4 Electron Paramagnetic Resonance (EPR) Spectroscopy

All samples for EPR were prepared by cutting the membranes to ± 2 mm wide,

followed by stacking up to ± 11 layers into Suprasil tubes. Around eleven layers were

needed so that the EPR lineshape could easily be observed at room temperature as

explained in Chapter 2, Section 2.1.3. All measurements were run on a Bruker EMX

(equipped with a Bruker Cryostat and an Oxford Instruments Cryospares temperature

P a g e | 103

controller) and a micro EMX at X band (~9.4 - 9.8 GHz) using 2 mW, 1 G modulation

amplitude, 10-20 scans, 40.96 ms time constant and conversion time, under non-

saturating conditions. During the experiments, the graphene laminates were rotated to get

two different orientations as described in Chapter 2 (Section 2.1.3). The g value and spin

concentration of the graphene laminates were calculated by using 2,2-diphenyl-1-

picrylhydrazyl (DPPH) as a standard.

3.1.5 Atomic Force Microscopy (AFM)

The average graphene flakes size and shape were analyzed by using a Bruker

Multimode 8 atomic force microscopy (AFM). AFM measurements were performed

according to the method described in Chapter 2 section 2.1.5. The AFM measurement and

analysis was supplied by Tong Jincheng, School of Chemistry, University of Manchester.

3.1.6 Raman Spectroscopy

A Renishaw Invia Raman spectrometer was used to determine the quality of the

graphene laminate. The sample for Raman measurements was prepared by placing the

laminate sheet onto the surface of a glass microscope preparative slide. The measurements

were taken by using a 514.5 nm laser excitation with 1 mW laser power, 100X NA0.85

objective lens and 2400 grooves/mm grating.

P a g e | 104

3.2 Results and Discussion

3.2.1 Observation of Electrochemical Exfoliated Graphite by Electron

Paramagnetic Resonance and Raman Spectroscopy

The electrochemical reaction involves reduction at the cathode and oxidation at

the anode. During the experiments, graphite foil was used for both the anode and cathodes.

Figure 3.1 shows the graphite foil at the cathode and anode after the electrochemical

process. The graphite foil at the anode was exfoliated as expected while the cathode

graphite foil showed a damaged surface. During the electrochemical exfoliation, gaseous

species were formed on both anode and cathode. According to previous reports, several

gases are released during the electrochemical exfoliation process such as SO2, O2 and H2

and are essential to the intercalation process [73, 201].

(a) (b)

Figure 3.1 The anode (a) and the cathode (b) after 30 seconds of the electrochemical

exfoliation process.

The EPR signal of the graphite foil of the anode after 30 seconds of the

electrochemical process shows a dramatic narrowing at both sample orientations (Figure

3.2) compared to the initial spectra. Figure 3.3 displays how the EPR signal became more

isotropic with the clear presence of a broad and intense narrow component. The presence

of the narrow component with a strong intensity indicates an increase in the number of

localized electron spins. The anodic spectra show similarity to the graphene laminates

P a g e | 105

samples discussed in Chapter 2. Initially, the graphite foil had a g value = 2.0031 ± 0.0004

with a peak-peak linewidth (𝐼𝑃𝑃) = 3.08 ± 0.22 mT for 𝐻⊥ and a g value = 2.0196 ± 0.0011

with 𝐼𝑃𝑃 = 5.21 ± 0.32 mT for 𝐻∥. After 30 seconds of the electrochemical process, the

graphite foil had a g value of the narrow component = 2.0029 ± 0.0001 with 𝐼𝑃𝑃 = 0.46 ±

0.02 mT and a g value of the broad component = 2.0020 ± 0.0008 with 𝐼𝑃𝑃 = 3.06 ± 0.14

mT for both orientation. Thus, the EPR spectra of graphite foil are consistent with vacancy

defect creation, as well as layer expansion due to intercalation within the graphite foil

which disrupt the interplay between layers and reduces the anisotropy. The results are in

agreement with the exfoliation mechanism proposed by [76] that stated the process

involves graphene oxidations, anion intercalation and expansion of graphite foil.

336 340 344 348 352

No

rma

lize

d I

nte

nsity (

a.u

.)

Magnetic Field (mT)

Figure 3.2 EPR spectra of the anode graphite foil before and after 30 seconds of the

electrochemical exfoliation process. The solid and dash lines represent the EPR spectra

before and after electrochemical exfoliation, respectively. The black and red colours

represent the EPR spectra at 𝐻⊥ and 𝐻∥ orientations, respectively.

P a g e | 106

336 340 344 348 352

Inte

nsity (

a.u

.)

Magnetic Field (mT)

Figure 3.3 The EPR spectrum of the anode graphite foil after 30 seconds of

electrochemical exfoliation process at the 𝐻⊥ orientation (solid black line). The green

dash line represents the Lorentzian line of the broad component; the blue dash line

represents the Lorentzian line of the narrow component. The solid purple line represents

the overall simulation result. The simulation was performed by using Easyspin [176].

Structural defects were confirmed by Raman spectroscopy shown in Figure 3.4.

The intensities of D (~1360 cm-1) and 2D’ (~3120 cm-1) bands increases after 30 seconds

of the electrochemical process. The D band was activated by short-range defects such as

vacancies and edges as the wavevector of the phonon which gives rise to the D band is

relatively long [149]. Although, the 2D’ band originates from the overtone (second-order

harmonic) of the phonon mode that gives rise to the D’ band and did not require defects

for momentum conservation [144], nevertheless, the 2D’ band was escalated after 30

seconds of electrochemical process. The D’ (1620 cm-1) band was present after the

electrochemical process which could indicate the occurrence of long-range defects such

as charged impurities adsorbed on the graphene sheet [149].

P a g e | 107

2600280030001400 1600

Inte

nsity (

a.u

.)

Raman Shift (cm-1)

before

after

Figure 3.4 Raman spectra of anode graphite foil before (black) and after (red) 30 seconds

of electrochemical exfoliation.

The cathode, after the electrochemical exfoliation process, showed defects on the

surface (Figure 3.1b). The EPR signal showed no significant changes at the 𝐻⊥

orientation. However, at the 𝐻∥ orientation, the EPR signal narrowed by 0.84 ± 0.18 mT

and the magnetic field resonance moved to a higher magnetic field at the constant

frequency by 0.37 ± 0.04 mT equivalent to the g value shifting towards the free electron

by 0.00072 ± 0.00011 (Figure 3.5). The differences indicate a change in the energy bands

and Fermi level [169, 178] and an increase in electron mobility [129]. These observations

lead to speculation that the graphite foil at the cathode was slightly expanded after the

electrochemical process due to cation intercalation and gaseous species formation. The

observation was in agreement with the proposed cathode exfoliation mechanism and the

production of a graphite intercalated compound [84, 197-199]. The narrow component

remains unobservable after the electrochemical process indicating that there was no

significant increase in the number of vacancy defects.

P a g e | 108

336 340 344 348 352

No

rma

lize

d I

nte

nsity (

a.u

.)

Magnetic Field (mT)

Figure 3.5 EPR spectra of cathode graphite foil before and after 30 seconds of the

electrochemical exfoliation process. The solid and dash lines represent the EPR spectra

before and after electrochemical exfoliation respectively. The black and red colours

represent the EPR spectra at 𝐻⊥ and 𝐻∥ orientations, respectively.

The Raman measurement after the electrochemical process showed a small

increase of D and 2D’ peaks which indicates significantly fewer defects compared to the

anode (Figure 3.6). Thus, was in agreement with the appearance of the graphite foil after

the electrochemical process shown in Figure 3.1.

P a g e | 109

1400 1600 2600 2800 3000

after

before

Inte

nsity (

a.u

.)

Raman Shift (cm-1)

Figure 3.6 Raman spectra of cathode graphite foil before (black) and after 30 seconds

(red) of electrochemical exfoliation.

3.2.2 Graphene flakes characterization

Figure 3.7 shows the AFM images of LPE graphene, EC graphene and rGO. The

LPE graphene and EC graphene flakes were nearly equal in size which is around 50-90

nm, while the rGO flakes size was around 1-2 µm. The result is likely due to the long

duration of sonication treatment of the LPE graphene solution and the EC graphene

solution leading to similar average flakes sizes. The topographic height of the LPE

graphene flakes was in the range of 0.7 – 1.6 nm, which shows the existence of thin layers

of graphene [2, 173, 174]. The EC graphene and rGO flakes have a thickness of around

1-3 nm.

P a g e | 110

1 m 1 m

(a) (b) (c)

Figure 3.7 The AFM images of (a) EC graphene, (b) rGO and (c) LPE graphene (taken

from Chapter 2).

3.2.3 Defect-induced paramagnetism

At room temperature, the LPE and EC graphene laminates both show two-

component EPR lineshapes; the presence of narrow and broad components. The rGO

graphene laminate, however, shows a single isotropic EPR lineshape which represents the

narrow component (Figure 3.8). As discussed in Chapter 2, the occurrence of the broad

component is thought to be due to the interlayer interactions of localized and itinerant

conduction electrons. The g value of the broad component is temperature-dependent and

affected by the thickness of the graphene laminates or stacking of graphene planes

(Chapter 2). The narrow component of LPE graphene laminate has a g value of 2.0031 ±

0.0002 (Chapter 2), while the narrow component of EC graphene laminate shows a g

value of 2.0035 ± 0.0001 and the rGO laminate has a g value of 2.0037 ± 0.0002. All of

the g values of the laminates are close to the free-electron value and are also in the range

where the mixture of carbon 𝑠𝑝2 and 𝑠𝑝3 -hybridized states have been frequently

observed [128, 179]. This indicates that the EPR signals are originating from carbon-

centred states.

P a g e | 111

340 344 348

Inte

nsity (

a.u

.)

Magnetic Field (mT)

200 250 300 350 400 450

Inte

nsity (

a.u

.)

Magnetic field (mT)

340 344 348 352

Inte

nsity (

a.u

.)

Magnetic Field (mT) (a) (b) (c)

Figure 3.8 EPR spectra at room temperature of the LPE graphene laminate (a), EC

graphene laminate (b) and rGO laminate (c). The black and red lines represent 𝐻⊥

orientation and 𝐻∥ orientation, respectively. The samples were 1.132 mg/cm2 of graphene

laminates.

The most striking differences are the relatively large line width of the broad

component from the EC graphene laminate compared to LPE graphene laminate (see

Figure 3.8) and the absence of a significant broad component observed from the rGO

laminate. The broad component of the EC graphene laminate has a linewidth of 63.64 ±

5.3 mT and there is no significant difference between the 𝐻∥ and 𝐻⊥ orientations. The

broad component linewidth for 1.132 mg/cm2 LPE graphene laminate was 4.07 ± 0.3 mT

at 𝐻∥ orientation (Chapter 2).

The Raman spectra of the laminates shown in Figure 3.9 show the broadened

peaks of D (~1360 cm-1), G (~1560 cm-1) and 2D (~2700 cm-1) bands for EC graphene

and rGO laminates which indicate a relatively increased amount of disorder compared to

the LPE graphene laminate. The indication was further strengthened by the relative

increase of intensity of the D’ (~1620 cm-1) band, the combination D+D’ (~2970 cm-1)

and 2D’ (~3120 cm-1) bands for EC graphene and rGO laminates (compared to LPE

graphene laminate). The D band intensity for a single layer of graphene has been related

to the amount of disorder [145]. However, for graphene laminates the D band intensity

P a g e | 112

could not be used to determine the amount of disorder, the reason was because of the

more and less damaged graphene areas could co-exist within the Raman probe causing a

significant error [202]. An increase in the D’ band (1620 cm-1) could indicate an increase

of long-range defects such as charged impurities adsorbed on the graphene sheet [149].

Figure 3.9 also shows that the G band of rGO band is blue-shifted, indicating that the

incident light gains energy after interacting with the vibrational modes of the graphene

and may be due to defects. It has been reported on single-layer graphene that the shifting

of G and 2D bands can be caused by dopants, strain or defects [202-207].

1200 1400 1600 2400 2600 2800 3000 3200

Inte

nsity (

a.u

)

Raman Shift (cm-1)

LPE graphene

laminate

EC graphene

laminate

RGO

laminate

Figure 3.9 Raman spectrum of rGO laminate (red), EC graphene laminate (blue) and LPE

graphene laminate (black).

The number of defects in graphene laminates could affect the EPR linewidth of

the broad component. An increased number of defects in laminates increases the amount

of disorder in graphene stacking. It has been shown in Chapter 2 that as the laminate’s

thickness increases, the broad component EPR linewidth increases, and thereby correlates

with the increase of graphene stacking disorder. An increase of defects and stacking

disorder could change the energy band and shift the Fermi level and therefore change the

P a g e | 113

linewidth of the broader component observed [169, 171, 177, 178]. Overall, the

complementary EPR and Raman experiments (Figures 3.8 - 3.9) suggest that the EPR

linewidth of the broad component broadens as the number of defects increases, which is

in agreement with Chapter 2 and Chehab et al. [177].

The rGO laminate showing, in contrast, a single narrow EPR lineshape may be

explained by the presence of a severely damaged structure causing the conduction

electron spins interactions within the rGO laminate to be confined or minimized within a

few layers. The annealing treatment of rGO may remove the oxygen species

functionalizations but cannot sufficiently repair the defects (i.e. missing carbon atoms,

holes in carbon network) [208, 209] consistent with the Raman result shown in Figure

3.9. Another possibility is that the defects increase the amount of disorder and cause the

EPR linewidth of the broad component to broaden and become unobservable.

Chapter 2 showed that the broad component of LPE graphene laminate behaves

similarly to the broad component of graphite: The g value of the broad component at 𝐻∥

is greater than the g value at 𝐻⊥. However, the g value of the broad component on EC

graphene laminate shows a dramatic change, the g value at 𝐻∥ is smaller than the g value

at 𝐻⊥. The result is not consistent with the results reported in [177] for graphite showing

that as the amount of disorder increases, and the g value anisotropy observed at 𝐻∥

increases. Figures 3.8 and 3.9 show that an increase of defects/stacking disorder does not

necessarily increase the g value of the broad component observed at the 𝐻∥ orientation,

however, it increases the broad component linewidth. The defects found within EC

graphene layers are most likely greater than the defects within the graphene layers of LPE

graphene laminate. The structural defects could have a big influence on the interaction of

spins between layers and could change the degeneracy of energy bands, shift the Fermi

level [177, 178]. According to Barbon et al. [129], the EPR bands of stacked graphene

P a g e | 114

were complex resonance phenomena involving charge carriers mobility and its interplay

with the spin-lattice relaxation time; the electrons mobility will be affected by flakes

dimension, interlayer interaction and disorder. The EPR spectra of EC graphene laminates

suggests that the electron-electron interaction of the broad component could show

massive changes due to defects and functionalization. Graphene produced from anode

exfoliation is known to be easily exposed to oxidation and functionalization [73]. Khaled

et al. [76] have investigated the graphene produced by using electrochemical exfoliation

in an aqueous solution of inorganic salt (specifically the salts used were (NH4)2SO4,

Na2SO4 and K2SO4). They observed the powder X-ray diffraction of exfoliated graphene

showed a slightly lower 2θ angle with large d-spacing compared to graphite indicating a

small amount of functional groups. The ultraviolet photoelectron spectra of exfoliated

graphene have shown a higher work function than pristine graphene due to

functionalization possibly by oxygen-containing functional groups [76]. In addition, X-

ray photoelectron spectroscopy analysis also detected the presence of approximately

5.5 % oxygen content [76].

3.2.4 Temperature dependence

Chapter 2 presented and discussed the temperature dependence of LPE graphene

laminates. The LPE graphene laminates showed the coupling and decoupling of localized

and itinerant conduction electron spins. The changes of coupled between states can be

observed through the variation of the temperature dependence of EPR linewidth.

Moreover, pronounced changes of the decoupling between localized and itinerant

conduction electron spins were observed at the 𝐻∥ orientation for the thick graphene

laminate (1.132 mg/cm2). The coupling of the localized and itinerant conduction electron

spins can cause the average EPR linewidth to narrow [20, 128]. However, decoupling of

the two can cause the average of the EPR linewidth to increase as the temperature

P a g e | 115

decreases [20]. In addition, EPR linewidth could also be affected by spin-lattice relaxation

[181].

Upon observing the temperature dependence of the linewidth for EC graphene and

rGO laminates, the results were different compared to the LPE graphene laminate. The

EPR linewidth of EC graphene and rGO laminates were temperature-dependent but were

not affected by sample orientation within the external magnetic field (Figures 3.10-3.11).

Figure 3.10 shows that the EPR linewidth of the EC graphene laminate increases as the

temperature decreases for both sample orientations which is consistent with a decoupling

state of localized and itinerant conduction electrons. The observations suggest that the

hexagonal graphene structure on the EC graphene laminate could have been altered due

to the production process (defects and functionalizations). The structure modification

increases the amount of disorder in the laminate (again, through defects and

functionalizations) which leads to the disruption of the conduction electron interaction

between layers, and results in a reduction of the exchange narrowing. In contrast, Figure

3.11 shows that rGO’s linewidth weakly decreases as the temperature is lowered and

increases at temperatures lower than 50 K. This behaviour has been observed before by

Diamantopoulou et al. [210] and attributed to a narrowing mechanism which effectively

reduces dipolar broadening and other inhomogeneous contributions to the resonance

width such g-anisotropy and unresolved hyperfine splitting, similar to -type defects

associated with sp2 clusters in amorphous carbon. The increase of linewidth at the lowest

temperatures could be attributed to the presence of a high concentration of spins.

Overall, the more isotropic with orientation behaviour suggests a chaotic

interlayer interaction similarly found in polycrystalline graphite [178] and glassy graphite

[211]. Again, the chemical and annealing treatment of rGO occurring during sample

preparation (Section 3.1.3) may have removed the functionalizations on the graphene

plane [208] causing defects and disorder. Similarly, functionalizations on graphene plane

P a g e | 116

potentially present in the EC graphene laminate may be associated with the disruption of

the interplay between localized and conduction electrons.

0 50 100 150 200 250 300

0.3

0.4

0.5

0.6

0.7

0.8

Lin

ew

idth

(m

T)

Temperature (K)

Figure 3.10. EPR linewidth of the EC graphene laminate on the variation of temperature.

Black dot represents 𝐻⊥. Red dot represents 𝐻∥.

0 50 100 150 200 250 300

0.52

0.54

0.56

0.58

0.60

0.62

0.64

Lin

ew

idth

(m

T)

Temperature (K)

Figure 3.11. EPR linewidth of the rGO laminate on the variation of temperature. Black

dot represents 𝐻⊥. Red dot represents 𝐻∥.

P a g e | 117

Magnetic susceptibility was judged by plotting 𝜒𝐸𝑃𝑅−1 vs temperature (Figures

3.12-3.13) as described in Chapter 2, Section 2.2.6. The magnetic susceptibility

behaviours were observed in the graphene laminate samples as judged by fitting the

Curie-Weis equation to the inverse susceptibility plots below 75 K. Above 75 K, where

the broad component exists, the Pauli contribution has been found to be significant due

to the dominant influence of conduction electrons [181].

Magnetic susceptibility can be affected by flake size, the number of layers in the

graphene sample [184] and the stacking disorder [161, 162]. The LPE graphene laminate

showed anisotropic behaviour of magnetic susceptibility, ferromagnetism and

antiferromagnetism co-existed within the laminate and were affected by sample

orientation in the external magnetic field (Chapter 2). The EC graphene laminate showed

antiferromagnetic behaviour with θ = -10.6 ± 2 K and -20 ± 1.1 K for both 𝐻⊥ and 𝐻∥

orientations, respectively, while the rGO laminate showed ferromagnetic behaviour with

θ = 7.3 ± 3.4 K and 7.4 ± 3.6 K for 𝐻⊥ and 𝐻∥ orientations, respectively. The magnetic

properties of graphene laminates are the result of the interplay between the electron-

electron interaction within individual graphene layers and between graphene layers [27,

161, 162, 184]. Based on theoretical studies, electron spins within a graphene layer

populating the quasi-localized states in the same sublattice are thought to be oriented

parallel to each other while the antiparallel arrangement is realized when electron spins

populate different sublattices [27, 162]. Thus, the net magnetic moment between

sublattices determines the magnetism behaviour in graphene. The stacking disorder of

graphene layers, as evident in our experiments, could increase the complexity making it

difficult to fully understand the magnetic moment of the laminates [161, 162]. It has been

predicted theoretically on a single layer of graphene that the zigzag edge spin’s interaction

formed so-called edge states that could align ferromagnetically or antiferromagnetically

[163, 212]. Another study on single-layer graphene stated that the spin interaction at the

P a g e | 118

zigzag edge can be controlled by an external electrical field and that the edge states are

spin-polarized [213]. In stacked graphene, the net magnetic ordering has been thought to

be affected by the energy difference between the different stacking sequences, local

magnetic moments and π state contribution to the edge magnetic moment with the lower

total energy giving the most stable form of magnetic ordering [163].

