You are on page 1of 227

저작자표시-비영리-변경금지 2.

0 대한민국

이용자는 아래의 조건을 따르는 경우에 한하여 자유롭게

l 이 저작물을 복제, 배포, 전송, 전시, 공연 및 방송할 수 있습니다.

다음과 같은 조건을 따라야 합니다:

저작자표시. 귀하는 원저작자를 표시하여야 합니다.

비영리. 귀하는 이 저작물을 영리 목적으로 이용할 수 없습니다.

변경금지. 귀하는 이 저작물을 개작, 변형 또는 가공할 수 없습니다.

l 귀하는, 이 저작물의 재이용이나 배포의 경우, 이 저작물에 적용된 이용허락조건


을 명확하게 나타내어야 합니다.
l 저작권자로부터 별도의 허가를 받으면 이러한 조건들은 적용되지 않습니다.

저작권법에 따른 이용자의 권리는 위의 내용에 의하여 영향을 받지 않습니다.

이것은 이용허락규약(Legal Code)을 이해하기 쉽게 요약한 것입니다.

Disclaimer
[UCI]I804:31001-200000371086

Doctoral Thesis

TWO-DIMENSIONAL MATERIAL-SUBSTRATE
INTERACTIONS FOR EPITAXIAL GROWTH

Leining Zhang

Department of Materials Science and Engineering

Ulsan National Institute of Science and Technology

2021

i
TWO-DIMENSIONAL MATERIAL-SUBSTRATE
INTERACTIONS FOR EPITAXIAL GROWTH

Leining Zhang

Department of Materials Science and Engineering

Ulsan National Institute of Science and Technology

ii
Two-dimensional Material-Substrate Interactions for
Epitaxial Growth

A thesis/dissertation submitted to
Ulsan National Institute of Science and Technology
in partial fulfillment of the
requirements for the degree of
Doctor of Philosophy

Leining Zhang

1/12/2021 of submission
Approved by

_________________________
Advisor
Feng Ding

iii
Two-dimensional Material-Substrate Interactions for
Epitaxial Growth

Leining Zhang

This certifies that the thesis/dissertation of Leining Zhang is approved.

1/12/2021 of submission

___________________________
Advisor: Feng Ding

___________________________
Zonghoon Lee: Thesis Committee Member #1

___________________________
Joonki Suh: Thesis Committee Member #2

___________________________
Moon-Ho Jo: Thesis Committee Member #3

___________________________
Gun-Do Lee: Thesis Committee Member #4;

iv
Abstract
Due to their unique dimensionality, two-dimensional (2D) materials exhibit many
excellent electronic, magnetic, mechanical and thermal properties that are absent from their 3D
counterparts, which makes them hold great potential in various applications, such as electronics,
optoelectronics and photovoltaics, etc. To maximize the advantages of 2D materials and realize their
practical applications in 2D devices, synthesizing single crystals of the 2D materials with a wafer scale
is highly required. Over the past decade, various strategies for synthesizing 2D materials have been
developed, and there are mainly two routes towards the fabrication of 2D single crystals especially with
a wafer scale via chemical vapor deposition (CVD): (i) nucleating only one 2D nucleus on the whole
substrate and growing it to wafer scale and (ii) seamless stitching of a large number of unidirectionally
aligned 2D material islands on a substrate. Although route (i) can produce wafer scale single crystalline
(WSSC) 2D materials with a very high quality, the synthesis process usually requires tens of hours and
delicate experimental setups. Route (ii) is much more cost-effective because the synthesis process is
mediated by the simultaneous growth of all the 2D islands, while it requires substrates that can template
the synthesis of unidirectionally aligned 2D islands.

For route (ii), an in-depth understanding on the alignment of 2D materials on substrates is of


critical importance for choosing proper substrates that can template the synthesis of unidirectionally
aligned 2D islands. Experimentally, various 2D material islands with both multi- and mono- alignment
orientations have been observed on different substrates, including unidirectionally aligned graphene
islands on Cu(111) and Cu(110) substrates, graphene islands with two orientations on Cu(100)
substrates, multi-orientations of hexagonal boron nitride (hBN) islands on all low-index Cu substrates,
unidirectionally aligned hBN islands on vicinal Cu(110) surfaces and on stepped Cu(111) surfaces,
unidirectionally aligned MoS2 islands on vicinal Au(111) surfaces, etc. Up to now, WSSC graphene,
hBN and MoS2 have been successfully synthesized by route (ii) on Cu(111) substrates, vicinal Cu(110)
or stepped Cu(111) surfaces, and vicinal Au(111) surfaces, respectively. In principle, the interactions
between 2D materials and their substrates are responsible for the epitaxial growth of the 2D materials
and the behaviors of the 2D materials on the substrates after growth. However, up to now, an in-depth
understanding on the alignment mechanism of 2D materials on substrates at atomic scale is still lacking.

In this dissertation, we carry out theoretical investigations on the alignment of 2D materials on


both high-symmetric and low symmetric transition metal (TM) surfaces. Firstly, we find that a high-
symmetric direction of a 2D material usually prefers to align along a high symmetric direction of the
substrate, which determines the most stable configuration of a 2D material island on the substrate.
Moreover, we reveal that the interplay between the symmetry of the 2D material and that of the substrate
determines the alignments of 2D islands on the substrate. To grow unidirectionally aligned 2D islands,

v
the most stable orientation of the 2D material on the substrate cannot be changed by any symmetrical
operation of the substrate, i.e., the symmetry group of the substrate should be a subgroup of that of the
2D material. Therefore, low symmetric TM surfaces are more promising for templating unidirectional
2D islands. This part is presented in Chapter 4.

We then further investigated the alignment of 2D materials on an arbitrary high-index low


symmetric TM substrate in Chapters 5-9. The high-index low symmetric FCC TM substrates are
classified into three categories according to their surface configuration, which are FCC{111}-based,
FCC{100}-based and FCC{110}-based low symmetric surfaces, and the alignment of various 2D
materials, including graphene, hBN and transition metal dichalcogenides (TMDCs) monolayers, on
these three types of low symmetric FCC substrates are systematically explored by density functional
theory (DFT) calculations. It is revealed that the interaction differences between various edges of a 2D
material and a unidirectional TM step edge determines the sole orientation of the 2D island along this
step edge. However, TM step edges usually present a meandering direction in real cases because of
surface roughness, the alignments of 2D islands along such meandering step edges are then investigated
by structural analysis. We find that the similar kink heights of substrate step edges and the edges of the
2D island are critical for the unidirectional alignment of 2D islands along meandering step edges.
Besides, the orientations of hBN and TMDCs on substrates are also affected by the ambient conditions
in experiments due to their binary compositions.

As a promising substrate for the epitaxial growth of graphene, Cu{111} foils attract lots of
attentions. By collaborating with experimental groups, a single crystalline Cu{111} foil with a size up
to 32 cm2 is realized using a contact-free annealing method. We theoretically revealed the transition
mechanism of single crystalline Cu{111} from polycrystalline Cu foils by classical molecular dynamic
(MD) simulations combined with DFT calculations in Chapter 10. It is found that the high annealing
temperature and Cu{112} grains in the raw Cu foils are essential for obtaining the single crystalline
Cu{111} foil.

After growth, 2D materials usually show novel moiréstructures and properties modulated by
the underlying substrate, and in Chapter 11 of this dissertation, the behaviors of graphene on various
TM substrates that have a large lattice mismatch with graphene are explored by the DFT calculations.
Depending on the rotation angles between graphene and the substrate, two kinds of graphene moiré
structures are revealed, i.e., highly corrugated graphene moiré superstructures under small rotation
angles and ultra-flat graphene layers under large rotation angles, and we further find that such different
behaviors of graphene are determined by the competition between the graphene-substrate interaction
and curvature energy of the graphene/TM superstructures. Compared with the ultra-flat ones, the
corrugated graphene/TM superstructures show anisotropic properties and are found to be capable of
templating size-tunable metal clusters. We finally formulate the evolution of the graphene-substrate

vi
interaction and curvature energy, and the morphology of graphene on various TM substrate can be
estimated effectively.

vii
viii
Table of Contents

Abstract ................................................................................................................................................... v

Table of Contents ................................................................................................................................... ix

List of Figures ...................................................................................................................................... xiv

List of Tables .................................................................................................................................... xxvii

Technical Terms and Abbreviations .................................................................................................. xxviii

Chapter 1 Research background on two-dimensional (2D) materials growth ........................................ 1

1.1 Chemical vapor deposition (CVD) growth of single crystalline graphene on a substrate ...... 2

1.1.1 Graphene growth via nucleation control ......................................................................... 3

1.1.2 Graphene growth via seamless stitching on single-crystal substrates ............................. 7

1.1.3 Graphene growth via seamless stitching on liquid substrates ....................................... 11

1.2 CVD growth of single-crystal hexagonal boron nitride (hBN) on a substrate ...................... 13

1.2.1 hBN growth via nucleation control ...................................................................................... 13

1.2.2 HBN growth via seamless stitching ..................................................................................... 15

1.3 The synthesis of transition metal dichalcogenide (TMDC) films ......................................... 19

1.4 Graphene moirésuperstructures on a substrate ........................................................................... 22

1.4.1 Structures of graphene moirésuperstructures ...................................................................... 23

1.4.2 Applications of graphene moirésuperstructures .................................................................. 24

1.5 Conclusions ................................................................................................................................. 27

1.6 Research motivation.................................................................................................................... 27

Chapter 2 Theoretical foundation on the growth of 2D materials......................................................... 30

2.1 CVD growth of single-crystal graphene ..................................................................................... 30

2.1.1 Nucleation of graphene on a substrate ................................................................................. 30

2.1.2 Mechanisms of graphene growing across substrate grain boundaries ................................. 33

2.1.3 Mechanisms of alignment and coalescence of graphene islands on a TM surface .............. 35

2.2 CVD growth of single-crystal hBN film ..................................................................................... 39

ix
2.2.1 Nucleation of hBN on a substrate ........................................................................................ 39

2.2.2 Mechanisms of hBN alignment on a TM surface ................................................................ 41

2.3 CVD growth of single-crystal monolayer TMDC films ............................................................. 46

2.3.1 Nucleation barrier of TMDC on a substrate......................................................................... 46

2.3.1 Mechanisms of TMDC alignment on a substrate ................................................................. 47

2.4 Conclusions and unsolved research problems ............................................................................ 49

Chapter 3 Methodology ........................................................................................................................ 51

3.1 Introduction ................................................................................................................................. 51

3.2 Introduction of Density functional theory (DFT)........................................................................ 51

3.2.1 The Born-Oppenheimer approximation ............................................................................... 52

3.2.2 The Hohenberg-Kohn theorem ............................................................................................ 53

3.2.3 The Kohn-Sham method ...................................................................................................... 54

3.2.4 Approximation to the exchange-correlation functional ....................................................... 57

3.3 Introduction to classical molecular dynamics (MD) simulation ................................................. 58

3.3.1 Interactions between atoms .................................................................................................. 59

3.3.2 Boundary conditions ............................................................................................................ 60

3.3.3 Ensemble .............................................................................................................................. 60

3.3.4 Time integration algorithm .................................................................................................. 61

Chapter 4 A general theory of 2D materials alignment on crystalline substrates with different


symmetries ............................................................................................................................................ 63

4.1 Introduction ................................................................................................................................. 63

4.2 The symmetries of 2D materials and substrate surfaces ............................................................. 63

4.3 Alignment of 2D materials on low-index high symmetric substrates ......................................... 66

4.3.1 Alignment of C6V-2D materials on high-symmetric substrates............................................ 67

4.3.2 Alignment of 3-fold symmetric 2D materials on high-symmetric substrates ...................... 69

4.4 Alignment of 2D materials on low symmetric substrates ........................................................... 73

4.4.1 Alignment of 6-fold symmetric 2D materials on low symmetric substrates ........................ 73

4.4.2 Alignment of C3V-2D materials on low symmetric substrates ............................................. 74

4.5 Summary ..................................................................................................................................... 75

x
Chapter 5 Alignment of graphene islands on low symmetric Cu substrates......................................... 77

5.1 Introduction ................................................................................................................................. 77

5.2 Structures of low symmetric Cu substrates ................................................................................. 78

5.3 Modeling and simulation methods .............................................................................................. 81

5.3.1 Modeling .............................................................................................................................. 81

5.3.2 Simulation methods.............................................................................................................. 83

5.4 Results and Discussions .............................................................................................................. 84

5.4.1 Alignment of graphene islands on Cu{111}-based low symmetric surfaces ....................... 84

5.4.2 Alignment of graphene islands on Cu{100}-based high-index low symmetric substrates .. 87

5.4.3 Alignment of graphene islands on Cu{110}-based high-index low symmetric surfaces .... 89

5.5 Summary ..................................................................................................................................... 90

Chapter 6 Single-crystal graphene grown on twinned Cu substrates .................................................... 92

6.1 Introduction ................................................................................................................................. 92

6.2 Experimental methods and results .............................................................................................. 92

6.2.1 Preparation of twinned Cu foils ........................................................................................... 92

6.2.2 Synthesis of unidirectional graphene islands on twinned Cu foils....................................... 94

6.3 Theoretical modeling and discussions ........................................................................................ 95

6.3.1 Modeling of twinned Cu foils .............................................................................................. 95

6.3.2 Alignment of graphene islands on twinned Cu substrates ................................................... 97

6.3.3 Application to other 2D materials ...................................................................................... 103

6.3.4 Anticipation of graphene alignment on twined hexagonal close-packed (HCP) metal substrate
.................................................................................................................................................... 104

6.4 Summary ................................................................................................................................... 106

Chapter 7 Alignment of hBN islands on low symmetric Cu substrates .............................................. 108

7.1 Introduction ............................................................................................................................... 108

7.2 Modeling and simulation methods ............................................................................................ 108

7.2.1 Modeling ............................................................................................................................ 108

7.2.2 Simulation methods............................................................................................................ 110

7.3 Results and discussions ............................................................................................................. 110

xi
7.3.1 Alignment of hBN islands on Cu{111}-based high-index low symmetric substrates ....... 112

7.3.2 Alignment of hBN islands on Cu{100}-based low symmetric substrates ......................... 116

7.3.3 Alignment of hBN islands on Cu{110}-based low symmetric surfaces ............................ 118

7.4 Experimental evidence .............................................................................................................. 119

7.5 Summary ................................................................................................................................... 120

Chapter 8 The effect of surface roughness on the alignment of grown 2D materials ......................... 122

8.1 Introduction ............................................................................................................................... 122

8.1 Modeling and discussions ......................................................................................................... 122

8.3 Summary ................................................................................................................................... 127

Chapter 9 Alignment of TMDC islands on vicinal Au(111) substrates .............................................. 128

9.1 Introduction ............................................................................................................................... 128

9.2 Calculation methods.................................................................................................................. 128

9.3 Results and Discussions ............................................................................................................ 128

9.4 Summary ................................................................................................................................... 131

Chapter 10 Formation mechanisms of large single-crystal Cu(111) foil by contact free annealing ... 133

10.1 Introduction ............................................................................................................................. 133

10.2 Experimental setups and results .............................................................................................. 133

10.3 Theoretical methods and discussions ...................................................................................... 137

10.3.1 Calculation methods......................................................................................................... 137

10.3.2 Modeling and discussions ................................................................................................ 137

10.4 Summary ................................................................................................................................. 140

Chapter 11 Rotated graphene moirésuperstructures on various transition metal surfaces ................. 141

11.1 Introduction ............................................................................................................................. 141

11.2 Modeling and calculation methods ......................................................................................... 142

11.2.1 Modeling .......................................................................................................................... 142

11.2.2 Calculation methods......................................................................................................... 143

11.3 Rotation-dependent graphene on the Ru(0001) surface .......................................................... 144

11.3.1 Structural evolution of graphene on the Ru(0001) surface .............................................. 144

11.3.2 Electronic properties of graphene/Ru(0001) superstructures ........................................... 146

xii
11.3.3 Applications of graphene/Ru(0001) superstructures as templates ................................... 147

11.4 Formation mechanisms of graphene moirésuperstructures on the Ru(0001) surface ............ 150

11.5 Rotated graphene moirésuperstructures on other transition metal surfaces ........................... 154

11.6 Summary ................................................................................................................................. 159

Chapter 12 Conclusions and further research plans ............................................................................ 161

References ........................................................................................................................................... 164

Acknowledgement .............................................................................................................................. 196

xiii
List of Figures
Figure 1.1 Three growth modes of graphene on a substrate. (a) Single nucleus growth. (b) Seamless
stitching of unidirectional islands. (c) Stitching of multi-orientated islands.

Figure 1.2 (a) Setup of graphene CVD growth. MFC is mass flow controllers. (b) Graphene growth
mechanisms on a Cu surface.

Figure 1.3 (a) Photo of a Cu foil enclosure. (b) Scanning electron microscope (SEM) image of graphene
on the inside of Cu enclosure. (c) Optical and SEM images of graphene domains on Cu. (d) Optical
image of graphene domains on oxygen rich-Cu. (e) Superimposed SEM and EBSD images of a
graphene domain grown across Cu grains.

Figure 1.4 (a) The effects of Cu pretreatments including surface etching (green), electropolishing (blue)
and backside oxidization (red) on graphene nucleation density and surface roughness of Cu substrates,
where CMP is chemical mechanically polishing. (b) Sketches of the three different Cu pretreatment.

Figure 1.5 (a) Sketch of local feedstock feeding of graphene grown on a Cu-Ni alloy. (b) Optical image
of grown graphene by local feedstock feeding. (c) SIMS line scan profiles showing the distribution of
carbon concentration on the substrate below the nozzle. (d) The set-up for advancing local control CVD
used for graphene growth. (e) Optical images of graphene nucleation at various buffer gas speeds on
Cu and Cu-Ni alloy substrates.

Figure 1.6 Synthesis of single-crystal metal substrate by thin film growth technique and as-grown
graphene. (a) Cross-sectional SEM image of the Ge(110) film on the Si(110) wafer. (b) SEM image of
graphene islands on the Ge(110) surface. (c) Optical image of graphene grown on a 5.08 cm Ge/Si
wafer. (d) High-resolution TEM image of the single-crystal graphene film. (e) Optical image of a 4-
inchs single-crystal Cu(111) thin film on sapphire. (f) STEM image of the interface regions of
Cu/sapphire. (g) Optical image of graphene grown on the Cu/sapphire substrate. (h) Optical image
showing the same orientation of graphene islands.

Figure 1.7 Synthesis of single-crystal Cu(111) substrate by annealing methods and as-grown graphene.
(a) Optical image of 6×3 cm2 single-crystal Cu foil with aligned graphene islands grown on it. (b)
Optical image of unidirectional graphene islands grown at point 4 in (a). (c) Optical microscopy images
of a polycrystalline graphene island and a single crystalline graphene island formed through coalescence
after UV treatment. (d) Schematic of the continuous production of single-crystal Cu(111) foil. (e)
Cu(111) foils with various graphene coverages. (f) Optical image of graphene covered Cu(111) foil
marked as 2 in (e). (g) Schematic of a Cu foil suspended by a quartz holder. (h) Photograph of the
obtained single-crystal Cu(111) foil. (i) SEM image of graphene islands on the obtained Cu(111) foil.

xiv
Figure 1.8 Synthesis of single-crystal high-index Cu substrates and as-grown graphene. (a) Optical
image of the eight representative high-index single-crystal Cu foils with a size of 35 ×21 cm2 after mild
oxidation. (b) SEM images of unidirectional graphene islands on four representative high-index single-
crystal Cu foils. (c) Photograph of the high-index Cu foils synthesized by strain-engineered annealing.
(d) Pie chart of the proportions of obtained high-index single-crystal Cu foils in 133 pieces. (e) EBSD
map of the obtained Cu(311) foil and SEM image of well-aligned graphene islands grown on it.

Figure 1.9 The growth of unidirectional graphene islands on liquid Cu surface. (a) Schematic of CVD
process for the synthesis of HGFs on liquid Cu surface. (b) SEM image of HGFs with an average size
of ~120 μm on liquid Cu. (c) SEM image of graphene domains on liquid Cu. (d) A histogram of the
distribution of graphene orientation angles. (e) Schematic of the self-assembly mechanism of the
graphene single crystals to form graphene SOS. (f) SEM image of the self-assembly SOS. (g) Schematic
depiction of the two distinct routes of graphene coalescence on liquid Cu surface. (h) Formation
energies of graphene grain boundaries as a function of the misorientation angle between graphene grains
at the two sides of the grain boundary.

Figure 1.10 Large-scale hBN synthesis by one nucleus. (a) Sketch of hBN growth on a Cu-Ni alloy.
(b-c) SEM images of hBN at different growth time. (d) Illustrations of hBN growth on 127 μm Cu
enclosure. (e-f) SEM images of hBN with ultralow nucleation density and a single crystalline hBN
domain.

Figure 1.11 The growth of hBN on a Ni(111)/sapphire substrate. (a) Schematic of hBN domains on the
Ni(111)/sapphire substrate. (b) Photograph of the grown hBN on the substrate. (c) Optical microscopy
image of the marked area in (b). (d) SEM image of the largest hBN domain.

Figure 1.12 The growth of WSSC hBN on a vicinal Cu(110) surface. (a) Optical image of the obtained
Cu foil after oxidation showing the vicinal Cu(110) single crystal has an area of 100 cm 2. (b) SEM
image of unidirectional hBN domains on the obtained substrate. (c, d) Polarized SHG mapping of two
aligned (left) and misaligned (right) hBN domains. (e) Low-magnification TEM image of the concave
corner in the joint area of two aligned hBN domains on a monolayer single-crystal graphene film. (f)
Representative HRTEM image of a uniform hBN/graphene moirépattern at the concave corner. The
orange and blue in the insert represent the diffraction pattern of hBN and graphene, respectively.

Figure 1.13 The growth of WSSC hBN on a highly stepped Cu(111) surface. (a-b) Schematic and
photography of as-grown two-inch hBN film on a Cu (111)/sapphire wafer. (c) Optical image of
unidirectional hBN islands on the obtained Cu(111) surface. (d-e) STM images of hBN/Cu showing the
obtained Cu(111) surfaces are highly stepped. (f) STM image showing moirépattern of hBN on the
obtained Cu(111) surface. (g) Atomic-scale STM image of the grown hBN film

xv
Figure 1.14 WSSC hBN grown on a liquid Au substrate. (a) Photograph of a WSSC hBN film
transferred on a SiO2-Si wafer. (b-e) SEM images showing the process of hBN film growth on the liquid
Au surface. (f) Illustration for the growth of single crystalline hBN film by self-collimated grains. (g)
Formation energy profile of grain boundaries of hBN as a function of the misaligned angles of hBN
domains.

Figure 1.15 Alignment of TMDC islands on various substrates. (a) Antiparallel MoS2 islands on the
Au(111) surface. (b) Antiparallel WSe2 islands on the Al2O3(0001) surface. (c) Antiparallel MoS2
islands on the GaN(0001) surface. (d) Unidirectional WS2 islands on a hBN film. (e) Antiparallel WS2
on a hBN film. (f) Atomic model demonstrating the difference between antiparallel TMDC islands on
a hBN film.

Figure 1.16 The growth of WSe2 on a vicinal Al2O3 (0001) surface. (a) Photography of the vicinal
Al2O3 (0001) wafer. (b) Optical microscopy image of WSe2 islands on the vicinal Al2O3 (0001) wafer
with SEM image inserted. (c) AFM image showing WSe2 islands nucleated near the step edge of the
substrate. (d-g) The formation of few-layer WSe2 by layer-over-layer overlapping between the layers
grown from adjacent (d-e) and nonadjacent (f-g) substrate steps.

Figure 1.17 The growth of WSSC MoS2 film on a vicinal Au(111) surface. (a) Illustration of the vicinal
Au(111) surface formation and MoS2 growth. (b) Photography of a MoS2 monolayer on the Au(111)/W
substrate. (c) SEM images showing MoS2 growth process at∼720 °C. (d) STM images of a continuous
MoS2 monolayer over single atomic steps of the vicinal Au(111). (e) SEM images of two type films
composed by misaligned (left) and aligned (right) MoS2 domains after water oxidation.

Figure 1.18 Graphene moirépatterns on FCC(100) surfaces. (a) STM topographic images of graphene
on the Cu(100) surface showing stripe pattern and the rhombic pattern. (b) STM topographic images of
graphene on the Ni(100) surface showing either stripe pattern or rhombic network depending on the
relative angle between graphene and the substrate, with the corresponding Fourier transforms shown in
inserts.

Figure 1.19 Atomically resolved STM images of graphene moirésuperstructures on (a) Ru(0001), (b)
Rh(111), (c) Ir(111) surfaces.

Figure 1.20 C 1s photoelectron spectrums of different graphene/TM superstructures and that of the
highly oriented pyrolytic graphite (HOPG) is used for comparison, with the side views of graphene/TM
superstructures are illustrated. MG means monolayer graphene.
Figure 1.21 Illustration of lattice relationships of graphene on the four high-symmetric sites of (a) an
FCC(111) and (b) an HCP(0001) surfaces.

xvi
Figure 1.22 Synthesis of TM clusters and assembly of Kagome lattices templated by graphene/TM
superstructure. (a-d) SEM images of (a-b) Pt clusters and (c-d) Pd clusters on the graphene/Ru(0001)
substrate. (e) STM image of Kagome lattice of FePc molecules on the graphene/Ru(0001) substrate
with a marked unit cell. (f) Atomic structure of the Kagome lattice in (e). (g) STM image of squire
lattice of CoPc molecules on the graphene/Ir(111) substrate. (h) Atomic structure of the squire lattice
in (g).
Figure 2.1 Nucleation of graphene. (a) The most stable structures of C11 and C12 on the Ni(111) surface.
(b) Formation energies of various carbon structures on the Ni(111) surface (c) The most stable structures
of C7 and C10 near a step edge of the Ni(111) surface. (d) Energies of carbon structures on a Ni(111)
terrace or near a step versus the cluster size (N). (e) Nucleation barrier (G*) and corresponding
nucleation size (N*) of graphene on a Ni(111) terrace or near a step. (f) Nucleation rate of graphene on
a Ni(111) terrace or near a step.

Figure 2.2 (a1) Formation energy profile (upper panel) of carbon clusters on different TM surfaces and
its second order derivative (lower panel). The structures of magic C21 and C24 clusters are shown in (a2)
and (a3). The simulated STM images of C21 on Rh(111) surface at the voltages of -1.0 V and 1.0 V are
shown in (a4) and (a5), respectively. Experimentally obtained STM images of a carbon cluster on
Rh(111) surfaces during CVD process are shown in (a6) an (a7). (b-c) Energies of C clusters with and
without dangling C atoms on the Ru(0001) surface.

Figure 2.3 Growth of graphene on polycrystalline TM substrates. (a-d) Illustrations of two paths of
graphene grow crossing the grain boundaries of a substrate. (e) Growth processes of the two paths of
graphene grow crossing a grain boundary of a substrate.