0 10 20 30 40 50 60 70 800 10 20 30 40 50 60 70 80

5

10

15

20

25

E

PR

-1 (

a.u

.)

Temperature (K)Temperature (K)

Figure 3.12. Curie-Weis behaviour of EC graphene laminate measured from 10-75 K.

Black dot represents data at 𝐻⊥. Red dot represents data at 𝐻∥. Black line represents the

Curie-Weis line fit for 𝐻⊥, θ = -10.6 ± 2 K. Red line represents the Curie-Weis line fit for

𝐻∥, θ = -20 ± 1.1 K.

P a g e | 119

0 10 20 30 40 50 60 70 80

0.04

0.08

0.12

0 10 20 30 40 50 60 70 80

E

PR

-1 (

a.u

.)

Temperature (K)

Figure 3.13. Curie-Weis behaviour of rGO laminate measured from 10-75 K. Black dot

represents data at 𝐻⊥. Red dot represents data at 𝐻∥. Black line represents the Curie-Weis

line fit for 𝐻⊥, θ = 7.3 ± 3.4 K. Red line represents the Curie-Weis line fit for 𝐻∥, θ = 7.4

± 3.6 K.

3.3 Conclusion

The EPR and Raman spectroscopic analysis on the anode and cathode graphite

foils showed the presence of defects and expansion. The results were in line with the

current mechanistic understanding of electrochemically prepared graphene. The

contribution of localized and conduction electrons for each type of graphene laminate

were identified and characterized. A narrow component was present in all samples and

associated with localized defects. The g value of the laminate’s narrow component was

2.0031 ± 0.0002 (LPE graphene laminate), 2.0035 ± 0.0001 (EC graphene laminate) and

2.0037 ± .0002 (rGO laminate). At room temperature, the EC graphene laminate has a

wide broad component while for the rGO laminate the broad component was

unobservable, contrasting with the appearance of the broad component on LPE graphene

P a g e | 120

laminate. Complementary Raman experiments showed the presence of defects, and

together with the EPR observations suggested that defects affected the interlayer electron-

electron interaction. Thus, the coupled and decoupled states of localized and itinerant

conduction electrons were influenced by defects and functionalizations in the different

material laminates. In addition, the intrinsic magnetic properties of EC and rGO laminate

could be derived from temperature-dependent studies and showed the influence of defects

and disorder.

P a g e | 121

4. CHAPTER FOUR

Paramagnetic Stability and Defect Creation of

Graphene Laminates Under Controlled Conditions and

Action of Laser

4.0 Introduction

Graphene has a featureless spectral response over a wide wavelength range of

electromagnetic absorption [13, 214-216]. The irradiation of graphene such as by using

laser, electron beam and ion bombardment can not only be used to create defects [217-

221] but can also be used for a wide range of applications: removal of oxygen-containing

groups [222], modification of structure [202], control of radical species [164] and

improvement of the regeneration of graphene-based sensors [223]. Thus, the light

irradiation of graphene has been characterized to some extent. Previous studies have used

Raman spectroscopy as a characterization tool.

Laser irradiation or ion bombardment on graphene can generate radicals or

vacancy defects [19, 219, 221, 224, 225]. These type of defects can modify the magnetic

properties of graphene as it is known that paramagnetism in graphene may come from

active defects (i.e. in-plane vacancy defects/dangling bonds, non-bonding edge defects)

[22-24]. Paramagnetic defects in graphene materials are often highly reactive and

promptly react with oxygen or hydrogen functional groups to create non-paramagnetic

groups. Kausteklis et al. [128] have demonstrated that ball-milled graphite powder stored

under inert conditions for several months immediately lost an order of magnitude of

paramagnetism after the graphite powder was exposed to air. Moreover, the electron

paramagnetic resonance (EPR) spectroscopic analysis of the graphite powder after

P a g e | 122

exposure to air showed relatively broad peak-to-peak linewidths compared to before

exposure. Other studies also report that vacancy defects were prone to self-reconstruction

and passivation by other atoms/molecules [28-30]. Paramagnetism in graphene was

strongly related to the interaction of localized and itinerant conduction electron spins [14-

21]. The interaction of the localized and itinerant conduction electron spins was evident

through two EPR components [129, 180].

Previous studies have generated defects using ion bombardment [19, 221] and

various types of the laser such as femtosecond [218, 220, 224-226], electron beam [219,

222] and nanosecond [217]. The last study investigated the impact of nanosecond pulsed

laser flashes at 532 and 266 nm on chemical vapour deposition (CVD) graphene. The

report stated the damage threshold of ~200 mJ/cm2 and found a strong fluorescence signal

from damaged areas due to the residues of oxidized graphene. The only study to our

knowledge to have used EPR spectroscopy was a study on ion bombardment [19]. The

report analysed the effect of ion bombardment applied to CVD graphene on Si substrate

in vacuum conditions. A vacancy defect was created after ion fluence of 10/nm2 and the

temperature-dependent measurements revealed antiferromagnetic correlations with a

Curie-Weiss temperature of -10 K. The EPR signal was found to significantly broaden

and decrease in intensity after the sample was exposed to air.

The present Chapter studies paramagnetic stability and defects induced by

nanosecond pulsed laser irradiation on graphene laminate samples. EPR spectroscopy

will be used to analyse any changes to the paramagnetism. Raman spectroscopy will

analyse the potential creation of defects.

P a g e | 123

4.1 Sample Preparation

4.1.1 Liquid phase exfoliation graphene laminate

Graphene dispersions were prepared by following a liquid-phase exfoliation

described in Chapter 2 (Section 2.1.1). The graphene laminates were prepared by filtering

the freshly made graphene dispersion using the method described in Section 2.1.1. The

laminates had 1.132 mg/cm2 of graphene.

4.1.2 Paramagnetic stability experiment using EPR

Four graphene laminates were used for the experiment. The sample was prepared

by cutting each laminate to ± 2 mm wide. The graphene laminate sheets were placed inside

four Suprasil EPR tubes. Two sample tubes were placed under vacuum overnight and

after that were filled with an argon atmosphere. The other two sample tubes were stored

under normal atmospheric condition. All four sample tubes were measured for EPR

activity typically every 6 days. The EPR measurements were performed on a Bruker

micro EMX X band (~9.4 GHz) spectrometer. The EPR measurements were carried out

by placing the z-axis of the samples 0o (𝐻∥) with respect to the magnetic field (H). All

EPR measurements were taken with 2 mW, 1 G modulation amplitude, 20 scans, 40.96

ms time constant and conversion time, under non-saturating conditions.

4.1.3 The aged graphene laminate experiment

The sample was prepared by cutting the laminate and a blank durapore membrane

to ± 2 mm wide. The laminate was stored under normal atmospheric conditions for several

months or until the EPR signal reduced significantly. After confirmation that the sample

P a g e | 124

only had a trace of EPR signal, one graphene laminate sheet was placed inside a Suprasil

EPR tube using 3-4 blank durapore membrane sheets in order to vertically stand to hold

the laminate. Afterwards, the EPR tube was placed under vacuum overnight and then was

filled with an argon atmosphere. The EPR tube was then placed into a sample holder in

front of the laser. The distance between the sample and the laser source was ± 25 cm

(Figure 4.1). The laser model was Radiant 355 LD made by OPOTEK and contained an

optical parametric oscillator (OPO) to allow tunable wavelengths from 210 to 2500 nm,

energy from 0 to 11.5 mJ, a repetition rate of 10 Hz, a pulse width of 5 ns and a beam

diameter of 6 mm. The laser wavelength was set for 270 nm and could generate 3.5 mJ

of power. The irradiation was carried out for 180 seconds, while the EPR measurement

was made every 30 seconds. The EPR measurements were performed as described in

Section 4.1.2 but with 40 scans. During the EPR measurements, the graphene laminate

was rotated to produce two different orientations as previously described in Chapter 2,

Section 2.1.3. The laser fluence (F) at the sample was estimated from the power (P) of the

laser and the beam area (a) inflicted on the sample (~ 2 x 6 mm2) with F = P/a.

Figure 4.1 (a) The experiment setup for the aged graphene laminate irradiation. (b) The

sample tube after 180 seconds of irradiation. (*) Marks the sample tube.

P a g e | 125

4.1.4 In-situ defect creation experiment studied using EPR

The sample for the in-situ experiment was prepared by cutting the laminate and a

blank durapore membrane to ± 2 mm wide. One graphene laminate sheet and 3-4 blank

durapore membrane sheets were placed inside a Suprasil tube allowing the graphene

laminate to vertically stand. The sample tube was then placed under vacuum overnight

and then filled with an argon atmosphere. For the experiments, the sample tube was placed

inside the EPR cavity. The distance of the laser beam source was ± 150 cm from the cavity.

During the irradiation, not all of the laser beam hit the sample because the cavity grid

partially obscured the laser beam, which made it impossible to calculate the laser fluence

and significantly reduced the effectiveness of the laser power. The Radiant 355 LD laser

was used as described in Section 4.1.3. The laser was set for three different wavelengths,

which were 270 nm, 660 nm and 800 nm. The laser could generate a strength of ± 3.5 mJ,

± 8.7 mJ and ± 6 mJ for each wavelength respectively. During the irradiation, the z-axis

of the samples faced the laser beam or the z-axis of the samples was positioned 90o (𝐻⊥)

with respect to the magnetic field. The irradiations were carried out at 100 K for 30

minutes and the EPR measurements were also performed at 100 K. The EPR

measurements were carried out by rotating the z-axis of the samples 0o (𝐻∥) with respect

to the magnetic field (H). The EPR measurements were performed as described in Section

4.1.2. The EPR instrument used were a Bruker EMX Plus X band (~9.4 GHz)

spectrometer equipped with Bruker cryostat and Oxford Instruments Cryospares

temperature controller.

P a g e | 126

4.1.5 Raman Spectroscopy

A Renishaw Invia Raman spectrometer was used to determine the quality of the

graphene laminate. The Raman measurements of the laminate were performed by using

the method described in Section 3.1.6.

4.1.6 X-ray Photoelectron Spectroscopy (XPS)

An Axis Ultra Hybrid (Kratos Analytical) was used to determine the chemical

composition of the graphene laminates and graphite. It used an Al K_alpha X-ray source

(1486.6 eV, 10 mA emission with 15 kV bias). Experiments typically happen at pressures

below 3E-8 mbar. A charge neutralizer is usually used to remove any effects of differential

charging. The samples were a ± 1 mm wide aged graphene laminate and a graphite flake.

The XPS data was supplied by Khaled Parvez, School of Chemistry, University of

Manchester.

4.2 Results and Discussion

4.2.1 Paramagnetic Stability

The disappearance of the EPR signal arising from graphene laminates could

indicate the possibility of passivation of the localized paramagnetic sites within the

samples [28-30] and provide an important understanding of this process. As the localized

spins could be diminished by passivation (bonding with other atoms/molecules or self-

reconstruction), the paramagnetic properties may alter.

Investigation on an aged graphene laminate in the normal atmosphere using ultra-

high vacuum XPS showed that the oxygen content had increased (Table 4.1). Comparison

P a g e | 127

to a graphite sample stored in oxygen showed a greater oxygen content in the laminate

sample, consistent with the greater surface area of the laminate reacting with oxygen. It

is also possible that the oxygen content shown by XPS in the laminates comes from

surface adsorption [227]. The XPS result suggests the possibility of passivation caused

by bonding with oxygen. However, outside from the possibility of oxygen passivation, a

self-reconstruction mechanism should be taken into consideration as well [29, 30]. There

could be a competition between the two mechanisms.

Elements Graphite

(%)

Graphene

laminate as

prepared (%)

Graphene laminate

after 38 days

(%)

C 93.78 88.35 83.7

O 6.22 7.29 12.09

N 0.84 0.52

F 3.52 3.69

5.

Table 4.1 Ultra-high vacuum XPS on graphite and graphene laminates (1.132 mg/cm2).

Figure 4.2 displays a single isotropic EPR lineshape comes from an aged graphene

laminate stored under normal atmospheric conditions for several months. The signal was

very weak and appeared to show only a single lineshape which represents the narrow

component. The graphene signal was observable although the contribution from the

background signals i.e. the EPR tube and the cavity was not small. The linewidth was

0.48 ± 0.02 mT, which was narrower than the 0.98 ± 0.08 mT measured for the narrow

component found from unaged graphene laminates at 𝐻∥ , while the g value 2.0026 ±

0.0002. For comparison, the g value of the narrow component measured on the unaged

graphene laminate was 2.0033 ± 0.0001. The broad component could not be observed and

may not be present although the signal-to-noise is very weak. The observed changes are

P a g e | 128

probably due to the lack of spin concentration which causes an increase of the distance

between spins and a restriction of the interlayer spin interactions. The reduction in spin

concentration most likely arises from the decrease in localized electron spins generated

from vacancy defects due to passivation [28-30]. In line with the XPS results above, in

this case, oxygen appears to be playing a role in the reduction of paramagnetic centres.

338 340 342 344 346 348 350 352

Inte

nsity (

a.u

)

Magnetic Field (mT)

Figure 4.2 EPR spectrum at room temperature of an aged graphene laminate. The black

and red signals represent the 𝐻⊥ and 𝐻∥ orientations, respectively. The blue line

represents the EPR background from the EPR tube and cavity.

It has been discussed in Chapters 2 and 3 that the presence of a broad EPR

component was associated with the interlayer interaction and sp2 electrons. Nevertheless,

Figure 4.2 suggests that the interlayer interaction on disordered stacked graphene such as

laminates may require an adequate amount of localized electrons to inject the spin into

the sp2 graphene backbone and trigger the effect. In order to investigate further, four

graphene laminate samples were prepared (Figure 4.3). Two samples were stored under

normal atmospheric conditions and two samples were stored under argon. The samples

were stored for 60 days and measured periodically (typically every 6 days). The

P a g e | 129

measurements were taken at the 𝐻∥ orientation to observe the narrow and broad

components, individually.

6.

338 340 342 344 346 348 350338 340 342 344 346 348 350

c1)

d1)

a1)

b1)

d)

b)

Inte

nsity (

a.u

.)

a)

Magnetic Field (mT)

c)

Figure 4.3 EPR lineshape evolution at room temperature and 𝐻∥ the orientation of

graphene laminate samples stored throughout 60 days. Samples stored under normal

atmospheric conditions (a and b); samples stored under argon (c and d). Time zero spectra

are shown in black. Spectra recorded at increasing duration are lighter in colour (black to

red to yellow). The a1, b1, c1 and d1 represent the lineshape at time zero (black) and

lineshape at 60th day (bright yellow).

P a g e | 130

The total EPR lineshape showed dramatic changes. Figure 4.3 shows that the EPR

lineshape changes over time with significant variation of the narrow component for both

samples stored in argon and in normal atmospheric conditions. Throughout the

observation time window, the intensity of the narrow component decreases and the

linewidth of the narrow component broadens for all samples. The narrow component for

the samples stored in argon broadens by 0.684 ± 0.0376 mT over the duration of the

experiment. The narrow component for the samples stored under normal atmospheric

conditions showed varying rates of broadening (Figure 4.3a-b). Throughout the

observation time, the narrow component of sample A (Figures 4.3a and a1) was broadened

by 0.3796 mT which was close to the broadening of sample B at day 24. At the end of the

observation time window, the narrow component of sample B (Figures 4.3b and b1) was

narrowed by 0.749 mT compared to the initial linewidth. Thus, there are differences in

the rates of linewidth decrease for the atmospheric samples even though they have

experienced the same apparent conditions. This suggests that the samples are very

sensitive and further experiments would be needed to isolate the difference between two

seemingly “like” samples.

Clearly, under both conditions, the narrow component decreases in intensity and

increases in linewidth as a function of time. The narrow component is associated with

localized electron spins [129, 180]. The decrease in intensity and the changing of the

linewidth could be attributed to passivation of vacancy defects (self-reconstruction or

bonding with other atoms) making the sample more diamagnetic. The increase of

linewidth may be associated with an increase of hyperfine/g value strain and disorder.

The g value of the narrow component for sample A, C and D remains unchanged

and is within the error of the measured g values reported in Chapter 2. The g value of the

narrow component for sample B remains the same until the 24th day of observation time.

Afterwards, the g value shifted toward lower values and on the last day, the g value was

P a g e | 131

2.0024. The shifting of g value toward lower values may be due to the weakening of

interlayer interactions and conduction electron influence through the potential restoration

of the sp2 framework. The broad component of sample A shows minor changes throughout

the experiment, while in sample B, the broad component exhibited a linewidth broadening,

g value shift toward lower fields and a decrease of intensity. At time zero up to 18 days,

the g value was practically unchanged and remains near a value of ~2.0150. On the 24th

day, the lineshape exhibited a dramatic change, which shows the narrow component

intensity greatly reduced and the broad component’s g value shifted to 2.0168. On the

following observation days, the broad component exhibited a linewidth broadening, a

reduced intensity and a shifted g value. A g value of 2.0189 was recorded at day 30th and

g value near ~2.0205 was recorded for rest of the observation days.

The samples stored in normal atmospheric conditions showed varying results for

the broad component with little decease of intensity for sample A and almost vanishing

intensity for sample B. This makes drawing conclusions difficult and further experiments

would be needed to elucidate the true trend. Interestingly, the broad components for the

argon samples showed only mild decreases over the time course of the experiment (Figure

4.3c-d). Thus, the spin population that the broad component represents remains

unchanged whereas the narrow component significantly decreases. This indicates that the

spin population giving rise to the broad component is almost unaltered by passivation.

Intriguingly, it suggests there is a population of spin that remain undisturbed by

passivation. However, the spin decay in sample B could give an important clue that at a

certain degree of passivation, in this case after the 24th day, the intensity of the broad

component starts to decrease along with the narrow component. This led to speculation

that the broad and narrow component can have the same or interrelated mechanism of

passivation

P a g e | 132

Figure 4.4 shows the passivation rate comparing the two different environments.

It indicates that samples stored in argon have a slower passivation rate. This is to be

expected because of the inert conditions. The only passivation path, in this case, would

be the self-reconstruction. The samples stored in normal atmospheric conditions possibly

follow two passivation paths which are the self-reconstruction and reaction with other

atoms (most likely with oxygen). The vacancy defects in graphene are known to be

sensitive to air exposure [19, 28-30, 128]. The spins concentration of both samples A and

B fell relatively at the same rate until the 24th day. This can be observed, outside of the

error bars, in Figure 4.4b. After that, the spins in sample B decay much faster than sample

A and the variability is reflected in an increase of the error bars. At the end of the

observation time window, sample A retains 67 % of the original spins while sample B

only had 15 % of the original spins.

0 10 20 30 40 50 60

20

40

60

80

100

Sp

in C

on

ce

ntr

atio

n (

%)

Time (days)

0.0 0.5 1.0 1.5 2.0

0.5

0.6

0.7

0.8

0.0 0.5 1.0 1.5 2.0

Spin

Con

ce

ntr

ation

(1

/log

(S %

))

Time (log (t) days)

Figure 4.4 The evolution of mean total spin concentration throughout storage time. a)

normal spin concentration vs time. b) 1/log (spin concentration) vs log time. The blue dot

represents samples stored in normal atmospheric conditions; the red dot represents

samples stored in argon.

a)

b)

P a g e | 133

4.2.2 Defect creation on an aged graphene laminate sample

The sample used for the experiment was an aged graphene laminate stored under

the normal atmospheric conditions for several months. Initially, the sample showed a

typical graphene laminate EPR signal with two EPR signals (narrow and broad

components) that can be easily identified at the 𝐻∥ orientation. However, after the ageing

process, the EPR signal was significantly reduced, the graphene laminate showed only a

weak EPR resonance as shown in Figure 4.5. In this experiment at 295 K, the laser beam

at 270 nm hit the laminate on ~ 2 x 6 mm2 of the area and gave a laser fluence of 29.2

mJ/cm2.

Figure 4.5 EPR spectra evolution of an aged graphene laminate after irradiation at 270

nm at room temperature. (a) The sample is positioned (𝐻⊥). (b) The sample is positioned

(𝐻∥).

P a g e | 134

Figure 4.5 shows a single isotropic EPR lineshape before irradiation for both

sample orientations. There was only observable a single lineshape which represents a

narrow component. The broad component could not be clearly discerned. After the

irradiation at 270 nm, the spin concentration started to increase. There was no significant

change of lineshape or formation of the broad component. The most likely cause of the

increase of signal is from breaking of 𝑠𝑝2 carbon-carbon bonds and radical formation

[164]. Again, the lack of broad component to the EPR signal may be related to the reduced

number of spins and thereby reducing interactions such as interlayer spin coupling. It is

of note that the laser beam was concentrated on one specific area and was not spread

throughout the entire graphene layers and thus the EPR signal must show the presence of

more than one environment (laser exposure and non-exposure).

The experiment was stopped after 150 seconds of irradiation time and the final

EPR intensity was increased by 10-12 times the original intensity. The peak-peak

linewidth (𝐼𝑃𝑃) for both orientations after 150 seconds of irradiation was ~0.8 mT which

was ~0.3 mT wider than the 𝐼𝑃𝑃 before the irradiation. Increasing the irradiation time did

not cause any further changes to the EPR signal. After the experiment, it was observed

that a small amount of carbon species, assumed to be amorphous carbon, were sticking

onto the EPR tube wall. This suggests that during the irradiation, the degradation

mechanism involves either material ablation or formation of disordered carbonaceous

species/amorphous carbon molecules [225]. This may supply a background to the

graphene laminate EPR signal.

P a g e | 135

1200 1400 1600 2600 2800

No

rma

lize

d I

nte

nsity (

a.u

.)

Raman Shift (cm-1)

before

after

Figure 4.6 Raman spectrum of graphene laminate before and after ultraviolet irradiation.

Raman spectroscopy of the laminate before and after irradiation on the affected

area provides supporting evidence of defect creation in graphene. Figure 4.6 shows the

broadening of D (~1354 cm-1) and G (~1581 cm-1) peak after irradiation indicating defects

on the structure [202]. In this experiment, however, the D peak intensity which should

indicate defects seems not to be affected. In this case, as discussed in Chapter 3 in the

case of graphene laminate, the reason could be because of the coexistence of more and

less damaged graphene areas within the Raman probe [202]. There is no Raman frequency

shift of the D and G peak. However, a blue shift occurs at the 2D (~2698 cm-1) peak,

shifting the 2D peak after the irradiation to a higher wavenumber by 20 cm-1. It has been

reported for graphene single layer – few layers that the shifting of G and 2D peaks can be

caused by dopants, strain or defects [202-207]. Further on a single layer graphene the n-

type doping and uniaxial strain most likely cause red-shifted 2D peak phenomena, while

the p-type doping could cause blue-shifted 2D peak phenomena [203]. In our experiment,

there are no dopants added to the sample. The reason for the blue shifting of 2D peak after

P a g e | 136

irradiation may be due to the breaking of 𝑠𝑝2 carbon-carbon bonds lead to the formation

of vacancy defects.

The complementary EPR and Raman spectroscopy indicate that the graphene

indeed receives damage from the laser light irradiation. However, the pulsed laser fluence

of 29.2 mJ/cm2 used in this study was far less than 250 mJ/cm2 theoretically predicted

damage threshold for a graphitic film ablation [226] or 200 mJ/cm2 damage threshold

measured for graphene ablation [217, 224]. This may be the reason why multiple

exposures were needed to degrade the graphene lattice by using 29.2 mJ/cm2 of pulsed

laser fluence for 150 seconds.

4.2.3 In-situ defect creation by irradiation at 270 nm, 660 nm and 800 nm

Graphene is known to have a wide spectral range of electromagnetic absorption

[13, 214-216]. In these experiments, three different laser wavelengths were chosen. The

three wavelengths were 270 nm, 660 nm and 800 nm. The EPR measurement was

performed at 100 K and at the 𝐻∥ sample orientation to observe both narrow and broad

components individually. It has already been described in Chapter 2 that the temperature-

dependent experiments show that the broad component is observable from room

temperature down to ~75 K and at the 𝐻∥ sample orientation.

P a g e | 137

340 344 348 340 344 348 340 344 348

b) d)

Magnetic Field (mT)

Norm

aliz

ed I

nte

nsity (

a.u

.)

270 nm

a) c)

660 nm

e)

800 nm

f)

Figure 4.7 Continuous-wave EPR spectrum of graphene laminate at 100 K and 𝐻∥

orientation. (a-b) 270 nm laser wavelength irradiation, (c-d) 660 nm laser wavelength

irradiation and (e-f) 800 nm laser wavelength irradiation. The black line represents the

graphene laminate before irradiation while the blue, green and red signal represents the

graphene laminate after 30 minutes of the irradiation.

The irradiation time was longer by 30 minutes compared to the aged irradiation

experiment in order to counter the laser power restriction because of the cavity grid of the

microwave resonator used in these low-temperature experiments. For each experiment,

two samples were prepared. Figure 4.7 shows the EPR spectra before and after irradiation.

The two Lorentzian lineshapes which represent narrow and broad components can be

observed clearly. A noticeable difference between the EPR lineshape before and after

irradiation is shown by the broad component, whereas only a mild change can be observed

for the narrow component. The peak-peak EPR intensity (𝐼𝑝𝑝) of the broad component

decreases in signal height and the EPR linewidth broadens after the irradiation. Also, the

broad component initially had an A/B ratio of ~1 and did not change after the irradiation.

The broad component linewidth was broadened by 0.13 ± 0.099 mT at 270 nm irradiation,

P a g e | 138

0.14 ± 0.031 mT at 660 nm irradiation and 0.68 ± 0.089 mT at 800 nm irradiation. The

broad component which has been discussed in Chapter 2 is believed to be representative

of the interaction between localized and itinerant conduction electron spins between

layers of graphene. It has been discussed in Chapter 2 and 3 that the broadening of the

broad component is linked to an increasing amount of disorder in graphene stacking

leading to changes in the electron-electron interaction.

Figure 4.8 displays the double integration area of the total EPR signal of the

samples after irradiation at different wavelengths. The 270 nm irradiation generates

~0.9 % more spins while the 800 nm irradiation generates 6.2 % spins more spins. Thus,

although the signal height diminishes, the actual double integral increases. The irradiation

could generate radicals due to the breaking of 𝑠𝑝2carbon-carbon bonds. The formation of

radicals was to be expected as the graphene laminate absorb the electromagnetic

irradiation at these wavelengths and can cause localized heating and eventual damage of

the structure by rapidly ionizing the graphene [225]. The result was also in agreement

with [164] which showed that graphene oxide (GO) in deionized water generated more

electron spins after UV irradiation. In addition, according to [164], the radicals formed

after UV irradiation were air-stable over a long period of time for both GO in the solution

state and in freeze-dried powders.

Interestingly, irradiation at 660 nm showed a decrease in electron spins by ~8.7 %

which indicates passivation. The passivation could be caused by energetically favourable

self-reconstruction. A previous report by ean et al. [29] shows that under electron beam

irradiation and in the absence of metals, healing could occur via reconstruction of the

hexagon structure.

P a g e | 139

270 660 800

-0.05

0.00

0.05

Are

a (

a.u

.)

Laser wavelength (nm)

Figure 4.8 Double integration area of the EPR line at 100 K and 𝐻∥ orientation. Positive

area means that the sample gains more electron spins after the irradiation; negative area

means that the sample loses electron spins after irradiation.

The irradiations did not significantly affect the narrow component. The amount of

defects/disorder, in this case, is suspected to be small compared to the aged graphene

laminate or to the electrochemically exfoliated graphene and reduced graphene oxide

samples discussed in Chapter 3. Moreover, it has to be realised that the effectiveness of

laser power was greatly reduced as it only focused on a small part of the laminate.

P a g e | 140

1200 1400 1600 1800 2600 2800

No

rma

lize

d I

nte

nsity (

a.u

.)

Raman Shift

270 nm

660 nm

800 nm

before

Figure 4.9 Raman spectrum of graphene laminate before (black line) and after 30 minutes

of irradiation using a 270 nm laser (blue line), 660 nm laser (green line) and 800 nm laser

wavelengths (red line).

The Raman spectroscopy shown in Figure 4.9 shows that the D (~1351 cm-1) and

G (~1583 cm-1) peaks remain unchanged. There is no noticeable frequency shift or

broadening of the D and G peaks. As discussed earlier, the D peak intensity cannot be

used as a definite parameter to determine the number of defects in graphene laminates as

the more and fewer damaged regions could coexist within the Raman probe [202].

Therefore, in this case, the added defects caused by irradiation is relatively small

compared to the aged irradiation experiment. The result is to be expected as the

effectiveness of the laser power has been greatly reduced. The Raman 2D (~2696 cm-1)

peak shows noticeable blue shifting. It had been reported that defects caused by laser

irradiation on single-layer graphene could cause a 2D peak frequency shift [202].

However, in the case of graphene laminate, such small shifting of the 2D peak could also

mean nothing due to the possibility of the small difference in the degree of disorder

between regions within the laminate’s surface. Nevertheless, the change in 2D peak is

P a g e | 141

consistent with the EPR spectra and with previous experiments and it may be related to

an increasing number of the broken 𝑠𝑝2 carbon-carbon bonds.

4.3 Conclusion

Ageing of graphene laminates showed a reduction in the EPR intensity with time

in both atmospheric and argon atmospheres indicating passivation. Further changes in the

electronic environment were observed through changes in linewidth. At the early stage of

passivation, the linewidth of the narrow component was broadened and its intensity was

decreased. As the passivation progressed, the electron spin interaction between layers was

weakened and the broad component was affected resulting in a reduction in intensity. This

is an important consideration for further use of graphene laminate materials as well as

storage. For example, use in electronic devices could cause deterioration due to the

environment. Laser irradiation of the aged sample caused an increase in the numbers of

spins whereas a reduction was observed for unaged samples. The defects created by the

laser could break the 𝑠𝑝2 carbon-carbon bonds and eventually destroy the honeycomb

structure of the graphene or could passivate the vacancy defects. Overall, a combination

of EPR and Raman spectroscopy can be useful to monitor defects. Further experiments

are needed to explore these initial findings in the view of using EPR spectroscopy to

monitor defects.

P a g e | 142

5. CHAPTER FIVE

Electron Paramagnetic Resonance Study of Fluorinated

Graphene Laminate

5.0 Introduction

The magnetic moments in graphene have been predicted and thought to be due to

the existence of active defects such as in-plane vacancy defects/dangling bonds and non-

bonding edge defects which could generate magnetic moments [22-24]. Several

experimental and theoretical findings have been presented [11, 27, 31-33, 35, 165-168,

212] which focused on the interaction of magnetic moments from a single layer graphene

point-of-view and only a few theoretical studies have discussed the implication of having

magnetic moments in a multilayer stacked graphene [161-163]. The studies in this field

have led to the possibility of using graphene in advanced applications [24, 157-159].

However, graphene typically has several shortcomings such as a zero bandgap and

chemical inertness [228] and thus, many functionalizations have been introduced to alter

graphene such as dispersion, chemical interaction and electronic properties [229-234].

Recently, among many graphene derivatives, fluorinated graphene (FGn) has

attracted a lot of interest due to its intriguing structures and properties [228, 235-237].

The introduction of fluorine into the hexagonal structure of graphene changes the C-C

bonds from sp2 hybridized to sp3 hybridized [229]. Owing to this structure, FGn is a

semiconductor with a wide optical bandgap (~3.8 eV) [229]. The bandgap exhibits a

strong dependence on the degree of fluorination and tuning of the bandgap from 0 to

~3.13 eV is thought to be possible with precise incorporation of fluorine [238]. Moreover,

FGn shows the potential for use as an atomically thin insulator owing to the high

electronegativity of fluorine (3.98) [229]. Further, FGn is regarded as an excellent cathode

P a g e | 143

material for lithium batteries and exhibits an ability to store and release a high energy

density [239, 240]. Recently, Peng et al. [240] have demonstrated the performance of a

FGn/sulphur hybrid cathode (an improvement to FGn cathode) in lithium/carbon fluoride

(Li/CFx) button cell batteries which showed a high energy density of 2341 Wh kg-1 and a

power density up to 13621 W kg-1 at 8 A g-1.

Chemically, the C-F bonds in FGn exhibit reactivity and susceptibility to nucleophilic

attack [235, 241-243]. This has made it possible to use FGn as a precursor to synthesize

different types of functionalized graphene under controlled conditions [235, 242, 244].

The magnetic properties of FGn with fluorine nanoridges show the potential as a room

temperature spintronics material, superconducting quantum interference device (SQUID)

spectroscopy revealed a strong coupling of magnetic states at the graphene-

fluorographene interface [32]. Further, Makarova et al. [32] predict that the spins at the

localized edge states are ferromagnetically ordered within each of the zigzag interfaces

whereas the spin interaction across a nanoridge is antiferromagnetic. The magnetism such

as ferromagnetism and anti-ferromagnetism in FGn had been discussed theoretically and

predicted previously [32, 237, 245-247] even though little experimental evidence has

been reported [32].

The present study reports an electron paramagnetic resonance (EPR)

spectroscopic investigation of FGn laminates produced by exfoliation of fluorinated

graphite (FG) in N-methyl-2-pyrrolidone (NMP) using sonication. The FGn laminates

were measured at 10-280 K and a Curie-Weis fit was made to understand the magnetic

behaviour. Hyperfine sublevel correlation (HYSCORE) spectroscopy was used to

measure hyperfine interactions.

P a g e | 144

5.1 Sample Preparation

5.1.1 Liquid phase exfoliation fluorinated graphene laminate

Fluorinated graphene (FGn) dispersions were prepared by following a liquid phase

exfoliation (LPE) method reported in previous work [61] with modifications. In detail,

several batches containing 3 mg/ml of fluorinated graphite (FG) (> 61 wt % F, Sigma

Aldrich) and 5 ml of N-methyl-2-pyrrolidone (NMP) (Sigma-Aldrich) were prepared. The

mixtures were bubbled with nitrogen for 1 minute and then sonicated for 6 days in a bath

sonicator (Hilsonic, 40 Hz and 600 W). The FGn dispersion was obtained after

centrifugation at 2000 rpm for 20 minutes to remove the unexfoliated flakes. The FGn

dispersion was then put inside a sealed glass bottle and stored in the fridge. The FGn

laminates were prepared by filtering the dispersion using durapore membranes (Merck

Millipore) which is EPR silent. Acetone (Sigma-Aldrich) was added into the graphene

dispersion before filtrating to flocculate the graphene flakes. FGn laminates were

prepared by filtering a certain amount of the dispersion onto a filter membrane. The

filtration was repeated until the solution was clear. The samples were prepared on a filter

membrane with an area of around ~ 1.77 cm2 (diameter ~ 1.5 cm). The laminates had

1.132 mg/cm2 of fluorinated graphene.

5.1.2 Electron Paramagnetic Resonance spectroscopy

All samples for electron paramagnetic resonance (EPR) were prepared by cutting

the membranes to ± 2 mm wide, followed by stacking up to 11 layers into Suprasil tubes.

The continuous-wave (CW) measurements were performed on a Bruker EMX X band

instrument equipped with a cryostat and an Oxford Instruments Cryospares temperature

controller. The experiments were carried out using 2 mW, 1 G modulation amplitude, 10

P a g e | 145

scans, 40.96 ms time constant and conversion time, under non-saturating conditions.

Sample orientation and quantitation were performed as described in Chapter 2. Hyperfine

sublevel correlation (HYSCORE) spectroscopy at 10 K was performed using a Bruker X

band ELEXSYS E580 spectrometer equipped with a cryostat and an Oxford Instruments

Cryospares temperature controller. The measurement was carried out at the 𝐻⊥ and 𝐻∥

orientations at 3486 G, 9.6993 GHz with pulse lengths of 16 ns for π/2 and 32 ns for π

and pulse delays of τ = 160 and 300 ns and t1 and t2 = 300 ns.

5.1.3 Fourier-Transform Infrared spectroscopy

A Thermo Scientific Nicolet iS5 Fourier-transform infrared (FTIR) spectrometer

equipped with iD5 attenuated total reflectance (ATR) was used to determine the chemical

bonds in the laminates. The sample for the measurements was prepared by placing the

laminate sheet onto the optical surface of the instrument.

5.1.4 Raman spectroscopy

A Renishaw Invia Raman spectrometer was used to analyse the laminate. The

Raman measurements were performed as described in Section 3.1.6 for 325 nm and 514.5

nm laser excitation.

P a g e | 146

5.2 Results and Discussion

5.2.1 Fluorinated graphene laminate

LPE FGn dispersion can be prepared with different solvents and can cause

different results. Figure 5.1 shows that the FGn dispersion prepared by using NMP solvent

had a black colour, while the FGn prepared in isopropanol : water (1:1) had a white colour.

Previously, Mazanek et al. [248] synthesized a series of FGn samples with various

contents of fluorine in an autoclave with a nitrogen/fluorine atmosphere at different

exposure times and temperatures. The result was three different colours of FGn with each

colour indicating the amount of fluorine contained in the bulk samples. The three FGn

samples were black with ~20 % of fluorine, brown with ~40 % of fluorine and white with

~50 % of fluorine. Therefore, the black colour of dispersion shown in Figure 5.1 had

lower fluorine content.

Figure 5.1 FGn dispersion prepared in (a) NMP and (b) isopropanol : water (1:1).

Although the black dispersion had less fluorine content, the black dispersion

showed good stability apparent through the observation that sedimentation was not

observed up until three weeks in storage while sedimentation of the white dispersion was

observed in less than a week. The stability of the FGn dispersion in NMP was in

agreement with the previously reported sonochemical preparation of FGn from FG in

NMP [249]. Further discussion will focus on the FGn produced in NMP solvent due to its

P a g e | 147

stability which indicates a nearly equal of FGn and solvent surface energies. The balance

of FGn and solvent surface energies affected the enthalpy of mixing which favours small

value for exfoliation to occur [61].

The FTIR analysis of the FGn shows three vibration region bands corresponding

to C-F (~1200 cm-1), C=C (~1669 cm-1) and C-H (~2936 cm-1) stretching vibration bands,

while the FG displays a single vibration region band which corresponds to the C-F

stretching vibrational band (Figure 5.2a). The C-F stretching vibration region is typically

located between 1000 – 1300 cm-1 [250]. Previously, Asanov et al. [251] have used

Yudanov et al. [252] ideas and proposed a schematic structure of C-F bonds in graphene.

The FTIR spectrum of fluorinated graphite intercalated compound (FGIC) with a matrix

composition of C2.5F showed a band with four components which corresponded to C-F

bonds depending on the local surrounding. The four components were tentatively

assigned to: a) CF3 at 1230 cm-1 corresponding to the vibrations of C-F group having

three C-F neighbours, b) CF2 at 1132 cm-1 corresponding to the vibrations of C-F group

having two C-F neighbours and one sp2 hybridized carbon atom, c) CF1 at 1095 cm-1

corresponding to the vibrations of C-F group having one C-F neighbour and two sp2

hybridized carbon atoms, d) CF at 1045 cm-1 corresponding to the vibrations of C-F group

having three sp2 hybridized carbon atoms. Other work on FG and fluorinated carbon

nanotubes have stated that these modes are related to the CF distribution over the carbon

skeleton, with the species of bulk CF assigned to the highest IR frequency and a single

C-F bond assigned to the lowest IR frequency [250, 253-255]. Figure 5.2b shows four

components which we tentatively assigned to the stretching vibrations of CF bonds with

different local surroundings. The four components and assignment are: CF3 at 1203 cm-1

(FG) and 1198 cm-1 (FGn), CF2 at 1117 cm-1 (FGn) and CF1 at 1068 cm-1 (FGn). Another

component of C-F vibrational stretching is assigned to CFedge located at 1313 cm-1 (FG)

and 1300 cm-1 (FGn), which may correspond to the C-F2 and C-F bonds located at the

P a g e | 148

graphene edge. The C-F vibrational modes and the emerging of C=C and C-H bands

suggest that the FGn produced has less fluorine content than FG. The defluorination could

happen during the production process and is possibly due to the NMP solvent molecules

interacting with the carbon atom of C-F bonds via dipolar-dipolar interaction causing the

release of energy for promoting the rupture of C-F bonds [241]. Noteworthy is that the

defluorination could reduce the bandgap and may open the possibility for tuning of the

desirable bandgap for specific applications [235, 241, 256].