Figure 2.4 Interaction between graphene and the Cu(111) substrate. (a) Optimized distance and binding
energies of graphene layer on the Cu(111) surface versus the rotation angle (θ). (b) Energy profiles of
C24 and C54 clusters versus the rotation angle.

Figure 2.5 Interaction of a graphene layer near a step edge of TM substrates. (a) Structures and
formation energies of graphene ZZ and AC edges attaching to a Ni(111) terrace and a Ni(111) step edge.
(b) Formation energies of various graphene edges and a Cu<110> step edge versus the orientation angle.

Figure 2.6 Theoretical analyses of the coalescence of two well-aligned or mis-aligned graphene islands.
(a-b) The process of adding C atoms to a concave corner between two well-aligned graphene islands.
(c) Schematic of the growth behavior at a concave corner between two well-aligned graphene islands.
Red arrows mark the kink sites and black arrows indicate growth direction. (d-e) The process and the
corresponding formation energy (ΔEF) profile of adding C atoms to a concave corner between two mis-
aligned graphene islands. (f) Schematic of kink formation at graphene edges near a concave corner
between two mis-aligned graphene islands.

xvii
Figure 2.7 Theoretical analysis of hBN nucleation. (a) Formation energy profiles of BN chains and BN
sp2 networks on the Cu(111) surface under a N-rich (∆μ = -0.91 eV), a medium(∆μ = 1.0 eV) or a B-
rich (∆μ = 3.0 eV) condition, where ∆μ is the chemical potential differences of boron and nitrogen. (b)
MD simulation of hBN nucleation and growth on the Ni(111) surface.

Figure 2.8 (a) Formation energies of various hBN edges on the Cu(111) surface as a function of the
chemical potential difference of boron and nitrogen (Δμ). (b) The most stable hBN edge corresponding
to Δμ. (c) Atomic configurations of the most stable edges corresponding to (b).

Figure 2.9 The alignment of h-BN islands on low-index TM substrates. (a) Atomic structures of hBN
on the Cu(111), Cu(110) and Cu(100), respectively. (b) The binding energies between hBN and Cu
substrate as a function of their relative angle.

Figure 2.10 (a-b) SEM images of grown hBN triangles on Cu(102) and Cu(103) surfaces with a scale
bar of 20 μm. Black dashed arrows represent the Cu[010] direction. (c-d) Energy profile of hBN on
Cu(102) and Cu(103) surfaces as a function of their relative angle. (e-f) Atomic configurations with the
lowest energy corresponding to (c-d).

Figure 2.11 Alignment mechanism of hBN on vicinal Cu(110) surfaces. (a) Atomic-resolution STM
image showing the interface between hBN edge and Cu substrate. (b) Atomic configuration of the
interface of a nitrogen-terminated hBN ZZ edge aligning along a Cu<211> step edge of a vicinal
Cu(110), γ is the relative angle between them. (c) Formation energy of hBN edges attaching to the
Cu<211> step edge as a function of γ. (d) Schematic of meandering step edges on a high-index substrate.
(e) Schematic of interface between a hBN island and a tilted step edge with kinks. (f) Atomic
configuration of the interface between a tilted hBN edge and a tilted step edge.

Figure 2.12 Alignment mechanism of hBN on highly stepped Cu(111) surfaces. (a) STM image of a
highly stepped Cu(111) surface and schematic showing the step edges is composed by two types of
segments (A and B). (b-c) Binding energy and atomic configurations of a B6N7 cluster attaching to the
two typical step edges of the Cu(111) surface.

Figure 2.13 (a-d) STEM images of various MoSe2 edges observed in experiments and corresponding
atomic configurations. (e-f) Formation energy of various MoSe2 edges as a function of the chemical
potential of Mo, μMo.

Figure 2.14 Alignment mechanism of TMDCs on the Al2O3(0001) surface. (a) Schematic of MoS2 on
the Al2O3(0001) surface. (b) Binding energy of MoS2 on the Al2O3(0001) surface as a function of the
orientation angle. (c) Stereo and top views of an aligned WS2 cluster on the Al2O3(0001) surface with
0ºrotation angle. (d) The comparison of aligned and misaligned WS2 clusters on the Al2O3(0001)
surface.

xviii
Figure 2.15 (a) Atomic configurations of two types of Au⟨110⟩ step edges, and illustration of three
typical directions of MoS2 with θ = 0°, 30° and 60° denoting MoZZ, AC, and SZZ edges, respectively. (b)
Contact energies of different MoS2 edge docking on the surface and step edges.

Figure 3.1 The self-consistency loop of the Kohn-Sham equations.

Figure 3.2 The schematic of the classical MD simulations.

Figure 4.1 Illustrations of the symmetries of various 2D materials and low-index high symmetric FCC
surfaces. (a) Graphene, (b) hBN, (c) MoS2, (d) FCC{111}, (e) FCC{100} and (f) FCC{110}. The
principal axes perpendicular to the 2D surfaces are denoted by the red circles, and the mirror planes
passing through the principal axes are denoted by dashed lines.

Figure 4.2 (a) Illustration of a high-index low symmetric surface. (b) Atomic models for the two types
of CS FCC{111}-based surfaces, whose surface indices are FCC{n n n+m} and FCC{n+m n-m n},
respectively. (c) Atomic models for the two types of CS FCC{100}-based surfaces, whose surface
indices are FCC{n m m} and FCC{n m 0}. (d) Atomic models for the two types of CS FCC{110}-based
surfaces, whose surface indices are FCC{n+m n+m m} and FCC{n n+m 0}. Dashed lines denote the
mirror symmetric plane of CS surfaces. For all the six types of CS symmetric FCC surfaces, n and m are
integers and n>>m.

Figure 4.3 Binding energy of graphene attaching to three high symmetric Cu surfaces, which are
Cu(111) (a), Cu(100) (b) and Cu(110) (c) surfaces.

Figure 4.4 Illustrations showing the alignment of C6V-graphene islands on Cu(111) (a) Cu(100), (b)
and Cu(110) (c) surfaces. ZZ edges aligning along Cu<110> directions are marked by blue lines.

Figure 4.5 Binding energies of hBN ZZN edge attaching to three low-index high symmetric Cu surfaces,
which are Cu(111) (a), Cu(100) (b) and Cu(110) (c) surfaces. Similar to the configurations in Figure.
4.3, hBN ZZ nanoribbons were used and the ZZB edges are passivated by hydrogen to avoid their
interaction with the Cu substrates.

Figure 4.6 (a) The configuration of a WS2 cluster on the high symmetric Au(111) surface, h is the
distance between WS2 and the Au(111) surface. (b) Comparison of the WS2 cluster with different
orientations on the Au(111) surface, the binding energies of the top-hcp and top-fcc structures are 0.653
and 0.645 eV/W atoms, respectively. (c) Binding energy of WS2 on the Au(111) surfaces in the unit
of eV/W atom, and the distance as a function of relative angles between Au<110> and ZZW directions.

Figure 4.7 Illustrations showing the alignment of C3V-2D material islands on FCC{111} (a), FCC{100}
(b) and FCC{110} (c) surfaces. ZZN edges aligning along <110> directions are marked by blue lines.

xix
The mirror planes of the 2D material and the substrates along different directions are denoted by dashed
lines.

Figure 4.8 Interfacial formation energies of different graphene edges attaching to the CS Cu{111}-
based surfaces (a), CS Cu{100}-based surfaces (b) and CS Cu{110}-based surfaces (c).

Figure 4.9 Interfacial formation energies of different hBN edges attaching to the CS Cu{111}-based
surfaces (a), CS Cu{100}-based surfaces (b) and CS Cu{110}-based surfaces (c) under the chemical
potential of nitrogen, 𝜇𝑁 =-9.575 eV.

𝑖×<segment>
Figure 5.1 The configurations of various low symmetric Cu surfaces, where SE<kink> represents

a step edge with a segment direction denoting in superscript and a kink direction denoting in subscript,
and i is the number of unit cells in the segment.

Figure 5.2 Sketches of three categories of low symmetic surfaces, i.e., Cu{111}-based (a), Cu{100}-
based (b) and Cu{110}-based (c) high-index low symmetric surfaces. Various surfaces with different
step edge direction are distinguished by different colors, and the straight step direction is marked by
dashed lines. (d-f) The evolution of step kink density versus step edge direction as a function of step
edge orientation for Cu{100}-based, Cu{100}-based and Cu{110}-based high-index low symmetric
surfaces, respectively. The Cu[110] direction is chosen as reference of 0°for all three categories of low
symmetic surfaces .

Figure 5.3 Four interfacial configurations of different types of graphene edges attaching to different
types of step edges. In (d), 𝑙1 and 𝑙2 represent the unit length of tilted graphene edge or tilted Cu step
edge; 𝑑1 and 𝑑2 are the kink heights of graphene edge and Cu step edge, 𝛾1 denotes the angle between
the tilted graphene edge and the edge segment, and 𝛾2 denotes the angle between the tilted Cu step edge
and its segment.

Figure 5.4 (a) Atomic model for calculating the formation energy of a graphene edge attaching to a
step edge. (b-c) Atomic models for calculating graphene edge energy.

Figure 5.5 (a) Formation energies and configurations of various graphene edges attaching to the two
straight steps of Cu{111}-based low symmetric surfaces. (b) Sketch of graphene islands along a curved
step varying from -90 ~ 90º.Graphene islands along straight steps all show same orientation.

Figure 5.6 (a) Matched interfaces of graphene edges attaching to different kinked Cu step edges. (b)
The intact evolution of graphene orientation (light lines) versus different step edges. Slight lines
showing kink density of the step edges.

xx
Figure 5.7 (a) Formation energies of various graphene edges attaching to the two straight step edges on
Cu{100}-based low symmetric surfaces. (b) Sketch of graphene islands along straight steps show 4
orientations. (c) Configurations of graphene ZZ and tilted edges attaching to tilted step edges. (d) The
evolution of graphene orientation versus different steps.

Figure 5.8 (a) Formation energies of various graphene edge attaching to three straight steps on
Cu{110}-based low symmetric surfaces. (b) Sketch of graphene islands along straight steps show 4
orientations. (c) Configurations of tilted graphene edges attaching to tilted step edges. (d) The evolution
of graphene orientation versus different steps.

Figure 6.1 (a) Photo of a flat Cu foil. (b) Optical image of the annealed flat Cu foil after oxidization.
(c) EBSD map of the annealed flat Cu foil. (d) Photo of a bended Cu foil. (e) Optical image of the
annealed bended Cu foil after oxidization. (f) EBSD map of the annealed bended Cu foil.

Figure 6.2 (a) Photo of a bended Cu foil with microhardness indent (shown in insert) in center area. (b)
Optical image of an oxidized twinned Cu foil. (c) EBSD maps of three representative areas of the
twinned Cu foil marked in (b), with scale bar of 500 m; (d) (111) Pole figure and inverse pole figure
of the twinned Cu foil. (e) HRTEM and SAED images of the grain boundary in the twinned Cu foil.

Figure 6.3 (a) Growth process of graphene on the twin Cu foil. (b) Optical images of graphene films
on the twin Cu foil after hydrogen etching. (c) Photo and EBSD map of the twin Cu foil. (d) LEED
patterns of 16 different areas of graphene on the twin Cu foil.

Figure 6.4 (a) The definition of the <111>/60o twin Cu foil, where θA (θB) represents the angles between
normal direction of the (111) boundary plane of the grain A (B) and that of the Cu foil, 𝜓A and 𝜓B
denote the rotate angle of grain A and B around their <111> coaxis. <111>/60o twin requires ∆𝜃 =
0° and ∆𝜓 =60°. (b) Atomic model of the twin Cu foil with a <111>/60o twin boundary.

Figure 6.5 (a) Sketch of a Cu foil in coordinates of specimen and an arbitrary Cu gain in the foil in
crystal coordinates. (b) Sketch of Euler angles of a Cu grain with anticlockwise rotations are defined to
be positive. (c) Sketches of three decomposed operations through Euler angles.

Figure 6.6 (a-b) Illustrations of the relationship between 6 <110> directions and Cu steps on Cu(542)
and Cu(611) surfaces. (c) Sketch of a <110> direction projected on the surface of Cu foil.

Figure 6.7 Map of the misangles of graphene islands on two sides of <111>/60o twined Cu foils. θ is
the angle between the normal direction of the (111) boundary plane with that of the Cu foil surface, 𝜓
is the rotate angle of the twin around their <111> coaxis.

xxi
Figure 6.8 (a) Three atomic configurations of <111>/60o twin Cu foils along the point line of the map.
(b) SEM images and atomic models of well-aligned graphene islands on surface of <111>/60o Cu twin
located in the point line of the map.

Figure 6.9 SEM images and atomic models of misaligned graphene islands on surface of <111>/60o
Cu twin.

Figure 6.10 (a, b) Sketches of C3V 2D materials aligning on <111>/60o twinned Cu foils with parallel
prominent <110> steps but different stair directions. (c) Map of the misangles of C3V 2D materials on
two side of <111>/60o twin foils. (d) Map of the misangles of C4V 2D materials on two side of <111>/60o
twin foils.

Figure 6.11 (a) Atomic model of HCP twin. (b) Sketch of the different directions in HCP crystal and
their coordinates in orthogonal system. (c) Three examples showing steps of different surfaces. γ is the
smallest angle between close-packed direction and the surface.

Figure 6.12 (a) Map of the misangles of graphene islands on two side of Co twins with <0001>/30º
twin boundary. (b) Atomic models of graphene islands on two sides of HCP twins with different
surfaces.

Figure 7.1 (a) Atomic model for calculating formation energy of a hBN edge attaching to a Cu step
edge of a low symmetric Cu substrate. (b) Atomic model for calculating edge energy of a pristine hBN
edge. (c) Edge energies of pristine hBN edges under different chemical potential of nitrogen, 𝜇𝑁 .

Figure 7.2 Various interfaces between a hBN edge and a Cu step edge. (a) Interfaces of three straight
hBN edge versus a straight step edge. (b) Interface of a tilted hBN edge versus a staight step edge. (c)
Interfaces of three straight hBN edges versus a tilted step edge. (d) Interface of a tilted hBN edge versus
a tilted step edge.

Figure 7.3 The orientation of hBN islands along straight step edges of Cu{111}-based surfaces. (a)
Configurations of three straight hBN edges attached to two straigth step edges. (b) Formation energies
of straight hBN edges attached to straigth step edges as a function of chemical potential of nitrogen,
𝜇𝑁 . (c) Illustrations of hBN islands along straight step edges at different 𝜇𝑁 ranges corresponding to
(b). Triangles denote the orientations of hBN islands with three ZZN edges and different orientations
are distinguished by colors.

Figure 7.4 The evolution of hBN orientation along step edges of Cu{111}-based surfaces. (a) The
orientation of hBN islands versus step edge direction at 𝜇𝑁 =-8.21~ -9.61 eV. (b) Illustration of hBN
i×<110> i×<211>
islands along SE<211> and SE<110> corresponding to (a). (c) The orientation of hBN islands versus
i×<211>
step edge direction at 𝜇𝑁 =-9.61~ -9.92 eV. (d) Illustration of hBN islands along SE<110>

xxii
corresponding to (b). (e) The orientation of hBN islands versus step edge direction at 𝜇𝑁 =-9.92~-10.84
eV. (f-h) The alignment of hBN islands along a curved step edge with direction varying from -180 to
180°at different 𝜇𝑁 ranges, corresponding to (a), (c) and (e), respectively. Pure colors at margin means
small orientation variations while gradient colors represent large orientation variations.

Figure 7.5 Orientation of hBN on Cu{100}-based low symmetric surfaces. (a) Formation energies of
straight hBN edges attached to two straigth step edges as a function of chemical potential of nitrogen, 𝜇𝑁
and the corresponding stable interfaces. (b) Alignment of hBN islands along a curved step edge
constructed by straight step edges at different 𝜇𝑁 range. (c) Algiment of hBN orientation along a curved
step edge varying from -180 to 180°at 𝜇𝑁 =-8.31~ -10.27 eV. (d) The orientation of hBN islands versus
step edge orientation at 𝜇𝑁 =-8.31~ -10.27 eV.

Figure 7.6 Orientation of hBN on Cu{110}-based surfaces. (a) Formation energies of straight hBN
edges attached to three straigth step edges as a function of chemical potential of nitrogen, 𝜇𝑁 and the
corresponding stable interfaces. (b) Alignment of hBN islands along a curved step edge varying from -
180 to 180°at 𝜇𝑁 =-8.31~-10.50 eV. (c) The orientation of hBN islands versus step edge orientation
at 𝜇𝑁 =-8.31~-10.50 eV.

Figure 7.7 Comparison between experimental and theorical results. (a-c) hBN growth on Cu{100}-
based low symmetric surfaces, Cu(102), Cu(103), Cu(5 0 11). (d-g) hBN growth on Cu{110}-based
low symmetric surfaces, Cu(110), Cu(14 4 15), Cu(14 1 15), Cu(8 2 9). (h) hBN growth on a Cu{111}-
based low symmetric surface, Cu(313). The upper panels show experimental observations and lower
panels show the corresponding atomic models.

Figure 8.1 (a) 3D AFM image and root mean square (RMS) roughness of Cu foil. (b) SEM image of a
Cu surface after graphene growth.

Figure 8.2 The effects of roughness on the surface topography of a Cu foil. (a) Atomic models of an
ideal high-index Cu surface which has a tilted angle of 𝛼 with its low-index terrace. (b) Different views
of a fluctuated Cu foil surfaces with 𝜃 denoting the fluctuation degree. (c) Model of constructing a
fluctuated Cu surface. (d) Relationship between Cu foil fluctuation degree and variation of Cu step edge
direction for various high-index Cu foils. (e) Top views of various high-index Cu surfaces under a same
surface fluctuation with Height profile colored. (e) Relationship between 𝜃 and 𝛼 regard to different
step variation.

Figure 8.3 (a) Variation of the step edge direction regarding to (𝜃, 𝛼). (b) Variation of n of the step
edge direction regarding to 𝜃/𝛼 with small 𝜃 and 𝛼.

Figure 8.4 Alignments of hBN islands on Cu(5 5 6) and Cu(10 10 17) surfaces under same surface
roughness of 𝜃 = 3.5º.

xxiii
Figure 9.1 Comparison of WS2 film on the pristine Au(111) surface and S-terminated Au(111) surface.
(a) Atomic configuration of the (√3 × √3) R30ºS-terminated Au(111) surface. (b) Formation energy
of WS2 film on the pristine Au(111) surface and on the (√3 × √3) R30ºS-terminated Au(111) surface
as a function of the chemical potential of sulfur. (c) Atomic configurations of WS2 film on the (√3 × √3)
R30ºS-terminated Au(111) surface with relative angles of 0ºand 30º.(d) Atomic configurations of WS2
film on the pristine Au(111) surface with relative angles of 0ºand 30º.

Figure 9.2 Total energies of different edges of a WS2 cluster attaching (a) Au<110> step edge and (b)
Au<211> step edge, the unit of the total energy is eV.

Figure 10.1 The synthesis of single-crystal Cu(111) foil by contact free annealing. (a) Diagrams of Cu
foil suspended on the quartz holder. (b) Photograph of the obtained Cu(111) foil with a size of 32 cm2.

Figure 10.2 The characterization of the annealed single-crystal Cu(111) foil. (a) Photograph of the
obtained single-crystal Cu(111) foil. (b) XRD 2𝜃 scans of P1-P3 areas of the annealed Cu foil showing
(111) crystallographic plane. (c-d) EBSD maps and (001) pole figures of different areas of the annealed
Cu foil.

Figure 10.3 Texture evolution of as-received Cu foils during contact-free annealing. (a) ODF section
with 𝜑2 = 45ºshowing the positions of typical textures of FCC metal. (b) EBSD and ODF maps of an
as-received Cu foil. (c) EBSD and ODF maps of the Cu foil annealed for 1 hour. (d) Photography,
EBSD and ODF maps of the Cu foil annealed for 2 hours.

Figure 10.4 (a) SEM image of a Cu grain with {112}<111> surface, and EBSD maps of the Cu grain
before and after contact-free annealing. (b) Surface energies map of different Cu surfaces.

Figure 10.5 Illustration of the movement of vacancies in Cu foil. (a) Diagram of the formation of SF in
FCC metal. (b) Atomic configurations showing stacking order of a perfect FCC metal and that of a SF.

Figure 10.6 MD simulations showing the evolution of the as-received Cu foils during contact-free
annealing. (a) Snapshots of MD simulations for annealing of a Cu grain. (b) Atomic configurations
showing orientational evolution of Cu grain during annealing.

Figure 10.7 The mechanism of grain rotation during annealing. (a) Sketch and atomic models showing
grain rotation induced by SF. (b) Relative energy of Cu gain rotation.

Figure 11.1 Modeling of the graphene/Ru(0001) superstructures. (a, b) Illustrations of the construction
of supercells of graphene and the Ru(0001) surface. (c) The sizes of graphene/Ru(0001) superstructures
(blue point) and the size of the moirépattern unit cells (lines) versus the rotation angles. The sizes of
graphene/Ru(0001) superstructures chosen in DFT calculations are marked by red circles.

xxiv
Figure 11.2 Evolution of graphene structures on the Ru(0001) surface with different rotation angles. (a)
Top views of graphene/Ru(0001) superstructures, where ATOP, FCC and HCP regions respectively are
marked by pink Ru atoms, hexagons and triangles, and graphene/Ru(0001) unit cells are denoted by
rhombus. (b) Perspective views of a few typical graphene/Ru(0001) superstructures. (c) The height
oscillation of above graphene on the Ru(0001) surface, in which the average distances between
graphene and substrate are marked by points.

Figure 11.3 C-C bond length profiles of graphene/Ru(0001) superstructures. (a) The atomic
configuration of C-C bond with a colored length profile. (b) The statistical result of C-C bond length,
with the longest, shortest and average bond length shown.

Figure 11.4 Electronic properties of graphene/Ru(0001) superstructures with four different rotation
angles. (a) The PDOS of the C and Ru atoms at different sites of graphene/Ru(0001) superstructures.
(b) CDD of graphene/Ru(0001) superstructures with an isosurface level of 0.004 e/bohr3. (c) Side view
and section of the CDD of the 7.59ºgraphene/Ru(0001) superstructure.

Figure 11.5 The adsorption energy of a Pt atom at different sites of graphene/Ru(0001) superstructures
versus the rotation angles.

Figure 11.6 The formation energies of Pt clusters of different size as a function of Pt atoms number.

Figure 11.7 Atomic configurations of 1- and 2-layer Pt clusters on various graphene/Ru(0001)


templates, with the corresponding formation energy of Pt clusters on graphene/Ru(0001) templates (EPt )
showing at the bottom.

Figure 11.8 Two models of graphene on the Ru(0001) surface with grey representing the graphene film
and blue representing the Ru(0001) substrate. In model I, h denotes the height oscillation degree and r
denotes the radius of the hump area.

Figure 11.9 (a) Lattice relationships of graphene on the four high-symmetric sites of the Ru(0001)
surface. (b) Binding energies of graphene with the four sites of the Ru(0001) surface as a function of
the distance. (c) Illustration of a graphene/Ru(0001) common cell with different lattice relationships
distinguished by colors. (d) Linear fitting of the curvature energies of corrugated graphene moiré
superstructures. (e) Fitting of the height of graphene layers on the Ru(0001) surface versus the length
of graphene/Ru(0001) common cell (L). (f) Curvature energies of graphene layers on the Ru(0001)
surface as a function of the rotation angle and the dashed line denotes the binding energy increase of
graphene/Ru(0001) with graphene changing from an ultra-flat to a corrugated structure.

Figure 11.10 Comparison the total energies of the ultra-flat and corrugated graphene layers on the
Ru(0001) substrate with larger rotation angles, and the unit of the total energy here is eV.

xxv
Figure 11.11 (a) Top and side views of the 0ºgraphene/Rh(111) superstructure. (b) Optimized binding
energies and distance of graphene lattice on different sites of the Rh(111) surface. (c) Curvature energies
of graphene layers on the Rh(111) surface as a function of the rotation angle and the dashed line denotes
the binding energy increase of graphene/Rh(111) with graphene changing from an ultra-flat to a
corrugated structure. (d) Height oscillation of graphene layers on the Rh(111) surface regarding to the
rotation angle by DFT calculations. (e) Stereo views of graphene/Rh(111) superstructures with different
rotation angles.

Figure 11.12 (a) Top and side views of the 0ºgraphene/Ir(111) superstructure. (b) Optimized binding
energies and distance of graphene lattice on different sites of the Ir(111) surface. (c) Curvature energies
of graphene layers on the Ir(111) surface as a function of the rotation angle and the dashed line denotes
the binding energy increase of graphene/ Ir(111) with graphene changing from an ultra-flat to a
corrugated structure. (d) Height oscillation of graphene layers on the Ir(111) surface regarding to the
rotation angle by DFT calculations. (e) Stereo views of graphene/ Ir(111) superstructures with different
rotation angles.

Figure 11.13 (a) Top and side views of the 0ºgraphene/Pt(111) superstructure. (b) Optimized binding
energies and distance of graphene lattice on different sites of the Pt(111) surface. (c) Curvature energies
of graphene layers on the Pt(111) surface as a function of the rotation angle and the dashed line denotes
the binding energy increase of graphene/Pt(111) with graphene changing from an ultra-flat to a
corrugated structure. (d) Height oscillation of graphene layers on the Pt(111) surface regarding to the
rotation angle by DFT calculations. (e) Stereo views of graphene/ Pt(111) superstructures with different
rotation angles.

Figure 11.14 (a, d) Different views of the 0º graphene/Re(0001) and the 0º graphene/Pd(111)
superstructure. (b, e) Optimized binding energies and distance of graphene lattice on different sites of
the Re(0001) and the Pd(111) surface. (c) Curvature energies of graphene layers on the Re(0001) and
the Pd(111) surface as a function of the rotation angle and the dashed line denotes the binding energy
increase of systems with graphene changing from an ultra-flat to a corrugated structure.

xxvi
List of Tables
Table 4.1 The compositions and surface indexes of the CS symmetric FCC surfaces.

Table 4.2 The alignment of 2D materials on different substrate based on rotation symmetry.

Table 5.1 Formation energies of graphene edges attaching to different Cu step edges of Cu{111}-based
low symmetric surfaces.

Table 6.1 The angles (°) between <110> direction and Cu surface.

Table 6.2 Experimentally obtained misalignment angles (∆𝐸𝑥𝑝 ) and theoretically predicted ones (∆𝑀𝑎𝑝 )
of graphene islands on 14 twinned Cu foils.

Table 7.1 Summary of the alignment of hBN on various high-index Cu surfaces.

Table 11.1 Summary of lattice parameters of graphene/TM superstructures and graphene behaviors on
different TM surfaces.

xxvii
Technical Terms and Abbreviations

2D two-dimensional
hBN hexagonal boron nitride
TMDCs transition metal dichalcogenides
vdW van der Waals
CVD chemical vapor deposition
APCVD ambient pressure chemical vapor deposition
PVD physical vapor deposition
MBE molecular beam epitaxial
MFC Mass flow controllers
WSSC wafer scale single crystalline
EBSD electron backscatter diffraction
SEM Scanning electron microscope
SIMS Secondary ion mass spectrometry
TEM transmission electron microscopy
STEM scanning transmission electron microscopy
LEED low-energy electron diffraction
TM transition metal
HGFs hexagonal graphene flacks
SOS super-ordered structure
SHG second-harmonic generation
AFM Atomic force microscopy
FCC faced centered cubic
HCP hexagonal close packed
HOPG highly oriented pyrolytic graphite
Pc phthalocyanines
MD Molecular dynamics
DFT Density functional theory
LDA local density approximation
LSDA local spin-density approximation
EAM embedded atom model
ZZ zigzag
AC armchair
ZZW tungsten terminated ZZ

xxviii
ZZB boron terminated ZZ
HRTEM High Resolution Transmission Electron Microscope
SAED selected area electron diffraction
TD transverse direction
RD rolling direction
ND normal direction
ODF orientation distribution function
XRD X-ray diffraction
LAMMPS Large-scale Atomic/Molecular Massively Parallel Simulator
VASP Vienna ab initio simulation Package
SF stacking fault
FPD Frank partial dislocation
SPD Shockley partial dislocation
PDOS partial density of states
CDD charge density difference

xxix
Chapter 1 Research background on two-dimensional (2D)
materials growth

The first mechanical exfoliation of graphene in 20041 raised enormous attention on two-
dimensional (2D) materials because of their excellent physical and electric properties.2-15 Up to now,
various 2D materials, including graphene, hexagonal boron nitride (hBN) monolayer, transition metal
dichalcogenides (TMDCs), FeSe monolayer and phosphorene, etc., have been broadly synthesized and
investigated. Different from the traditional thin films, 2D materials are composed of one (graphene and
hBN) or a few (TMDCs, FeSe and phosphorene) atomic layers, and intra-layer atoms are covalently
bonded with each other, while between layers are bonded through weak van der Waals (vdW)
interaction.

Due to their unique atomic configurations, 2D materials show many excellent properties that
are absent from their 3D counterparts, which make them hold great potential in various applications. To
maximize the applications of the 2D materials in different areas, synthesis of wafer scale single
crystalline (WSSC) 2D materials at industrial scale are highly needed. Generally, the synthesis methods
for 2D materials can be classified into top-down and bottom-up methods. Top-down methods obtain a
2D material from its bulk by exfoliation or thermal ablation from its bulk counterpart, and therefore the
size and quality of the obtained 2D material are limited by the quality of its bulk counterpart.16-20 In
contrast, bottom-up methods, including chemical vapor deposition (CVD),21-30 physical vapor
deposition (PVD)31-35 and molecular beam epitaxial (MBE),36-41 show greater potential in the synthesis
of WSSC 2D materials at industrial scale.

Considering their ultra-thin structures and low bending stiffness, the growth of 2D materials by
bottom-up methods must be supported by substrates. The growth of a 2D material on a substrate was
usually called an epitaxial process in literatures,42-47 because the scenery of a substrate templated growth
of a 2D material is very similar to the epitaxial grow of thin films.48-51 The epitaxial relationship between
a thin film and its substrate originates from the strong chemical bonding between them, therefore
seeking a proper substrate which has a similar lattice constant with the overlayer thin film is critical to
avoid the formation of dislocations in the overlayer material.52-56 On contrary, during the growth of a
2D material, the interaction between the 2D material and the substrate is the weak vdW interaction,
which is generally one or two orders of magnitude lower than the chemical bonding, and as a result, the
exact lattice registry between the 2D material and the underlying substrate is usually absent.57-59
However, highly orientated 2D islands were broadly observed on single crystalline substrates, no matter
whether there is lattice matching between 2D materials and their substrates or not, and therefore the
substrate templated 2D material growth can still be considered as an epitaxial process.

1
Using hBN as an example, the three possible growth modes of a 2D material templated by a
substrate are shown in Figure 1.1: (i) single nucleus growth, (ii) seamless stitching of unidirectional 2D
material islands, (iii) stitching multi-orientated 2D material islands. Obviously, WSSC 2D materials
can be realized by the former two modes, while the latter one results in polycrystalline 2D material
films.

Figure 1.1 Three growth modes of hBN on a substrate. (a) Single nucleus growth. (b) Seamless stitching
of unidirectional islands. (c) Stitching of multi-orientated islands.

1.1 Chemical vapor deposition (CVD) growth of single crystalline graphene on a substrate

Since the successful synthesis of graphene on Ni and Cu surfaces by the CVD method,60,61 this
method rapidly become the most popular approach in graphene synthesis because large-sized and
transferable graphene can be obtained at a low cost.60-65 A common experimental setup for graphene
CVD growth is presented in Figure 1.2(a), in which a Cu foil is put into a quartz vacuum chamber and
located in a tube furnace as the growth substrate and catalyst. Mass flow controllers (MFC) and pressure
control system are used to regulate experimental conditions, such as the partial pressures of buffer gas
and carbon feedstocks. Figure 1.2(b) illustrates the mechanism of graphene growth on a Cu surface,
where C feedstocks adsorb on the Cu catalyst and decompose into active C species, and the active C
species then diffuse and assemble on the substrate and finally form graphene.

With more than 10 years of efforts, graphene CVD synthesis gradually becomes mature, and
especially in recent years WSSC graphene films have been reported in succession.66-69 As introduced
above, there are mainly two growth modes for single crystalline graphene growth, which are (i) single
nucleus growth and (ii) seamless stitching of unidirectional 2D material islands. In the following
sections, the current progress of graphene CVD growth based on the two routes will be introduced.

2
Figure 1.2 (a) Setup of graphene CVD growth. MFC is mass flow controllers.70 (b) Graphene growth
mechanisms on a Cu surface.71

1.1.1 Graphene growth via nucleation control

In early stage, metal foils for graphene CVD growth are usually polycrystalline and most
researches of high-quality graphene growth are focused on the route (i) -- nucleation control, and here
the endeavor along this direction is summarized in Figure 1.3. In 2011, Ruoff’s group adopted Cu
enclosures in a low-pressure CVD system as substrates to grow graphene, as shown in Figure 1.3(a). 72
It was found that graphene can grow on both sides of the Cu enclosure, a much lower nucleation density
and larger graphene domain size on the inside surface is observed and the largest graphene island is up
to 0.5 mm (see Figure 1.3(b)). The authors claimed that the low partial pressure of methane and
improved growth environment inside the Cu enclosure are the main reasons for the low nucleation
density and large graphene size. In 2012, Yan et al. synthesized a monolayer single crystalline graphene
domain with an area of ~4.5 mm2 on a polycrystalline Cu foil in a controlled pressure CVD system.71
By optimizing H2/CH4 flow rate, only one graphene nucleus was achieved on the centimeter scale
substrate. The pretreatments of Cu foil, such as electrochemical polishing and high-pressure annealing,
were also proved to be critical to suppress the nucleation density. The hexagonal shape, straight edges
and well-identified 120ºcorners in Figure 1.3(c) all proved the high quality of the obtained graphene
domains. Afterwards, Ruoff’s group realized centimeter scale single crystalline graphene islands by
introducing oxygen in the CVD growth process.73 Oxygen is found to be able to passivate the active
sites of Cu surface and hence reduce the graphene nucleation density, and it was also found that oxygen
can improve the growth rate and shift the growth kinetics from attachment-limited to diffusion-limited

3
growth, resulting in a dendritic island shape, as shown in Figure 1.3(d). The electron backscatter
diffraction (EBSD) map in Figure 1.3(e) showed a hexagonal graphene island growing across several
Cu grains, confirming that a polycrystalline Cu substrate can be used to synthesize WSSC graphene
through the nucleation control method.

Figure 1.3 (a) Photo of a Cu foil enclosure.72 (b) Scanning electron microscope (SEM) image of
graphene on the inside of Cu enclosure. (c) Optical and SEM images of graphene domains on Cu.71 (d)
Optical image of graphene domains on oxygen rich-Cu.73 (e) Superimposed SEM and EBSD images of
a graphene domain grown across Cu grains.

In 2016, Braeuninger-Weimer et al. systematically explored the effects of impurity, surface


roughness and oxygen on graphene nucleation.74 Figure 1.4 shows the effects of three types of
pretreatments on the surface roughness of Cu substrates and on graphene nucleation density, i.e., (i)
surface etching of the Cu substrate to remove contaminations, (ii) electropolishing to reduce Cu surface
roughness and (iii) oxidization of the backside of the Cu foil. It was found that increasing the surface
etching time can effectively remove carbon residues on the Cu substrates, and simultaneously the
graphene nucleation density is reduced by ~ 2 orders of magnitude, despite the substrate surface
roughness increases from 300 nm to 550 nm. Electropolishing can not only remove the surface
contaminations but also reduce the Cu surface roughness. An electropolishing time of 70 s removed the
surface contaminations with the surface roughness remained at 300 nm, resulting in a decreasing of the
graphene nucleation density by 2 orders of magnitude. Further increasing the electropolishing time and
employing chemical mechanically polishing, the Cu substrate surface roughness can be significantly
reduced to 3 nm, and the graphene nucleation density is further reduced by 1 order of magnitude.
Oxidization of the backside of the Cu substrate and further annealing in Ar atmosphere were used to
investigate the effect of oxygen. After oxidization and annealing, the surface roughness of the Cu

4
substrate remains, while graphene nucleation density is significantly reduced to ~ 10-2 mm-2. These
results suggest that surface roughness might be not the main factor influencing the graphene nucleation
density. Except for passivating active nucleation sites on Cu substrates, oxygen pretreatment was also
found to be able to effectively remove carbon residues (mainly amorphous or graphitic carbon) on the
Cu substrates and lead to a homogeneous distribution of carbon on the Cu substrate surface.

Figure 1.4 (a) The effects of Cu pretreatments including surface etching (green), electropolishing
(blue) and backside oxidization (red) on graphene nucleation density and surface roughness of Cu
substrates, where CMP is chemical mechanically polishing. (b) Sketches of the three different Cu
pretreatment.74

Despite of many years of effort, the growth of single crystalline graphene islands to a centimeter
scale from a single nucleus by traditional CVD process generally takes a very long time, e.g., 12 hours,
which is neither efficient nor economic.

In 2016, Wu et al. realized the synthesis of ~1.5-inch single crystalline graphene monolayer on
a polycrystalline Cu-Ni alloy substrate within 2.5 hours by using a local feeding technique in CVD
process (Figure 1.5(a-c)), which opened the door of synthesizing WSSC graphene.66 Cu-Ni alloy
combines the advantages of low carbon solubility of Cu and superior catalytic efficiency of Ni and is
proved to be an ideal catalyst for graphene growth. In addition, Ni was also found to facilitate the fast
diffusion of C atoms in the substrate and hence improve the growth rate of graphene films. Secondary
ion mass spectrometry (SIMS) in Figure 1.5(c) showed that carbon species only concentrate around the
feeding center below the nozzle with a radius of about 500-1000 μm, which is the typical distance
between neighboring graphene nuclei in traditional CVD experiments, and consequently only one

5
graphene nucleus can be realized on the whole substrate in this study. Two years later, the local feeding
setup is further improved by adopting a movable substrate, as shown in Figure 1.5(d), which allows for
the continuous roll-to-roll graphene growth on both Cu and Cu-Ni alloy polycrystalline substrates.75
Fig. 1.5(e) shows that multi-nucleation of graphene near the graphene growth frontier on a Cu substrate
can be fully suppressed when the H2-Ar buffer gas flow rate was increased to 32 cm·s-1, and at this flow
rate, single crystalline graphene growth can also be realized on Cu-Ni alloy substrates.

Figure 1.5 (a) Sketch of local feedstock feeding of graphene grown on a Cu-Ni alloy.66 (b) Optical
image of grown graphene by local feedstock feeding. (c) SIMS line scan profiles showing the
distribution of carbon concentration on the substrate below the nozzle. (d) The set-up for advancing
local control CVD used for graphene growth.75 (e) Optical images of graphene nucleation at various
buffer gas speeds on Cu and Cu-Ni alloy substrates.

To date, controlling the graphene nucleation density in a traditional CVD system is still
challenging, although great progresses have been made and some critical factors that affect the graphene
nucleation were found. Local feeding of carbon precursor, on the other hand, is an effective method of
controlling the nucleation of graphene. However, the efficiency of this method in growing WSSC
graphene is still limited, because of the rather low growth rate of graphene. Usually, a few hours’
experimental time is required to obtain one piece of inch sized single crystalline graphene.

6
1.1.2 Graphene growth via seamless stitching on single-crystal substrates

As compared to the route (i) of a single nucleation on a wafer scale substrate, route (ii) of
seamless stitching of unidirectional graphene islands presents two obvious advantages: the growth time
is much shorter due to millions of islands growing simultaneously; the growth condition is not that strict
because nucleation suppression is no longer required. Therefore, the seamless stitching method is more
promising for large scale production, and in this case to make all graphene domains to be
unidirectionally is critical. As explained in Figure 1.1, proper single crystalline substrates are required.
In last few years, the fabrication of WSSC metal substrates and the growth of graphene single crystals
on them have been reported, which will be summarized in this section.

In 2014, Lee et al. reported the synthesis of WSSC graphene film on a Ge(110) surface by
seamless stitching for the first time.68 Through solid phase epitaxial growth, a single crystalline Ge(110)
film with a thickness of ~3 μm was first grown on a Si(110) wafer. Figure 1.6(a) shows the cross-section
SEM image of the grown Ge(110) film on the Si(110) wafer, the ultra-flat surface of the Ge(110) film
and the absence of voids indicate the high quality of the Ge(110) film. The grown graphene islands on
the Ge(110) surface were found to be unidirectionally aligned along the Ge[1̅10] direction (see Figure
1.6(b)), and eventually coalesce into a wafer scale single crystal covering the entire substrate with a
lateral size of ~5.08 cm (see Figure 1.6(c)). The high-resolution transmission electron microscopy
(TEM) image in Figure 1.6(d) reveals the very high quality of the grown graphene film without
noticeable defects. Similarly, Dai et al. also realized unidirectional growth of graphene on several
vicinal Ge(110) surfaces, and both experimental characterizations and theoretical simulations proved
that the graphene armchair (AC) edge preferred to align along the step edges of the Ge substrate that
are parallel to the [1̅10] direction.76

Compared to Ge substrates, Cu substrates are more economical for the massive synthesis of
graphene, and researchers have made great achievements in the production of WSSC Cu(111) films and
the seamless stitching of unidirectional graphene islands grown on them. Figure 1.6(e) presents a 4-inch
single crystalline Cu(111) thin film obtained by PVD on an oxygenated sapphire substrate.77 The
atomically resolved scanning transmission electron microscopy (STEM) image in Figure 1.6(f) reveals
the epitaxial relationship at the interface between Cu and Al2O3. Figure 1.6(g-h) show optical
microscopy images of graphene grains grown on the 4-inch Cu(111) surface, from which we can see
that more than 99% graphene grains are unidirectionally aligned all over the Cu(111) surface, and the
zigzag (ZZ) direction of graphene is found to be parallel to the Cu<110> direction.

7
Figure 1.6 Synthesis of single-crystal metal substrate by thin film growth technique and as-grown
graphene. (a) Cross-sectional SEM image of the Ge(110) film on the Si(110) wafer.68 (b) SEM image
of graphene islands on the Ge(110) surface. (c) Optical image of graphene grown on a 5.08 cm Ge/Si
wafer. (d) High-resolution TEM image of the single-crystal graphene film. (e) Optical image of a 4-
inchs single-crystal Cu(111) thin film on sapphire.77 (f) STEM image of the interface regions of
Cu/sapphire. (g) Optical image of graphene grown on the Cu/sapphire substrate. (h) Optical image
showing the same orientation of graphene islands.

Except for the deposition methods discussed above, single crystalline Cu substrates have also
been obtained through annealing of polycrystalline Cu foils. In 2015, Nguyen et al. obtained a 6×3 cm2
single crystalline Cu foils by repetitively annealing and chemical-mechanical polishing a
polycrystalline Cu foil and found that ~98% graphene islands grown on the obtained Cu(111) foil were
unidirectionally aligned, as shown in Figure 1.7(a-b).78 Figure 1.7(c) displayed the coalesced two
misaligned (left) and aligned (right) graphene islands after UV irradiation treatment. Two grain
boundary lines can be clearly observed for the coalesced two misaligned graphene islands and each
grain boundary corresponds to a sharp concave corner. On the contrary, no grain boundary was observed
in the coalesced aligned graphene islands and the concave corner between the two graphene islands
become dull during growth. It is worth noting that such dull concave corner has been proved to be an
indicator of the seamless stitching of well aligned 2D materials.79

8
Figure 1.7 Synthesis of single-crystal Cu(111) substrate by annealing methods and as-grown graphene.
(a) Optical image of 6×3 cm2 single-crystal Cu foil with aligned graphene islands grown on it.78 (b)
Optical image of unidirectional graphene islands grown at point 4 in (a). (c) Optical microscopy images
of a polycrystalline graphene island and a single crystalline graphene island formed through coalescence
after UV treatment. (d) Schematic of the continuous production of single-crystal Cu(111) foil.69 (e)
Cu(111) foils with various graphene coverages. (f) Optical image of graphene covered Cu(111) foil
marked as 2 in (e). (g) Schematic of a Cu foil suspended by a quartz holder. (h) Photograph of the
obtained single-crystal Cu(111) foil. (i) SEM image of graphene islands on the obtained Cu(111) foil.67

In 2017, Xu et al. reported a temperature-gradient-driven annealing strategy to transform a


polycrystalline Cu foil into a single crystalline Cu(111) foil, as demonstrated in Figure 1.7(d).69 The Cu
polycrystalline foil was fed into the heating zone that is located at the central area of the furnace tube
continuously by two rollers, and as a result, a temperature gradient was induced at the Cu foil near the
heating area. Molecular dynamics (MD) simulations proved that grain boundaries in the Cu foil can be
driven towards the high temperature zone. Using this continuous annealing method, single crystalline
Cu(111) foils with a size up to (5×50) cm2 were achieved successfully. Furthermore, the obtained single
crystalline Cu(111) foils have been used as substrates to synthesize meter-sized single crystalline

9
graphene films by the seamless coalescence of unidirectionally aligned graphene islands on them.
Figure 1.7(e) shows the obtained Cu(111) foils with different graphene coverages. To demonstrate the
alignment of graphene islands on it, the border of a graphene film (marked as 2 in Figure 1.7(e)) is
shown in Figure 1.7(f). One year later, Ruoff’s group also realized the production of single crystalline
Cu(111) foils with a size of 32 cm2 by the contact-free annealing of commercial polycrystalline Cu foils,
and similarly graphene islands grown on the obtained Cu(111) foils were also proved to be
unidirectional, as shown in Figure 1.7(g-i).67

In 2019, high-index transition metal (TM) surfaces were found to show a great potential for the
growth of unidirectionally aligned 2D material islands.80-84 The synthesis of TM foils with high-index
surfaces attracts great attention in the field of epitaxial growth of 2D materials. By a seeded growth
technique, Wu et al. realized the production of more than 30 kinds of single crystalline high-index Cu
surfaces with a size of about (30×20) cm2, and Figure 1.8(a) shows several representative samples.85
The possibility of using such high-index surfaces as substrates for the epitaxial growth of graphene was
also explored, as shown in Figure 1.8(b), graphene islands grown on Cu(112), Cu(113), Cu(133) and
Cu(223) surfaces all showed unidirectional alignment. Besides, this seeded abnormal grain growth
method was also proved to be suitable for the growth of single crystalline Ni foils with high-index
surfaces, which may be able to serve as substrates for the growth of wafer-scale multi-layer graphene
films with a high crystallinity in the future. Right after that, Li et al. also successfully transformed
commercial decimeter-sized polycrystalline Cu foils into a series of high-index Cu single crystals
through a strain-engineered anomalous grain growth technique (see Figure 1.8(c-d)), and the growth of
unidirectionally aligned graphene islands on high-index Cu surfaces was also realized (see Figure
1.8(e)).86

10
Figure 1.8 Synthesis of single-crystal high-index Cu substrates and as-grown graphene. (a) Optical
image of the eight representative high-index single-crystal Cu foils with a size of 35 ×21 cm2 after mild
oxidation.85 (b) SEM images of unidirectional graphene islands on four representative high-index
single-crystal Cu foils. (c) Photograph of the high-index Cu foils synthesized by strain-engineered
annealing.86 (d) Pie chart of the proportions of obtained high-index single-crystal Cu foils in 133 pieces.
(e) EBSD map of the obtained Cu(311) foil and SEM image of well-aligned graphene islands grown on
it.

1.1.3 Graphene growth via seamless stitching on liquid substrates

The epitaxial growth of graphene films was realized on not only the solid single crystalline TM
substrates but also liquid substrates. Well-aligned graphene islands on liquid metal surfaces were
reported by several studies, despite of the unclear alignment mechanisms.

Figure 1.9(a) shows the schematics of graphene growth on the liquid Cu surface supported by
a W substrate.87 In general, hexagonal graphene flacks (HGFs) distributed on the liquid Cu surface with
random orientations when the surface was not fully covered. However, the HGFs were found to
automatically align into a compact packing arrangement with the increase of coverage, until perfectly
ordered hexagonal graphene islands with a uniform size was achieved, as shown in Figure 1.9(b). Well

11
aligned graphene islands on liquid Cu surface supported on a Mo substrate have also been reported (see
Figure 1.9(c-d)).88 In this work, the authors claimed that the epitaxial relationship between graphene
and the Cu lattice is the main reason of the unidirectional alignment, because the cooled Cu surfaces
showed (110) texture with the cover of graphene films. Zeng et al. proposed that the super-ordered
structure (SOS) of graphene islands is driven by a mutual electrostatic force between adjacent graphene
islands.89 To get a uniform electric field on the liquid Cu surface, the anisotropic electrostatic potential
of the hexagonal graphene island facilitated the rotation of graphene islands to match with their
neighbors (see Figure 1.9(e-f)).

Figure 1.9 The growth of unidirectional graphene islands on liquid Cu surface. (a) Schematic of CVD
process for the synthesis of HGFs on liquid Cu surface.87 (b) SEM image of HGFs with an average size
of ~120 μm on liquid Cu. (c) SEM image of graphene domains on liquid Cu.88 (d) A histogram of the
distribution of graphene orientation angles. (e) Schematic of the self-assembly mechanism of the
graphene single crystals to form graphene SOS.89 (f) SEM image of the self-assembly SOS. (g)
Schematic depiction of the two distinct routes of graphene coalescence on liquid Cu surface. 90 (h)
Formation energies of graphene grain boundaries as a function of the misorientation angle between
graphene grains at the two sides of the grain boundary.

It should be noted that, Dong et al. reported that two neighboring graphene islands on a liquid
Cu surface may coalescence into either a single crystal or a polycrystal with a ~ 30o twin grain boundary

12
as illustrated in Figure 1.9(g).90 The graphene grain boundary formation energy profile as a function of
the misorientation angle in Figure 1.9(h) shows two local minima at 0º and 30º respectively.
Consequently, on the liquid Cu surface two neighboring graphene islands can rotate freely to form a
single crystal or a polycrystal with a 30o twin grain boundary to minimize the system energy during
coalescence. Based on this, the author further predicted a family of 30 kinds of twinned polycrystalline
graphene islands that might be formed on a liquid Cu surface, which was verified by experimental
observations. A close inspection on Figure 1.9(f) shows that some of the graphene islands (highlighted
by red circles) are indeed polycrystals and their shapes are in consistent with those predicted by the
theoretical study.

Above discussions reviewed the synthesis of WSSC graphene by both route (i) and route (ii).
For route (i), different methods have been employed to control graphene nucleation, such as
pretreatments of substrate, using oxygen during CVD, etc. Among them, local feeding of carbon
precursor is considered to be a most effective way of controlling graphene nucleation and WSSC
graphene films can be synthesized in hours by this method. Route (ii) is more efficient than route (i)
because of the multi-nucleus and simultaneous growth, whereas proper single crystalline substrates are
required for this route. Currently, WSSC TM substrates have been obtained by either annealing
polycrystalline metal foils or depositing metal film on a single crystalline wafer. Liquid metal surfaces
were also proved to be good substrates for WSSC graphene synthesis, although the mechanism of
graphene unidirectional alignment on liquid metal surface is still under debate.

1.2 CVD growth of single-crystal hexagonal boron nitride (hBN) on a substrate

hBN has the same hexagonal lattice structure with graphene but with boron and nitrogen atoms
alternatively occupying the honeycomb lattice sites, therefore it is also known as white graphene.
Compared to the zero bandgap of graphene, hBN shows a large bandgap of ~ 5.5 eV, and exhibits even
higher thermal and chemical stabilities than graphene,91-94 and thus has a wide range of potential
applications in dielectrics, such as dielectric,95,96 sensor,97,98 deep ultraviolet light-emitting,99,100 and
proton transportation,101,102 etc. By combining with other 2D materials to form heterojunctions,
electronic devices with high performances could be built.103-106

1.2.1 hBN growth via nucleation control

Similar with graphene, the efforts in the growth of WSSC hBN film in early age are mainly
focused on the route (i), i.e., nucleation control.107-109 The experimental trials are also similar with those
adopted in graphene growth, such as pretreating substrates,110,111 using enclosures,112,113 introducing

13
oxygen,114,115 etc. It was reported that adding 10-20 atom % Ni into a Cu substrate can reduce hBN
nucleation density to ~ 60 mm-2 (see Figure 1.10(a-b)) and the size of grown hBN single crystals can
be up to 7500 μm2 (see Figure 1.10(c)), and the well-defined triangular shape of hBN domains implies
their single crystallinity.116 Cu enclosures with a thickness of 127 μm were also proved to be able to
reduce hBN nucleation density to about 1 mm-2 in a low pressure CVD experiment, and the largest hBN
single crystal with a lateral size of 300 μm was achieved (see Figure 1.10 (d-f)).117 In addition, it was
also demonstrated that the peroxidization of Cu substrates can suppress the hBN nucleation density by
~ 3 orders of magnitude from 106 to 103 mm-2.110

Figure 1.10 Large-scale hBN synthesis by one nucleus. (a) Sketch of hBN growth on a Cu-Ni alloy.116
(b-c) SEM images of hBN at different growth time. (d) Illustrations of hBN growth on 127 μm Cu
enclosure.117 (e-f) SEM images of hBN with ultralow nucleation density and a single crystalline hBN
domain.

In 2017, Meng et al. studied the growth of hBN islands on a single crystalline Ni(111) film that
was deposited on a sapphire(0001) substrate by the ion beam sputtering method, as illustrated in Figure
1.11(a).118 It was found that the high quality Ni(111) film with a smooth surface is responsible for the
growth of millimeter scale hBN islands, and the largest hBN island reached a lateral size up to ~ 0.6
mm under the optimized experimental conditions (see Figure 1.11(b-d)), which is five times larger than
previous studies. It is worth noting that the obtained hBN islands show two equally distributed opposite
orientations, which inevitably results in the formation of grain boundaries during the coalescence of the
hBN islands.