1000 1500 2000 2500 3000 3500 4000

%T

ransm

itta

nce (

a.u

.)

Wavenumbers (cm-1)

a)

C-F

C=C

C-H

1000 1500 2000 2500 3000 3500 4000Wavenumbers (cm-1)

1000 1200 1400

%T

ransm

itta

nce (

a.u

.)

Wavenumbers (cm-1)1000 1200 1400

Wavenumbers (cm-1)

1117

1068

1198

1203

1300

1313b)

CF3

CF2

CF1CFedge

Figure 5.2 a) FTIR spectra of FGn (orange) and FG (black). b) Four components which

correspond to stretching vibrations of C-F bonds with different local surroundings. CF3,

CF2 and CF1 annotation assigns the bonds, which have three, two and one C-F

neighbours. CFedge annotation assigns the bonds located at the graphene edges which

may be attributed to C-F2 and C-F bonds.

P a g e | 149

The Raman analysis was conducted by using a laser at 325 nm and 514.5 nm and

the results are depicted in Figure 5.3. The FGn shows an intense luminescence which

affected the baseline shape and led to the suppression of the D and G band intensities.

The luminescence is common for FGn and had been reported previously [228, 248, 257].

The presence of the 2D band could not be detected due to a high degree of structural

disorder in FGn [228]. The G band of FGn, as shown in Figure 5.3b, was shifted to a

higher wavenumber by ~12 cm-1 indicating that the incident light gains energy after

interacting with the vibrational modes of the FGn and this has been associated with the

C-F bonds at hexagonal rings of graphene [202, 204, 248]. The Raman result of FGn

characterized by the occurrence of D and G bands using 514.5 nm laser light indicates

that the graphene was only partially fluorinated [229], suggesting that defluorination

happens during the LPE process and also in agreement with the FTIR spectra. Typically,

the Raman signals using 514.5 nm laser light for fully fluorinated graphene were

suppressed due to the energy of the laser being smaller than the bandgap [229].

1500 2000 2500 3000

Inte

nsity (

a.u

.)

Raman Shift (cm-1)

a)

1200 1400 1600 2400 2800 3200

Inte

nsity (

a.u

.)

Raman Shift (cm-1)

b)

Figure 5.3 a) Raman spectra of FGn with 325 nm laser. b) Raman spectra of LPE

graphene (black) and FGn (blue) with 514.5 nm laser.

P a g e | 150

5.2.2 Paramagnetism of fluorinated graphene laminate

The magnetic properties of FGn laminate were characterized by using EPR

spectroscopy. The FGn laminate exhibited an isotropic paramagnetic signal with an

average g value of 2.0028 ± 0.0001, which was close to the free-electron value indicating

that the observed signal did not come from transition metal ions. The value was found to

be within the range of previously reported for carbon EPR signals (2.0022 – 2.0035) [258].

Thus indicating that the EPR signal observed predominately associates with the C related

dangling bonds. The FGn laminates exhibited an average spin concentration of 3.3895 x

1017 spin/g. The spin concentration was calculated by comparing the double integrated

intensity area of the EPR signal against a known DPPH standard at room temperature

[175]. The EPR lineshape of the FGn laminate at room temperature can be simulated by

using a single Lorentzian component indicating a homogeneously broadened EPR

resonance possibly arising from electron-electron interactions [21]. Figure 5.4 shows that

both orientations can be simulated by using a single Lorentzian component with the A/B

ratio of 1. Despite, the previous FTIR results (Figure 5.2a) showing the possibility of sp2

hybridized carbon atoms in the FGn sample the presence of conduction electrons was not

directly detected. In the FGn laminate, the broad component was not present possibly due

to a high concentration of C-F bonds on the graphene skeleton causing the evolution of

the degeneracy of the energy bands and the Fermi level position [169, 178]. In the case

of FGn, the Fermi level is thought to be shifted down into the valence bands indicating

that the F atoms act as hole dopants [238].

P a g e | 151

340 344 348 352

Inte

nsity (

a.u

.)

Magnetic Field (mT)

Figure 5.4 The EPR lineshape of FGn laminate at 𝐻⊥ (solid black) and 𝐻∥ (solid blue)

simulated using a single Lorentzian lineshape (dash purple and orange). The simulation

was performed using easypin [176].

5.2.3 Temperature-dependence of the EPR resonance

The temperature dependence experiments revealed, again, that the EPR linewidth

of FGn laminate was not affected by the external magnetic field orientation. The linewidth

behaves isotropically regardless of the laminate orientation. This behaviour reminds us of

the electrochemical exfoliated (EC) graphene and reduced graphene oxide (RGO)

laminates in Chapter 3. Figure 5.5 displays the evolution of peak-peak linewidth (Ipp)

throughout the temperature in the range (room temperature – 10 K). The results show that

the Ipp increases as the temperature decreases for 𝐻⊥ and 𝐻∥ orientations. The average Ipp

from room temperature down to 10 K was shifted from ~0.49 to ~0.77 mT for the 𝐻∥

orientation and from ~0.49 to ~0.71 mT for the 𝐻⊥ orientation. This is consistent with a

decoupled state of localized and itinerant conduction electrons [14, 20]. The increase of

the linewidth at low temperature may represent the decrease in the exchange rate due to

P a g e | 152

the trapping of conduction electrons in some isolated regions. This results in a reduction

of the motional narrowing which determines the linewidth at high temperatures [259].

0 50 100 150 200 250 300

0.5

0.6

0.7

Lin

ew

idth

(m

T)

Temperature (K)

Figure 5.5 The evolution of EPR linewidth on the variation of temperature. The black

rectangle represents the 𝐻⊥ orientation and the red triangle represents the 𝐻∥ orientation.

The linewidth trend was similar to the EC graphene laminate (Chapter 3) and it is

possible that functionalization has caused significant disruption to the mobility of

conduction electrons as if the conduction electrons have been confined within isolated

regions, influencing coupling and realizing that the broad component was not present.

The magnetic susceptibility ( 𝜒𝐸𝑃𝑅 ) is directly proportional to the double

integrated EPR intensity. The average evolution of 𝜒𝐸𝑃𝑅 at the various temperatures was

found to be isotropic. Again, this was similar to as observed for other laminates with

massive defects such as EC graphene and RGO laminates (Chapter 3). Interestingly, in

the case of FGn laminate, two different types of magnetic ordering were found in two

different temperature regions. Figure 5.6 shows the two temperature regions that hold a

magnetic susceptibility trend; the area between 280 – 210 K and a lower temperature

region between 100 – 10 K.

P a g e | 153

0 50 100 150 200 250 300

2

4

6

8

E

PR (

a.u

.)

Temperature (K)

Figure 5.6 The evolution of double integrated EPR intensity ( 𝜒𝐸𝑃𝑅 ) on a wide

temperature range (300 – 10 K). Black rectangle represents the 𝐻⊥ orientation and red

rectangle represents the 𝐻∥ orientation.

The 𝜒𝐸𝑃𝑅 of FGn laminate was understood by plotting the inverse of the double

integration of the EPR intensity (𝜒𝐸𝑃𝑅−1) vs temperature. The 𝜒𝐸𝑃𝑅 could be described

through the use of the Curie-Weis equation (Chapter 2). The Curie-Weis fit at the lower

temperature region and at the two field orientations showed antiferromagnetic behaviour

with a nearly similar Curie-Weis temperature. The 𝐻⊥ orientation had a Curie-Weis

temperature θ = -12.2 ± 3.1 K, while the 𝐻∥ orientation gave θ = -11.5 ± 2.7 K. Figure 5.7

displays the Curie-Weis fit for the average 𝜒𝐸𝑃𝑅 in the range of 100 – 10 K.

P a g e | 154

0 50 100 150 200 250 300

2

4

6

8

E

PR (

a.u

.)

Temperature (K)

Figure 5.7 The Curie-Weis line fit for 𝐻⊥ orientation (solid black line) and 𝐻∥ orientation

(solid red line) at the temperature range of 100 – 10 K. The black rectangle and red

triangle represent 𝜒𝐸𝑃𝑅 at the 𝐻⊥ and 𝐻∥ orientations, respectively.

The higher temperature region of the average 𝜒𝐸𝑃𝑅 showed ferromagnetic

ordering at both field orientations. The Curie-Weis fit in the range of 280 – 230 K for both

orientations is displayed in Figure 5.8. The Curie-Weis fit for 𝐻⊥ orientation was θ =

199.9 ± 25.3 K while Curie-Weis line fit for 𝐻∥ orientation was θ = 199.7 ± 32.2 K. The

result is interesting since the addition of C-F bonds to the graphene skeleton was able to

give a magnetic ordering at around ~199 K while pure graphene was only able to show

magnetic ordering at relatively low temperature [(0 > (TC or TN) > 9.2 K) see Chapter 2

and 3].

Graphene-based materials with magnetic ordering near room temperature would

be attractive for a lot of applications. Moreover, the electronic and magnetic properties of

FGn can be tuned and have been found to be strongly dependent on the degree of

fluorination [238, 245, 246]. A theoretical study conducted by Liu et al. revealed that a

P a g e | 155

precise increase in fluorine content was able to tune the bandgap from 0 to ~3.13 eV and

cause a transformation of graphene from nonmagnetic semimetal to

nonmagnetic/magnetic metal, or to magnetic/nonmagnetic semiconductor [238].

200 220 240 260 280 300

1.0

1.5

2.0

EP

R (

a.u

.)

Temperature (K)

Figure 5.8 The Curie-Weis line fit for 𝐻⊥ orientation (solid black line) and 𝐻∥ orientation

(solid red line) at the temperature range of 280 – 230 K. The black rectangle and red

triangle represent 𝜒𝐸𝑃𝑅 at the 𝐻⊥ and 𝐻∥ orientations, respectively.

As mentioned previously, the magnetism such as ferromagnetism and anti-

ferromagnetism has been predicted to exist in FGn [32, 237, 245-247]. The fully

fluorinated graphene system was thought to consist of C-C, C-F and F-F bonds with a

strong orbital interaction [245]. The C-C and F-F bonds can create energy bands leading

to hybridized valence bands by the C-F bonds. This might lead to the existence of hole

doping and ferromagnetism under certain concentrations and distributions [246]. eheng

et al. have calculated that systems with a different number of fluorine atoms located in

the Brillouin sublattice A on one graphene side and in the sublattice B on the other side

to be magnetic and found the magnetic moment increases with the imbalance number

[247]. A double-side attachment of fluorine atoms along the zigzag direction of the

P a g e | 156

graphene plane was thought to break the uniformity of the π-system of graphene leading

to the π-electrons of carbon atoms neighbouring with a C-F chain aligning along the chain

causing the electron spins to interact ferromagnetically [32]. In this current work, we

speculate that the exchange from the ferromagnetic to the antiferromagnetic alignment

was probably caused by energetically favourable conditions [163] similar to the stacked

graphene laminate (Chapter 2).

5.2.4 HYSCORE spectroscopy

The interaction of carbon paramagnetic centres with other nuclei i.e. fluorine in

FGn was found not to be evident from visual inspection of the CW linewidth suggesting

that potential hyperfine interactions were within the EPR linewidth. HYSCORE

spectroscopy was used to probe for small hypefine interactions (> 20 MHz) and was

carried out at 10 K at the 𝐻⊥ (Figure 5.9) and 𝐻∥ orientations (Figure 5.10). The spectrum

shows three intense anti-diagonal peaks belonging to three nuclei coupled to the electron

spin exhibiting weak hyperfine couplings that only occur in the (+,+) quadrant. An intense

anti-diagonal peak centred at ~13.8 MHz is characterized by two ridges along the diagonal

line in the range of 10 – 17.5 MHz corresponding to the hyperfine interaction between

19F nuclei and the electron spin. Simulation using Easyspin [176] was made to measure

the hyperfine values and a typical simulation is displayed in Figure 5.11. The fluorine

hyperfine couplings constant (A) was found to be ~2 MHz (0.07 mT) along the 𝐻⊥ and

𝐻∥ orientations, respectively. This is the first time this has been measured in FGn to the

best of our knowledge. The observed A values were lower compared to the A values (4.5

-8.6 MHz) reported using HYSCORE for 19F nuclei in 40SiO2-30PbF2-30CdF2 glass

doped with Cu2+ ions at the perpendicular orientation and 10 K [260].

P a g e | 157

The intense anti-diagonal peak centred at ~14.7 MHz corresponds to the nuclear

magnetic resonance (NMR) Larmour frequency of 1H. The appearance of 1H was in

agreement with FTIR result of FGn (Figure 5.9a). The proton diagonal peak is not

accompanied by ridges indicating that the hyperfine interaction between proton nuclei

and electron spin is not resolved in the HYSCORE spectra recorded. Proton hyperfine

HYSCORE ridges have been observed on graphene nanoribbons (GNRs) edges decorated

with protons. Rao et al. reported A (1H) to be round ~25 MHz (0.89 mT) on GNRs with

proton decorated edges [16]. Further experiments using different tau values as well as

Matched-HYSCORE and electron-nuclear double resonance (ENDOR) experiments

would be needed to fully characterise the proton hyperfine splitting of the laminates. It is

noted that the value has to be less than the linewidth of ~0.7 mT.

The Larmour frequency of 13C was ~3.7 MHz and was characterized by two small

ridges arising about its centre. The ridges indicated a hyperfine interaction between the

electron spin and 13C nuclei and the average hyperfine couplings constant, A (13C), was

around ~1 MHz (0.04 mT) along the 𝐻⊥ and 𝐻∥ orientations. The A (13C) parameter

observed was close to the A (13C) value reported for different coals between 1.2 – 3 MHz

(0.04 – 0.11 mT) [261] or to A (13C) < 4 MHz reported for GNRs [16].

P a g e | 158

0 5 10 15 200

5

10

15

20

13C

19F

1H

2 (

MH

z)

1 (MHz)

a

0 10 200

10

20

2 (

MH

z)

1 (MHz)

1H

19F

13C

b

Figure 5.9 The HYSCORE 2D plot spectrum measured at 10 K and the 𝐻⊥ orientation

on FGn in frequency coordinates. a) τ = 160 ns. b) τ = 300 ns.

P a g e | 159

0 10 200

10

20

2 (

MH

z)

1 (MHz)

a

1H

19F

13C

0 10 200

10

20

2 (

MH

z)

1 (MHz)

b

1H

19F

13C

Figure 5.10 The HYSCORE 2D plot spectrum measured at 10 K and the 𝐻∥ orientation

on FGn in frequency coordinates. a) τ = 160 ns. b) τ = 300 ns.

P a g e | 160

Figure 5.11 The HYSCORE simulation 2D plot spectrum measuring 13C, 19F and 1H

resonances in frequency coordinates with A (13C) = 1 MHz, A (19F) = 2 MHz and A (1H)

= 0.6 MHz. a) τ = 160 ns. b) τ = 300 ns.

5.3 Conclusion

The FGn prepared by sonication of FG in NMP exhibited a stable dispersion and a

blackening due to defluorination. The FTIR result displayed four vibrational modes

assigned to CF groups with different local surroundings. The samples showed an EPR

P a g e | 161

signal with a g value of 2.0028 and linewidths of 0.5-0.7 mT across the temperature range.

This is the first time FGn has been characterised using EPR to the best of our knowledge.

The Curie-Weis fit of the magnetic susceptibility behaviour showed two temperature

regions, which show the magnetic moments to couple ferromagnetically and

antiferromagnetically. The magnetic coupling, as observed in our sample, is thought to be

caused by the CF distribution on both sides of the graphene plane and edges. The

imbalanced number of the CF groups on both sides increases the magnetic moment and

the interaction between π electron and neighbouring a C-F chain at the edges may cause

the electron spins to interact ferromagnetically. The exchange of magnetic coupling at the

low-temperature region was thought due to energetically favourable conditions. A

HYSCORE spectrum directly showed, for the first time, that the electron spin is

delocalized both on 1H, 19F and 13C. This path the way for future experiments to

potentially correlate defects with hyperfine splitting.

P a g e | 162

6. CHAPTER SIX

Conclusions and Future Work

6.0 Conclusions

The magnetic properties of graphene (i.e. LPE graphene, EC graphene and rGO)

and its derivative (i.e FGn) in the form of laminates have been studied by using electron

paramagnetic resonance (EPR) and Raman spectroscopy. Graphene laminates consist of

randomly stacked graphene layers.

The difference of magnetic properties between the graphene and its derivative are

distinguishable using continuous-wave (CW) EPR spectroscopy. The g values of the

laminates showed the presence of carbon centred spins. The g values of the laminates

were close to the free-electron value indicating that the signals did not come from

transition metal ions, and were within the range previously reported for carbon signals

[128, 179, 258]. Typically, the CW EPR spectra revealed the presence of two resonances

with narrow and broad line widths. The g value of the narrow component was temperature

independent. In less defective graphene laminate (Chapter 2), the intensity of the narrow

component was affected by the external magnetic field orientation and at near room

temperature the intensity was stronger at the 𝐻⊥orientation. The linewidth of the narrow

component at 𝐻∥ increased on lowering of the temperature revealing a reduction in the

exchange narrowing mechanism. The narrow component was assigned to spins generated

from localized electrons in vacancy defects.

The broad component in graphene laminate was attributed to spins arising from

the interplay of electrons between graphene layers in less defective graphene laminate.

The broad component can be understood from the similarity with the anisotropic

P a g e | 163

component of graphite at the 𝐻∥ orientation, which is affected by conduction electron

mobility, the degeneracy energy bands and Fermi level position. Typically, the broad

component was unobservable in more defective graphene laminates. The broad

component was temperature-dependent. The X band CW EPR spectra of less defective

graphene laminates (Chapter 2) at the 𝐻∥ orientation showed the g value increased while

the intensity was diminished and the linewidth was broadened as the temperature

decreased. The broad component was unobservable below ~70 K. Defects such as

vacancy, topological and functionalization cause a disruption in the conduction electron

mobility and displacement in the Fermi level position. As a consequence, the g value and

the linewidth of the broad component were affected.

Magnetic susceptibility experiments of the graphene laminates revealed a

complex mechanism involving stacking disorder, vacancy defect locations, topological

defects and functionalization. The magnetic susceptibility in less defective graphene

laminate tends to be anisotropic (Chapter 2) whereas in the more defective graphene it

behaves isotropically (Chapter 3 and 5). Magnetic ordering was observed on all graphene

laminate samples near the absolute zero temperature and follows the Curie-Weiss law.

The Curie-Weiss fit of the magnetic susceptibility of FGn laminates (Chapter 5) showed

two temperature regions, which show the magnetic moments to couple ferromagnetically

and antiferromagnetically. This indicates that the introduction of defects and or

functionalization may alter the magnetic susceptibility behaviour.

The investigation on the aged graphene laminate samples (Chapter 4) revealed a

passivation mechanism. The passivation was thought due to self-reconstruction and

bonding with other atoms i.e. oxygen. The narrow component was immediately affected

by the passivation while the broad component was affected in the later stage. At the early

stage of passivation, the linewidth of the narrow component was broadened and its

P a g e | 164

intensity was decreased. As the passivation progressed, the electron spin interaction

between layers was weakened and the broad component was affected resulting in a

reduction in intensity.