14
Figure 1.11 The growth of hBN on a Ni(111)/sapphire substrate.118 (a) Schematic of hBN domains on
the Ni(111)/sapphire substrate. (b) Photograph of the grown hBN on the substrate. (c) Optical
microscopy image of the marked area in (b). (d) SEM image of the largest hBN domain.

Up to now, the size of grown hBN single crystals by nucleation control is still limited in
millimeter scale. The local feeding method has not been reported in the field of hBN growth yet, despite
of its success in the synthesis of WSSC graphene.

1.2.2 HBN growth via seamless stitching

The synthesis of WSSC hBN films via the route (ii), i.e., seamless coalescence of
unidirectionally aligned hBN islands, is also challenging because no low-index TM surfaces can
template the growth of unidirectionally aligned hBN islands. In general, there are two antiparallel
alignments of triangular hBN islands on both the Cu(111) and Cu(110) surfaces and four dominant
orientations of hBN on the Cu(100) surface.113,119-122 Until 2016, Li et al. for the first time observed the
growth of hBN islands with only orientation on both Cu(102) and Cu(103) surfaces, paving a promising
route for the synthesis of WSSC hBN films.80 After that, great progress has been made for unidirectional
hBN islands growth together with WSSC hBN synthesis on high-index TM surfaces.

In 2019, Wang et al. reported the fabrication of 100 cm2 single crystalline Cu foils that show a
tilted angle of about 1ºdeviating from the ideal Cu(110) surface, and further realized the growth of
WSSC hBN by seamless coalescence of unidirectional hBN islands on the obtained vicinal Cu(110)
surfaces, as shown in Figure 1.12(a-b).84 The coalescence of two aligned hBN islands was examined by
different techniques. The polarized second-harmonic generation (SHG) mapping in Figure 1.12(c)
confirms the seamless stitching of the two aligned hBN domains, as there is no boundary line in the
coalescence area. On the contrary, in Figure 1.12(d), two dark lines representing grain boundaries are

15
presented in the coalescence area of two misaligned hBN islands. To characterize the quality of the
stitching area between the two aligned hBN domains at atomic scale, the grown hBN was transferred
onto a single-crystal graphene TEM grid. It is known that moirépatterns caused by overlapping of two
periodic lattices are very sensitive to the rotation angle of the two lattices, therefore they are widely
used to identify the orientation of the two lattices. From the high-resolution TEM image in Figure 1.12(f)
that was taken from the concave corner of the joint area between two aligned hBN domains, a consistent
moirépattern is observed, implying the exact same orientation of the two hBN domains. Besides, the
seamless stitching of two aligned hBN islands was also confirmed by etching and UV treatment in this
work. In the next year, this research team successfully synthesized WSSC Cu foils with various high-
index surfaces, as introduced in Figure 1.8, and the growth of unidirectional hBN domains was verified
on some of the high-index Cu surfaces.

Figure 1.12 The growth of WSSC hBN on a vicinal Cu(110) surface.84 (a) Optical image of the obtained
Cu foil after oxidation showing the vicinal Cu(110) single crystal has an area of 100 cm2. (b) SEM
image of unidirectional hBN domains on the obtained substrate. (c, d) Polarized SHG mapping of two
aligned (left) and misaligned (right) hBN domains. (e) Low-magnification TEM image of the concave
corner in the joint area of two aligned hBN domains on a monolayer single-crystal graphene film. (f)
Representative HRTEM image of a uniform hBN/graphene moirépattern at the concave corner. The
orange and blue in the insert represent the diffraction pattern of hBN and graphene, respectively.

Right after Wang’s study, a statistical study about the alignment of hBN islands on more than
100 different high-index Cu surfaces was reported, and it turned out that on 30 vicinal Cu surfaces, hBN

16
islands show only one orientation,123 further confirming the great potential of vicinal Cu surfaces for
templating unidirectional hBN growth.

The epitaxial growth of WSSC hBN monolayer was also successfully realized on a 2-inch
Cu(111) wafer with step edges (see Figure 1.13(a-b)).124 By annealing a 500-nm-thick polycrystalline
Cu film on the c-plane of a sapphire wafer at a temperature range of 1040 ~ 1070 °C, a single crystalline
Cu(111) film with atomic step edges on surface was produced (see Figure 1.13(d-e)). The grown hBN
islands on the obtained Cu(111) surface were found to be unidirectionally aligned with an orientation
consistency ratio of more than 99.6% (see Figure 1.13(c)). STM image in Figure 1.13(f) shows that the
consistent moirépatterns even exist over the Cu step edges, and atomic-scale STM image in Figure
1.13(g) shows that the lattice constant of the obtained hBN is 0.25 nm, which agrees well with the
theoretical value.125

Figure 1.13 The growth of WSSC hBN on a highly stepped Cu(111) surface.124 (a-b) Schematic and
photography of as-grown two-inch hBN film on a Cu (111)/sapphire wafer. (c) Optical image of
unidirectional hBN islands on the obtained Cu(111) surface. (d-e) STM images of hBN/Cu showing the
obtained Cu(111) surfaces are highly stepped. (f) STM image showing moirépattern of hBN on the
obtained Cu(111) surface. (g) Atomic-scale STM image of the grown hBN film.

Except for solid single crystalline TM substrates, the growth of WSSC hBN film by seamless
stitching has also been achieved on liquid Au substrates (see Figure 1.14(a-e)).126 It was proposed that
the electrostatic interaction between B and N atoms enables the rotation and alignment of the hBN
islands and then results in the seamless stitching (Figure 1.14(f)). As inferred from Dong’s study, the
grain boundary energy of graphene is a critical factor for the coalescence of graphene islands on liquid

17
Cu surface,90 the grain boundary energy of hBN may also affect the coalescence behavior of hBN on
liquid Au surface. Compared to graphene, the grain boundary energies of hBN are distinctively
higher.127,128 From Figure 1.14(g), we see that there is only one deep local minimum in the grain
boundary energy profile for hBN, indicating that hBN islands on liquid Au surface tend to seamlessly
coalesce into a single crystal from the viewpoint of minimizing the grain boundary energy.120 Moreover,
it should be noted that the hBN flakes on the liquid Au surface show a round shape, which is different
from the triangular hBN islands grown on solid TM substrates. Actually, graphene islands with a
circular shape have also been observed on liquid Cu surface,129,130 whereas the mechanism of the
formation of circular 2D islands remains unclear until now.

Figure 1.14 WSSC hBN grown on a liquid Au substrate. (a) Photograph of a WSSC hBN film
transferred on a SiO2-Si wafer.126 (b-e) SEM images showing the process of hBN film growth on the
liquid Au surface. (f) Illustration for the growth of single crystalline hBN film by self-collimated grains.
(g) Formation energy profile of grain boundaries of hBN as a function of the misaligned angles of hBN
domains.120

In summary, the synthesis of WSSC hBN films via the route (i) is still a great challenge, while
for the route (ii), choosing a proper substrate is of critical importance. To date, high-index TM and
liquid Au surfaces show great potentials in synthesizing WSSC hBN films.

18
1.3 The synthesis of transition metal dichalcogenide (TMDC) films

Different from graphene and hBN, a TMDC monolayer shows a three-atom thickness, with a
middle layer of TM atoms sandwiched between two chalcogen atomic layers.131-133 Different
combinations of TM and chalcogen elements result in more than 100 types of stable TMDC materials,
e.g., MoS2, WS2, MoSe2, WSe2, MoTe2, etc., and a huge number of possible TMDC alloys,134,135 e.g.,
WS2xSe2–2x136 and Mo1−xWxSe2137, etc. Moreover, due to their three-atom-layer structure, TMDCs have
different polymorphs and phases.138-142 For example, the TM atoms between chalcogen atoms may
arrange in either trigonal prismatic lattice (2H or 3R phase) or octahedral lattice (1T phase), 143,144 and
the 1T phase also has different derivatives that are distinguished by different bonding styles of TM
atoms.145-147 The variety in structures gives rise to a large family of TMDC materials with various
electronic properties. The 2H MoS2 monolayer is a semiconductor, while it becomes metallic after a
phase transition from 2H to 1T.148,149 The bandgap of MoS2 undergoes a transition from indirect to direct
when it is scaled down from a bulk to a monolayer structure.150,151 The wide range of electronic
properties of TMDCs make them promising building blocks for the next-generation electronics and
optoelectronics.152-159

To realize the industrial applications of TMDCs, the synthesis WSSC TMDC materials via the
two routes, (i) nucleation control and (ii) seamless stitching, has been wildly explored. Similar to hBN,
WSSC TMDC monolayers via the route (i) has not been reported yet. It is found to be very challenging
to reach a very low nucleation density (only one nucleus over a large area) and the continuous nucleation
during the growth of TMDCs is hardly to be avoided. Currently, the nucleation density of TMDCs is
typically thousands per mm2 so far.160-162 The reasons for the difficulties in lowering the nucleation
density of TMDCs will be presented in Chapter 2.

For the route (ii), a proper substrate that can template the growth of unidirectionally aligned
TMDC islands is prerequisite. However, most of the currently reported studies on the TMDC synthesis
use amorphous (e.g. quartz) or high-symmetric crystalline substrates (e.g. sapphire, mica, GaN and TM,
etc.). On an amorphous quartz substrate, the grown MoS2 islands were found to be randomly
orientated.163-165 On high-symmetric crystalline substrates, such as Au(111),166-168 sapphire(0001),169-171
mica,160 and GaN(0001),172,173 triangular TMDC islands usually present two opposite orientations, as
shown in Figure 1.15(a-c). Until 2018, Lee et al. reported unidirectionally aligned WS2 islands on hBN
films (Figure 1.15(d)).126 Afterwards, unidirectionally aligned MoS2 islands on hBN films was also
realized.174 However, the mechanisms of the well-aligned TMDC islands on hBN films are still unclear,
and it should be noted that antiparallel WS2 islands grown on hBN were also observed (Figure
1.15(e)).175 As demonstrated in Figure 1.15(f), rotating the triangular TMDC island on hBN by 60o
actually results in a different environment of the edges of the TMDC island on the hBN lattice. Thus,

19
the reason for the appearance of antiparallel TMDC islands grown on a hBN surface needs further
exploration.

Figure 1.15 Alignment of TMDC islands on various substrates. (a) Antiparallel MoS2 islands on the
Au(111) surface.168 (b) Antiparallel WSe2 islands on the Al2O3(0001) surface.169 (c) Antiparallel MoS2
islands on the GaN(0001) surface.172 (d) Unidirectional WS2 islands on a hBN film.126 (e) Antiparallel
WS2 on a hBN film.175 (f) Atomic model demonstrating the difference between antiparallel TMDC
islands on a hBN film.

The growth of TMDCs on high-index surfaces with steps has also been explored by several
groups.176-179 In 2015, Chen et al. reported the synthesis of unidirectional WSe2 islands by step-edge-
guided nucleation on a vicinal sapphire (0001) surface, as shown in Figure 1.16.178 Parallel steps with
a height of only ~0.2 nm were formed on the (0001) surface of the Al2O3 wafer after a high temperature
(950 oC) annealing in Ar/H2 atmosphere. The grown WSe2 islands on the obtained Al2O3 wafer showed
a trapezoid shape with a clear alignment along the step edges (Figure 1.16(b)). Atomic force microscopy
(AFM) image in Figure 1.16(c) confirms that WSe2 islands nucleate near the substrate step edges and
are parallel to the step edges. Nevertheless, such well-aligned WSe2 islands did not seamlessly coalesce
into a single crystalline film. Instead, the WSe2 islands on the uphill terraces grow over across those on
the downhill terraces, resulting in overlapping grain boundaries in the synthesized WSe2 films, as
illustrated in Figure 1.16(d-g).

20
Figure 1.16 The growth of WSe2 on a vicinal Al2O3 (0001) surface.178 (a) Photography of the vicinal
Al2O3 (0001) wafer. (b) Optical microscopy image of WSe2 islands on the vicinal Al2O3 (0001) wafer
with SEM image inserted. (c) AFM image showing WSe2 islands nucleated near the step edge of the
substrate. (d-g) The formation of few-layer WSe2 by layer-over-layer overlapping between the layers
grown from adjacent (d-e) and nonadjacent (f-g) substrate steps.

Very recently, centimeter scale MoS2 single crystals were successfully synthesized by seamless
stitching of unidirectional MoS2 islands on a vicinal Au(111) surface, as shown in Figure 1.17.179 A
vicinal Au(111) single crystalline film was obtained by annealing a polycrystalline Au film deposited
on a W foil, unidirectional MoS2 islands were then synthesized on the obtained vicinal Au(111) surface
via ambient-pressure CVD growth. Statistical results showed that ~ 98% triangular MoS2 islands are
unidirectionally aligned, and after ~8 mins of growth, a continuous MoS2 single crystalline film was
formed (see Figure 1.17(c)). The high crystallinity of the continuous MoS2 film was confirmed by onsite
STM measurements. Figure 1.17(d) displayed that a continuous monolayer MoS2 film covers the Au
surface and the MoS2 lattice keeps intact even over the atomic step edges of the Au substrate. Moreover,
water oxidation experiments were used to identify the grain boundaries and it turned out that there are
no grain boundaries in the monolayer MoS2 film obtained from the coalescence of unidirectionally
aligned MoS2 islands, as demonstrated in Figure1.17(e).

21
Figure 1.17 The growth of WSSC MoS2 film on a vicinal Au(111) surface.179 (a) Illustration of the
vicinal Au(111) surface formation and MoS2 growth. (b) Photography of a MoS2 monolayer on the
Au(111)/W substrate. (c) SEM images showing MoS2 growth process at∼720 °C. (d) STM images of a
continuous MoS2 monolayer over single atomic steps of the vicinal Au(111). (e) SEM images of two
type films composed by misaligned (left) and aligned (right) MoS2 domains after water oxidation.

In this section, the experimental progresses on synthesizing WSSC TMDC monolayers are
presented. The nucleation density of TMDCs during CVD is typically over thousands per mm2 and to
synthesize WSSC TMDC monolayers via route (i) seems formidable until now. On the contrary, the
progresses on the seamless stitching of well-aligned TMDC islands is more promising, and centimeter
scale single crystalline MoS2 films have been achieved on a vicinal Au(111) surface, implying the
critical roles of high-index substrates in controlling the alignment of TMDC islands.

1.4 Graphene moirésuperstructures on a substrate

TMs are not only good substrates for the growth of single crystalline 2D materials, but also
bring new structures and properties for 2D materials.180-185 For example, moirépatterns are broadly
observed in the system of a 2D material grown on a TM substrate due to the overlapping of the two
mismatched regular lattices.186-192 Such moirépatterns are very sensitive to the relative angles between

22
the 2D material and the substrate, and therefore are generally used to identify the qualities of the grown
2D materials, as we inferred in Figures 1.12(e-f) and 1.13(f).

Depending on the lattice relationships between graphene and the underlying TM substrates,
graphene moiré patterns present different shapes and sizes. On the Cu(100) and Ni(100) substrate,
graphene moirépatterns show either striped or rhombic shapes, as shown in Figure 1.18,193-196 while on
a (111) surface of a faced centered cubic (FCC) crystal or a (0001) surface of a hexagonal close packed
(HCP) crystal, they always show a regular rhombic shape.197-202 In this thesis, the structures, properties
and applications of graphene moirésuperstructures on various TM surfaces are also explored.

Figure 1.18 Graphene moirépatterns on FCC(100) surfaces. (a) STM topographic images of graphene
on the Cu(100) surface showing stripe pattern and the rhombic pattern.194 (b) STM topographic images
of graphene on the Ni(100) surface showing either stripe pattern or rhombic network depending on the
relative angle between graphene and the substrate, with the corresponding Fourier transforms shown in
inserts.195

1.4.1 Structures of graphene moirésuperstructures

Graphene moiré pattern is not only a visual effect for graphene/TM systems, especially on
active TM surfaces, like Ru(0001),203,204 Rh(111),205,206 Pb(111),207 Pt(111)208,209 and Ir(111),210,211 the
graphene lattice shows regular height variation in the periodicity of the moiré pattern. Figure 1.19
presents atomically resolved STM images of graphene on Ru(0001), Rh(111) and Ir(111) surfaces.
Obviously, the different contrasts indicate different height of graphene lattice on the substrate, and here
we call such corrugated graphene layers as graphene moirésuperstructures.

Besides, many experimental and theoretical studies revealed that the corrugation degree (the
height variation) of graphene moirésuperstructures highly depends on the rotation angle of graphene

23
on the substrate.212-215 On the Pt(111) surface, graphene films present a corrugation degree of 0.5~0.8
Å at small rotation angles of 2º~ 6º,while become ultra-flat at large rotation angles of 14º~30 º.208 On
the Ir(111) substrate, graphene layer with a smaller rotation angle corresponds a larger graphene
corrugation, and that with zero rotation angle shows the largest corrugation degree.216

Figure 1.19 Atomically resolved STM images of graphene moirésuperstructures on (a) Ru(0001), 204
(b) Rh(111),205 (c) Ir(111)197 surfaces.

1.4.2 Applications of graphene moirésuperstructures

Except for identifying the quality of the grown graphene as introduced in above, the unique
corrugated graphene moirésuperstructures lead to inhomogeneous properties of the graphene layers,
which brings about potential applications in electronic devices.198,217,218 The inserts of Figure 1.20
illustrate the height profiles of graphene monolayer on different TM substrate. Compared with the
weakly corrugated graphene layers on the Pt(111) and Ir(111) surfaces, the C 1s photoelectron
spectrums of the strongly corrugated graphene on the Rh(111) and Ru(0001) surfaces exhibit a split
double-peak, indicating two distinct types of bondings in graphene.219 Besides, it was revealed that the
highly corrugated graphene moirésuperstructure on the Ru(0001) surface presents quantum-dot-like
behaviors with confined effects in both lateral and vertical directions, and the hump region of the
superstructure has a higher local work function than other regions.220,221

24
Figure 1.20 C 1s photoelectron spectrums of different graphene/TM superstructures and that of the
highly oriented pyrolytic graphite (HOPG) is used for comparison, with the side views of graphene/TM
superstructures are illustrated. MG means monolayer graphene.219

Moreover, it is found that the adsorption abilities of graphene/TM superstructures are site-
dependent, making them good templates for the synthesis of metal clusters and the assembly of organic
molecules.222-230 According to the lattice relationship between graphene and the underlying TM
substrate, a unit cell of graphene/TM superstructure can be divided into four regions, namely ATOP,
FCC and HCP and Bridge sites,197,231 as illustrated in Figure 1.21.

Figure 1.21 Illustration of lattice relationships of graphene on the four high-symmetric sites of (a) an
FCC(111) and (b) an HCP(0001) surfaces.

25
Interestingly, the adsorption ability of a graphene/TM superstructure for different TM metals
are distinguishable.224,228,232,233 For instance, on the graphene/Ru(0001) template, Pt, Ru, Ir, Ti and Rh
atoms prefer to form small dispersed nanoclusters at FCC sites, while Pd, Au, Ag, Cu and Co prefer to
form large islands by covering different regions with similar coverages, here the formed Pt cluster and
Pd islands are displayed in Figure 1.22(a-d).234,235 Graphene/TM superstructures have also been used in
templating the assembly of organic molecules,236-239 it was revealed that Fe phthalocyanines (Pc)
molecules can duplicate the lattice of the graphene/Ru(0001) superstructure and form a Kagome lattice
that is composed of four neighbored lattice points located in interlaced triangles (see Figure 1.22(e-
f)),240 whereas CoPc molecules self-assemble into a nearly squire lattice on the graphene/Ir(111) surface
despite of the hexagonal corrugated template (see Figure 1.22(g-h)).241

Figure 1.22 Synthesis of TM clusters and assembly of Kagome lattices templated by graphene/TM
superstructure. (a-d) SEM images of (a-b) Pt clusters and (c-d) Pd clusters on the graphene/Ru(0001)
substrate.234 (e) STM image of Kagome lattice of FePc molecules on the graphene/Ru(0001) substrate
with a marked unit cell.240 (f) Atomic structure of the Kagome lattice in (e). (g) STM image of squire
lattice of CoPc molecules on the graphene/Ir(111) substrate.241 (h) Atomic structure of the squire lattice
in (g).

In summary, modulated by TM substrates, graphene displays corrugated moirésuperstructures


and inhomogeneous surface property. Depending on the type of the substrate and the rotation angle on
the substrate, graphene moiré superstructures are distinct. The versatile moiré superstructures in
graphene make it a good candidate in templating the synthesis of metal clusters and organic molecular

26
frameworks. In addition, the inhomogeneity in the electronic structure of graphene induced by the moiré
superstructures also hold promising applications in nanoelectronics.

1.5 Conclusions

The excellent physical and chemical properties of 2D materials make them hold great potentials
in a vast range of applications. To achieve their industrial applications, obtaining WSSC 2D materials
is of critical importance. To date, there are mainly two routes developed in synthesizing WSSC 2D
materials, which are (i) formation and growth of only one nucleus of the 2D material on the whole
substrate and (ii) seamless stitching of unidirectionally aligned 2D material islands on the substrate.
After synthesis, 2D materials on substrates show distinct structures and properties due to the interactions
with the substrates, which further widens the applications of the 2D materials. In this chapter, the growth
of WSSC 2D materials by CVD methods via the two routes are reviewed first, and then the structures,
properties and potential applications of graphene moiré superstructures on TM substrates are also
introduced.

To date, WSSC graphene has been successfully synthesized on polycrystalline Cu and Cu-Ni
substrates via route (i) by the local feeding technique, however this route usually takes hours and
requires a delicate experimental setup. In contrast, route (ii) is more cost-effective for industrial
production, because multiple nuclei can grow simultaneously. By this route, choosing an appropriate
substrate that can guide the growth of unidirectionally aligned 2D islands is essential, and up to date,
the synthesis of WSSC graphene, hBN and TMDC films has been successfully realized by this route.

In the past decade, CVD growth of graphene on almost all types of TM substrates has been
explored, and it was found that on most TM substrates, graphene shows regular corrugations in the
periodicity of the moirépattern, and such corrugated moirésuperstructures are very sensitive to the
rotation angle of graphene on the substrate and the type of the substrates. Besides, graphene/TM
superstructures have been broadly used as templates for the synthesis of metal clusters and the assembly
of large organic molecules.

1.6 Research motivation

Having reviewed the current processes of 2D material growth, it is clear that the most promising
method to synthesize WSSC 2D materials is by the seamless stitching of unidirectionally aligned 2D
islands on a single crystalline substrate, in this case, to figure out the alignment mechanisms of 2D
materials on a substrate and to obtain single crystalline substrates are essential.

27
The alignment of 2D material islands on a substrate is determined by the interactions between
the 2D material and the substrate, which means that an in-depth understanding on the interaction 2D
materials and substrates is of critical importance for choosing appropriate substrates to synthesize
WSSC 2D materials.

In early stage, researchers mainly focused on synthesizing 2D materials on low-index TM


surfaces, but only graphene islands grown on the Ge(110) and Cu(111) surfaces show unidirectional
orientation, while unidirectional hBN and TMDC islands have never been reported on low-index
surfaces. Until 2018, it is found that high-index surfaces are more promising for the growth of
unidirectionally aligned 2D islands because the existence of step edges breaks the substrate symmetries,
and right after, great success has been made in the growth of WSSC hBN and TMDC. It should be
emphasized that the number of low-index substrates is limited, i.e., there are only three types of low-
index surfaces for FCC TMs, namely (100), (110) and (111). Deviated from low-index surfaces, there
are, in principle, countless high-index surfaces. However, up to now, a systematic understanding on the
alignment of 2D islands on high-index substrates at atomic scale is lacking and the understanding on
the interactions of 2D materials and high-index surface is far from completion. Therefore, in this thesis,
we first present a systematic theoretical study on the alignment of various 2D materials on high-index
substrates, aiming to reveal the alignment mechanism at atomic scale and provide guidelines for
experimental designs on the synthesis of WSSC 2D materials.

To realize the method of seamless coalescence of unidirectionally aligned 2D islands to


synthesize WSSC 2D materials, single crystalline substrates are usually required. As reviewed in this
chapter, there have been several strategies to fabricate WSSC substrates, such as depositing TMs on a
single crystalline wafer, temperature-gradient-driven annealing of polycrystalline TM foils into a single
crystal, seeded growth of single crystalline TM foils from polycrystals. In collaboration with
experimental groups, we have reported a contact-free annealing method of transforming polycrystalline
TM foils into single crystals. In this thesis, we will also present the mechanism of the contact-free
annealing method at atomic scale.

After the growth of graphene on a substrate that shows mismatching lattice constants with
graphene, the graphene layer shows corrugated moirésuperstructures due to the non-uniform graphene-
TM interaction. Although the corrugated moiré superstructures are revealed to be sensitive to the
rotation angle by many experimental observations, most theoretical studies only focused on the moiré
superstructure with a 0o rotation angle, and the systematic study on the evolution of graphene moiré
superstructure and its properties with respect to the rotation angle is rare. Such a study is important for
various applications of the corrugated structures, such as to grow size-tunable metal clusters. Moreover,
it was reported that graphene shows an ultra-flat structure at large rotation angles, but the formation

28
mechanisms of the corrugated and ultra-flat graphene structures have not been explored yet. At the final
part of this thesis, all these questions will be studies at atomic scale by extensive theoretical calculations.

29
Chapter 2 Theoretical foundation on the growth of 2D materials

For the growth of 2D materials by CVD, TM surfaces are the most popular substrates because
they not only serve as templates but also play important roles in catalyzing the decomposition of the
feedstock molecules into active precursors, allowing the active precursors to diffuse on it and facilitating
the incorporation of the active precursors to the edges of 2D films.242,243 As introduced in Chapter 1, by
using TM surfaces as substrates, WSSC graphene has been realized via both two routes: (i) single
nucleus growth and (ii) seamless stitching of unidirectional aligned 2D islands, while WSSC hBN and
TMDCs have only been synthesized by the route (ii). A deep insight into the growth mechanisms of 2D
materials is of crucial importance for understanding the growth behaviors of 2D materials through the
two routes, and further for providing guidelines to experimentalists to synthesize WSSC 2D materials.

To date, extensive theoretical studies have been carried out to study the nucleation and growth
mechanisms of graphene244-254 and hBN,111,255-261 while the exploration on the synthesis mechanism of
TMDCs is very limited.43,262,263 In this Chapter, the theoretical progress in understanding the synthesis
mechanisms of 2D materials will be reviewed.

2.1 CVD growth of single-crystal graphene

2.1.1 Nucleation of graphene on a substrate

By DFT calculations, Gao et al. explored the stabilities of small carbon structures with different
configurations and different sizes on the Ni(111) surface to investigate the nucleation of graphene, as
illustrated in Figure 2.1(a-b).244 Compared with a carbon ring, a carbon chain of the same size always
show a slightly lower formation energy because both ends of the carbon chain are well passivated by
the TM surface. More importantly, carbon chains are more stable than carbon sp2 networks when the
number of carbon atoms is less than 13, while carbon sp2 networks become more stable once the number
of carbon atoms exceeds 13.

30
Figure 2.1 Nucleation of graphene. (a) The most stable structures of C11 and C12 on the Ni(111)
surface.244 (b) Formation energies of various carbon structures on the Ni(111) surface (c) The most
stable structures of C7 and C10 near a step edge of the Ni(111) surface.245 (d) Energies of carbon
structures on a Ni(111) terrace or near a step versus the cluster size (N). (e) Nucleation barrier (G*) and
corresponding nucleation size (N*) of graphene on a Ni(111) terrace or near a step. (f) Nucleation rate
of graphene on a Ni(111) terrace or near a step.

The effect of step edges on the nucleation of graphene has also been studied, as displayed in
Figure 2.1(c-d), both carbon chains and carbon networks near a Ni step edge are more stable than those
on a Ni(111) terrace,245 implying that graphene nucleation near a step edge of the substrate is more
preferred. Besides, it was found that the critical number of carbon atoms for the most stable carbon
cluster changing from a sp chain to a sp2 network is reduced to 10 near a step edge. Due to its mono-
element-composition, the nucleation barrier (G*) of graphene is only a function of the chemical potential
difference (Δμ) of carbon precursors on a TM surface in relative to the graphene film, and from Figure
2.1(e) we see that the nucleation barrier of graphene on a Ni(111) terrace is always larger than that near
a Ni step edge, and the critical nucleation size of graphene near a Ni step edge is smaller than that on a
terrace area of the substrate. In general, a high nucleation barrier leads a low nucleation rate and thus a

31
low nucleation density. Figure 2.1(f) shows the nucleation rates of graphene on a terrace and near a step
edge, the smaller Δμ value, the lower nucleation rate. For example, at Δμ =0.2 eV, the nucleation rate
of graphene on the Ni(111) terrace is smaller than 10-12 cm-2s-1, while that near a step edge is several
orders of magnitude higher (~10-3 cm-2s-1). Therefore, under a near equilibrium condition and on TM
surfaces with a low density of step edges, it is possible to obtain only one graphene nucleus over the
whole substrate and make it grow in to a WSSC graphene film, which has been realized by Wu et al. in
2016.66

The stabilities of carbon clusters with the number of carbon atoms in a range of 16 ~ 26, which
are possible precursors during graphene nucleation, were also explored by DFT calculations.246,247 The
formation energy profile and the corresponding second order derivatives of C16~26 clusters on 4 different
TM surfaces are shown in Figure 2.2 (a1). Obviously, C21 and C24 are the two most stable carbon clusters
regardless the types of the substrate.246 Figures 2.2 (a2) and (a3) present the atomic configuration of the
most stable C21 and C24 clusters, where the C24 cluster is composed of 7 hexagons while the C21 consists
of 4 hexagons and 3 pentagons. The curved planar structure of the C21 cluster induced by carbon
pentagons results in a better passivation of the cluster edge by the TM substrates, because a graphene
edge tends to stand upright on the TM substrate.264 To testify the configuration of the C21 cluster, its
scanning tunneling microscopy (STM) images on the Rh(111) surface at voltages of -1.0 V and 1.0 V
were simulated, which are well consistent with experimental observations of carbon clusters on Rh(111)
surface at the initial stage of graphene growth, as demonstrated in Figures 2.2 (a4-7).

Figure 2.2 (a1) Formation energy profile (upper panel) of carbon clusters on different TM surfaces and
its second order derivative (lower panel). The structures of magic C21 and C24 clusters are shown in (a2)
and (a3). The simulated STM images of C21 on Rh(111) surface at the voltages of -1.0 V and 1.0 V are
shown in (a4) and (a5), respectively. Experimentally obtained STM images of a carbon cluster on
Rh(111) surfaces during CVD process are shown in (a6) an (a7).246 (b-c) Energies of C clusters with
and without dangling C atoms on the Ru(0001) surface. 247

32
It should be noted that, a branched C24 carbon cluster with three dangling carbon atoms
respectively attaching to the three pentagons of the C21 structure was found to be more stable than the
C24 cluster purely consisting of carbon hexagonal rings on both the Ru(0001) and Rh(111) surfaces.247
As exhibited in Figure 2.2(b-c), changing three of the six hexagons at the edge of a C24 cluster consisting
of only hexagonal rings to three pentagons with three dangling carbon atoms decreased the formation
energy of the carbon cluster by 1.99 eV. Similarly, attaching one carbon atom to each of the three
pentagons of the C21 cluster can decrease the total energy of the systems by 2.6 eV. Therefore, the
branched C24 carbon cluster with three dangling carbon atoms is also a possible precursor for graphene
nucleation. It can be seen that carbon pentagons might be essential to construct stable carbon clusters
on various TM surfaces because of the curved planar structure induced by them, and in fact, such carbon
pentagons were frequently observed in MD simulations of graphene growth based on different force
fields.245,248,250

Above discussions show that the stabilities of various carbon structures, including carbon
chains, carbon rings and sp2 carbon networks, on various TM metal surfaces have been extensively
studied, and the transition from carbon chains to sp2 carbon networks in the nucleation process of
graphene growth has been revealed by theoretical studies. It was found the nucleation barrier is
dependent of the chemical potential of carbon and a low nucleation density can be reached by adjusting
the ambient environment, which proves that it is possible to synthesize WSSC graphene from formation
and growth of only one graphene nucleus in the whole substrate surface.

2.1.2 Mechanisms of graphene growing across substrate grain boundaries

As reviewed in Section 1.1.1, the growth of a WSSC graphene film from one nucleus has been
realized on polycrystalline TM surfaces, suggesting that graphene can grow across the substrate grain
boundaries without changing its crystallographic orientation. In this case, the epitaxial relationship
between graphene and the underlying TM substrate should be reexamined. Provided that the epitaxial
relationship between graphene and the TM substrate is strictly maintained, the grown graphene copies
the grain texture of the substrate and graphene grain boundaries should be formed just above of the
grain boundaries of the substrate, as illustrated in the path 1 of Figure 2.3(a-b), and consequently a
polycrystalline graphene film is formed on the polycrystalline surface, which contradicts with the
experimental observations.265

33
Figure 2.3 Growth of graphene on polycrystalline TM substrates. (a-d) Illustrations of two paths of
graphene grow crossing the grain boundaries of a substrate.265 (e) Growth processes of the two paths of
graphene grow crossing a grain boundary of a substrate.266 (f) The corresponding formation energy
profiles of two paths in (e).

Due to the weak vdW interaction between graphene and the substrate, there might be no enough
driving force for the formation of a graphene grain boundary to copy the substrate surface texture when
graphene grows across the substrate grain boundary. As a result, the graphene lattice is maintained when
passing through the grain boundaries of the substrate, as illustrated in the path 2 of Figure 2.3(a, c-d),
suggesting that the growth of graphene on a TM surface is not a strict epitaxial process. It can be seen
that a deep insight into the interactions between graphene and TM substrates is the key to understand
the mechanism of graphene growth.

DFT calculations have shown that the interactions between graphene and most TM surfaces are
very weak and not sensitive to the orientation of graphene on the substrate. 265 Dong et al. studied the
epitaxial and non-epitaxial graphene growth on various TM surfaces, and found that the binding energy
difference between epitaxial and non-epitaxial graphene on the TM surface is not large enough to
facilitate the formation of grain boundaries in graphene to maintain its epitaxial growth when it grows
across the substrate grain boundaries.266 As shown in Fig. 2.3(e-f), for the epitaxial graphene growth on
a polycrystalline Cu substrate (path 1), the formation energy increases sharply when a grain boundary
is formed in graphene, while the non-epitaxial graphene growth (path 2) has no obvious formation
energy increase when graphene grows across a grain boundary of the substrate. This study clearly
proves that graphene growth on a TM substrate is not a strict epitaxial process and in principle it is
possible to synthesize WSSC graphene on a polycrystalline substrate by the nucleation control method.

34
Experimentally, single crystalline graphene islands growing over grain boundaries or even step edges
of a substrate without forming any corresponding line defects have been widely observed.267-272

In summary, we conclude that graphene growing across grain boundaries of a TM substrate


without changing its orientation is energetically favorable, which offers possibilities for synthesizing
WSSC graphene on polycrystalline TM substrates. Similarly, single crystalline hBN and MoS2 flakes
growing over grain boundaries of the substrate have also been observed.117,273 However, it should be
reminded that synthesizing WSSC 2D materials through the nucleation control method also requires a
high nucleation barrier of the 2D material, so that only one nucleus can be formed on the whole substrate.
Currently it is still challenging to control the nucleation densities of hBN and TMDC, which will be
explained in following sections.

2.1.3 Mechanisms of alignment and coalescence of graphene islands on a TM surface

The route of synthesizing WSSC graphene films via seamless stitching of large number of
graphene islands requires all graphene islands have the same orientation on a wafer-scale
substrate.68,69,77,78,274 In general, a single crystalline TM substrate is required for this method. Here, we
discuss the effect of the interaction between graphene and a substrate on the orientation of graphene
islands on the substrate.

Experimentally, unidirectionally aligned graphene islands were broadly observed on single


crystalline TM surfaces, especially on the Cu(111) surface, graphene islands generally present a
hexagonal shape with their ZZ edges aligning along Cu<110> directions.42,275 By DFT calculations,
Zhang et al. explored the aligned mechanism of graphene islands on Cu(111) surface and the
interactions between the Cu(111) substrate and graphene bulk/edge were studied.265 Figure 2.4(a)
reveals that the graphene bulk - Cu(111) interaction is strongest at the 0ºrotation angle with graphene
ZZ direction paralleling to the Cu<110> direction, meanwhile the distance between graphene film and
the Cu(111) surface reaches the minimum value. It is noted that the interaction between the graphene
bulk and Cu(111) surface is not sensitive to the rotation angle, showing a binding energy variation
within 5 meV per carbon atom. When a graphene island is grown on the Cu(111) surface, the graphene
edge also interacts with the Cu(111) surface. Figure 2.4(b) shows the binding energy profiles of C24 and
C54 clusters to the Cu(111) surface as functions of the rotational angle, where the edges of the carbon
clusters are passivated by the Cu substrate rather than hydrogen. It is obvious that the interaction is very
sensitive to the rotation angle, and the rotation barrier increases significantly with the size of the cluster,
indicating the orientation of a graphene island is determined in the early stage of its growth if its edges
are passivated by the substrate. In addition, from this study, we can also know that the vdW interaction
between the graphene bulk and the substrate will determine the orientation of a graphene island when
its edges are well-passivated by hydrogen rather than the substrate.

35
Figure 2.4 Interaction between graphene and the Cu(111) substrate.265 (a) Optimized distance and
binding energies of graphene layer on the Cu(111) surface versus the rotation angle (θ). (b) Energy
profiles of C24 and C54 clusters versus the rotation angle.

In practice, it is difficult to obtain atomic flat substrate surfaces at macroscale, instead step
edges on Cu substrates were broadly observed during the graphene CVD growth.276-281 Therefore, the
effect of the step edges of substrates on the graphene orientation cannot be ignored. Due to the compact
atomic arrangement and high stability, step edges along <110> directions are mostly formed on low-
index surfaces of an FCC TM foil.124,282 Gao et al. theoretically compared graphene nucleation on a flat
Ni(111) terrace and near a Ni<110> step edge.245 As shown in Figure 2.5(a), the formation energies of
both graphene ZZ and AC edges attaching to the Ni<110> step edge are much lower than those attaching
to a flat terrace, suggesting that the nucleation of graphene near a step edge is more energetically
preferred. Yuan et al. further explored the formation energies of various graphene edges attaching to a
Cu<110> step edge.283 As shown in Figure 2.5(b), the lowest formation energy appears at the 0ºrotation
angle, where the graphene ZZ edge aligns along a Cu<110> step edge, which is attributed to the
straightness of both the graphene ZZ edge and Cu<110> step edge and the small lattice mismatch
between them.

36
Figure 2.5 Interaction of a graphene layer near a step edge of TM substrates. (a) Structures and
formation energies of graphene ZZ and AC edges attaching to a Ni(111) terrace and a Ni(111) step
edge.245 (b) Formation energies of various graphene edges and a Cu<110> step edge versus the
orientation angle.283

Except for the alignment of graphene islands on a TM surface, the coalescence behaviors of
well-aligned and mis-aligned graphene islands have also been revealed, as shown in Figure 2.6.79 The
attachment of C atoms to a concave corner formed from the coalescence of two well-aligned graphene
islands leads to a decrease of the formation energy (Figure 2.6(a)). Figure 2.6(b) shows the
transformation of a sharp concave corner to a dull concave corner by sequentially attaching C atoms to
low-energy sites of the concave corner formed from the coalescence of two well-aligned graphene
islands, which can be used as an indicator of the grain boundary-free coalescence between two aligned
graphene islands, as introduced in Figure 1.7. On the contrary, the concave corner formed from the
coalescence of two mis-aligned graphene islands will remain sharp during the growth process. As shown
in Figures 2.6(d-e), the coalescence of two mis-aligned graphene islands leads to the formation of a
grain boundary between them and the concave corner corresponds to one end of the grain boundary.
Adding C atoms to the concave corner between two mis-aligned graphene islands always results in an
increase of the formation energy, while the structures with attaching C atoms to the two ZZ edges of
the concave conner show lower formation energies, indicating that the growth of concave corner
between mis-aligned graphene islands is governed by growth of the two ZZ edges of the concave corner
(Figure 2.6(f)) and as a result, the sharp concave structure remains unchanged. It should be noted that
the coalescence behavior of graphene also applies to other 2D materials, as also demonstrated in this
study.

37
Figure 2.6 Theoretical analyses of the coalescence of two well-aligned or mis-aligned graphene
islands.79 (a-b) The process of adding C atoms to a concave corner between two well-aligned graphene
islands. (c) Schematic of the growth behavior at a concave corner between two well-aligned graphene
islands. Red arrows mark the kink sites and black arrows indicate growth direction. (d-e) The process
and the corresponding formation energy (ΔEF) profile of adding C atoms to a concave corner between
two mis-aligned graphene islands. (f) Schematic of kink formation at graphene edges near a concave
corner between two mis-aligned graphene islands.

In summary, the orientation of a graphene island on a substrate is determined by both the


interaction between the graphene bulk and the substrate and the interaction between graphene edges
and the substrate. In addition, the coalescence behaviors of well-aligned and mis-aligned graphene
islands are distinguishable from the growth behavior of the concave corners formed from the
coalescence of neighboring graphene islands. Above theoretical studies proved that on an ideal (111)
surface of an FCC TM, a graphene island prefers to align its ZZ direction aligning along the <110>
direction of the substrate if the graphene edges have been passivated by hydrogen. In addition, if the
graphene growth is far from equilibrium and the graphene edges are passivated by the substrate, multi-
orientations of graphene islands will appear, because of the high rotation barriers of small graphene
clusters on the substrate. In practice, step edges appear on the substrates and it was found that the

38
graphene ZZ edge prefers to align along the substrate <110> step edges. Currently, we can see that, to
ensure graphene islands that have a C6V symmetry to show a unidirectional orientation, Cu substrates
that have only one <110> direction or several <110> directions with an 60o misorientation angle with
respect to each other are required. Comparing with Cu foils with high-index surfaces, low-index Cu
foils with high symmetries, such as Cu(111), Cu(110) and Cu(100), are more easier to be obtained by
annealing polycrystalline Cu foils because of their lower surface energies. Among them, the Cu(111)
surface has three equivalent <110> directions with an 60o misorientation angle and the Cu(110) has
only one <110> direction, implying that both surfaces can be used to template the growth of
unidirectional graphene islands. In fact, The synthesis of WSSC graphene films by the seamless
stitching method have already been realized on single crystalline Cu(111) surfaces.67,69,78 In contrast,
the grown graphene islands on the Cu(100) surface that has two orthogonal <110> directions generally
shows two orientation.284-287 In Chapter 4, we will present a general theory on the epitaxial growth of
2D materials on an arbitrary substrate.

Last but not least, we would like to note that the strength of graphene-substrate interaction also
plays an important role in the coalescence of aligned graphene islands. For example, on the chemically
active Ni(111) substrate, the coalescence may induce a line defect despite the graphene islands are
unidirectional, because the relatively strong interaction between graphene and the Ni(111) surface
prevents the translational sliding of the graphene islands on the substrate.288,289 In contrast, on the
Cu(111) or Ge(110) surfaces that have a weak interaction with graphene, the unidirectionally aligned
graphene islands can seamlessly merge into a perfect single crystal, which were discussed in Chapter 1.

2.2 CVD growth of single-crystal hBN film

2.2.1 Nucleation of hBN on a substrate

Compared to graphene, the nucleation of hBN is more complicated due to its binary
composition, i.e., the chemical potentials of both boron and nitrogen can affect the nucleation and
shapes of hBN, furthering inducing different shapes of grown hBN films.107,290-295 Liu et al. investigated
the stabilities of small BN structures, including BN chains and sp2 BN networks, on Cu(111) surface
by DFT calculations.255 As shown in Figure 2.7(a), Bn-1Nn chains and clusters are more stable than BnNn
and BnNn-1 structures at a N-rich condition (∆μ = -0.91 eV, which is the chemical potential difference
between boron and nitrogen), while BnNn-1 chains and clusters becomes more stable at a B-rich
condition (∆μ = 3.0 eV). Moreover, similar to graphene nucleation, BN networks always become more
stable than BN chains when the number of B and N atoms is larger than 13, regardless of the ambient

39
condition. It is worth noting that due to the unfavorite of B-B and N-N bonds,127,296,297 there are no
pentagonal rings in the stable BN networks, which is different from graphene clusters.

MD simulations were also carried out to explore the nucleation and growth process of hBN.256
Figure 2.7(b) presents that on the Ni(111) surface, boron and nitrogen monomers combine together to
form BN chains, and the chains then link with each other to form “Y” junctions, where hexagonal rings
begin to form, and sp2 hBN networks are gradually produced around the formed hexagonal rings.
Besides, MD simulations also revealed that the morphology of the formed hBN networks were affected
by the concentrations of boron and nitrogen. Under concentration ratios of boron to nitrogen to be 1:1
and 1:2, the formed hBN network shows a triangular shape with boron-terminated ZZ edges, while
decreasing the ratio to 1:2.5 ~ 1:5 results in the appearance of nitrogen-terminated ZZ edges.

Figure 2.7 Theoretical analysis of hBN nucleation. (a) Formation energy profiles of BN chains and BN
sp2 networks on the Cu(111) surface under a N-rich (∆μ = -0.91 eV), a medium(∆μ = 1.0 eV) or a B-
rich (∆μ = 3.0 eV) condition, where ∆μ is the chemical potential differences of boron and nitrogen.255
(b) MD simulation of hBN nucleation and growth on the Ni(111) surface.256

For a 2D nucleus, the nucleation barrier mainly originates from the edge formation energy, and
a high edge formation energy suggests a low nucleation rate. For hBN, the formation energies of edges
with unbalanced stoichiometry are dependent on the chemical potentials of nitrogen and boron and there
are reconstructed edge structures.296-300 Zhao et al. theoretically investigated several hBN edges with

40
different structures and calculated their formation energies on various TM substrates, and the results on
the Cu(111) are shown in Figure 2.8.301 Depending on the chemical potential difference of boron and
nitrogen (∆μ), hBN edge shows different configurations, and the edge formation energy as well as the
nucleation barrier can be very low in some ∆μ ranges, implying that a low nucleation density of hBN is
difficult to achieve in experiments.119,302

Figure 2.8 (a) Formation energies of various hBN edges on the Cu(111) surface as a function of the
chemical potential difference of boron and nitrogen (Δμ).301 (b) The most stable hBN edge
corresponding to Δμ. (c) Atomic configurations of the most stable edges corresponding to (b).

2.2.2 Mechanisms of hBN alignment on a TM surface

Compared to the C6V graphene, the C3v hBN islands usually present a triangle shape and
generally show multi-orientation on low-index TM surfaces.303-305 As shown in Figure 2.9(a), hBN
islands on the Cu(111), Cu(110) and Cu(100) surfaces have two, two and four equivalent orientations,
respectively, consistent with the binding energy profiles as functions of the rotation angle of a triangular
hBN cluster on these surfaces (Figure 2.9(b)), which show that there are two maxima in binding energy
profile in one periodicity of the hBN/Cu(111) system and four maxima in the binding energy profile in
one periodicity of the hBN/Cu(100) system.113 Previous DFT studies also demonstrated that the vdW
interaction between the hBN bulk and the FCC(111) surface also has two local maxima, corresponding
to the lattice relationship with N atoms located on the top of metal atoms and B atoms on the fcc and
hcp sites, respectively, which are the same as those illustrated in Figure 2.9(a).304,306-308 Therefore, it can
be inferred that a ZZ direction of a hBN island prefers to align along the <110> direction of low-index
FCC TM surfaces.

41
Figure 2.9 The alignment of h-BN islands on low-index TM substrates.113 (a) Atomic structures of hBN
on the Cu(111), Cu(110) and Cu(100), respectively. (b) The binding energies between hBN and Cu
substrate as a function of their relative angle.

Different from low-index surfaces, a high-index surface normally possesses a lower symmetry
because the existence of atomic step edges and/or kinks at the step edges. Interestingly, hBN islands
with unidirectional orientation are observed broadly on such low symmetric surfaces.123 In 2016,
unidirectionally aligned hBN islands were first reported on Cu(102) and Cu(103) surfaces (see Figure
2.10(a-b)), and DFT calculations in Figure 2.10(c-d) confirmed that there is only one minimum in the
binding energy profiles of a triangular hBN island on these two surfaces as a function of the island
rotation angle, corresponding to the configurations with one ZZ edge of hBN paralleling to the step
edge direction (see Figure 2.10(e-f)).80 It is worth noting that on the high-index Cu(102) and Cu(103)
surfaces, the step edges are along Cu<010> direction, which is different with the case of hBN on low-
index Cu surfaces, where one of ZZ directions of hBN prefers to align along the Cu<110> direction.
Obviously, step edges on a substrate play a critical role in determining the alignment of grown hBN
islands.

42
Figure 2.10 (a-b) SEM images of grown hBN triangles on Cu(102) and Cu(103) surfaces with a scale
bar of 20 μm.80 Black dashed arrows represent the Cu[010] direction. (c-d) Energy profile of hBN on
Cu(102) and Cu(103) surfaces as a function of their relative angle. (e-f) Atomic configurations with the
lowest energy corresponding to (c-d).

In Chapter 1, Figure 1.12 and Figure 1.13 present the experimental success in WSSC hBN
synthesis by seamless stitching of unidirectionally aligned hBN islands on a WSSC vicinal Cu(110)
surface and a WSSC Cu(111) surface with a high concentration of step edges, respectively, and we
introduce the alignment mechanisms of hBN islands in these two works.

For hBN grown on a vicinal Cu(110) surface, an atomic-resolution STM image (Figure 2.11(a))
proved that the hBN island shows a ZZ edge at the interface between the grown hBN and the obtained
surface, and low-energy electron diffraction (LEED) pattern confirmed that the step edges on the
obtained surface is along Cu<211> direction.84 We performed theoretical calculations to explore the
formation energies of interfaces between various hBN edges with the Cu<211> step edge. As shown in
Figure 2.11(b-c), where γ denotes the relative angle between a hBN ZZ direction and the Cu<211> step
edge, there is only one local minimum located at γ = 0º, where the nitrogen-terminated ZZ edge of hBN
is parallel with the Cu<211> step edge, confirming the unique orientation of hBN islands on the
obtained vicinal Cu(110) substrate.

43
Figure 2.11 Alignment mechanism of hBN on vicinal Cu(110) surfaces. (a) Atomic-resolution STM
image showing the interface between hBN edge and Cu substrate.84 (b) Atomic configuration of the
interface of a nitrogen-terminated hBN ZZ edge aligning along a Cu<211> step edge of a vicinal
Cu(110), γ is the relative angle between them. (c) Formation energy of hBN edges attaching to the
Cu<211> step edge as a function of γ. (d) Schematic of meandering step edges on a high-index
substrate.309 (e) Schematic of interface between a hBN island and a tilted step edge with kinks. (f)
Atomic configuration of the interface between a tilted hBN edge and a tilted step edge.

Practically, step edges on a substrate cannot be always straight, instead curved step edges with
varying directions are frequently observed, especially under a high temperature for CVD growth of 2D
materials.310-312 The effect of such curved step edges on the alignment of hBN islands has also been
investigated.309 Figure 2.11(d) illustrates the meandering step edges on a substrate, due to the variation
of step edge direction, hBN islands along such step edges may show different orientations. At the atomic
scale, a meandering step edge consists of straight atomic segments and kinks. As shown in Figure
2.11(e), only if there is a tilted hBN edge which has kinks complementary to the meandering step edge,
the orientation of hBN islands and therefore their unidirectional alignment can be maintained during
the growth. From the atomic configuration in Figure 2.11(f), it can be seen that the kink height
difference between hBN ZZ edge and Cu<211> step edge is only 0.08 Å, which is the reason for the
tight docking of hBN ZZ edge to the Cu<211> step edge. For Cu step edges along other directions on
vicinal Cu(110) surfaces, their kink heights show much larger difference with that of ZZ edge of hBN,
and therefore step edges along Cu<211> direction is the key to ensure unidirectionally aligned hBN
islands. Besides, due to the existence of step edge, the seamless coalescence of unidirectionally aligned
hBN islands on vicinal Cu surfaces requires that the step height is close to the sp2 B-N bond length
(1.44 Å), and the step height of Cu<211> step on vicinal Cu(110) surfaces is 1.27 Å. Therefore,

44
seamless coalescence of hBN islands on such kind of Cu surfaces can be realized, which has been
verified by the MD simulations.

The alignment mechanism of hBN islands on a highly stepped Cu(111) surface was also
studied.124 As discussed above, triangular hBN islands should have two alignment orientations on an
ideal Cu(111) surface. As shown in Figure 2.12, the STM image shows that the step edges on the
obtained Cu(111) surface are composed of two typical step-edge terminations (A and B). By DFT
calculations, the binding energies of a B6N7 cluster attaching to these two types of step edges and to a
flat Cu(111) terrace are compared, it turns out that the binding energy of the two most stable
configurations, NIBII (60°) and NIBIII (0°), attaching to the Cu(111) terrace are almost degenerate,
whereas a difference of ~ 0.23 eV appears when attaching to the two types of step edges. Besides, the
binding energy difference can further increase in proportional to the size of the hBN cluster, making
the NIBII configuration to be the dominant orientation on the highly stepped Cu(111) surface.

Figure 2.12 Alignment mechanism of hBN on highly stepped Cu(111) surfaces.124 (a) STM image of a
highly stepped Cu(111) surface and schematic showing the step edges is composed by two types of
segments (A and B). (b-c) Binding energy and atomic configurations of a B6N7 cluster attaching to the
two typical step edges of the Cu(111) surface.