The electron-electron interactions in graphene laminate can be understood as the

interaction between localized and conduction electrons and as the interaction of electrons

between graphene layers. Thus, the interplay between layers, such as in graphite, generate

complex energy bands, which affects the magnetic properties of the laminate. The full

characterisation in this study provides a reference for future studies using graphene

laminates. This is much needed considering the mass of EPR studies of graphite materials

(Chapter 1).

Overall, the results show that defects and disorder in the laminates, i.e. vacancy

defects, topological defects and functionalization, have spectroscopic signatures

highlighted through temperature, spin concentration and external magnetic field

orientation depending on the degree of disorder and the type of defects or

functionalization. This characterisation and investigation are therefore useful for a further

electronic understanding of graphene laminates (i.e. on how to tune the magnetic

properties of stacked graphene compounds) and monitoring of stability in different

applications (i.e. quality control in graphene production, and graphene ink electronic and

spintronic applications etc.).

P a g e | 165

6.1 Future work

The study on graphene laminates provides a base to allow further study. In

particular, further experiments could be carried out using different wavelengths and

powers of laser light as well as different methods of defect creation such as ion

bombardment. In this case, temperature-dependent EPR experiments will be important to

fully understand the magnetic properties of the material. Furthermore, longer irradiation

times are required so that the differences in the temperature-dependent experiments can

be clearly observed.

The current study presents results showing temperature-dependent experiments

from room temperature down to 5 K which was necessary in order to compare the results

with other materials. However, it is also interesting to have knowledge of the magnetic

properties at above room temperature especially because many applications are

performed at room temperature and above. In this case, we hope to find, if any, a magnetic

ordering near room temperature especially from the FGn laminate because we observe a

decrease of EPR intensity from room temperature down to 280 K. It is likely that the EPR

signal at higher temperatures is very weak but the growing renaissance of advanced EPR

instrumentation such as involving rapid-scan [262] offers much hope for greater

sensitivity.

The current work provides a reference point for further study of the magnetic

properties of laminates made of novel materials. In order to explore more of the magnetic

properties of graphene-based material, as an example of a suggested target is a laminate

made from a mixture of graphene and other 2D materials i.e. FGn and hexagonal boron

nitride (h-BN). In 2D h-BN, B atoms and N atoms are alternately arranged to form a

honeycomb structure (Figure 6.1). The B-N bond length is 1.45 Å and forms through sp2

hybridization. The interlayer of h-BN is connected through weak van der Waals forces.

P a g e | 166

In fact, a pilot study of a solution possibly containing monolayers of h-BN was produced

using sonication-assisted LPE and made into laminates. However, due to the fact that the

EPR S/N was very weak and almost comparable to the background signal, it was not

investigated further. This work could be continued through variation in the preparation

of the samples.

Figure 6.1 (A) Two layers of h-BN with B atoms are on top of the N atoms. (B)

A unit cell of the honeycomb structure of h-BN with Bravais lattice vectors. Taken from

[263].

P a g e | 167

References

1. Radzig, A.A. and Smirnov, B.M., Reference data on atoms, molecules, and ions.

Chemical Physics. Vol. 31. 1985: Springer-Verlag Berlin Heidelberg.

2. Novoselov, K.S., Geim, A.K., Morozov, S.V., Jiang, D., ehang, Y., Dubonos, S.V.,

Grigorieva, I.V., and Firsov, A.A., Electric Field Effect in Atomically Thin Carbon

Films. Science, 2004. 306(5696): p. 666-669.

3. Dresselhaus, M.S. and Dresselhaus, G., Intercalation compounds of graphite.

Advances in Physics, 2002. 51(1): p. 1-186.

4. Krishnan, A., Dujardin, E., Treacy, M.M.J., Hugdahl, J., Lynum, S., and Ebbesen,

T.W., Graphitic cones and the nucleation of curved carbon surfaces. Nature, 1997.

388(6641): p. 451-454.

5. Geim, A.K. and Novoselov, K.S., The rise of graphene. Nature Materials, 2007.

6: p. 183-191.

6. Lee, J.U., Yoon, D., and Cheong, H., Estimation of Young’s Modulus of Graphene

by Raman Spectroscopy. Nano Letters, 2012. 12(9): p. 4444-4448.

7. Lee, C., Wei, X., Kysar, J.W., and Hone, J., Measurement of the Elastic Properties

and Intrinsic Strength of Monolayer Graphene. Science, 2008. 321(5887): p. 385-

388.

8. Balandin, A.A., Ghosh, S., Bao, W., Calizo, I., Teweldebrhan, D., Miao, F., and

Lau, C.N., Superior thermal conductivity of single-layer graphene. Nano Letters,

2008. 8(3): p. 902-907.

9. Novoselov, K.S., Geim, A.K., Morozov, S.V., Jiang, D., Katsnelson, M.I.,

Grigorieva, I.V., Dubonos, S.V., and Firsov, A.A., Two-dimensional gas of

massless Dirac fermions in graphene. Nature, 2005. 438: p. 197-200.

P a g e | 168

10. Neto, A.H.C., Guinea, F., Peres, N.M.R., Novoselov, K.S., and Geim, A.K., The

electronic properties of graphene. Rev. Mod. Phys., 2009. 81(1): p. 109-162.

11. Makarova, T. and Palacio, F., Carbon Based Magnetism: An Overview of the

Magnetism of Metal Free Carbon-based Compounds and Materials. 2006:

Elsevier Science.

12. Jia, X., Campos-Delgado, J., Terrones, M., Meunier, V., and Dresselhaus, M.S.,

Graphene edges: a review of their fabrication and characterization. Nanoscale,

2011. 3: p. 86-95.

13. Nair, R.R., Blake, P., Grigorenko, A.N., Novoselov, K.S., Booth, T.J., Stauber, T.,

Peres, N.M.R., and Geim, A.K., Fine Structure Constant Defines Visual

Transparency of Graphene. Science, 2008. 320(5881): p. 1308.

14. Ćirić, L., Sienkiewicz, A., Náfrádi, B., Mionić, M., Magrez, A., and Forró, L.,

Towards electron spin resonance of mechanically exfoliated graphene. physica

status solidi (b), 2009. 246(11-12): p. 2558-2561.

15. Augustyniak-Jabłokow, M.A., Tadyszak, K., Maćkowiak, M., and Yablokov, Y.V.,

EPR evidence of antiferromagnetic ordering in single-layer graphene. physica

status solidi (RRL) – Rapid Research Letters, 2011. 5(8): p. 271-273.

16. Rao, S.S., Jammalamadaka, S.N., Stesmans, A., Moshchalkov, V.V., Tol, J.v.,

Kosynkin, D.V., Higginbotham-Duque, A., and Tour, J.M., Ferromagnetism in

Graphene Nanoribbons: Split versus Oxidative Unzipped Ribbons. Nano Letters,

2012. 12(3): p. 1210-1217.

17. Akbari-Sharbaf, A., Cottam, M.G., and Fanchini, G., A combined theoretical-

experimental investigation of paramagnetic centres in chemically exfoliated

graphene nanoribbons. Journal of Applied Physics, 2013. 114(2): p. 024309.

P a g e | 169

18. Augustyniak-Jabłokow, M.A., Tadyszak, K., Maćkowiak, M., and Lijewski, S.,

ESR study of spin relaxation in graphene. Chemical Physics Letters, 2013. 557: p.

118-122.

19. Just, S., eimmermann, S., Kataev, V., Büchner, B., Pratzer, M., and Morgenstern,

M., Preferential antiferromagnetic coupling of vacancies in graphene on SiO2:

Electron spin resonance and scanning tunneling spectroscopy. Physical Review

B, 2014. 90(12): p. 125449.

20. Náfrádi, B., Choucair, M., and Forró, L., Spin lifetime of itinerant electrons in

chemically synthesized graphene multi-layers. Carbon, 2014. 74: p. 346-351.

21. Tampieri, F., Silvestrini, S., Ricco, R., Maggini, M., and Barbon, A., A

comparative electron paramagnetic resonance study of expanded graphites and

graphene. Journal of Materials Chemistry C, 2014. 2(38): p. 8105-8112.

22. Nair, R.R., Sepioni, M., Tsai, I.L., Lehtinen, O., Keinonen, J., Krasheninnikov,

A.V., Thomson, T., Geim, A.K., and Grigorieva, I.V., Spin-half paramagnetism in

graphene induced by point defects. Nat Phys, 2012. 8(3): p. 199-202.

23. Castro Neto, A.H., Kotov, V.N., Nilsson, J., Pereira, V.M., Peres, N.M.R., and

Uchoa, B., Adatoms in graphene. Solid State Communications, 2009. 149(27–28):

p. 1094-1100.

24. Han, W., Kawakami, R.K., Gmitra, M., and Fabian, J., Graphene spintronics. Nat

Nano, 2014. 9(10): p. 794-807.

25. Liu, Y., Tang, N., Wan, X., Feng, Q., Li, M., Xu, Q., Liu, F., and Du, Y.,

Realization of ferromagnetic graphene oxide with high magnetization by doping

graphene oxide with nitrogen. Scientific Reports, 2013. 3: p. 2566.

26. Palacios, J.J., Fernández-Rossier, J., and Brey, L., Vacancy-induced magnetism in

graphene and graphene ribbons. Physical Review B, 2008. 77(19): p. 195428.

P a g e | 170

27. Yazyev, O.V. and Helm, L., Defect-induced magnetism in graphene. Physical

Review B, 2007. 75(12): p. 125408.

28. Kunstmann, J., Ozdogan, C., Quandt, A., and Fehske, H., Stability of edge states

and edge magnetism in graphene nanoribbons. Phys. Rev. B, 2011. 83: p. 045414.

29. ean, R., Ramasse, Q.M., Bangert, U., and Novoselov, K.S., Graphene Reknits Its

Holes. Nano Letters, 2012. 12(8): p. 3936-3940.

30. He, K., Robertson, A.W., Fan, Y., Allen, C.S., Lin, Y.-C., Suenaga, K., Kirkland,

A.I., and Warner, J.H., Temperature Dependence of the Reconstruction of Zigzag

Edges in Graphene. ACS Nano, 2015. 9(5): p. 4786-4795.

31. Osipov, V.Y., Shames, A.I., Enoki, T., Takai, K., Endo, M., Hayashi, T., Kaburagi,

Y., and Vul, A.Y., Magnetic and EPR studies of edge-localized spin

paramagnetism in multi-shell nanographites derived from nanodiamonds.

Diamond and Related Materials, 2009. 18(2): p. 220-223.

32. Makarova, T.L., Shelankov, A.L., eyrianova, A.A., Veinger, A.I., Tisnek, T.V.,

Lähderanta, E., Shames, A.I., Okotrub, A.V., Bulusheva, L.G., Chekhova, G.N.,

Pinakov, D.V., Asanov, I.P., and Šljivančanin, Ž., Edge state magnetism in zigzag-

interfaced graphene via spin susceptibility measurements. Scientific Reports,

2015. 5: p. 13382.

33. Wang, S., Talirz, L., Pignedoli, C.A., Feng, X., Müllen, K., Fasel, R., and Ruffieux,

P., Giant edge state splitting at atomically precise graphene zigzag edges. Nature

Communications, 2016. 7: p. 11507.

34. Koshino, M. and Ando, T., Diamagnetism in disordered graphene. Physical

Review B, 2007. 75(23): p. 235333.

35. Sepioni, M., Nair, R.R., Rablen, S., Narayanan, J., Tuna, F., Winpenny, R., Geim,

A.K., and Grigorieva, I.V., Limits on Intrinsic Magnetism in Graphene. Physical

Review Letters, 2010. 105(20): p. 207205.

P a g e | 171

36. Rao, S.S., Stesmans, A., van Tol, J., Kosynkin, D.V., Higginbotham-Duque, A.,

Lu, W., Sinitskii, A., and Tour, J.M., Spin Dynamics and Relaxation in Graphene

Nanoribbons: Electron Spin Resonance Probing. ACS Nano, 2012. 6(9): p. 7615-

7623.

37. Tadyszak, K., Maćkowiak, M., Augustyniak-Jabłokow, M.A., and Strzelczyk, R.,

FMR evidence of ferromagnetic correlations at zigzag edge states in single-layer

graphene. Journal of Molecular Structure, 2014. 1076: p. 31-34.

38. Novoselov, K.S., Fal′ko, V.I., Colombo, L., Gellert, P.R., Schwab, M.G., and

Kim, K., A roadmap for graphene. Nature, 2012. 490: p. 192-200.

39. Bae, S., Kim, H., Lee, Y., Xu, X., Park, J.-S., eheng, Y., Balakrishnan, J., Lei, T.,

Kim, H.R., Song, Y.I., Kim, Y.-J., Kim, K.S., Özyilmaz, B., Ahn, J.-H., Hong,

B.H., and Iijima, S., Roll-to-roll production of 30-inch graphene films for

transparent electrodes. Nature Nanotechnology, 2010. 5: p. 574-578.

40. Neerinck, D.G. and Vink, T.J., Depth profiling of thin ITO films by grazing

incidence X-ray diffraction. Thin Solid Films, 1996. 278(1): p. 12-17.

41. Novoselov, K.S., Falko, V.I., Colombo, L., Gellert, P.R., Schwab, M.G., and Kim,

K., A roadmap for graphene. Nature, 2012. 490(7419): p. 192-200.

42. Ma, P., Salamin, Y., Baeuerle, B., Josten, A., Heni, W., Emboras, A., and Leuthold,

J., Plasmonically Enhanced Graphene Photodetector Featuring 100 Gbit/s Data

Reception, High Responsivity, and Compact Size. ACS Photonics, 2019. 6(1): p.

154-161.

43. Xia, F., Mueller, T., Lin, Y.-m., Valdes-Garcia, A., and Avouris, P., Ultrafast

graphene photodetector. Nature Nanotechnology, 2009. 4: p. 839-843.

44. Echtermeyer, T.J., Britnell, L., Jasnos, P.K., Lombardo, A., Gorbachev, R.V.,

Grigorenko, A.N., Geim, A.K., Ferrari, A.C., and Novoselov, K.S., Strong

P a g e | 172

plasmonic enhancement of photovoltage in graphene. Nature Communications,

2011. 2: p. 458.

45. Liu, C.-H., Chang, Y.-C., Norris, T.B., and ehong, e., Graphene photodetectors

with ultra-broadband and high responsivity at room temperature. Nature

Nanotechnology, 2014. 9: p. 273-278.

46. Kim, K., Choi, J.-Y., Kim, T., Cho, S.-H., and Chung, H.-J., A role for graphene

in silicon-based semiconductor devices. Nature, 2011. 479: p. 338-344.

47. Xu, J.-L., Li, X.-L., He, J.-L., Hao, X.-P., Wu, Y.-e., Yang, Y., and Yang, K.-J.,

Performance of large-area few-layer graphene saturable absorber in femtosecond

bulk laser. Appl. Phys. Lett., 2011. 99: p. 261107.

48. Sun, e., Hasan, T., Torrisi, F., Popa, D., Privitera, G., Wang, F., Bonaccorso, F.,

Basko, D.M., and Ferrari, A.C., Graphene Mode-Locked Ultrafast Laser. ACS

Nano, 2010. 4(2): p. 803-810.

49. Ju, M.J., Kim, J.C., Choi, H.-J., Choi, I.T., Kim, S.G., Lim, K., Ko, J., Lee, J.-J.,

Jeon, I.-Y., Baek, J.-B., and Kim, H.K., N-Doped Graphene Nanoplatelets as

Superior Metal-Free Counter Electrodes for Organic Dye-Sensitized Solar Cells.

ACS Nano, 2013. 7(6): p. 5243-5250.

50. Ahn, H.-J., Kim, I.-H., Yoon, J.-C., Kim, S.-I., and Jang, J.-H., p-Doped three-

dimensional graphene nano-networks superior to platinum as a counter electrode

for dye-sensitized solar cells. Chem. Commun., 2014. 50: p. 2412-2415.

51. Hu, L.-H., Wu, F.-Y., Lin, C.-T., Khlobystov, A.N., and Li, L.-J., Graphene-

modified LiFePO4 cathode for lithium ion battery beyond theoretical capacity.

Nature Communications, 2012. 4: p. 1687.

52. ehang, H., Guo, H., Li, A., Chang, X., Liu, S., Liu, D., Wang, Y., ehang, F., and

Yuan, H., High specific surface area porous graphene grids carbon as anode

P a g e | 173

materials for sodium ion batteries. Journal of Energy Chemistry, 2019. 31: p. 159-

166.

53. Li, e.-F., ehang, H., Liu, Q., Sun, L., Stanciu, L., and Xie, J., Fabrication of High-

Surface-Area Graphene/Polyaniline Nanocomposites and Their Application in

Supercapacitors. ACS Applied Materials & Interfaces, 2013. 5(7): p. 2685-2691.

54. ehang, L., ehang, F., Yang, X., Long, G., Wu, Y., ehang, T., Leng, K., Huang, Y.,

Ma, Y., Yu, A., and Chen, Y., Porous 3D graphene-based bulk materials with

exceptional high surface area and excellent conductivity for supercapacitors.

Scientific Reports, 2013. 3: p. 1408.

55. Bonaccorso, F., Lombardo, A., Hasan, T., Sun, e., Colombo, L., and Ferrari, A.C.,

Production, Processing and Placement of Graphene and Two Dimensional

Crystals. Materials Today, 2012. 15: p. 564-589.

56. Moldt, T., Eckmann, A., Klar, P., Morozov, S.V., ehukov, A.A., Novoselov, K.S.,

and Casiraghi, C., High-Yield Production and Transfer of Graphene Flakes

Obtained by Anodic Bonding. ACS Nano, 2011. 5(10): p. 7700-7706.

57. Miyamoto, Y., ehang, H., and Tománek, D., Photoexfoliation of Graphene from

Graphite: An Ab Initio Study. Physical Review Letters, 2010. 104(20): p. 208302.

58. Dhar, S., Barman, A.R., Ni, G.X., Wang, X., Xu, X.F., eheng, Y., Tripathy, S.,

Ariando, Rusydi, A., Loh, K.P., Rubhausen, M., Neto, A.H.C., Őzyilmaz, B., and

Venkatesan, T., A new route to graphene layers by selective laser ablation. AIP

Advances, 2011. 1: p. 022109.

59. Lotya, M., Hernandez, Y., King, P.J., Smith, R.J., Nicolosi, V., Karlsson, L.S.,

Blighe, F.M., De, S., Wang, e., McGovern, I.T., Duesberg, G.S., and Coleman,

J.N., Liquid Phase Production of Graphene by Exfoliation of Graphite in

Surfactant/Water Solutions. J. Am. Chem. Soc., 2009. 131(10): p. 3611-3620.

P a g e | 174

60. Maragó, O.M., Bonaccorso, F., Saija, R., Privitera, G., Gucciardi, P.G., Iatì, M.A.,

Calogero, G., Jones, P.H., Borghese, F., Denti, P., Nicolosi, V., and Ferrari, A.C.,

Brownian Motion of Graphene. ACS Nano, 2010. 4(12): p. 7515-7523.

61. Hernandez, Y., Nicolosi, V., Lotya, M., Blighe, F.M., Sun, e., De, S., McGovern,

I.T., Holland, B., Byrne, M., Gun'Ko, Y.K., Boland, J.J., Niraj, P., Duesberg, G.,

Krishnamurthy, S., Goodhue, R., Hutchison, J., Scardaci, V., Ferrari, A.C., and

Coleman, J.N., High-yield production of graphene by liquid-phase exfoliation of

graphite. Nat Nano, 2008. 3(9): p. 563-568.

62. Shin, Y., Prestat, E., ehou, K.-G., Gorgojo, P., Althumayri, K., Harrison, W., Budd,

P.M., Haigh, S.J., and Casiraghi, C., Synthesis and Characterization of Composite

Membranes made of Graphene and Polymer of Intrinsic Porosity. Carbon, 2016.

102: p. 357-366.

63. Yang, H., Hernandez, Y., Schlierf, A., Felten, A., Eckmann, A., Johal, S., Louette,

P., Pireaux, J.-J., Feng, X., Mullen, K., Palermo, V., and Casiraghi, C., A simple

method for graphene production based on exfoliation of graphite in water using

1-pyrenesulfonic acid sodium salt. Carbon, 2012. 53: p. 357-365.