From above discussions, it can be seen that although hBN islands have at least two most stable
alignment orientations on all low-index surfaces of a FCC TM foil, unidirectionally aligned hBN islands
can be realized on low symmetric high-index TM surfaces due to the strong binding between one of the
edges of the hBN island and the step edge of the substrate. Therefore, instead of low-index surfaces,
high-index surfaces might be more promising in synthesizing WSSC 2D materials, as will be revealed
in Chapter 4.

45
2.3 CVD growth of single-crystal monolayer TMDC films

2.3.1 Nucleation barrier of TMDC on a substrate

Similar to hBN, the nucleation of TMDCs is also dependent of the ambient condition because of
their binary compositions. Moreover, TMDCs have several different phases, e.g., 2H, 3R, 1T, 1T′,
1T′′,138-140 making their nucleation and growth more complicated. For example, 1T phase nucleation in
a 2H phase MoS2 film were observed by experiments,313,314 and DFT calculations also revealed the
phase transition of MoS2 between 1T and 2H.315-317

Wu et al. systematically explored the edge structures of MoSe2 and their formation energies by
both theoretical calculations and experimental techniques.318 As displayed in Figure 2.13, the most
stable edge structures and their formation energies are dependent of the chemical potential of Mo.
Besides, because the TM atoms at the edge can be well-passivated by chalcogen atoms, the edge
formation energies of TMDC materials are usually smaller than those of graphene and hBN, indicting
a high nucleation density of TMDC materials, and therefore controlling the nucleation density of
TMDCs during CVD is quite challenging.160,319,320

Figure 2.13 (a-d) STEM images of various MoSe2 edges observed in experiments and corresponding
atomic configurations.318 (e-f) Formation energy of various MoSe2 edges as a function of the chemical
potential of Mo, μMo.

46
2.3.1 Mechanisms of TMDC alignment on a substrate

Because of the difficulties in reducing nucleation density of TMDCs, the route of seamless
stitching of unidirectionally aligned TMDC islands seems to be a better choice in producing WSSC
TMDC films. As summarized in Section 1.3, TMDC islands generally align randomly on amorphous
surfaces, and present multi-orientations on high-symmetric crystalline substrates,160,163-173 which is
similar to hBN. However, the exploration on the alignment of various TMDC islands on substrates is
quite limited until now.

Figure 2.14 Alignment mechanism of TMDCs on the Al2O3(0001) surface. (a) Schematic of MoS2 on
the Al2O3(0001) surface.321 (b) Binding energy of MoS2 on the Al2O3(0001) surface as a function of the
orientation angle. (c) Stereo and top views of an aligned WS2 cluster on the Al2O3(0001) surface with
0ºrotation angle.322 (d) The comparison of aligned and misaligned WS2 clusters on the Al2O3(0001)
surface.

Ji et al. investigated the alignment of MoS2 on Al2O3(0001) surface by both theoretical


calculations and experimental observations.321 Figure 2.14(a-b) shows the binding energy profile of the
MoS2 layer and the Al2O3(0001) surface as a function of the rotation angle of MoS2 with respect to the
Al2O3(0001) surface. The two local minima at 0ºand 60ºof binding energy profile agree well with the
experimental observations showing that the ZZ direction of the MoS2 lattice prefers to align along the

47
high symmetric <21̅1̅0> direction of the Al2O3(0001) substrate, resulting in two antiparallel orientations
of MoS2 islands on the substrate. In Figure 2.14(c-d), a WS2 cluster structure was also used to compare
the orientation-dependent interaction between TMDC and the Al2O3(0001) surface, which also confirms
that the ZZ direction of TMDC parallel with the <21̅1̅0> direction of the Al2O3(0001) is energetically
more favorable.322

Very recently, motivated by the success of well-aligned hBN islands on high-index Cu surfaces,
unidirectionally aligned MoS2 islands was first realized on a vicinal Au(111) surface (see Figure 1.17),
and the alignment mechanism of MoS2 islands on the obtained vicinal Au(111) surface is presented in
Figure 2.15.179 Similar with the case of highly stepped Cu(111) surface, there are also two types of
Au<110> step edges on an Au(111) surface and the stabilities of various MoS2 edges attaching to these
two steps and on the Au(111) terrace are compared by DFT calculations. It was found that the MoZZ
edge docking to the B-type step edge is more stable than other configurations, which is responsible for
the only one dominant orientation of MoS2 islands. Nevertheless, a systematical understanding on the
mechanism of TMDC alignment is still absent on both low-index and high-index TM surfaces up to
now.

Figure 2.15 (a) Atomic configurations of two types of Au⟨110⟩ step edges, and illustration of three
typical directions of MoS2 with θ = 0°, 30° and 60° denoting MoZZ, AC, and SZZ edges, respectively.179
(b) Contact energies of different MoS2 edge docking on the surface and step edges.

In this section, we have explicitly demonstrated that:

(i) WSSC 2D materials can be realized from one nucleus formed on a wafer-scale
polycrystalline substrate because a 2D island can maintain its crystallinity when growing across grain
boundaries of the underlying substrate. By this route, WSSC graphene has been produced successfully,
though WSSC hBN and TMDCs has not realized by this route because of the great challenge in
suppressing their nucleation densities.

48
(ii) WSSC 2D materials can also be obtained by the seamless stitching of unidirectionally
aligned 2D islands. In this case, the selection of wafer-scale single crystalline substrates to ensure a
unidirectional alignment of 2D islands is essential. Currently, C6V Cu(111) and C2V Ge(110) surfaces
have been proved to be good substrate for the growth of unidirectionally aligned C6V graphene islands.
C3V 2D materials have been found to show at least two orientations on all low-index TM surfaces. On
the contrary, high-index TM surfaces show a great potential in templating the growth of unidirectionally
aligned low symmetric 2D materials due to the existence step edges on such substrates. By the seamless
stitching method, WSSC graphene, hBN and TMDCs films have all been successfully synthesized on
appropriate substrates.

2.4 Conclusions and unsolved research problems

In this Chapter, the current theoretical understanding on the synthesis of 2D materials by the
two routes, i.e., (i) single nucleus growth and (ii) seamless stitching of unidirectional 2D material islands,
are reviewed. Although great achievements have been made, there are still many scientific puzzles
regarding these two routes remain unsolved.

For the method of single nucleus growth, both theoretical and experimental studies confirmed
the continuous lattice structure of a 2D material without the formation of any line defects when the 2D
material grows across grain boundaries of an underlying TM substrate, which is contributed by the weak
interactions between the 2D material and its substrate. Therefore, polycrystalline substrates can be used
to synthesize WSSC 2D materials by this method. It should be noted that, to grow a WSSC 2D material
on a substrate by this method, a very low nucleation density, i.e., a high nucleation barrier, is necessary.
However, DFT studies showed that only the mono-elemental graphene can reach a low nucleation
density, while hBN and TMDCs cannot because of their binary composition and low edge formation
energies, which explains why only WSSC graphene has been realized by the method of single nucleus
growth up to now.

For the method of seamless stitching of unidirectionally aligned 2D islands to achieve WSSC
2D materials, a single crystalline substrate that can template the growth of unidirectionally aligned 2D
islands is required. Due to the high symmetries of the low-index substrate surfaces, 2D materials on a
low-index TM substrate generally show multiple orientations, corresponding to the multiple equivalent
local maxima in the binding energy profile between the 2D material and the substrate. In contrast, high-
index TM substrates have a low symmetry and are found to be more suitable for the growth of
unidirectionally aligned 2D islands. Despite it is evidenced that the orientation of a 2D island on a high-
index TM substrate is dominated by the interaction between the edge of the 2D island and the step edge
of the TM substrate, the theoretical understandings on the alignment of 2D islands on high-index TM
substrates are very limited, because the existed theoretical studies in this field are all based on the limited

49
experimental observations. In principle, the number of high-index surfaces is countless. An
experimental investigation on the alignment of 2D islands on all high-index TM substrates are
impossible. Therefore, a systematic theoretical study on the alignment mechanism of 2D materials on
various high-index TM substrates is urgent, which is very helpful for the screening of appropriate
substrates for growing unidirectionally aligned 2D islands and thus various WSSC 2D materials.

After growth, the corrugation behavior of a graphene layer on a TM substrate is also a study-
worthy phenomenon. Even though extensive experiments have observed that the corrugation of
graphene layer is highly dependent of the orientations of graphene layers and the types of the underlying
TM substrates, which is discussed in Section 1.4, the formation mechanisms of such orientation-
dependent corrugated graphene layers and the interactions between rotated graphene and TM substrates
have been barely revealed. Moreover, the effect of the orientation angle on the applications of the
corrugated graphene moirésuperstructures that may guide the experimental designs for templating the
synthesis of quantum dots or molecular frameworks has never been explored by theoretical methods.

50
Chapter 3 Methodology
3.1 Introduction

In this Chapter, the two methods that are mostly used in this dissertation, i.e., DFT and classical
MD simulations, are introduced. Both methods are very powerful and popular for dealing with many-
body systems in computational materials science. DFT is a quantum mechanical theory for calculating
the electronic structures of materials by approximating Schrödinger equations, while classical MD is
based on empirical or semi-empirical potentials to simulate the motions of atoms to study thermal or
mechanical properties of a system by solving Newton's equation.

3.2 Introduction of Density functional theory (DFT)

According to Heisenberg uncertainty principle,323 the position and velocity of a microscopic


particle cannot be exactly measured simultaneously, and therefore the postulated wavefunction (𝜓) is
introduced to describe the state that includes all possible information of an isolated system in quantum
mechanics, and the Schrödinger equation that describes the wavefunction of a quantum-mechanical
system is also proposed.324

The time-independent Schrödinger equation for the energy eigenfunctions and eigenvalues of
a molecular system is written as:

̂ 𝜓(𝑞𝑖 , 𝑞𝛼 ) = 𝐸𝜓(𝑞𝑖 , 𝑞𝛼 )
𝐻 (3.1)

̂ is the Hamiltonian operator of the system, 𝑞𝑖 and 𝑞𝛼 refer to the coordinate set of the electron and
𝐻
the atomic nuclei of the system. Without considering the spin-orbit and other relativistic interaction and
assuming the nuclei and electrons to be point masses, the Hamiltonian operator consists of five terms,
which are the kinetic energy of the nuclei, the kinetic energy of the electrons, the potential energy of
the repulsion between the nuclei, the potential energy of the attractions between the nuclei and the
electrons and the potential energy of the repulsion between the electrons:

̂ = 𝑇̂𝑁 + 𝑇̂𝑒 + 𝑉̂𝑁𝑁 + 𝑉̂𝑁𝑒 + 𝑉̂𝑒𝑒


𝐻

ℏ2 ℏ2 𝑍𝛼 𝑍𝛽 𝑒 2 𝑍𝛼 𝑒 2 𝑒2 (3.2)
=− ∑ ∇2𝛼 − ∑ ∇2𝑖 + 𝑘𝑒 ∑ ∑ − 𝑘𝑒 ∑ ∑ + 𝑘𝑒 ∑ ∑
2𝑚𝛼 2𝑚𝑒 𝑟𝛼𝛽 𝑟𝛼𝑖 𝑟𝑖𝑗
𝛼 𝑖 𝛼=1 𝛽>𝛼 𝛼 𝑖 𝑖=1 𝑗>𝑖

where α and β refer to the nuclei and i and j refer to the electrons, m represents the mass and Z represents
1
the atomic number of the corresponding nucleus, 𝑘𝑒 is the Coulomb’s constant which is equal to 4𝜋𝜀
0

51
and 𝜀0 is the electric constant, r denotes the corresponding distance between two particles. The
𝜕2 𝜕2 𝜕2
Laplacian operator ∇2 ≡ 𝜕𝑥 2 + 𝜕𝑦2 + 𝜕𝑧2 .

Since the electron-electron, nucleus-nucleus and electron-nucleus couplings makes the precise
solution of the Schrödinger equation of a many-body system to be impossible, appropriate
approximations were then proposed.

3.2.1 The Born-Oppenheimer approximation

Due to the huge difference in the masses of the atomic nuclei and the electrons, the motions of
the nuclei are much slower than that of the electrons, the Born-Oppenheimer approximation is then
proposed to separate the motions of atomic nuclei and electrons, and therefore the wave function of the
system can be decoupled as:

𝜓(𝑞𝑖 , 𝑞𝛼 ) = 𝜓𝑒𝑙 (𝑞𝑖 , 𝑞𝛼 )𝜓𝑁 (𝑞𝛼 ) (3.3)

To a good approximation, the nuclei of the system is considered to be fixed, therefore the kinetic
energy of the nuclei is omitted and the Schrödinger equation for the electrons of the system is:

̂𝑒𝑙 + 𝑉̂𝑁𝑁 )𝜓𝑒𝑙 (𝑞𝑖 , 𝑞𝛼 ) = 𝑈𝜓𝑒𝑙 (𝑞𝑖 , 𝑞𝛼 )


(𝐻 (3.4)

̂𝑒𝑙 = 𝑇̂𝑒 + 𝑉̂𝑁𝑒 + 𝑉̂𝑒𝑒 , and 𝑈 = 𝐸𝑒𝑙 + 𝑉̂𝑁𝑁 , where 𝐸𝑒𝑙 is the
in which the purely electronic Hamiltonian 𝐻
purely electronic energy and 𝑉̂𝑁𝑁 is the internuclear repulsion.

Due to the term of 𝑉̂𝑁𝑁 is only dependent of the nuclear configuration, the electronic
Schrödinger equation for a given nuclear configuration can be written as:

̂𝑒𝑙 𝜓𝑒𝑙 (𝑞𝑖 , 𝑞𝛼 ) = 𝐸𝑒𝑙 𝜓𝑒𝑙 (𝑞𝑖 , 𝑞𝛼 )


𝐻 (3.5)

and a set of electronic wave functions and their corresponding electronic energies can be got for each
of assumed nuclei configurations.

Since the electronic energy depends parametrically on the coordinates of the atomic nuclei, i.e.,
when the nuclei change their positions slightly, the electrons will adjust the coordination immediately,
as well as the electronic wave functions and their corresponding electronic energies. As the nuclei move
gradually, the electronic energy changes smoothly as a function of the nuclear configuration, 𝑈(𝑞𝛼 ),
which is in effect the potential energy for the nuclear motion.

Therefore, the Schrödinger equation for the nuclei of the system is:

̂𝑁 𝜓𝑁 (𝑞𝛼 ) = 𝐸𝜓𝑁 (𝑞𝛼 )


𝐻 (3.6)

52
̂𝑁 = 𝑇̂𝑁 + 𝑈(𝑞𝛼 ), and since both the nuclear energy and the
and the Hamiltonian for the nuclei 𝐻
electronic energy are included, the energy eigenvalue E in Eq. (3.6) is in fact the total energy of the
system.

3.2.2 The Hohenberg-Kohn theorem

The wave functions of a system with n electrons has 3n spatial coordinates, the Schrödinger
equations for electrons of this system is a 3n dimensional problem and solving the Schrödinger equation
for real materials is in practice a hassle, which promotes the search for functions with fewer variables
than the wave functions. In 1964, Pierre Hohenberg and Walter Kohn proved that the total energy of a
molecule and its other ground state properties are determined by the ground state electron density,
𝑛0 (𝑥, 𝑦, 𝑧).325

As introduced above, the pure electronic Hamiltonian consists of the kinetic energy of the
electrons, the potential energy of the attractions between the nuclei and the electrons and the potential
energy of the repulsion between the electrons, in DFT, the potential energy of the attraction between
the nuclei and the ith electron is called the external potential acting on the ith electron, denoted by
𝑣𝑒𝑥𝑡 (𝑟⃑𝑖 ), because it is induced by external charges applied to the electron.

The first Hohenberg-Kohn theorem states that the external potential and hence the ground state
electronic energy is a unique functional of the electron density, 𝐸0 = 𝐸0 [𝑛0 (𝑟⃑)]. To prove that 𝑛0 (𝑟⃑)
determines the external potential, we assume there are two systems of electrons trapped in two external
potentials, 𝑣𝑎 and 𝑣𝑏 , having the same ground state electron density, 𝑛0 (𝑟⃑), and the differences between
the two external potentials are more than a constant. The Schrödinger equations for the two systems are:

̂𝑎 𝜓0,𝑎 = (𝑇̂ + 𝑈
𝐻 ̂ + 𝑣𝑎 )𝜓0,𝑎 = 𝐸0,𝑎 𝜓0,𝑎 (3.7)

̂𝑏 𝜓0,𝑏 = (𝑇̂ + 𝑈
𝐻 ̂ + 𝑣𝑏 )𝜓0,𝑏 = 𝐸0,𝑏 𝜓0,𝑏 (3.8)

̂𝑎 and 𝐻
where 𝐻 ̂𝑏 are the electronic Hamiltonians corresponding to the two external potentials, 𝜓0,𝑎 and

𝜓0,𝑏 and 𝐸0,𝑎 and 𝐸0,𝑏 are the normalized ground state wavefunctions and energies for the electronic
Hamiltonians of the two systems. 𝑇̂ and 𝑈
̂ denote the kinetic operator and operator of the repulsion
between the electrons, respectively.

̂𝑎 − 𝐻
Due to the two systems have the same ground state, 𝜓0,𝑎 = 𝜓0,𝑏 = 𝜓0 , we get (𝐻 ̂𝑏 )𝜓0 =
(𝑣𝑎 − 𝑣𝑏 )𝜓0 , implying the difference between the electronic Hamiltonians is only in the external
̂ |𝜙⟩ > 𝐸0 if the ground state is nondegenerate,
potentials. According to the variation theorem that ⟨𝜙|𝐻
here 𝜓0,𝑏 is regarded as a trial function and we get:

53
̂𝑎 |𝜓0,𝑏 ⟩ = ⟨𝜓0,𝑏 |𝐻
⟨𝜓0,𝑏 |𝐻 ̂𝑎 − 𝐻
̂𝑏 |𝜓0,𝑏 ⟩ + ⟨𝜓0,𝑏 |𝐻
̂𝑏 |𝜓0,𝑏 ⟩
(3.9)
= ∫ 𝑛0 (𝑟⃑)[𝑣𝑎 (𝑟⃑) − 𝑣𝑏 (𝑟⃑)] 𝑑𝑟⃑ + 𝐸0,𝑏 > 𝐸0,𝑎

Exchanging the subscripts a and b, we then get:

∫ 𝑛0 (𝑟⃑)[𝑣𝑏 (𝑟⃑) − 𝑣𝑎 (𝑟⃑)] 𝑑𝑟⃑ + 𝐸0,𝑎 > 𝐸0,𝑏 (3.10)

By adding the Eq.(3.9) and (3.10), we finial get:


𝐸0,𝑏 + 𝐸0,𝑎 > 𝐸0,𝑎 + 𝐸0,𝑏 (3.11)

which is definitely false, indicating that the initial assumption that the two systems with different
external potentials have the same ground state electron density is not true. Therefore, the ground state
electron density uniquely determines the external potential and also the electronic Hamiltonian, which
means that the ground state electronic energy is a unique functional of the electron density and is
dependent of the external potential.

The second Hohenberg-Kohn theorem states that the electron density that minimize the energy
of the overall functional is the true ground state electron density, and the proof is as follows.

̂𝑒𝑙 = 𝑇̂𝑒 + 𝑉̂𝑒𝑒 + 𝑉̂𝑁𝑒 for


Taking the average of equation of the purely electronic Hamiltonian 𝐻
the ground state, we get:

𝐸0 = 𝑇̅[𝑛0 ] + 𝑉̅𝑒𝑒 [𝑛0 ] + 𝑉̅𝑁𝑒 [𝑛0 ] = 𝐹[𝑛0 ] + ∫ 𝑛0 𝑣(𝑟⃑) 𝑑𝑟⃑ (3.12)

where the functional 𝐹[𝑛0 ] ≡ 𝑇̅[𝑛0 ] + 𝑉̅𝑒𝑒 [𝑛0 ] and is universal and independent of the external
potential.

Using the wavefunction 𝜓𝑡𝑟 corresponding to the electron density 𝑛𝑡𝑟 as a trial variation
functional for the purely electronic Hamiltonian, from variation theorem we then get:

̂ |𝜓𝑡𝑟 ⟩ = 𝐹[𝑛𝑡𝑟 ] + ∫ 𝑛𝑡𝑟 𝑣(𝑟⃑) 𝑑𝑟⃑ > 𝐸0


𝐸𝑡𝑟 = ⟨𝜓𝑡𝑟 |𝐻 (3.13)

implying that there is no trial electron density can give a lower ground state energy that the true ground
state electron density, 𝑛0 .

3.2.3 The Kohn-Sham method

Despite that the Hohenberg-Kohn theorem can reduce the problem of a many-body system with
n electrons from a 3n dimensional problem to a 3 dimensional one by using the ground state electron
density instead of the wavefunctions to calculate the ground state properties of the system, it neither

54
provide a way to find the ground state electron density nor tell how to practically calculate the ground
state energy from the ground state electron density because of the existence of the unknown functional
𝐹[𝑛0 ]. Until 1965, the Kohn-Sham method was proposed and made DFT to be a practical tool for
calculations.326

Kohn and Sham considered a fictitious system consisting of n noninteracting electrons, and the
ground state electron density of such a noninteracting system, 𝑛(𝑟⃑), is equal to that of the real system,
𝑛(𝑟⃑) = 𝑛0 (𝑟⃑). According to the first Hohenberg-Kohn theorem, the external potential, 𝑣𝑠 (𝑟⃑), can be
uniquely determined since 𝑛𝑠 (𝑟⃑) is defined. In the noninteracting system, the electrons have no
interaction with one another and therefore the electronic Hamiltonian can be written as:

ℏ2 2
̂𝑠 = ∑ ( −
𝐻 ∇ + 𝑣𝑠 (𝑟⃑𝑖 )) ≡ ∑ ℎ̂𝑖𝐾𝑆 (3.14)
2𝑚𝑒 𝑖
𝑖 𝑖

here, ℎ̂𝑖𝐾𝑆 is the one-electron Kohn-Sham Hamiltonian. Because the electrons of the fictitious system
are independent, the ground state wavefunction of the fictitious system can be obtained by a slater
determinant of one-electron orbitals:

𝜑1 (𝑟⃑1 ) … 𝜑1 (𝑟⃑𝑛 )
1
𝜓0 (𝑟⃑) = | ⋮ ⋱ ⋮ | (3.15)
√𝑛! 𝜑 (𝑟⃑ ) … 𝜑 (𝑟⃑ )
𝑛 1 𝑛 𝑛

and then the expression of the electron density can be simplified to:

𝑛(𝑟⃑) = ∑|𝜑𝑖 (𝑟⃑)|2 (3.16)


𝑖

where 𝜑𝑖 (𝑟⃑) is the eigenfunctions of the ith electron of the fictitious system.

Kohn and Sham rewrote the equation of the ground state electronic energy as:

𝐸0 = 𝑇𝑠 [𝑛(𝑟⃑) ] + 𝑈𝐻 [𝑛(𝑟⃑) ] + ∫ 𝑣𝑒𝑥𝑡 (𝑟⃑) 𝑛(𝑟⃑) 𝑑𝑟⃑ + 𝐸𝑥𝑐 [𝑛(𝑟⃑)] (3.17)

where 𝑇𝑠 is the Kohn-Sham kinetic energy, 𝑈𝐻 is the classical electrostatic interelectronic repulsion
energy (Hartree energy) and 𝐸𝑥𝑐 is exchange-correlation energy.

According to the second Hohenberg-Kohn theorem that the ground state energy can be found
by varying the electron density, equivalently, the ground state energy can also be found by varying the
Kohn-Sham orbitals which determine the electron density. Therefore, the Kohn-Sham kinetic energy
can be calculated by expressing in terms of the Kohn-Sham orbitals 𝜓𝑖 (𝑟⃑):

55
ℏ2
𝑇𝑠 [𝑛(𝑟⃑)] = − ∑⟨𝜓𝑖 (𝑟⃑)| ∇2𝑖 |𝜓𝑖 (𝑟⃑)⟩ (3.18)
2𝑚
𝑖

The Hartree energy can be obtained by:

1 𝑛(𝑟⃑)𝑛(𝑟⃑ ′ )
𝑈𝐻 [𝑛(𝑟⃑)] = ∬ 𝑑𝑟⃑𝑑𝑟⃑ ′ (3.19)
2 |𝑟⃑ − 𝑟⃑ ′ |

The exchange-correlation energy is contributed by two parts, which are the difference of the
electronic kinetic energy between the fictitious noninteracting system and the real interacting system
and the differences between the Hartree energy and the real interelectronic repulsion energy:

𝐸𝑥𝑐 [𝑛(𝑟⃑)] = Δ𝑇[𝑛(𝑟⃑)] + Δ𝑉𝑒𝑒 [𝑛(𝑟⃑)]


(3.20)
= 𝑇[𝑛(𝑟⃑)] − 𝑇𝑠 [𝑛(𝑟⃑)] + 𝑉𝑒𝑒 [𝑛(𝑟⃑)] − 𝑈𝐻 [𝑛(𝑟⃑)]

By substituting the terms in Eq.(3.17) with the corresponding expressions, the equation of the
ground state electronic energy becomes:

ℏ2 1 𝑛(𝑟⃑)𝑛(𝑟⃑ ′ )
𝐸0 = − ∑⟨𝜓𝑖 (𝑟⃑)| ∇2𝑖 |𝜓𝑖 (𝑟⃑)⟩ + ∬ 𝑑𝑟⃑𝑑𝑟⃑ ′ + ∫ 𝑣𝑒𝑥𝑡 (𝑟⃑) 𝑛(𝑟⃑) 𝑑𝑟⃑
2𝑚 2 |𝑟⃑ − 𝑟⃑ ′ | (3.21)
𝑖

+ 𝐸𝑥𝑐 [𝑛(𝑟⃑)]

and the Kohn-Sham orbitals that minimize the ground state electronic energy satisfy the Kohn-Sham
equations:

ℏ2 2
(− ∇ + 𝑣𝑒𝑓𝑓 (𝑟⃑) ) 𝜑𝑖 (𝑟⃑) = 𝜀𝑖 𝜑𝑖 (𝑟⃑) (3.22)
2𝑚 𝑖

and the expression of the effective potential 𝑣𝑒𝑓𝑓 (𝑟⃑) is:

𝑛(𝑟⃑ ′ ) ′
𝛿𝐸𝑥𝑐 [𝑛(𝑟⃑)]
𝑣𝑒𝑓𝑓 (𝑟⃑) = ∫ 𝑑𝑟
⃑ + 𝑣𝑒𝑥𝑡 (𝑟
⃑) + (3.23)
|𝑟⃑ − 𝑟⃑ ′ | 𝛿𝑛(𝑟⃑)

𝛿𝐸𝑥𝑐 [𝑛(𝑟⃑)]
where 𝛿𝑛(𝑟⃑)
is the exchange-correlation potential. Clearly, neither the exchange-correlation energy

nor the exchange-correlation potential is known, therefore, a good approximation to the exchange-
correlation term is the key to calculate the many-body system by the Kohn-Sham method, which means
that the DFT calculation can give the exact solution for the may-body problems since the exchange-
correlation functional is known.