64. Khan, U., O'Neill, A., Lotya, M., De, S., and Coleman, J.N., High-Concentration

Solvent Exfoliation of Graphene. Small, 2010. 6(7): p. 864-871.

65. Khan, U., O’Neill, A., Porwal, H., May, P., Nawaz, K., and Coleman, J.N., Size

selection of dispersed, exfoliated graphene flakes by controlled centrifugation.

Carbon, 2012. 50(2): p. 470-475.

66. Lotya, M., King, P.J., Khan, U., De, S., and Coleman, J.N., High-Concentration,

Surfactant-Stabilized Graphene Dispersions. ACS Nano, 2010. 4(6): p. 3155-

3162.

P a g e | 175

67. O’Neill, A., Khan, U., Nirmalraj, P.N., Boland, J., and Coleman, J.N., Graphene

Dispersion and Exfoliation in Low Boiling Point Solvents. The Journal of Physical

Chemistry C, 2011. 115(13): p. 5422-5428.

68. Smith, R.J., King, P.J., Wirtz, C., Duesberg, G.S., and Coleman, J.N., Lateral size

selection of surfactant-stabilised graphene flakes using size exclusion

chromatography. Chemical Physics Letters, 2012. 531: p. 169-172.

69. Coleman, J.N., Liquid Exfoliation of Defect-Free Graphene. Accounts of

Chemical Research, 2013. 46(1): p. 14-22.

70. Skaltsas, T., Ke, X., Bittencourt, C., and Tagmatarchis, N., Ultrasonication

Induces Oxygenated Species and Defects onto Exfoliated Graphene. The Journal

of Physical Chemistry C, 2013. 117(44): p. 23272-23278.

71. Bracamonte, M.V., Lacconi, G.I., Urreta, S.E., and Foa Torres, L.E.F., On the

Nature of Defects in Liquid-Phase Exfoliated Graphene. The Journal of Physical

Chemistry C, 2014. 118(28): p. 15455-15459.

72. Li, L., Reich, S., and Robertson, J., Defect energies of graphite: Density-

functional calculations. Physical Review B, 2005. 72(18): p. 184109.

73. Abdelkader, A.M., Cooper, A.J., Dryfe, R.A.W., and Kinloch, I.A., How to get

between the sheets: a review of recent works on the electrochemical exfoliation of

graphene materials from bulk graphite. Nanoscale, 2015. 7(16): p. 6944-6956.

74. Yang, S., Lohe, M.R., Müllen, K., and Feng, X., New-Generation Graphene from

Electrochemical Approaches: Production and Applications. Advanced Materials,

2016. 28(29): p. 6213-6221.

75. Abdelkader, A.M., Kinloch, I.A., and Dryfe, R.A.W., Continuous Electrochemical

Exfoliation of Micrometer-Sized Graphene Using Synergistic Ion Intercalations

and Organic Solvents. ACS Applied Materials & Interfaces, 2014. 6(3): p. 1632-

1639.

P a g e | 176

76. Parvez, K., Wu, e.-S., Li, R., Liu, X., Graf, R., Feng, X., and Müllen, K.,

Exfoliation of Graphite into Graphene in Aqueous Solutions of Inorganic Salts.

Journal of the American Chemical Society, 2014. 136(16): p. 6083-6091.

77. Beck, F., Jiang, J., and Krohn, H., Potential oscillations during galvanostatic

overoxidation of graphite in aqueous sulphuric acids. Journal of Electroanalytical

Chemistry, 1995. 389(1): p. 161-165.

78. Beck, F., Junge, H., and Krohn, H., Graphite intercalation compounds as positive

electrodes in galvanic cells. Electrochimica Acta, 1981. 26(7): p. 799-809.

79. Goss, C.A., Brumfield, J.C., Irene, E.A., and Murray, R.W., Imaging the incipient

electrochemical oxidation of highly oriented pyrolytic graphite. Analytical

Chemistry, 1993. 65(10): p. 1378-1389.

80. M. K, P.K., Shanthini, S., and Srivastava, C., Electrochemical exfoliation of

graphite for producing graphene using saccharin. RSC Advances, 2015. 5(66): p.

53865-53869.

81. Xia, e.Y., Pezzini, S., Treossi, E., Giambastiani, G., Corticelli, F., Morandi, V.,

eanelli, A., Bellani, V., and Palermo, V., The Exfoliation of Graphene in Liquids

by Electrochemical, Chemical, and Sonication-Assisted Techniques: A Nanoscale

Study. Advanced Functional Materials, 2013. 23(37): p. 4684-4693.

82. Chen, C.-H., Yang, S.-W., Chuang, M.-C., Woon, W.-Y., and Su, C.-Y., Towards

the continuous production of high crystallinity graphene via electrochemical

exfoliation with molecular in situ encapsulation. Nanoscale, 2015. 7(37): p.

15362-15373.

83. Wu, W., ehang, C., and Hou, S., Electrochemical exfoliation of graphene and

graphene-analogous 2D nanosheets. Journal of Materials Science, 2017. 52(18):

p. 10649-10660.

P a g e | 177

84. Huang, H., Xia, Y., Tao, X., Du, J., Fang, J., Gan, Y., and ehang, W., Highly

efficient electrolytic exfoliation of graphite into graphene sheets based on Li ions

intercalation–expansion–microexplosion mechanism. Journal of Materials

Chemistry, 2012. 22(21): p. 10452-10456.

85. Cooper, A.J., Wilson, N.R., Kinloch, I.A., and Dryfe, R.A.W., Single stage

electrochemical exfoliation method for the production of few-layer graphene via

intercalation of tetraalkylammonium cations. Carbon, 2014. 66: p. 340-350.

86. Wang, J., Manga, K.K., Bao, Q., and Loh, K.P., High-Yield Synthesis of Few-Layer

Graphene Flakes through Electrochemical Expansion of Graphite in Propylene

Carbonate Electrolyte. Journal of the American Chemical Society, 2011. 133(23):

p. 8888-8891.

87. Yang, Y., Lu, F., ehou, e., Song, W., Chen, Q., and Ji, X., Electrochemically

cathodic exfoliation of graphene sheets in room temperature ionic liquids N-butyl,

methylpyrrolidinium bis(trifluoromethylsulfonyl)imide and their electrochemical

properties. Electrochimica Acta, 2013. 113: p. 9-16.

88. Hummers, W.S. and Offeman, R.E., Preparation of Graphitic Oxide. Journal of

the American Chemical Society, 1958. 80(6): p. 1339-1339.

89. Li, D., Müller, M.B., Gilje, S., Kaner, R.B., and Wallace, G.G., Processable

aqueous dispersions of graphene nanosheets. Nature Nanotechnology, 2008. 3: p.

101.

90. Paredes, J.I., Villar-Rodil, S., Martínez-Alonso, A., and Tascón, J.M.D., Graphene

Oxide Dispersions in Organic Solvents. Langmuir, 2008. 24(19): p. 10560-10564.

91. Gao, X., Jang, J., and Nagase, S., Hydrazine and Thermal Reduction of Graphene

Oxide: Reaction Mechanisms, Product Structures, and Reaction Design. The

Journal of Physical Chemistry C, 2010. 114(2): p. 832-842.

P a g e | 178

92. Chua, C.K. and Pumera, M., The reduction of graphene oxide with hydrazine:

elucidating its reductive capability based on a reaction-model approach.

Chemical Communications, 2016. 52(1): p. 72-75.

93. Guex, L.G., Sacchi, B., Peuvot, K.F., Andersson, R.L., Pourrahimi, A.M., Ström,

V., Farris, S., and Olsson, R.T., Experimental review: chemical reduction of

graphene oxide (GO) to reduced graphene oxide (rGO) by aqueous chemistry.

Nanoscale, 2017. 9(27): p. 9562-9571.

94. Muda, M.R., Ramli, M.M., Isa, S.S.M., Jamlos, M.F., Murad, S.A.e., Norhanisah,

e., Isa, M.M., Kasjoo, S.R., Ahmad, N., Nor, N.I.M., and Khalid, N., Fundamental

study of reduction graphene oxide by sodium borohydride for gas sensor

application. AIP Conference Proceedings, 2017. 1808(1): p. 020034.

95. Moon, I.K., Lee, J., Ruoff, R.S., and Lee, H., Reduced graphene oxide by chemical

graphitization. Nature Communications, 2010. 1: p. 73.

96. Gómez-Navarro, C., Meyer, J.C., Sundaram, R.S., Chuvilin, A., Kurasch, S.,

Burghard, M., Kern, K., and Kaiser, U., Atomic Structure of Reduced Graphene

Oxide. Nano Letters, 2010. 10(4): p. 1144-1148.

97. ehang, Y., Li, D., Tan, X., ehang, B., Ruan, X., Liu, H., Pan, C., Liao, L., ehai,

T., Bando, Y., Chen, S., Cai, W., and Ruoff, R.S., High quality graphene sheets

from graphene oxide by hot-pressing. Carbon, 2013. 54: p. 143-148.

98. Rozada, R., Paredes, J.I., Villar-Rodil, S., Martínez-Alonso, A., and Tascón,

J.M.D., Towards full repair of defects in reduced graphene oxide films by two-step

graphitization. Nano Research, 2013. 6(3): p. 216-233.

99. Cabrero-Vilatela, A., Weatherup, R.S., Braeuninger-Weimer, P., Caneva, S., and

Hofmann, S., Towards a general growth model for graphene CVD on transition

metal catalysts. Nanoscale, 2016. 8(4): p. 2149-2158.

P a g e | 179

100. Lin, Y.-C., Jin, C., Lee, J.-C., Jen, S.-F., Suenaga, K., and Chiu, P.-W., Clean

Transfer of Graphene for Isolation and Suspension. ACS Nano, 2011. 5(3): p.

2362-2368.

101. Kim, K.S., ehao, Y., Jang, H., Lee, S.Y., Kim, J.M., Kim, K.S., Ahn, J.-H., Kim,

P., Choi, J.-Y., and Hong, B.H., Large-scale pattern growth of graphene films for

stretchable transparent electrodes. Nature, 2009. 457: p. 706.

102. Pirkle, A., Chan, J., Venugopal, A., Hinojos, D., Magnuson, C.W., McDonnell, S.,

Colombo, L., Vogel, E.M., Ruoff, R.S., and Wallace, R.M., The effect of chemical

residues on the physical and electrical properties of chemical vapor deposited

graphene transferred to SiO2. Applied Physics Letters, 2011. 99(12): p. 122108.

103. ehang, Y., ehang, L., and ehou, C., Review of Chemical Vapor Deposition of

Graphene and Related Applications. Accounts of Chemical Research, 2013.

46(10): p. 2329-2339.

104. Frigeri, P., Seravalli, L., Trevisi, G., and Franchi, S., Molecular Beam Epitaxy: An

Overview, in Reference Module in Materials Science and Materials Engineering.

2016, Elsevier.

105. Lopes, J.M.J. and Vignaud, D., Chapter 21 - Molecular Beam Epitaxy of

Graphene and Hexagonal Boron Nitride, in Molecular Beam Epitaxy (Second

Edition), M. Henini, Editor. 2018, Elsevier. p. 487-513.

106. Maeda, F. and Hibino, H., Growth of few-layer graphene by gas-source molecular

beam epitaxy using cracked ethanol. physica status solidi (b), 2010. 247(4): p.

916-920.

107. Maeda, F. and Hibino, H., Molecular beam epitaxial growth of graphene and

ridge-structure networks of graphene. Journal of Physics D: Applied Physics,

2011. 44(43): p. 435305.

P a g e | 180

108. Wang, L., Meric, I., Huang, P.Y., Gao, Q., Gao, Y., Tran, H., Taniguchi, T.,

Watanabe, K., Campos, L.M., Muller, D.A., Guo, J., Kim, P., Hone, J., Shepard,

K.L., and Dean, C.R., One-Dimensional Electrical Contact to a Two-Dimensional

Material. Science, 2013. 342(6158): p. 614-617.

109. Wang, J., Ma, F., and Sun, M., Graphene, hexagonal boron nitride, and their

heterostructures: properties and applications. RSC Advances, 2017. 7(27): p.

16801-16822.

110. Summerfield, A., Davies, A., Cheng, T.S., Korolkov, V.V., Cho, Y., Mellor, C.J.,

Foxon, C.T., Khlobystov, A.N., Watanabe, K., Taniguchi, T., Eaves, L., Novikov,

S.V., and Beton, P.H., Strain-Engineered Graphene Grown on Hexagonal Boron

Nitride by Molecular Beam Epitaxy. Scientific Reports, 2016. 6: p. 22440.

111. Dabrowski, J., Lippert, G., Schroeder, T., and Lupina, G., Role of defects in the

process of graphene growth on hexagonal boron nitride from atomic carbon.

Applied Physics Letters, 2014. 105(19): p. 191610.

112. Chechik, V., Carter, E., and Murphy, D., Electron Paramagnetic Resonance. 2016:

OUP Oxford.

113. eavoisky, E., Relaxation of liquid solutions for perpendicular fields. J. Phys.

USSR, 1945. 9: p. 211–216

114. Chechik, V. eprsimulator.org. [cited 2019; Available from:

http://www.eprsimulator.org/.

115. Sechovský, V., Magnetism in Solids: General Introduction, in Encyclopedia of

Materials: Science and Technology, K.H.J. Buschow, et al., Editors. 2001,

Elsevier: Oxford. p. 5018-5032.

116. Tilley, R.J.D., Magnetic Solids, in Understanding Solids. 2004. p. 363-390.

117. Lund, A., Shiotani, M., and Shimada, S., Principles and Applications of ESR

Spectroscopy. 2011: Springer.

P a g e | 181

118. Weber, R.T., Elexsys E 580 User's Manual. 2005.

119. Fujita, N., Matsumoto, D., Sakurai, Y., Kawahara, K., Ago, H., Takenobu, T., and

Marumoto, K., Direct observation of electrically induced Pauli paramagnetism in

single-layer graphene using ESR spectroscopy. Scientific Reports, 2016. 6: p.

34966.

120. Marciano, O., Gonen, S., Levy, N., Teblum, E., Yemini, R., Nessim, G.D.,

Ruthstein, S., and Elbaz, L., Modulation of Oxygen Content in Graphene Surfaces

Using Temperature-Programmed Reductive Annealing: Electron Paramagnetic

Resonance and Electrochemical Study. Langmuir, 2016. 32(44): p. 11672-11680.

121. Kempiński, M., Florczak, P., Jurga, S., Śliwińska-Bartkowiak, M., and Kempiński,

W., The impact of adsorption on the localization of spins in graphene oxide and

reduced graphene oxide, observed with electron paramagnetic resonance.

Applied Physics Letters, 2017. 111(8): p. 084102.

122. Yazyev, O.V., Emergence of magnetism in graphene materials and nanostructures.

Reports on Progress in Physics, 2010. 73(5): p. 056501.

123. Magda, G.e., Jin, X., Hagymási, I., Vancsó, P., Osváth, e., Nemes-Incze, P.,

Hwang, C., Biró, L.P., and Tapasztó, L., Room-temperature magnetic order on

zigzag edges of narrow graphene nanoribbons. Nature, 2014. 514: p. 608-611.

124. Saremi, S., RKKY in half-filled bipartite lattices: Graphene as an example. Phys.

Rev. B, 2007. 76: p. 184430.

125. Neto, A.H.C., Kotov, V.N., Nilsson, J., Pereira, V.M., Peres, N.M.R., and Uchoa,

B., Adatoms in Graphene. Solid State Communications, 2009. 149(27-28): p.

1094-1100.

126. Náfrádi, B., Forró, L., and Choucair, M., Electron spin lifetime in chemically

synthesized graphene sheets. physica status solidi (b), 2014. 251(12): p. 2521-

2524.

P a g e | 182

127. Márkus, B.G., Simon, F., Chacón-Torres, J.C., Reich, S., Szirmai, P., Náfrádi, B.,

Forró, L., Pichler, T., Vecera, P., Hauke, F., and Hirsch, A., Transport, magnetic

and vibrational properties of chemically exfoliated few-layer graphene. physica

status solidi (b), 2015. 252(11): p. 2438-2443.

128. Kausteklis, J., Cevc, P., Arčon, D., Nasi, L., Pontiroli, D., Mazzani, M., and Riccò,

M., Electron paramagnetic resonance study of nanostructured graphite. Physical

Review B, 2011. 84(12): p. 125406.

129. Barbon, A. and Brustolon, M., An EPR Study on Nanographites. Applied

Magnetic Resonance, 2012. 42(2): p. 197-210.

130. Sercheli, M.S., Kopelevich, Y., Silva, R.R.d., Torres, J.H.S., and Rettori, C.,

Evidence for internal field in graphite: a conduction electron spin-resonance

study. Solid State Communications, 2002. 121(9-10): p. 579-583.

131. Sercheli, M.S., Kopelevich, Y., Silva, R.R.d., Torres, J.H.S., and Rettori, C.,

Conduction electron spin resonance evidence for internal field in graphite.

Physica B: Condensed Matter, 2002. 320(1-4): p. 413-415.

132. Dyson, F.J., Electron Spin Resonance Absorption in Metals. II. Theory of Electron

Diffusion and the Skin Effect. Phys. Rev., 1955. 98: p. 349.

133. Tommasini, M., Castiglioni, C., eerbi, G., Barbon, A., and Brustolon, M., A joint

Raman and EPR spectroscopic study on ball-milled nanographites. Chemical

Physics Letters, 2011. 516(4–6): p. 220-224.

134. Raman, C.V. and Krishnan, K.S., A New Type of Secondary Radiation. Nature,

1928. 121: p. 501-502.

135. Qian, C.-G., ehu, S., Feng, P.-J., Chen, Y.-L., Yu, J.-C., Tang, X., Liu, Y., and Shen,

Q.-D., Conjugated Polymer Nanoparticles for Fluorescence Imaging and Sensing

of Neurotransmitter Dopamine in Living Cells and the Brains of Zebrafish Larvae.

ACS Appl. Mater. Interfaces, 2015. 7(33): p. 18581-18589.

P a g e | 183

136. Aggarwal, A.V., Thiessen, A., Idelson, A., Kalle, D., Würsch, D., Stangl, T.,

Steiner, F., Jester, S.-S., Vogelsang, J., Höger, S., and Lupton, J.M., Fluctuating

exciton localization in giant π-conjugated spoked-wheel macrocycles. Nature

Chemistry, 2013. 5: p. 964-970.

137. Lichtman, J.W. and Conchello, J.-A., Fluorescence microscopy. Nature Methods,

2005. 2: p. 910-919.

138. Reichenbacher, M. and Popp, J., Challenges in Molecular Structure

Determination. 2012: Springer.

139. Ferraro, J.R., Nakamoto, K., and Brown, C.W., Introductory Raman Spectroscopy.

2 ed.: Academic Press.

140. Smith, E. and Dent, G., Introduction, Basic Theory and Principles, in Modern

Raman Spectroscopy. 2019, John Wiley & Sons Ltd: The Atrium, Southern Gate,

Chichester, West Sussex, PO19 8SQ, UK. p. 1-20.

141. Casiraghi, C., Hartschuh, A., Qian, H., Piscanec, S., Georgi, C., Fasoli, A.,

Novoselov, K.S., Basko, D.M., and Ferrari, A.C., Raman Spectroscopy of

Graphene Edges. Nano Letters, 2009. 9(4): p. 1433-1441.

142. Eckmann, A., Felten, A., Verzhbitskiy, I., Davey, R., and Casiraghi, C., Raman

study on defctive graphene: Effect of the exitation energy, type and amount of

defects. Phys. Rev. B, 2013. 88: p. 035426.

143. Rao, F.B., Almumen, H., Fan, e., Li, W., and Dong, L.X., Inter-sheet-effect-

inspired graphene sensors: design, fabrication and characterization.

Nanotechnology, 2012. 23(10): p. 105501.

144. Ferrari, A.C., Meyer, J.C., Scardaci, V., Casiraghi, C., Lazzeri, M., Mauri, F.,

Piscanec, S., Jiang, D., Novoselov, K.S., Roth, S., and Geim, A.K., Raman

Spectrum of Graphene and Graphene Layers. Phys. Rev. Lett., 2006. 97: p.

187401.

P a g e | 184

145. Tuinstra, F. and Koenig, J.L., Raman Spectrum of Graphite. J. Chem. Phys., 1970.

53: p. 1126.

146. Thomsen, C. and Reich, S., Double Resonant Raman Scattering in Graphite. Phys.

Rev. Lett., 2000. 85: p. 5214.

147. Ferrari, A.C. and Basko, D.M., Raman spectroscopy as a versatile tool for

studying the properties of graphene. Nature Nanotechnology, 2013. 8: p. 235-246.

148. Malard, L.M., Pimenta, M.A., Dresselhaus, G., and Dresselhaus, M.S., Raman

spectroscopy in graphene. Physics Reports, 2009. 473: p. 51-87.

149. Venezuela, P., Lazzeri, M., and Mauri, F., Theory of double-resonant Raman

spectra in graphene: Intensity and line shape of defect-induced and two-phonon

bands. Physical Review B, 2011. 84(3): p. 035433.

150. Eckmann, A., Felten, A., Mishchenko, A., Britnell, L., Krupke, R., Novoselov,

K.S., and Casiraghi, C., Probing the Nature of Defects in Graphene by Raman

Spectroscopy. Nano Letters, 2012. 12(8): p. 3925-3930.

151. Haar, S., Gemayel, M.E., Shin, Y., Melinte, G., Squillaci, M.A., Ersen, O.,

Casiraghi, C., Ciesielski, A., and Samorì, P., Enhancing the Liquid-Phase

Exfoliation of Graphene in Organic Solvents upon Addition of n-Octylbenzene.

Scientific Reports, 2015. 5: p. 16684.

152. Carozo, V., Almeida, C.M., Ferreira, E.H.M., Cançado, L.G., Achete, C.A., and

Jorio, A., Raman Signature of Graphene Superlattices. Nano Letters, 2011.

11(11): p. 4527-4534.

153. Carozo, V., Almeida, C.M., Fragneaud, B., Bedê, P.M., Moutinho, M.V.O.,

Ribeiro-Soares, J., Andrade, N.F., Filho, A.G.S., Matos, M.J.S., Wang, B.,

Terrones, M., Capaz, R.B., Jorio, A., Achete, C.A., and Cançado, L.G., Resonance

effects on the Raman spectra of graphene superlattices. Phys. Rev. B, 2013. 88(8):

p. 085401.

P a g e | 185

154. Haar, S., Ciesielski, A., Clough, J., Yang, H., Mazzaro, R., Richard, F., Conti, S.,

Merstorf, N., Cecchini, M., Morandi, V., Casiraghi, C., and Samorì, P., A

Supramolecular Strategy to Leverage the Liquid-Phase Exfoliation of Graphene

in the Presence of Surfactants: Unraveling the Role of the Length of Fatty Acids.

Small, 2015. 11(14): p. 1691-1702.

155. Ciesielski, A., Haar, S., Gemayel, M.E., Yang, H., Clough, J., Melinte, G., Gobbi,

M., Orgiu, E., Nardi, M.V., Ligorio, G., Palermo, V., Koch, N., Ersen, O.,

Casiraghi, C., and Samorì, P., Harnessing the Liquid-Phase Exfoliation of

Graphene Using Aliphatic Compounds: A Supramolecular Approach. Angew.

Chem. Int. Ed. Engl., 2014. 53(39): p. 10355-10361.

156. Ferrari, A.C., Bonaccorso, F., Fal'ko, V., Novoselov, K.S., Roche, S., Boggild, P.,

Borini, S., Koppens, F.H.L., Palermo, V., Pugno, N., Garrido, J.A., Sordan, R.,

Bianco, A., Ballerini, L., Prato, M., Lidorikis, E., Kivioja, J., Marinelli, C.,

Ryhanen, T., Morpurgo, A., Coleman, J.N., Nicolosi, V., Colombo, L., Fert, A.,

Garcia-Hernandez, M., Bachtold, A., Schneider, G.F., Guinea, F., Dekker, C.,

Barbone, M., Sun, e., Galiotis, C., Grigorenko, A.N., Konstantatos, G., Kis, A.,

Katsnelson, M., Vandersypen, L., Loiseau, A., Morandi, V., Neumaier, D., Treossi,

E., Pellegrini, V., Polini, M., Tredicucci, A., Williams, G.M., Hee Hong, B., Ahn,

J.-H., Min Kim, J., eirath, H., van Wees, B.J., van der eant, H., Occhipinti, L., Di

Matteo, A., Kinloch, I.A., Seyller, T., Quesnel, E., Feng, X., Teo, K., Rupesinghe,

N., Hakonen, P., Neil, S.R.T., Tannock, Q., Lofwander, T., and Kinaret, J., Science

and technology roadmap for graphene, related two-dimensional crystals, and

hybrid systems. Nanoscale, 2015. 7(11): p. 4598-4810.

157. Tombros, N., Jozsa, C., Popinciuc, M., Jonkman, H.T., and van Wees, B.J.,

Electronic spin transport and spin precession in single graphene layers at room

temperature. Nature, 2007. 448(7153): p. 571-574.

P a g e | 186

158. Han, W. and Kawakami, R.K., Spin Relaxation in Single-Layer and Bilayer

Graphene. Physical Review Letters, 2011. 107(4): p. 047207.

159. Dlubak, B., Martin, M.-B., Deranlot, C., Servet, B., Xavier, S., Mattana, R.,

Sprinkle, M., Berger, C., De Heer, W.A., Petroff, F., Anane, A., Seneor, P., and

Fert, A., Highly efficient spin transport in epitaxial graphene on SiC. Nat Phys,

2012. 8(7): p. 557-561.

160. Huertas-Hernando, D., Guinea, F., and Brataas, A., Spin-orbit coupling in curved

graphene, fullerenes, nanotubes, and nanotube caps. Physical Review B, 2006.

74(15): p. 155426.

161. Yazyev, O.V., Magnetism in Disordered Graphene and Irradiated Graphite.

Physical Review Letters, 2008. 101(3): p. 037203.

162. Oleg, V.Y., Emergence of magnetism in graphene materials and nanostructures.

Reports on Progress in Physics, 2010. 73(5): p. 056501.

163. Lee, H., Son, Y.-W., Park, N., Han, S., and Yu, J., Magnetic ordering at the edges

of graphitic fragments: Magnetic tail interactions between the edge-localized

states. Physical Review B, 2005. 72(17): p. 174431.

164. Hou, X.-L., Li, J.-L., Drew, S.C., Tang, B., Sun, L., and Wang, X.-G., Tuning

Radical Species in Graphene Oxide in Aqueous Solution by Photoirradiation. The

Journal of Physical Chemistry C, 2013. 117(13): p. 6788-6793.

165. Slota, M., Keerthi, A., Myers, W.K., Tretyakov, E., Baumgarten, M., Ardavan, A.,

Sadeghi, H., Lambert, C.J., Narita, A., Müllen, K., and Bogani, L., Magnetic edge

states and coherent manipulation of graphene nanoribbons. Nature, 2018.

557(7707): p. 691-695.

166. Wang, Y., Huang, Y., Song, Y., ehang, X., Ma, Y., Liang, J., and Chen, Y., Room-

Temperature Ferromagnetism of Graphene. Nano Letters, 2009. 9(1): p. 220-224.

P a g e | 187

167. Matte, H.S.S.R., Subrahmanyam, K.S., and Rao, C.N.R., Novel Magnetic

Properties of Graphene: Presence of Both Ferromagnetic and Antiferromagnetic

Features and Other Aspects. The Journal of Physical Chemistry C, 2009. 113(23):

p. 9982-9985.

168. Ney, A., Papakonstantinou, P., Kumar, A., Shang, N.-G., and Peng, N., Irradiation

enhanced paramagnetism on graphene nanoflakes. Applied Physics Letters, 2011.

99(10): p. 102504.

169. Wagoner, G., Spin Resonance of Charge Carriers in Graphite. Phys. Rev., 1960.

118: p. 647.

170. Matsubara, K., Tsuzuku, T., and Sugihara, K., Electron spin resonance in graphite.

Physical Review B, 1991. 44(21): p. 11845-11851.

171. Beuneu, F., l’Huillier, C., Salvetat, J.P., Bonard, J.M., and Forró, L., Modification

of multiwall carbon nanotubes by electron irradiation: An ESR study. Physical

Review B, 1999. 59(8): p. 5945-5949.

172. Eaton, S.S. and Eaton, G.R., Relaxation Times of Organic Radicals and Transition

Metal Ions, in Distance Measurements in Biological Systems by EPR, L.J. Berliner,

G.R. Eaton, and S.S. Eaton, Editors. 2000, Springer US: Boston, MA. p. 29-154.

173. Gupta, A., Chen, G., Joshi, P., Tadigadapa, S., and Eklund, Raman Scattering from

High-Frequency Phonons in Supported n-Graphene Layer Films. Nano Letters,

2006. 6(12): p. 2667-2673.

174. Cameron, J.S., Ashley, D.S., Andrew, J.S., Joseph, G.S., and Christopher, T.G.,

Accurate thickness measurement of graphene. Nanotechnology, 2016. 27(12): p.

125704.

175. Gareth R. Eaton, Sandra S. Eaton, David P. Barr, and Ralph T. Weber, Standard

Samples, in Quantitative EPR. 2010, Springer: Vienna.

176. Stoll, S. Easyspin. [cited 2019 ]; Available from: http://www.easyspin.org/.

P a g e | 188

177. S. Chehab, K. Guérin, J. Amiell, and Flandrois, S., Magnetic properties of mixed

graphite containing both hexagonal and rhombohedral forms. Eur. Phys. J. B,

2000. 13(2): p. 235-243.

178. Singer, L.S. and Wagoner, G., Electron Spin Resonance in Polycrystalline

Graphite. The Journal of Chemical Physics, 1962. 37(8): p. 1812-1817.

179. Collins, M., Barklie, R.C., Anguita, J.V., Carey, J.D., and Silva, S.R.P.,

Characterisation of defects in thin films of hydrogenated amorphous carbon.

Diamond and Related Materials, 2000. 9(3): p. 781-785.

180. Smith, C.I., Miyaoka, H., Ichikawa, T., Jones, M.O., Harmer, J., Ishida, W.,

Edwards, P.P., Kojima, Y., and Fuji, H., Electron Spin Resonance Investigation of

Hydrogen Absorption in Ball-Milled Graphite. The Journal of Physical Chemistry

C, 2009. 113(14): p. 5409-5416.

181. Osipov, V.Y., Shames, A.I., Enoki, T., Takai, K., Endo, M., Hayashi, T., Kaburagi,

Y., and Vul, A.Y., Magnetic and EPR studies of edge-localized spin

paramagnetism in multi-shell nanographites derived from nanodiamonds.

Diamond and Related Materials, 2009. 18(2–3): p. 220-223.

182. Yi, M., Liang, S., Liu, L., Shen, e., eheng, Y., ehang, X., and Ma, S., Investigating

the Nature of Graphene-Based Films Prepared by Vacuum Filtration of Graphene

Dispersions. Journal of Nanoscience and Nanotechnology, 2014. 14(7): p. 4969-

4975.

183. Partoens, B. and Peeters, F.M., From graphene to graphite: Electronic structure

around the K point. Physical Review B, 2006. 74(7): p. 075404.

184. Matte, H.S.S.R., Subrahmanyam, K.S., and Rao, C.N.R., Novel Magnetic

Properties of Graphene: Presence of Both Ferromagnetic and Antiferromagnetic

Features and Other Aspects. The Journal of Physical Chemistry C, 2009.

0(proofing): p. null.

P a g e | 189

185. Coroş, M., Pogăcean, F., Măgeruşan, L., Socaci, C., and Pruneanu, S., A brief

overview on synthesis and applications of graphene and graphene-based

nanomaterials. Frontiers of Materials Science, 2019. 13(1): p. 23-32.

186. Lee, Y., Bae, S., Jang, H., Jang, S., ehu, S.-E., Sim, S.H., Song, Y.I., Hong, B.H.,

and Ahn, J.-H., Wafer-Scale Synthesis and Transfer of Graphene Films. Nano

Letters, 2010. 10(2): p. 490-493.

187. Eda, G., Fanchini, G., and Chhowalla, M., Large-area ultrathin films of reduced

graphene oxide as a transparent and flexible electronic material. Nature

Nanotechnology, 2008. 3: p. 270.

188. Aunkor, M.T.H., Mahbubul, I.M., Saidur, R., and Metselaar, H.S.C., The green

reduction of graphene oxide. RSC Advances, 2016. 6(33): p. 27807-27828.

189. Ortega Amaya, R., Matsumoto, Y., Diaz-Torres, E., Gutiérrez Lazos, C., Perez

Guzman, M.A., and Ortega-López, M., Green Routes for Graphene Oxide

Reduction and Self- Assembled Graphene Oxide Micro- and Nanostructures

Production. 2017. p. 129-151.

190. De Silva, K.K.H., Huang, H.H., Joshi, R.K., and Yoshimura, M., Chemical

reduction of graphene oxide using green reductants. Carbon, 2017. 119: p. 190-

199.

191. Backes, C., Szydłowska, B.M., Harvey, A., Yuan, S., Vega-Mayoral, V., Davies,

B.R., ehao, P.-l., Hanlon, D., Santos, E.J.G., Katsnelson, M.I., Blau, W.J.,

Gadermaier, C., and Coleman, J.N., Production of Highly Monolayer Enriched

Dispersions of Liquid-Exfoliated Nanosheets by Liquid Cascade Centrifugation.

ACS Nano, 2016. 10(1): p. 1589-1601.

192. Ejigu, A., Miller, B., Kinloch, I.A., and Dryfe, R.A.W., Optimisation of

electrolytic solvents for simultaneous electrochemical exfoliation and

P a g e | 190

functionalisation of graphene with metal nanostructures. Carbon, 2018. 128: p.

257-266.

193. Achee, T.C., Sun, W., Hope, J.T., Quitzau, S.G., Sweeney, C.B., Shah, S.A., Habib,

T., and Green, M.J., High-yield scalable graphene nanosheet production from

compressed graphite using electrochemical exfoliation. Scientific Reports, 2018.

8(1): p. 14525.

194. Wang, H., Wei, C., ehu, K., ehang, Y., Gong, C., Guo, J., ehang, J., Yu, L., and

ehang, J., Preparation of Graphene Sheets by Electrochemical Exfoliation of

Graphite in Confined Space and Their Application in Transparent Conductive

Films. ACS Applied Materials & Interfaces, 2017. 9(39): p. 34456-34466.

195. Munuera, J.M., Paredes, J.I., Enterría, M., Pagán, A., Villar-Rodil, S., Pereira,

M.F.R., Martins, J.I., Figueiredo, J.L., Cenis, J.L., Martínez-Alonso, A., and

Tascón, J.M.D., Electrochemical Exfoliation of Graphite in Aqueous Sodium

Halide Electrolytes toward Low Oxygen Content Graphene for Energy and

Environmental Applications. ACS Applied Materials & Interfaces, 2017. 9(28): p.

24085-24099.

196. Huang, X., Li, S., Qi, e., ehang, W., Ye, W., and Fang, Y., Low defect

concentration few-layer graphene using a two-step electrochemical exfoliation.

Nanotechnology, 2015. 26(10): p. 105602.

197. Liu, L., ehou, H., Cheng, R., Yu, W.J., Liu, Y., Chen, Y., Shaw, J., ehong, X.,

Huang, Y., and Duan, X., High-Yield Chemical Vapor Deposition Growth of High-

Quality Large-Area AB-Stacked Bilayer Graphene. ACS Nano, 2012. 6(9): p.

8241-8249.

198. ehong, Y.L. and Swager, T.M., Enhanced Electrochemical Expansion of Graphite

for in Situ Electrochemical Functionalization. Journal of the American Chemical

Society, 2012. 134(43): p. 17896-17899.

P a g e | 191

199. ehou, M., Tang, J., Cheng, Q., Xu, G., Cui, P., and Qin, L.-C., Few-layer graphene

obtained by electrochemical exfoliation of graphite cathode. Chemical Physics

Letters, 2013. 572: p. 61-65.

200. Krivenko, A.G., Manzhos, R.A., Kotkin, A.S., Kochergin, V.K., Piven, N.P., and

Manzhos, A.P., Production of few-layer graphene structures in different modes of

electrochemical exfoliation of graphite by voltage pulses. Instrumentation Science

& Technology, 2019. 47(5): p. 535-544.

201. ehang, W., eeng, Y., Xiao, N., Hng, H.H., and Yan, Q., One-step electrochemical

preparation of graphene-based heterostructures for Li storage. Journal of

Materials Chemistry, 2012. 22(17): p. 8455-8461.

202. Stubrov, Y., Nikolenko, A., Strelchuk, V., Nedilko, S., and Chornii, V., Structural

Modification of Single-Layer Graphene Under Laser Irradiation Featured by

Micro-Raman Spectroscopy. Nanoscale Research Letters, 2017. 12(1): p. 297.

203. Iqbal, M.W., Singh, A.K., Iqbal, M.e., and Eom, J., Raman fingerprint of doping

due to metal adsorbates on graphene. Journal of Physics: Condensed Matter, 2012.

24(33): p. 335301.

204. Iqbal, M.W., Iqbal, M.e., Khan, M.F., Jin, X., Hwang, C., and Eom, J.,

Modification of the structural and electrical properties of graphene layers by Pt

adsorbates. Science and technology of advanced materials, 2014. 15(5): p.

055002-055002.

205. eheng, X., Chen, W., Wang, G., Yu, Y., Qin, S., Fang, J., Wang, F., and ehang, X.-

A., The Raman redshift of graphene impacted by gold nanoparticles. AIP

Advances, 2015. 5(5): p. 057133.

206. Mueller, N.S., Heeg, S., Alvarez, M.P., Kusch, P., Wasserroth, S., Clark, N.,

Schedin, F., Parthenios, J., Papagelis, K., Galiotis, C., Kalbáč, M., Vijayaraghavan,

A., Huebner, U., Gorbachev, R., Frank, O., and Reich, S., Evaluating arbitrary

P a g e | 192

strain configurations and doping in graphene with Raman spectroscopy. 2D

Materials, 2017. 5(1): p. 015016.

207. Wu, J.-B., Lin, M.-L., Cong, X., Liu, H.-N., and Tan, P.-H., Raman spectroscopy

of graphene-based materials and its applications in related devices. Chemical

Society Reviews, 2018. 47(5): p. 1822-1873.

208. Mao, S., Pu, H., and Chen, J., Graphene oxide and its reduction: modeling and

experimental progress. RSC Advances, 2012. 2(7): p. 2643-2662.

209. Kim, K.H., Yang, M., Cho, K.M., Jun, Y.-S., Lee, S.B., and Jung, H.-T., High

quality reduced graphene oxide through repairing with multi-layered graphene

ball nanostructures. Scientific Reports, 2013. 3: p. 3251.

210. Diamantopoulou, Α., Glenis, S., eolnierkiwicz, G., Guskos, N., and Likodimos,

V., Magnetism in pristine and chemically reduced graphene oxide. Journal of

Applied Physics, 2017. 121(4): p. 043906.

211. Mizushima, S., Electron spin resonance in graphite damaged heavily in reactor.

Carbon, 1968. 6(1): p. 13-17.

212. Yazyev, O.V. and Katsnelson, M.I., Magnetic Correlations at Graphene Edges:

Basis for Novel Spintronics Devices. Physical Review Letters, 2008. 100(4): p.

047209.

213. Son, Y.-W., Cohen, M.L., and Louie, S.G., Half-metallic graphene nanoribbons.

Nature, 2006. 444: p. 347.

214. ehu, J., Yan, S., Feng, N., Ye, L., Ou, J.-Y., and Liu, Q.H., Near unity ultraviolet

absorption in graphene without patterning. Applied Physics Letters, 2018.

112(15): p. 153106.

215. Mak, K.F., Ju, L., Wang, F., and Heinz, T.F., Optical spectroscopy of graphene:

From the far infrared to the ultraviolet. Solid State Communications, 2012.

152(15): p. 1341-1349.

P a g e | 193

216. Abergel, D.S.L. and Fal’ko, V.I., Optical and magneto-optical far-infrared

properties of bilayer graphene. Physical Review B, 2007. 75(15): p. 155430.