56
3.2.4 Approximation to the exchange-correlation functional

In this section, various approximations to the exchange-correlation functional will be discussed.


Initially, the local density approximation (LDA) is proposed based on a jellium system constituted by a
homogeneous electron gas. In this system, the electron density varies extremely slowly with position,
and therefore the exchange-correlation functional is given by:

𝐿𝐷𝐴 [𝑛(𝑟
𝐸𝑥𝑐 ⃑)] = ∫ 𝑛(𝑟⃑)𝜀𝑥𝑐 (𝑟⃑) 𝑑𝑟⃑ (3.24)

Since the exchange-correlation potential is also a function of the electron density, the Kohn-Sham
equation can be solved by a self-consistent-field calculation, and the Kohn-Sham scheme is shown in
Figure 3.1.

Figure 3.1 The self-consistency loop of the Kohn-Sham equations.

The main disadvantage of the LDA is that it cannot distinguish the spatial orbitals of electrons
with opposite spins paired with each other, therefore, the local spin-density approximation (LSDA) is
then proposed:

𝐿𝑆𝐷𝐴 𝛼 (𝑟
𝐸𝑥𝑐 [𝑛 ⃑), 𝑛𝛽 (𝑟⃑)] = ∫ 𝑓(𝑛𝛼 (𝑟⃑), 𝑛𝛽 (𝑟⃑)) 𝑑𝑟⃑ (3.25)

57
where 𝛼 and 𝛽 denote spin up and spin down. LSDA gives better results for open-shell molecules and
magnetic systems.327,328 However, both LDA and LSDA are based on the model of a homogeneous
electron gas and appropriate for systems where the electron density is a slowly varying function of
position. For inhomogeneous systems, functionals that go beyond the LDA and LSDA are needed, one
popular way to do this is by inducing gradients of the electron density in the integrand, which is called
the generalized-gradient approximation (GGA), and the exchange-correlation functional then becomes:

𝐺𝐺𝐴 𝛼 (𝑟
𝐸𝑥𝑐 [𝑛 ⃑), 𝑛𝛽 (𝑟⃑)] = ∫ 𝑓(𝑛𝛼 (𝑟⃑), 𝑛𝛽 (𝑟⃑), ∇𝑛𝛼 (𝑟⃑), ∇𝑛𝛽 (𝑟⃑)) 𝑑𝑟⃑ (3.26)

Compared with the local density functionals, gradient-corrected functionals is semi-local because they
involved the values of the electron density not only at the point 𝑟⃑ but also at in an infinitesimal
neighborhood of 𝑟⃑.

3.3 Introduction to classical molecular dynamics (MD) simulation

MD is based on laws of thermal dynamics, statistics and Newtonian mechanics to predict


various properties of N-body systems by simulating the motion of atoms, which is broadly applied in
materials science and chemical physics nowadays. Since the classical MD calculation was successfully
used to simulate the phase transition process for a hard sphere system in 1957,329 it is acknowledged as
one of the most powerful methods to study the structures of materials at the atomic level.

Figure 3.2 shows the schematics of a classical MD simulation, and there are basically four
steps: (1) Build an initial configuration and set preliminary conditions, e.g., atomic positions, atomic
velocities, interactions between atoms, boundary conditions, and set temperature and pressure
conditions;

(2) Optimize the initial configuration to avoid possibly unreasonable structures, in general, a relative
stable configuration with a low total energy is the basis for executing a MD simulation;

(3) Perform the time integration of MD calculations under the pre-defined constraints, including
calculating the forces exerted on each atom by differentiating the potential energy surface, and then
solving the motion equations of the system to obtain the atomic positions and velocities at the next
moment;

(4) Analyze and extract relevant physical quantities of the system.

58
Figure 3.2 The schematic of the classical MD simulations.

3.3.1 Interactions between atoms

During MD simulations, choosing an appropriate potential to describe the interactions between


atoms is critical for the accuracy of the simulated results. In fact, it is impossible to determine a potential
function that universally describes different systems because the interactions between different atoms
are quite different, and therefore, various empirical or semi-empirical potentials have been proposed.
Initially, pair potentials that treat the atomic interactions as only a function of the relative position
between two atoms were popular, however, in the actual multi-atom system, the existence of one atom
must affect the interactions of other atoms. Therefore, potential energy surfaces between atoms have
been gradually developed from the pair ones to many-body ones, and currently various many-body
potentials have been proposed to describe the interactions between atoms for various system. In this
thesis, MD simulation was performed in studying the formation mechanisms of the Cu(111) foil and
embedded atom model (EAM) potential was adopted to describe Cu-Cu interactions.330

The EAM potential is widely used in metal systems, the potential energy of the ith atom is:

1
𝐸𝑖 = 𝐹𝛼 (∑ 𝜌𝛽 (𝑟𝑖𝑗 )) + ∑ 𝜙𝛼𝛽 (𝑟𝑖𝑗 ) (3.27)
2
𝑗≠𝑖 𝑗≠𝑖

59
where 𝑟𝑖𝑗 is the distance between the ith and jth atoms, 𝜌𝛽 represents the contribution to the electron
charge density from the jth atom of type 𝛽 at the location of the ith atom, the embedding function 𝐹𝛼
represents the energy required to place the ith atom of type 𝛼 into the electron cloud, and 𝜙𝛼𝛽 is a pair
potential function. Clearly, EAM potential is a many-body potential because the electron charge density
is a summation over many atoms. A cutoff radius is needed in this potential, beyond which the
interactions between atoms can be ignored.

3.3.2 Boundary conditions

Since the macroscopic properties of materials are determined by a huge number of particles, to
accurately simulate the behaviors of a real material, a very large system is required. However, due to
the limitations of computer memory and computing power, MD simulations can only deal with a limited
number of atoms, and hence an appropriate boundary condition is necessary to mimic the bulk materials.
At present, there are mainly two boundary conditions, which are periodic boundary conditions and non-
periodic boundary conditions.

Periodic boundary condition is very useful for simulating bulk materials via a system with a
limited number of atoms in MD simulations. A simulation box containing dozens to tens of thousands
of atoms is built as a simulation unit cell, the periodic boundary condition then creates an infinite
number of identical boxes around this unit cell, and in this case, once there are atoms moving out of the
simulation box from one side, there must be a corresponding number of atoms entering the box from
the opposite side. Periodic boundary condition is broadly used to simulate the macroscopic properties
of real materials by using a limited number of atoms, greatly saving the calculation time and improving
the calculation efficiency. However, the periodic boundary condition is not always applicable to every
direction of the system, especially the low-dimensional materials. For instance, nonperiodic boundary
condition is usually required in investigating zero-dimensional nanoparticles, one-dimensional
nanowires or 2D materials. In fact, the condition of combining periodic and nonperiodic boundaries is
commonly adopted in the simulated system depending on the specific requirements of the studied
materials.

3.3.3 Ensemble

Ensemble is a statistical concept, which is a collection of systems with the same state. Ensemble
controls thermodynamic quantities of the studied system, such as atomic number (N), pressure (P),
volume (V), temperature (T), energy (E) and enthalpy (H), and it can be classified into microcanonical
ensemble (NVE), canonical ensemble (NVT), grand canonical ensemble (μVT), Isothermal–isobaric
ensemble (NPT) and Isoenthalpic–isobaric ensemble (NPH), etc.

60
NVE ensemble is the one with constant N, V and E, which corresponds to an adiabatic process
and has no heat and mass exchange with the outside environment. In NVT ensemble, N, V and T of the
system are conserved, which can be regarded as an isolated system placed in a thermostat and is
generally suitable for the MD simulation at a constant temperature. To ensure a constant temperature,
thermal exchange with the environment is allowed and a variety of thermostat algorithms have been
developed, such as Nosé-Hoover thermostat, Berendsen thermostat, etc. For μVT ensemble, μ, V and T
of the system are kept constant, and is usually used for the system that can exchange both energy and
mass with the environment. NPT ensemble keeps N, P and T of the system unchanged, and therefore,
besides a thermostat, a barostat is also needed to keep a constant pressure. NPH ensemble requires N,
P and H of the systems to be constant, which is difficult to realize and therefore is not common in
practical MD simulations.

3.3.4 Time integration algorithm

MD simulations require time integration of the motion equation. During simulation, the motion
equation is discretized into finite difference equations by the finite difference algorithm, and common
finite difference algorithms include Verlet and Gear predictor-corrector methods. The Verlet integration
is the mostly frequently adopted one in MD simulations, which is based on computing the Taylor
expansions to the third order for the atomic position at different time directions:331

1 1
𝑟⃑(𝑡 − Δ𝑡) = 𝑟⃑(𝑡) − 𝑣⃑(𝑡)Δ𝑡 + 𝑎⃑(𝑡)Δ𝑡 2 − 𝑏⃑⃑(𝑡)Δ𝑡 3 + 𝒪(Δ𝑡 4 ) (3.28)
2 6

1 1
𝑟⃑(𝑡 + Δ𝑡) = 𝑟⃑(𝑡) + 𝑣⃑(𝑡)Δ𝑡 + 𝑎⃑(𝑡)Δ𝑡 2 + 𝑏⃑⃑(𝑡)Δ𝑡 3 + 𝒪(Δ𝑡 4 ) (3.29)
2 6

where 𝑣⃑(𝑡) = 𝑟⃑ ′ (𝑡) is the velocity, 𝑎⃑(𝑡) = 𝑟⃑ ′′ (𝑡) is the acceleration, and 𝑏⃑⃑(𝑡) = 𝑟⃑ ′′′ (𝑡) is the third
derivative of the position regarding to the time.

Adding the two expansions, we get:

𝑟⃑(𝑡 + Δ𝑡) = 2𝑟⃑(𝑡) − 𝑟⃑(𝑡 − Δ𝑡) + 𝑎⃑(𝑡)Δ𝑡 2 + 𝒪(Δ𝑡 4 ) (3.30)

It can be seen that the Verlet integration only gives the atomic positions but not the velocities, and so
some improved algorithms were proposed, among them, the Velocity-Verlet algorithm is more
commonly used and the expression of position and velocity are as follows:332

1
𝑟⃑(𝑡 + Δ𝑡) = 𝑟⃑(𝑡) + 𝑣⃑(𝑡)Δ𝑡 + 𝑎⃑(𝑡)Δ𝑡 2 (3.31)
2

1
𝑣⃑(𝑡 + 𝛥𝑡) = 𝑣⃑(𝑡) + (𝑎⃑(𝑡) + 𝑎⃑(𝑡 + 𝛥𝑡))𝛥𝑡 (3.32)
2

61
It should be noted that the choice of time step is very important for MD simulations, which
significantly affects the efficiency and accuracy of the calculations. A smaller step can reduce the
calculation error and obtain more accurate results, but it also increases the simulation time and reduces
the calculation efficiency. A safe time step is usually set to be about 1/10 of the vibration period of the
atoms in the simulated system.

62
Chapter 4 A general theory of 2D materials alignment on
crystalline substrates with different symmetries
4.1 Introduction

The alignment of a 2D material grown on a crystalline substrate is highly dependent of the


symmetries of the 2D material and the substrate. At the early stage, high symmetric low-index TM
surfaces are widely used as substrates in the CVD growth of 2D materials, because their low surface
energies make them to be easily synthesized. Extensive experimental observations revealed that 2D
material islands usually show multi orientations on high-symmetric TM surfaces. It is not difficult to
image that if the number of symmetrical axes of a substrate is more than that of a 2D material, then the
symmetrical operation performed on the substrate with respect to its symmetrical axes will bring new
orientations of the grown 2D island on the substrate, therefore, an essential condition of achieving
unidirectionally aligned 2D islands on a crystalline substrate is that the symmetry group of the substrate
should be a subgroup of that of the 2D material, and thus high-index low symmetric surfaces probably
are suitable substrates to realize the growth of unidirectionally aligned 2D islands and further achieve
WSSC 2D materials.

It should be noticed that the most stable orientation of a 2D material islands on the substrate is
determined by the 2D material – substrate interactions. Within one periodic range of a 2D material
rotating on a substrate, 2D material will show only one preferential orientation if there is only one local
minimum in the binding energy profile for the 2D material adsorbed on the substrate.

In this Chapter, FCC TM surfaces are chosen as the substrates to reveal the alignment of 2D
materials. There are three types of high symmetric low-index FCC TM surfaces, namely FCC(111),
FCC(100) and FCC(110) surfaces. Except these three high symmetric ones, all others are low symmetric
surfaces with Cs or C1 symmetry. In the following, the symmetries of various 2D materials and various
FCC surfaces will be introduced first, and then the alignment of 2D materials on these surfaces are
revealed. We propose that, if a high-symmetric direction of the 2D materials prefers to align along a
high symmetric direction of the substrate and the symmetric group of the substrate is a subgroup of that
of the 2D materials, there is only one most stable alignment orientation of the 2D materials grown on
the substrate.333

4.2 The symmetries of 2D materials and substrate surfaces

The symmetries of 2D materials can be denoted by Cn and Cnv (The symmetric operations with
respect to the in-plane that perpendicular to the principal rotation axis are not considered here, because

63
there are no in-plane symmetric operations for the system of a 2D material on a substrate. As will be
proved later, this simplification does not affect our conclusions). Cn represents that a 2D material will
coincide with its original structure when it is rotated with respect to its principal axis by 2π/n (n = 1, 2,
3, 4 and 6), which is to say that this principal axis is a n-fold rotation axis of the 2D material. If there
are n mirror planes passing through the n-fold rotation axis of the 2D material, the symmetry of the 2D
material will increase to be Cnv.

Figure 4.1 illustrates the symmetries of three types of 2D materials, including graphene, hBN
and MoS2 and three high symmetric low-index FCC surfaces, all the three 2D materials and the three
low-index FCC surfaces have n-fold symmetries with n > 1. Strictly speaking, FCC(111) substrate is
C3v symmetric because of its ABC-stacking structure. However, because the interaction between the 2D
material and the substrate is mainly contributed by the top atomic layer of the FCC(111)
substrate,188,334,335 the FCC(111) surface can be considered to be a C6v symmetric substrate. Figure 4.1
shows the principal axes and mirror planes of the three 2D materials and the three low-index FCC
surfaces, as marked by the red circles and dash lines, respectively. Compared with graphene, hBN and
TMDCs show mirror symmetries only with respect to AC directions due to their binary compositions.
It should be noticed that although the TMDC materials has a D3h symmetry because of its three-atom
layered structure is symmetric respect to the central atomic layer, its symmetric operations are as same
as the C3V material in the 2D plane direction.

Figure 4.1 Illustrations of the symmetries of various 2D materials and low-index high symmetric FCC
surfaces. (a) Graphene, (b) hBN, (c) MoS2, (d) FCC{111}, (e) FCC{100} and (f) FCC{110}. The
principal axes perpendicular to the 2D surfaces are denoted by the red circles, and the mirror planes
passing through the principal axes are denoted by dashed lines.

64
From the view of atomic configurations, the high-index low symmetric FCC TM surfaces are
constructed by large numbers of high-symmetric terraces and steps, as demonstrated in Figure 4.2(a).
According to the types of the high-symmetric terraces, the low symmetric surfaces can be classified
into three categories: FCC{111}-based, FCC{100}-based and FCC{110}-based surfaces. Because the
high-index low symmetric surfaces are cut from an FCC TM along an arbitrary orientation that is
deviated from the three low-index surfaces, in principle the number of high-index low symmetric
surfaces is infinite. All the high-index low symmetric surfaces only have 1-fold rotational symmetry,
while depending on the step edge structures of the surfaces, there are six types of CS symmetric surfaces
with step edges along specific directions, which are presented in Figure 4.2(b). Every type of the low
symmetric surfaces has two kinds of CS ones, i.e., FCC{n n n+m} with <110> step edges and FCC{n+m
n-m n} with <211> step edges for FCC{111}-based surfaces; FCC{n m m} with <110> step edges and
FCC{n m 0} with <100> step edges for FCC{100}-based surfaces; and FCC{n+m n+m m} with <110>
step edges and FCC{n n+m 0} with <100> step edges for FCC{100}-based surfaces. All the six CS
symmetric surfaces have straight step edges, and the mirror symmetric plane of these surface is
perpendicular to the step edge direction.

Figure 4.2 (a) Illustration of a high-index low symmetric surface. (b) Atomic models for the two types
of CS FCC{111}-based surfaces, whose surface indices are FCC{n n n+m} and FCC{n+m n-m n},
respectively. (c) Atomic models for the two types of CS FCC{100}-based surfaces, whose surface
indices are FCC{n m m} and FCC{n m 0}. (d) Atomic models for the two types of CS FCC{110}-based
surfaces, whose surface indices are FCC{n+m n+m m} and FCC{n n+m 0}. Dashed lines denote the
mirror symmetric plane of CS surfaces. For all the six types of CS symmetric FCC surfaces, n and m are
integers and n>>m.

65
The deduction of these six types of CS-surfaces is as follows. Because a low symmetric surface
can be considered to be composed of large terrace surfaces and small step surfaces, and the intersection
lines of the terrace surfaces and the step surfaces determine the direction of the step edges. For example,
a CS surface with <110> step edges can be constructed by two surfaces which have intersecting lines
along <110> direction. Table 4.1 lists the compositions of the six types of CS surfaces and their surface
indexes.

Table 4.1 The compositions and surface indexes of the CS symmetric FCC surfaces.

Step edge direction of


Surfaces Compositions of the surface Surface indexes
the surface
FCC{111}- FCC<110> n×FCC(111) + m×FCC(001) FCC{n n n+m}
based FCC<211> n×FCC(111) + m×FCC(11̅0) FCC{n+m n-m n}
FCC{100}- FCC<110> n×FCC(001) + m×FCC(111) FCC{n m m}
based FCC<100> n×FCC(001) + m×FCC(010) FCC{n m 0}
FCC{110}- FCC<110> n×FCC(110) + m×FCC(111) FCC{n+m n+m m}
based FCC<100> n×FCC(110) + m×FCC(010) FCC{n n+m 0}
Notes: n and m are integers, and n>>m.

Because of the anisotropy of the C1 symmetric surfaces, the islands of a 2D material growing
on a C1 symmetric surface will show only one most preferential orientation, and detailed studies of
graphene and hBN islands on various C1 symmetric surfaces will be introduced in Chapters 5 and 7,
respectively. CS symmetric and low-index surfaces have mirror symmetric planes and rotational
symmetry axes, and in this Chapter, the alignment of 2D material islands on the low-index high
symmetric FCC surfaces and the six types of CS symmetric FCC surfaces are explored from the aspect
of rotation and mirror symmetries.

4.3 Alignment of 2D materials on low-index high symmetric substrates

From the aspect of the rotation symmetry, the total number of orientations of a 2D material on
a substrate, N, can be estimated by the solutions of

2𝜋 2𝜋 2𝜋 𝑛 (4.1)
(𝑖 × 𝑛 − 𝑗 × 𝑛 )⁄(𝑛 ) = 𝑖 × 𝑛 2𝐷 − 𝑗
𝑆𝑢𝑏 2𝐷 2𝐷 𝑆𝑢𝑏

where i and j represent integers, 𝑛𝑆𝑢𝑏 and 𝑛2𝐷 are the rotational fold numbers of the substrate and the
2𝜋 2𝜋
2D material rotating in the range of 2π, respectively. 𝑛 and 𝑛 then represent the smallest rotational
𝑆𝑢𝑏 2𝐷

periodicity of the substrate and the 2D material rotating back to their original orientation, therefore the

66
2𝜋 2𝜋 2𝜋
orientation of the 2D material will not be changed if 𝑖 × 𝑛 −𝑗×𝑛 is zero or integer times of 𝑛 ,
𝑆𝑢𝑏 2𝐷 2𝐷

Otherwise new equivalent orientations of the 2D material will appear, and the number of equivalent
orientations depends on the number of the different solutions of formula (4.1). Using a C3-2D material
on a C4-substrate as an example, the formula (4.1) becomes 3⁄4 𝑖 − 𝑗, and depending on i and j, there
are totally four solutions: 1/4, 1/2, 3/4 and 1, therefore, four equivalent orientations of the C3 2D material
can be obtained on the C4 substrate. Table 4.2 lists the number of alignment orientations of different 2D
materials on various high symmetric substrates based on the rotation symmetric relationship between
2D materials and their substrates.

Table 4.2 The alignment of 2D materials on different substrate based on rotation symmetry.

Rotation Symmetries
Solutions Orientations
2D materials Substrate
C6 1 1
C C ½, 1 2
6 4
C2 1 1
C6 ½, 1 2
C3 C4 ¼, ½, ¾, 1 4
C2 ½, 1 2

From the aspect of mirror symmetry, the alignment of a 2D material on a substrate highly
depends on the geometrical feature of the system of the 2D material adsorbed on the substrate in their
most stable configuration. If the mirror symmetry planes of the substrate coincide with those of the 2D
materials, the orientation of the 2D material will not be changed when the substrate is reflected
regarding to any of its mirror symmetry planes, otherwise, new orientations of the 2D material will
appear. Therefore, it is necessary to explore the most preferential orientation of a 2D material on a
substrate. In the following parts, the most stable configuration of the three 2D materials on different
low-index high symmetric substrates are first explored by DFT calculations, and based on the most
stable orientations, the alignment of 2D materials on substrates are also studied.

4.3.1 Alignment of C6V-2D materials on high-symmetric substrates

Graphene is a representative C6V 2D material. The orientation of graphene on a low-index high


symmetric Cu substrate is determined by the interactions between graphene and the substrate, which
includes graphene edge - Cu substrate interaction and graphene bulk - Cu substrate interaction. DFT
calculations have revealed that the orientation of graphene on a Cu substrates is inherited from its
nucleation stage, during which the strong graphene edge-Cu substrate interaction dominates the

67
graphene-substrate interaction, and the high rotation barrier makes the rotation of large graphene
domains impossible.265

Here, to figure out the most stable orientation of graphene on the three low-index high
symmetric Cu substrates, the binding energies of graphene ZZ edges, which are the slowest propagating
edges in graphene and thus are the edges of a graphene island during its growth, attaching to the three
low-index Cu substrates with different alignment orientations were compared, as shown in Figure 4.3.
The definition of the edge binding energy is:

𝐸𝐵 = (𝐸2𝐷 + 𝐸𝑆𝑢𝑏 − 𝐸𝑇 )⁄𝐿𝐸𝑑𝑔𝑒 , (4.2)


where E2D and Esub are energies of the 2D material nanoribbon and the substrate, respectively. ET is the
total energy of the 2D nanoribbon attaching to the substrate. LEdge is the length of the 2D nanoribbon. It
is noted that one of the two edges of the nanoribbon is passivated by hydrogen and thus has negligible
interaction with the substrate in these models.

Figure 4.3 Binding energy of graphene attaching to three high symmetric Cu surfaces, which are
Cu(111) (a), Cu(100) (b) and Cu(110) (c) surfaces.

Due to the small lattice mismatch (~3.7%) between graphene ZZ direction and Cu<110>
direction (2.460 Å for graphene ZZ lattice and 2.556 Å for Cu<110> lattice) and the densely packed
Cu<110> atomic arrange, the alignment orientation with graphene ZZ edge attaching along the Cu<110>
direction is always more energetically favorable than other alignment orientations for all the three low-
index high symmetric Cu substrates (Figure 4.3). Therefore, for graphene growing on low-index high
symmetric Cu substrates, the most stable alignment orientation satisfies that a graphene ZZ direction
is parallel to a Cu<110> direction.

Based on the most stable alignment orientation of graphene on low-index high symmetric Cu
substrates, i.e., graphene ZZ // Cu<110>, we obtained all the possible alignments of a graphene island

68
on these Cu substrates, as illustrated in Figure 4.4. On the Cu(111) surface (Figure 4.4(a)), when ZZ
directions of graphene align along Cu<110> directions, its AC directions are also parallel to Cu<211>
directions, that is to say all symmetry axes and planes of the Cu(111) substrate are coincident with those
of graphene. Therefore, there is only one most stable alignment orientation of graphene on the Cu(111)
surface. On the Cu(100) surface (Figure 4.4(b)), when the most stable alignment orientation of graphene
ZZ // Cu<110> is satisfied (structure 0 in Figure 4.4(b)), the mirror symmetry planes along Cu<100>
of the Cu(100) substrate are not coincident with any mirror symmetry planes of graphene and thus a
new orientation marked as “1” appears, which shows an 45°misorientation angle with respect to the
original orientation. Therefore, graphene has two most stable alignment orientations on the Cu(100)
surface and they are equivalent to each other. The Cu(110) surface has only two mirror symmetry planes
and both of them coincide with those of graphene in the most stable alignment configuration, as
illustrated in Figure 4.3(c). Therefore, graphene on the Cu(110) surface has only one most stable
alignment orientation.

Figure 4.4 Illustrations showing the alignment of C6V-graphene islands on Cu(111) (a) Cu(100), (b)
and Cu(110) (c) surfaces. ZZ edges aligning along Cu<110> directions are marked by blue lines.

4.3.2 Alignment of 3-fold symmetric 2D materials on high-symmetric substrates

hBN has the same honeycomb lattice structure as graphene, but it shows a lower symmetry
(C3V) due to its binary composition. Both experimental observations and theoretical calculations have
shown that CVD grown hBN islands show triangular shapes enclosed by nitrogen terminated ZZ (ZZN)
edges under most experimental conditions.336-338 Therefore, the interaction between ZZN edge of hBN
and a substrate plays a critical role in determining the alignment orientation on this substrate.

By DFT calculations, we explored the binding strength of the hBN ZZN edge on low-index
high symmetric Cu surfaces with different alignment orientations. Figure 4.5(a-c) shows the atomic
configurations and corresponding binding energies. Due to the smaller lattice mismatch (~1.8%)
between ZZN direction of hBN and Cu<110> direction (2.510 Å for hBN ZZN lattice and 2.556 Å for

69
Cu<110> lattice) and the closely packed Cu<110> atomic arrays, the most stable alignment orientation
of hBN growing on these substrates satisfies that a ZZN direction of hBN is parallel to a Cu<110>
direction, which is as same as graphene growing on the three high symmetric Cu surfaces. As already
discussed in Chapter 2, previous DFT study already shows that the weak vdW interaction between the
hBN bulk and the Cu(111) surface is strongest when a ZZ direction of hBN is aligned along a Cu<110>
direction.113 Therefore, it can be seen that both the strong interaction between hBN ZZN edge and the
substrate and the weak interaction between hBN bulk and the substrate favor the alignment of hBN ZZ
direction along Cu<110> direction.

Figure 4.5 Binding energies of hBN ZZN edge attaching to three low-index high symmetric Cu surfaces,
which are Cu(111) (a), Cu(100) (b) and Cu(110) (c) surfaces. Similar to the configurations in Figure.
4.3, hBN ZZ nanoribbons were used and the boron terminated ZZ (ZZB) edges are passivated by
hydrogen to avoid their interaction with the Cu substrates.