217. Kiisk, V., Kahro, T., Kozlova, J., Matisen, L., and Alles, H., Nanosecond laser

treatment of graphene. Applied Surface Science, 2013. 276: p. 133-137.

218. David, L., Feldman, A., Mansfield, E., Lehman, J., and Singh, G., Evaluating the

thermal damage resistance of graphene/carbon nanotube hybrid composite

coatings. Scientific Reports, 2014. 4: p. 4311.

219. Malekpour, H., Ramnani, P., Srinivasan, S., Balasubramanian, G., Nika, D.L.,

Mulchandani, A., Lake, R.K., and Balandin, A.A., Thermal conductivity of

graphene with defects induced by electron beam irradiation. Nanoscale, 2016.

8(30): p. 14608-14616.

220. Gil-Villalba, A., Meyer, R., Giust, R., Rapp, L., Billet, C., and Courvoisier, F.,

Single shot femtosecond laser nano-ablation of CVD monolayer graphene.

Scientific Reports, 2018. 8(1): p. 14601.

221. Yeo, S., Han, J., Bae, S., and Lee, D.S., Coherence in defect evolution data for the

ion beam irradiated graphene. Scientific Reports, 2018. 8(1): p. 13973.

222. Yang, Y., Chen, L., Li, D.-Y., Yi, R.-B., Mo, J.-W., Wu, M.-H., and Xu, G.,

Controllable reduction of graphene oxide by electron-beam irradiation. RSC

Advances, 2019. 9(7): p. 3597-3604.

223. Schedin, F., Geim, A.K., Morozov, S.V., Hill, E.W., Blake, P., Katsnelson, M.I.,

and Novoselov, K.S., Detection of individual gas molecules adsorbed on

graphene. Nature Materials, 2007. 6: p. 652.

224. Roberts, A., Cormode, D., Reynolds, C., Newhouse-Illige, T., LeRoy, B.J., and

Sandhu, A.S., Response of graphene to femtosecond high-intensity laser

irradiation. Applied Physics Letters, 2011. 99(5): p. 051912.

P a g e | 194

225. Currie, M., Caldwell, J.D., Bezares, F.J., Robinson, J., Anderson, T., Chun, H.,

and Tadjer, M., Quantifying pulsed laser induced damage to graphene. Applied

Physics Letters, 2011. 99(21): p. 211909.

226. Jeschke, H.O., Garcia, M.E., and Bennemann, K.H., Theory for the Ultrafast

Ablation of Graphite Films. Physical Review Letters, 2001. 87(1): p. 015003.

227. Araujo, P.T., Terrones, M., and Dresselhaus, M.S., Defects and impurities in

graphene-like materials. Materials Today, 2012. 15(3): p. 98-109.

228. Feng, W., Long, P., Feng, Y., and Li, Y., Two-Dimensional Fluorinated Graphene:

Synthesis, Structures, Properties and Applications. Advanced Science, 2016. 3(7):

p. 1500413.

229. Nair, R.R., Ren, W., Jalil, R., Riaz, I., Kravets, V.G., Britnell, L., Blake, P.,

Schedin, F., Mayorov, A.S., Yuan, S., Katsnelson, M.I., Cheng, H.-M., Strupinski,

W., Bulusheva, L.G., Okotrub, A.V., Grigorieva, I.V., Grigorenko, A.N.,

Novoselov, K.S., and Geim, A.K., Fluorographene: A Two-Dimensional

Counterpart of Teflon. Small, 2010. 6(24): p. 2877-2884.

230. Balog, R., Jørgensen, B., Nilsson, L., Andersen, M., Rienks, E., Bianchi, M.,

Fanetti, M., Lægsgaard, E., Baraldi, A., Lizzit, S., Sljivancanin, e., Besenbacher,

F., Hammer, B., Pedersen, T.G., Hofmann, P., and Hornekær, L., Bandgap opening

in graphene induced by patterned hydrogen adsorption. Nature Materials, 2010.

9: p. 315.

231. Guan, e., Ni, S., and Hu, S., Band gap opening of graphene by forming a

graphene/PtSe2 van der Waals heterojunction. RSC Advances, 2017. 7(72): p.

45393-45399.

232. Kang, W. and Li, S., Preparation of fluorinated graphene to study its gas

sensitivity. RSC Advances, 2018. 8(41): p. 23459-23467.

P a g e | 195

233. Rivera, L.M., García, G., and Pastor, E., Novel graphene materials for the oxygen

reduction reaction. Current Opinion in Electrochemistry, 2018. 9: p. 233-239.

234. Kansara, V., Patil, R., Tripathi, R., Jha, P.K., Bahadur, P., and Tiwari, S.,

Functionalized graphene nanosheets with improved dispersion stability and

superior paclitaxel loading capacity. Colloids and Surfaces B: Biointerfaces,

2019. 173: p. 421-428.

235. Chronopoulos, D.D., Bakandritsos, A., Pykal, M., ebořil, R., and Otyepka, M.,

Chemistry, properties, and applications of fluorographene. Applied Materials

Today, 2017. 9: p. 60-70.

236. Karlický, F., Kumara Ramanatha Datta, K., Otyepka, M., and ebořil, R.,

Halogenated Graphenes: Rapidly Growing Family of Graphene Derivatives. ACS

Nano, 2013. 7(8): p. 6434-6464.

237. Bulusheva, L.G. and Okotrub, A.V., 8 - Electronic Structure of Fluorinated

Graphene, in New Fluorinated Carbons: Fundamentals and Applications, O.V.

Boltalina and T. Nakajima, Editors. 2017, Elsevier: Boston. p. 177-213.

238. Liu, H.Y., Hou, e.F., Hu, C.H., Yang, Y., and ehu, e.e., Electronic and Magnetic

Properties of Fluorinated Graphene with Different Coverage of Fluorine. The

Journal of Physical Chemistry C, 2012. 116(34): p. 18193-18201.

239. Amatucci, G.G. and Pereira, N., Fluoride based electrode materials for advanced

energy storage devices. Journal of Fluorine Chemistry, 2007. 128(4): p. 243-262.

240. Peng, S., Yan, S., Wang, N., Nan, W., Wang, J., Chen, X., Wang, C., Qi, X., and

Dai, S., Fluorinated graphene/sulfur hybrid cathode for high energy and high

power density lithium primary batteries. RSC Advances, 2018. 8(23): p. 12701-

12707.

241. Wang, X., Wang, W., Liu, Y., Ren, M., Xiao, H., and Liu, X., Controllable

defluorination of fluorinated graphene and weakening of C–F bonding under the

P a g e | 196

action of nucleophilic dipolar solvent. Physical Chemistry Chemical Physics,

2016. 18(4): p. 3285-3293.

242. Lai, W., Liu, J., Luo, L., Wang, X., He, T., Fan, K., and Liu, X., The Friedel–

Crafts reaction of fluorinated graphene for high-yield arylation of graphene.

Chemical Communications, 2018. 54(72): p. 10168-10171.

243. Medveď, M., eoppellaro, G., Ugolotti, J., Matochová, D., Lazar, P., Pospíšil, T.,

Bakandritsos, A., Tuček, J., ebořil, R., and Otyepka, M., Reactivity of

fluorographene is triggered by point defects: beyond the perfect 2D world.

Nanoscale, 2018. 10(10): p. 4696-4707.

244. Pumera, M. and Sofer, e., Towards stoichiometric analogues of graphene:

graphane, fluorographene, graphol, graphene acid and others. Chemical Society

Reviews, 2017. 46(15): p. 4450-4463.

245. Nguyen, D.K., Lin, Y.-T., Lin, S.-Y., Chiu, Y.-H., Tran, N.T.T., and Fa-Lin, M.,

Fluorination-enriched electronic and magnetic properties in graphene

nanoribbons. Physical Chemistry Chemical Physics, 2017. 19(31): p. 20667-

20676.

246. Tran, N.T.T., Nguyen, D.K., Glukhova, O.E., and Lin, M.-F., Coverage-dependent

essential properties of halogenated graphene: A DFT study. Scientific Reports,

2017. 7(1): p. 17858.

247. eheng, Y., Wan, X., Tang, N., Feng, Q., Liu, F., and Du, Y., Magnetic properties

of double-side partially fluorinated graphene from first principles calculations.

Carbon, 2015. 89: p. 300-307.

248. Mazánek, V., Jankovský, O., Luxa, J., Sedmidubský, D., Janoušek, e., Šembera,

F., Mikulics, M., and Sofer, e., Tuning of fluorine content in graphene: towards

large-scale production of stoichiometric fluorographene. Nanoscale, 2015. 7(32):

p. 13646-13655.

P a g e | 197

249. Gong, P., Wang, e., Wang, J., Wang, H., Li, e., Fan, e., Xu, Y., Han, X., and Yang,

S., One-pot sonochemical preparation of fluorographene and selective tuning of

its fluorine coverage. Journal of Materials Chemistry, 2012. 22(33): p. 16950-

16956.

250. Mallouk, T., Hawkins, B.L., Conrad, M.P., eilm, K., Maciel, G.E., Bartlett, N.,

Gillespie, R.J., and Day, P., Raman, infrared and n.m.r. studies of the graphite

hydrofluorides C<sub>x<sub>F</sub>1-

&#x3b4;</sub>(HF)<sub>&#x3b4;</sub>(2 &#x2264; x &#x2264; 5).

Philosophical Transactions of the Royal Society of London. Series A,

Mathematical and Physical Sciences, 1985. 314(1528): p. 179-187.

251. Asanov, I.P., Bulusheva, L.G., Dubois, M., Yudanov, N.F., Alexeev, A.V.,

Makarova, T.L., and Okotrub, A.V., Graphene nanochains and nanoislands in the

layers of room-temperature fluorinated graphite. Carbon, 2013. 59: p. 518-529.

252. Yudanov, N.F. and Chernyavskii, L.I., Model for the structures of intercalation

compounds based on graphite fluoride. Journal of Structural Chemistry, 1988.

28(4): p. 534-541.

253. Claves, D. and Rossignol, J., Fluorine addition to single-wall carbon nanotubes

revisited. Chemical Physics Letters, 2009. 468(4): p. 231-233.

254. Claves, D., Spectroscopic study of fluorinated carbon nanostructures. New

Journal of Chemistry, 2011. 35(11): p. 2477-2482.

255. Chamssedine, F. and Claves, D., Three different modes of fluorine chemisorption

at the surface of single wall carbon nanotubes. Chemical Physics Letters, 2007.

443(1): p. 102-106.

256. Ho, K.-I., Huang, C.-H., Liao, J.-H., ehang, W., Li, L.-J., Lai, C.-S., and Su, C.-

Y., Fluorinated Graphene as High Performance Dielectric Materials and the

Applications for Graphene Nanoelectronics. Scientific Reports, 2014. 4: p. 5893.

P a g e | 198

257. Jeon, K.-J., Lee, e., Pollak, E., Moreschini, L., Bostwick, A., Park, C.-M.,

Mendelsberg, R., Radmilovic, V., Kostecki, R., Richardson, T.J., and Rotenberg,

E., Fluorographene: A Wide Bandgap Semiconductor with Ultraviolet

Luminescence. ACS Nano, 2011. 5(2): p. 1042-1046.

258. Barklie, R.C., Characterisation of defects in amorphous carbon by electron

paramagnetic resonance. Diamond and Related Materials, 2001. 10(2): p. 174-

181.

259. Tadyszak, K., Augustyniak-Jabłokow, M.A., Więckowski, A.B., Najder-

Kozdrowska, L., Strzelczyk, R., and Andrzejewski, B., Origin of electron

paramagnetic resonance signal in anthracite. Carbon, 2015. 94: p. 53-59.

260. Silva, I.D.A., Donoso, J.P., Magon, C.J., Tambelli, C.E., Santagneli, S.H., Ribeiro,

S.J.L., Silva, M.A.P., Chiesa, M., and Rodrigues, A.C.M., Magnetic Resonance

and Conductivity Study of Lead–Cadmium Fluorosilicate Glasses and Glass-

Ceramics. The Journal of Physical Chemistry C, 2018. 122(11): p. 6288-6297.

261. Ikoma, T., Ito, O., Tero-Kubota, S., and Akiyama, K., HYSCORE Study on Coal

Radicals. Energy & Fuels, 1998. 12(6): p. 1363-1368.

262. Möser, J., Lips, K., Tseytlin, M., Eaton, G.R., Eaton, S.S., and Schnegg, A., Using

rapid-scan EPR to improve the detection limit of quantitative EPR by more than

one order of magnitude. Journal of Magnetic Resonance, 2017. 281: p. 17-25.

263. Topsakal, M., Aktürk, E., and Ciraci, S., First-principles study of two- and one-

dimensional honeycomb structures of boron nitride. Physical Review B, 2009.

79(11): p. 115442.

P a g e | 199

Appendix A

Supporting Figures

Figure A1 Graphene dispersions (a) and graphene laminates (b)

Figure A2 The schematic of xyz plane of graphene laminate after it cut into ±2 mm width.

Rectangular blocks represent the pole faces of the magnet

b) a)

P a g e | 200

Figure A3 Representative Raman spectra obtained from graphene dispersions.

344 346 348 350 352 354 356

a single laminate

11 laminates stackNorm

aliz

ed Inte

nsity (

a.u

.)

Magnetic Field (mT)

a)

342 344 346 348 350 352 354 356 358

a single laminate

11 laminates stackNorm

aliz

ed Inte

nis

ty (

a.u

.)

Magnetic Field (mT)

b)

Figure A4 Comparison of EPR lineshape of thin graphene laminates (0.566 mg/cm2) at

10 K. The solid black line represents a single laminate; the blue dot represents 11

laminates stack. (a) represents 𝐻⊥. (b) represents 𝐻∥.

1200 1400 1600 2400 2600 2800 3000

FLG

Inte

nsity (

a.u

.)

Raman Shift (cm-1)

SLG

Thick Layer

P a g e | 201

338 340 342 344 346 348 350 352

Magnetic Field (mT)

Inte

nsity (

a.u

.)

295 K

10 K

20 K

30 K

40 K

50 K

75 K

100 K

150 K

200 K

250 K

Figure A5 EPR spectra of thick graphene laminates (0.113 mg/cm2) at various

temperatures. The black line represents 𝐻⊥. the red line represents 𝐻∥.

P a g e | 202

338 340 342 344 346 348 350 352

Inte

nsity (

a.u

.)

250 K

Magnetic Field (mT)

295 K

200 K

150 K

100 K

75 K

50 K

40 K

30 K

20 K

10 K

Figure A6 EPR spectra of thick graphene laminates (1.132 mg/cm2) at various

temperatures. The black line represents 𝐻⊥. The red line represents 𝐻∥.

0 50 100 150 200 250 3000.2

0.3

0.4

0.5

0.6

0.7a)

Lin

ew

idth

(m

T)

Temperature (K)

0 50 100 150 200 250 3000.8

1.0

1.2

1.4

1.6

b)

Lin

ew

idth

(m

T)

Temperature (K)

Figure A7 EPR linewidth of the narrow component for the thick graphene laminates

(0.566 mg/cm2). a) Red square represents 𝐻∥. b) Black square represents 𝐻⊥.

P a g e | 203

338 340 342 344 346 348 350-0.4

-0.2

0.0

0.2

0.4

0.6

BBIn

tensity (

a.u

.)

Magnetic Field (mT)

A A

Figure A8 EPR spectra of a graphite flake at room temperature. The black line represents

𝐻⊥. The red line represents 𝐻∥. The blue line measured the distance between the peak and

the baseline (green line). A and B assigned to the peak distance above and below the

baseline respectively. (A B)⁄⊥

= 1.75 and (A B)⁄∥ = 2.77.

P a g e | 204

10 20 30 40 50 60

0.22

0.24

0.26

0.28

Double

Inte

gra

tion

of In

tensity (

a.u

.)

Temperature (K)

Figure A9 Double integration of EPR intensity from the graphite flake on the variation

of temperature at 𝐻⊥ orientation.

340 344 348

Inte

nsity (

a.u

.)

Magnetic Field (mT)

Figure A10 EPR spectra of graphene laminate and various sources of potential

background. The solid black line represents graphene laminate (1.132 mg/cm2) on

membrane filter at 𝐻⊥; green dot represents membrane filter; red dot represents scotch

tape; blue dot represents suprasil tube.

P a g e | 205

0 2 4 6 8 10 12

10 K

12 K

15 K

20 K

30 K

50 K

70 K

Time (s)

Figure A11 Two pulse decay traces to measure the electron spin-spin relaxation (T2) of

a graphene laminate from 10 – 70 K at 𝐻⊥ . Red lines represents fits to determine

relaxation rates.

0 20 40 60 80 100 120 140

70 K

50 K

30 K

20 K

15 K

12 K

10 K

Time (s)

Figure A12 Inversion recovery traces to measure electron spin-lattice relaxation (T1) of

graphene laminates from 10 – 70 K at 𝐻⊥ . Red lines represents fits to determine

relaxation rates.

P a g e | 206

Appendix B

Temperature Error Measurement

The temperature sensor equipped in Bruker EPR EMX spectrometer was located

below the sample tube position. As a consequence, there was a temperature difference. In

order to observe the error caused by the console reading, we use the external temperature

sensor mounted inside a 4 mm quartz EPR tube to compare the difference in temperature

reading. The external sensor use was a Lakeshore Cryotronics Cernox sensor (in the CX-

AA canister package) calibrated between 1.4 K and 325 K. The external sensor was

connected to an Oxford Instruments Mercury ITC. The reading difference in various

temperature is presented in Figure B1.

0 50 100 150 200

0

4

8

12

The D

iffe

rence in T

em

pera

ture

readin

g (

K)

Temperature (K)

Err

or

/ K

4.2 0.02

7 0.21

10 0.14

20 0.5

30 0.65

40 0.8

50 1.6

60 2.15

80 3.35

100 5

150 10

200 12.5

Temp Error

Figure B1. The difference in temperature reading between the console reading and a

thermocouple placed at the tube location. The inset table shows the discrepancy.

P a g e | 207

Appendix C

Easyspin Simulation

C.1 CW EPR Code Simulation

% Fitting a two-component spectrum

%====================================================================

clear

% Import data

[B, spc] = textread('90D-anox3.txt','%f %f');

plot(B,spc);

%The parameter

Sys1.g = [2.03874]; %any g value

Sys2.g = [2.03756]; %any g value

Sys1.lwpp = [0 0.267675]; %any linewidth value

Sys2.lwpp = [0 1.97241]; %any linewidth value

Sys1.weight = 0.120377; %weight between two components

Sys2.weight = 1-Sys1.weight;

% How much the fitting algorithm can vary the parameter.

Vary1.g = [0.00002];

Vary2.g = [0.00005];

Vary1.lwpp = [0 0.00001];

Vary2.lwpp = [0 0.00005];

Vary1.weight = 0.00002;

%Experimental details

Exp.mwFreq = 9.886323; %EPR frequency

Exp.Range = [335 355];%data point range

Exp.nPoints = 6000; %number of data points

Exp.Temperature = 295; %in K

B = linspace(Exp.Range(1),Exp.Range(2),Exp.nPoints); % field axis

% Calling the fitting function

SimOpt.Method = 'perturb';

FitOpt.Method = 'simplex fcn'; %fit data as it is

%FitOpt.Method = 'simplex int'; % simplex algorithm, integrals of spectra

esfit('pepper',spc,{Sys1,Sys2},{Vary1,Vary2},Exp,SimOpt,FitOpt);

%to directly export the result

%[fitparams,spc] =

esfit('pepper',spc,{Sys1,Sys2},{Vary1,Vary2},Exp,SimOpt,FitOpt);

%data = [B(:) spc(:)];

%save ('result.txt','data','-ascii');

P a g e | 208

C.2 HYSCORE Code Simulation

clear, clf

Sys.Nucs = '13C,19F,1H'; Sys.A_ = [1 1 1; 2 2 2; 0.6 0.6 0.6];

Exp.Sequence = 'HYSCORE'; Exp.Field = 348.6; Exp.Freq = 9.6993; Exp.tau = 0.300; Exp.dt = 0.0200; Exp.nPoints = 512;


Recommended