Different from graphene and hBN, the edges of TMDCs are usually self-passivated by
chalcogen, as already revealed by both experimental and theoretical studies.318,339 In addition, due to the
three-atomic-thick structure of TMDCs, their binding stiffness is much higher than that of graphene and
hBN and, consequently the edges of TMDC islands do not bend towards the substrate surface. Therefore,
the alignment orientation of TMDCs on a substrate should be determined by the weak interaction
between the TMDC bulk and the substrate surface.

Here, to explore the orientation a TMDC island on FCC TM(111) substrates, we use a triangular
WS2 cluster with three tungsten terminated ZZ (ZZW) edges that are passivated by S on the Au(111)
surface as an example, the interactions between the WS2 cluster and the Au(111) surface under different
rotation angles are calculated by DFT, as shown in Figure 4.6(a), where the rotation angle is defined to
be 0°when a ZZW direction is parallel to a <110> direction of the Au(111) substrate.

70
The binding energy of the WS2 cluster on the Au(111) surface is defined as:

𝐸𝐵 = (𝐸𝐶𝑙𝑢𝑠𝑡𝑒𝑟 + 𝐸𝑆𝑢𝑏 − 𝐸𝑇 )⁄𝑁𝑊 , (4.3)


where ECluster and Esub are the energies of the WS2 cluster and the substrate, respectively. ET is the total
energy of the system with the WS2 cluster adsorbed on the Au(111) surface. NW is the number of W
atoms in the WS2 cluster.

Figure 4.6 (a) The configuration of a WS2 cluster on the high symmetric Au(111) surface, h is the
distance between WS2 and the Au(111) surface. (b) Comparison of the WS2 cluster with different
orientations on the Au(111) surface, the binding energies of the top-hcp and top-fcc structures are 0.653
and 0.645 eV/W atoms, respectively. (c) Binding energy of WS2 on the Au(111) surfaces in the unit
of eV/W atom, and the distance as a function of relative angles between Au<110> and ZZW directions.

Firstly, we testify that the Au(111) substrate can be regarded to be C6V symmetric for the system
of WS2/Au(111). The WS2 cluster with 0° and 60° orientation angles on the Au(111) surface is
compared, as shown in Figure 4.6(b). The binding energy difference of the two orientations is only
0.008 eV/W atom, which means that the effect of the second- and third- layer Au atoms on the WS2-
Au(111) interaction is negligible and the Au(111) substrate can be viewed to be C6V symmetric for the
WS2/Au(111) system. Therefore the periodic range for the rotation angle of the WS 2 cluster on the
Au(111) substrate can be limited in 0-60°. Figure 4.6(c) shows the binding energies and the distance
between the triangular WS2 cluster on the Au(111) substrate, and obviously, the WS2-Au(111)
interaction is strongest when a ZZ direction of WS2 is along a Au<110> direction.

71
Figure 4.7 Illustrations showing the alignment of C3V-2D material islands on FCC{111} (a), FCC{100}
(b) and FCC{110} (c) surfaces. ZZN edges aligning along <110> directions are marked by blue lines.
The mirror planes of the 2D material and the substrates along different directions are denoted by dashed
lines.

Based on the most stable alignment principle of 3-fold symmetric 2D materials on the three
low-index high symmetric Cu surfaces, i.e., ZZ // Cu<110>, we revealed the alignment of 3-fold
symmetric 2D material islands on FCC(111), FCC(100) and FCC(110) surfaces, as shown in Figure
4.7(a-c). C3V (D3H) 2D materials, such as hBN (MoS2), are mirror symmetric with respect to only AC
directions. On the FCC(111) surface (Figure 4.7(a)), the mirror symmetry planes along AC directions
of the 2D material only correspond to the mirror planes along the FCC<211> directions, while no mirror
symmetry planes of the 2D material correspond to the FCC<110> ones. Consequently, a new orientation
labeled as “1” appears and thus there are totally two most stable orientations that have an 60°
misorientation angle for the 2D material on the FCC(111) surface. The FCC(100) surface has two mirror
planes along the two FCC<110> directions and another two mirror planes along the two FCC<100>
directions. Among these four mirror planes, only one of them is coincident with a mirror plane of the
2D material, and thus the other three mirror planes result in three new alignment orientations of the 2D
material, labelled as “1, 2, 3” respectively, as shown in Figure 4.7(b). The FCC(110) surface has two
mirror symmetry planes along the FCC<110> and FCC<100> directions, respectively. Only the
FCC<100> mirror plane coincides with one AC mirror plane of the 2D material, and therefore there
will be two alignment orientations for a 3-fold symmetric 2D material on the FCC(110) surface, as
shown in Figure 4.7(c).

72
4.4 Alignment of 2D materials on low symmetric substrates

To get unidirectional alignment of 2D material islands on a substrate, it requires that the most
stable orientation of the 2D materials cannot be changed by any symmetric operations on the substrate,
which means that (i) the symmetry group of the substrate should be a subgroup of that of the 2D material;
(ii) all mirror planes passing the principal rotation axis of the substrate should be coincident with those
of the 2D material. Therefore, low symmetric substrates offer more opportunities for the growth of
unidirectionally aligned 2D material islands.

4.4.1 Alignment of 6-fold symmetric 2D materials on low symmetric substrates

Here, graphene is still used as a representative 6-fold symmetric 2D material to demonstrate the
most stable orientations of 6-fold symmetric 2D materials on the six types of CS Cu surfaces. By DFT
calculations, the interfacial formation energies of graphene nanoribbons with ZZ and AC edges
attaching to these surfaces are compared, respectively, as presented in Figure 4.8. Here, only ZZ and
AC edges of graphene attaching to the CS symmetric Cu surfaces are considered because only ZZ and
AC edges have a straight edge configuration, which bind to the straight step edges of the Cu substrates
stronger than other tilted graphene edges with kinks.

Figure 4.8 Interfacial formation energies of different graphene edges attaching to the CS Cu{111}-
based surfaces (a), CS Cu{100}-based surfaces (b) and CS Cu{110}-based surfaces (c).

The interfacial formation energy of a 2D materials’ edge attaching to a step edge of the CS
substrate is defined as:

𝐸𝐹 = 𝐸𝑒𝑑𝑔𝑒 − 𝐸𝐵 − 𝐸𝑣𝑑𝑊 , (4.4)

73
where 𝐸𝑒𝑑𝑔𝑒 is edge energy of a pristine edge of a 2D material, 𝐸𝐵 is the binding energy of a 2D material
edge attaching to a step edge of the CS substrate and 𝐸𝑣𝑑𝑊 is vdW interaction between the 2D material
nanoribbon and the substrate. The lower the interfacial formation energy, the more stable the alignment
orientation of the 2D material on the substrate.

For the C6V graphene growing on the two types of CS symmetric Cu{111}-based substrates
(Figure 4.8(a)), due to the perfect matching configuration between the Cu<110> step edge and the
graphene ZZ edge interface as well as between the Cu<211> step edge and the graphene AC edge, the
graphene ZZ edge prefers to attach to Cu<110> step edges while Cu<211> step edges prefer to be
docked by graphene AC edge. For graphene growing on the CS symmetric Cu{100}-based and
Cu{110}-based substrates, graphene ZZ edges attaching to all Cu step edges is more energetically
favorable than graphene AC edges, as shown in Figure 4.8 (b) and (c).

Above calculations infer that there will be only one most preferential alignment orientation for
a 6-fold symmetric 2D material grown on a high-index low symmetric substrate, because of the strong
binding between the edges of the 2D material and the step edges of the substrate. In the next Chapter, a
systematic study will be presented on the alignment of graphene on high-index low symmetric 2D
materials.

4.4.2 Alignment of C3V-2D materials on low symmetric substrates

Here, hBN is used as the representative 3-fold symmetric 2D material to demonstrate the
alignment of such 2D materials on the six types of CS Cu substrate, and Figure 4.9 shows the
configurations of hBN nanoribbons with ZZN and AC edges attaching to these substrates. It can be seen
that the configurations with hBN ZZN edge attaching to Cu step edges have a mirror symmetry respect
to the mirror plane of the substates, while the ones with hBN AC edge attaching to Cu step edges will
result in two equivalent hBN orientations that are antiparallel to each other. Therefore, only ZZN edge
attaching to the Cu step edges is beneficial to the growth of unidirectionally aligned hBN islands on the
CS Cu surfaces.

74
Figure 4.9 Interfacial formation energies of different hBN edges attaching to the CS Cu{111}-based
surfaces (a), CS Cu{100}-based surfaces (b) and CS Cu{110}-based surfaces (c) under the chemical
potential of nitrogen, 𝜇𝑁 =-9.575 eV.

Different from graphene, the formation energies of hBN edges and the interfacial formation
energies of hBN edges attaching to Cu step edges depend on the ambient condition, i.e., the chemical
potential of nitrogen (or boron), 𝜇𝑁 (𝜇𝐵 ). 𝜇𝑁 equals to -8.31 eV by using N2 as the reference while
equals to -10.84 eV by using α-boron as the reference (derived by 𝜇𝑁 = 𝜖𝐵𝑁 − 𝜇𝐵 = −10.84 eV, where
𝜖𝐵𝑁 denotes energy of a BN-pair in hBN structure). Here, the average of 𝜇𝑁 = −9.575 eV is chosen,
which represents most experimental conditions, and the calculated interfacial formation energies of
ZZN and AC edges attaching to various step edges of the six CS-Cu surfaces are given in Figure 4.9.
We can find that the interface with hBN ZZN edges attaching to Cu step edges are always more
energetically favored than that with AC edge, resulting in only one hBN orientation on the six types of
CS-Cu surfaces.

4.5 Summary

On a low-index high symmetric substrate, 2D materials usually show multiply orientations


depending on both the symmetries of the substrate and the 2D material. By DFT calculations, we found
that a high-symmetric direction of a 2D material prefers to align along a high symmetric direction of
the 2D material. We established a general strategy to obtain the equivalent but different alignment
orientations for a 2D materials on an arbitrary substrate by employing their rotation and mirror
symmetries. Based on this, we found that a 6-fold symmetric 2D material shows only one preferential
orientation on FCC(111) surfaces and FCC(110) surfaces, two equivalent but different orientations on

75
FCC(100) surfaces. 3-fold symmetric 2D materials such as hBN and TMDCs show two equivalent but
different orientations on FCC(111) surfaces, four equivalent but different orientations on FCC(100)
surfaces and two equivalent but different orientations on FCC(110) surfaces.

To obtain unidirectionally aligned 2D islands on a substrate, the orientation of a 2D islands


should not be changed by any symmetry operations of the substrate. Therefore, we proposed that
unidirectionally aligned 2D islands can only be obtained on a substrate, if the symmetry group of the
substrate is a subgroup of that of the 2D material. Consequently, high-index low symmetric substrates,
including C1 and CS ones, are more promising for the epitaxy growth of 2D materials than low-index
high symmetric substrates. Due to anisotropy of the C1 surfaces, there is only one local maximum for
the interaction strength between a 2D material and a C1 substrate, and thus C1 symmetric surfaces are
highly recommended for the epitaxy growth of 2D materials. On CS symmetric surfaces, the islands of
2D materials show one or two orientations, depending on the configurations of the interfaces. Among
them, C6V 2D materials always show one orientation on CS symmetric Cu surfaces, while C3V 2D
materials on CS symmetric surfaces can show one preferential orientation or two preferential
orientations antiparallel to each other, which depends on ambient conditions. In the next Chapter, we
present a systematic theoretical study on the alignment of graphene on high-index low symmetric 2D
materials.

76
Chapter 5 Alignment of graphene islands on low symmetric Cu
substrates
5.1 Introduction

Since 2008, the success of graphene CVD growth on TM surfaces60,61 boosts the explorations
for the epitaxial growth of graphene on various TM surfaces. Previous experimental and theoretical
studies mainly focus on investigating the alignment of graphene islands on low-index high symmetric
TM surfaces, because such surfaces usually have low surface energies and are easy to be produced by
annealing TM foils or thin film growth techniques. On the 6-fold symmetric Cu(111) surface, the 6-fold
symmetric graphene islands all show the same orientation,69,77,78,340 while on the 4-fold symmetric
Cu(100) surface, graphene islands show two preferential orientations with equivalent lattice
relationships but different rotation angles of 0 ±2ºand 30 ±2º.284,341,342 By using standard low-pressure
CVD method, it was reported that graphene growth on the 2-fold symmetric Cu(110) surface and a
reconstructed Cu(n10) surface derived from Cu(100) show two alignment orientations. 343 On the high-
index Cu(n10) surface graphene islands adopted two equivalent orientations with ± 8ºmisorientation
angles from the Cu[010] mirror axis, while on the Cu(110) surface, graphene islands preferred two
equivalent orientations with ±5ºmisorientation angles from the Cu[001] mirror axis, which is conflict
with above analysis that only one orientation on Cu(110) surface and may be caused by surface
reconstruction.

As discussed in Chapter 4, high-index low symmetric substrates are more promising in


templating the unidirectional alignment of 2D material islands due to the existence of step edges on
such substrates, and well-aligned graphene islands on high-index substrates were indeed observed
experimentally. Graphene nanoribbons with an alignment ratio of about 91% were grown on a vicinal
Ge(100) surface with ~ 9ºmiscut angle with respect to the Ge(100) surface, and the larger miscut angle,
the larger alignment ratio.344 Dai et al. reported the synthesis of unidirectionally aligned graphene
islands on the Ge(110) surfaces with intrinsic steps, and DFT calculations revealed that the alignment
behavior of graphene was induced by the strong interfacial coupling between graphene edges and steps
edges.76 Very recently, more than 30 kinds of single crystalline high-index Cu foils were obtained by a
seeded growth technique and epitaxial growth of single crystalline graphene films were also confirmed
on some of these high-index Cu foils.85

In principle, the number of high-index TM (like Cu) substrates is infinite. It is impossible to


investigate the alignment behaviors of 2D materials on all the high-index TM substrate by experiments.
An in-depth understanding on the alignment of a 2D material on an arbitrary substrate is of critical
importance for the controllable synthesis of the 2D material. In this Chapter, we present a systematic

77
study on the alignment mechanism of graphene islands on an arbitrary low symmetric Cu surface by
theoretical calculations.

5.2 Structures of low symmetric Cu substrates

In this section, the structures of various low symmetric Cu surface will be introduced. Except
the three high symmetric surfaces, i.e., C6V Cu(111), C4V Cu(100) and C2V Cu(110), all other Cu surfaces
are low symmetric with CS or C1 symmetries, which are composed of a large number of high symmetric
terraces connected step edges along an unidirectional direction (Figure 5.1). According to the terrace
configuration, low symmetric surfaces can be classified into three categories, named Cu{111}-based,
Cu{100}-based and Cu{110}-based low symmetric surfaces. The differences of low symmetric
surfaces of the same category are mainly from the structures and densities of their step edges.

There are two types of step edges: straight steps and tilted steps (see Figure 5.1), according to
their atomic configurations. The edge atoms of straight steps have the exact same coordination number
while that of tilted steps show a different coordination environment. To characterize different step edges
of low symmetric surfaces, the straight steps are considered as basic vectors and any tilted steps can be
constructed by two vectors with different ratios. As shown in Figure 5.1, for Cu{111}-based low
symmetric surfaces, there are two step vectors along Cu<110> and Cu<211> directions, respectively.
Tilted steps are composed by either <110> segments with <211> kinks or <211> segments with <110>
kinks. Here, a segment means that the length of the component is larger than its corresponding unit
length and a kink corresponds its unit length. For Cu{100}-based low symmetric surfaces, the two step
vectors are Cu<110> and Cu<100>, respectively, and tilted steps can be constructed by either <110>
segments with <100> kinks or <100> segments with <110> kinks. In contrast with the above two
categories, step edges on Cu{110}-based low symmetric surfaces have three types of straight steps and
results in three step vectors: Cu<110>, Cu<211> and Cu<100>. Tilted steps can be classified into four
types: <110> segments with <211> kinks, <211> segments with <110> kinks, <211> segments with
𝑖×<segment>
<100> kinks, and <100> segments with <211> kinks. To make it simply, we use SE<kink> to
describe an arbitrary step edge, in which the directions of segment and kink composed to the step edge
are shown in the superscript and subscript, respectively, and i denotes the times of unit length for the
<segment>
segment. It is worth noting that high-index low symmetric surfaces with straight steps, SE0 ,
are generally CS symmetry that shows mirror symmetry with respect to the direction perpendicular to
the straight step except the C1 Cu{110}-based low symmetric surface with SE0<211> steps, while
surfaces with tilted steps are all C1 symmetric.

78
𝑖×<segment>
Figure 5.1 The configurations of various low symmetric Cu surfaces, where SE<kink> represents

a step edge with a segment direction denoting in superscript and a kink direction denoting in subscript,
and i is the number of unit cells in the segment.

Although step edges (either straight or tilted steps) on high-index low symmetric surfaces are
ideally unidirectional with a constant direction, they are usually highly fluctuated and are composed of
various segments along different directions in practice, especially under the high temperatures for 2D
materials CVD synthesis.276,345 Therefore, the effect of the step edge variation on the alignment of 2D
materials grown on the substrates is also critical and will be addressed in this chapter.

Figure 5.2 sketched the structures of Cu{111}-based, Cu{100}-based, and Cu{110}-based low
symmetric surfaces, respectively, with step edge directions ranging from -180~180°, which can show
the structures of all types of step edges composed of different step edge segments and kinks. In Figure
5.2(a), the atomic configurations of three Cu{111}-based low symmetric surfaces are displayed at the
outer side, where the same terrace but different step edges can be clearly seen. Due to all low symmetric
surfaces has the exact Cu[110] direction, therefore here the Cu[110] step edge is chosen as reference of

79
0°for all three categories of low symmetic surfaces, and other step edge dierction are describe by the
their relative angles with the Cu[110] direction.

Figure 5.2 Sketches of three categories of low symmetic surfaces, i.e., Cu{111}-based (a), Cu{100}-
based (b) and Cu{110}-based (c) high-index low symmetric surfaces. Various surfaces with different
step edge direction are distinguished by different colors, and the straight step direction is marked by
dashed lines. (d-f) The evolution of step kink density versus step edge direction as a function of step
edge orientation for Cu{100}-based, Cu{100}-based and Cu{110}-based high-index low symmetric
surfaces, respectively. The Cu[110] direction is chosen as reference of 0°for all three categories of low
symmetic surfaces .

Due to the 6-fold symmetric the Cu(111) surface, the periodicity of the step edge orientation is
0°~60°. Considering the mirror symmetry with respect to the Cu<110> and Cu<211> directions, the
step edge orientation can be further restricted into 0°~30°, which varies from Cu<110> to Cu<211>
direction. Figure 5.2 (d) shows the evolution of the kink density as a function of the Cu step edge
orientation. At 0°and 30°, the two SE0<110> and SE0<211> have no kinks and therefore with a kink
density of 0. With the step orientation deviating from the straight steps, the kink density keeps
increasing. Depending on the main components of the tilted step edges, they can be further classified
𝑖×<110> 𝑖×<211>
into <110> dominated steps (SE<211> ) and <211> dominated steps (SE<110> ). At 19.11°, the tilted
<110> <211>
step (SE<211> or to say SE<110> ) is composed of <110> kink and <211> kink that are alternatively

80
connected and thus shows the highest density of kinks. Figures 5.2 (b) and (e) show the structure of
Cu{100}-based high-index low symmetric substrates and the kink density profile as a function of the
Cu step edge orientation, respectively. Due to C4V symmetry of the Cu(100) surface, the non-repeating
range of the step edge orientation for the kink density profile is 0°~45°, which varies from one Cu<110>
<110>
direction to its neighboring Cu<100> direction. The highest kink density appears at SE<100> or
<100>
SE<110> at a step edge direction of 26.57°. Due to the C2V symmetry of the Cu(110) surface, the non-
repeating range of step edge orientation on Cu{110}-based high-index low symmetric substrates is
0°~90°, which varies from one Cu<110> direction to the neighboring Cu<112> direction and then to
the neighboring Cu<001> direction. In the 0°~35.26°, 35.26°~70.53°, and 70.53°~90°ranges, step
𝑖×<110> 𝑖×<211> 𝑖×<211>
edges are dominated by <110> segments (SE<211> ), <211> segments (SE<110> and SE<100> ),
𝑖×<100>
and <100> segments (SE<211> ), respectively.

5.3 Modeling and simulation methods

5.3.1 Modeling

To explore the alignment orientation of graphene islands on an arbitrary Cu step, we proposed


a general model to describe different interfaces between graphene and Cu steps, as shown in Figure 5.3.
Similar with Cu step edges, graphene edges also have two types: straight edges (including ZZ edge and
AC edge) and tilted edges. Any tilted edges can be viewed to be constructed by straight edge segments
connected by kinks, which is similar to Cu step edges. For example, the tilted graphene edge in Figure
5.3 (b) and (d) is composed by ZZ segments with a length of two ZZ sites connected by AC kinks.

Since both Cu step edges and graphene edges have two types, the interface of a graphene edge
attaching to a Cu step edge have four possible scenarios: (i) a straight graphene edge (either ZZ or AC)
attaching to a straight Cu step edge (see Figure 5.3(a)); (ii) a tilted graphene edge attaching to a straight
Cu step edge (see Figure 5.3(b)); (iii) a straight graphene edge (either ZZ or AC) attaching to a tilted
Cu step edge(see Figure 5.3(c)); (iv) a tilted graphene edge attaching to a tilted Cu step edge(see Figure
5.3(d)). When the steps of Cu substrate are straight, interface (i) shows obvious advantages over
interface (ii) due to the stronger bonding between straight graphene edges and Cu step edges, which has
already been proved by previous studies. By DFT calculations, it was reported that along the straight
Cu<110> steps of Cu(111) surface, the interfaces between graphene ZZ (AC) edges and the Cu<110>
steps are energetically more favorable than interfaces between other possible graphene edges and the
Cu<110> step edges.282 When Cu steps are not straight, the stabilities of interfaces (iii) and (iv) need
further investigation. Here, for the most common interface (iv), to increase its stability, the segment of
the two tilted edges should match with each other as much as possible, i.e., the directions of graphene

81
edge and step edge are parallel (𝑙1 // 𝑙2) and the unit lengths of the two edges are close to each other, and
using 𝑙 to denote the unit length of tilted graphene edge or tilted Cu step edge, i.e., 𝑙1≈ 𝑙2=𝑙. However,
due to the lattice mismatch between graphene and Cu substrate, the kink heights of tilted graphene edge
and that of tilted step edge are different, which induces a small misorientation angle (∆𝛾) between the
graphene edge segment and the step edge segment.

Figure 5.3 Four interfacial configurations of different types of graphene edges attaching to different
types of step edges. In (d), 𝑙1 and 𝑙2 represent the unit length of tilted graphene edge or tilted Cu step
edge; 𝑑1 and 𝑑2 are the kink heights of graphene edge and Cu step edge, 𝛾1 denotes the angle between
the tilted graphene edge and the edge segment, and 𝛾2 denotes the angle between the tilted Cu step edge
and its segment.

According to the geometrical relationship of interface (iv), the misangle can be derived as
followed:

𝑑2 𝑑1
∆𝛾 = 𝛾2 − 𝛾1 = asin − asin (5.1)
𝑙 𝑙

where 𝑑1 and 𝑑2 are the kink heights of graphene edge and Cu step edge, respectively; 𝛾1 is the angle
between the tilted graphene edge and the edge segment, and similarly 𝛾2 is the angle between the tilted
Cu step edge and its segment. Obviously, the misorientation angle, ∆𝛾, will change the orientation of
graphene islands gradually when the Cu step edge deviates from the straight step edge direction.

82
Moreover, the larger kink height difference, the larger misorientation angle and therefore the larger
orientation change.

Figure 5.4 (a) Atomic model for calculating the formation energy of a graphene edge attaching to a
step edge. (b-c) Atomic models for calculating graphene edge energy.

In order to compare the stabilities of different interfaces quantitatively, we build atomic models
with a graphene nanoribbon attaching to a step edge of Cu substrate to mimic the interface between a
graphene edge and a Cu step edge (Figure 5.4). The other edge of the graphene nanoribbon is terminated
by hydrogen to avoid its interaction with the Cu terrace. Therefore, the formation energies of a graphene
edge attaching to a Cu step edge can be defined as: 𝐸𝑓 = 𝜀𝑒𝑑𝑔𝑒 − 𝐸𝐵 , where 𝜀𝑒𝑑𝑔𝑒 represents the edge
energy of a pristine graphene edge and 𝐸𝐵 is the binding energy between the edge of the graphene
nanoribbon and the Cu substrate step edges.

The binding energy (𝐸𝐵 ) is thus calculated by: 𝐸𝐵 = (𝐸𝐶𝑢 + 𝐸𝐺𝑟𝑎 − 𝐸𝑇𝑜𝑡 − 𝐸𝑣𝑑𝑊 )/𝐿, where
𝐸Cu , 𝐸Gra , 𝐸Tot and 𝐸vdW are the energies of the Cu substrate slab, graphene ribbon, graphene ribbon
attaching to the Cu substrate and the vdW interaction between graphene ribbon and the Cu substrate,
respectively. L is the length of the super cell along the step edge direction.

The models used to calculate the formation energy of a pristine graphene edge are shown in
Figure 5.4 (b-c). Structure II in Figure 5.4(b) is a graphene nanoribbon with its two edges passivated by
hydrogen, and the structure I in Figure 5.4(c) is half of the structure II and only one of its edges is
passivated by hydrogen. Thus, 𝜀𝑒𝑑𝑔𝑒 can be obtained by: 𝜀𝑒𝑑𝑔𝑒 = (2𝐸𝐼 − 𝐸𝐼𝐼 )/𝐿′, where 𝐸𝐼𝐼 and 𝐸𝐼
are the total energies of the wo graphene ribbons with same length, 𝐿′.

5.3.2 Simulation methods

The calculations are performed by DFT-D3 method with VASP.346,347 The exchange-correlation
functional is treated by GGA,348 and the interaction between valence electrons and ion cores is treated

83
by PAW.349 Periodic boundary conditions are applied. Along the step edge or graphene edge direction,
the length difference between graphene edge and step edge is limited to 3.0%, and to avoid strain in
graphene the graphene lattice length is kept unchanged. Along the direction perpendicular to the step
direction, the distance between two steps is set to be ~20 Å for all structures. Along the out of plane
direction, the vacuum spacing between neighboring images is set to be larger than 12 Å. For structural
optimization, the force criteria on each atom is less than 0.01 eV/Å, and an energy convergence of 10-4
eV is used.

5.4 Results and Discussions

Based on the interfacial model of Figure 5.3, graphene alignment along all possible step edges
of three categories of high-index low symmetric Cu substrates (Cu{111}-based, Cu{100}-based and
Cu{110}-based substrates) are explored systematically. The alignment of graphene islands along curved
steps with varying directions are also studied. This study provides a comprehensive understanding on
the alignment of graphene on an arbitrary substrate and we believe it can help experiments for better
design in the synthesis of unidirectional graphene islands and further WSSC graphene film.

5.4.1 Alignment of graphene islands on Cu{111}-based low symmetric surfaces

To figure out graphene alignment on Cu{111}-based low symmetric substrates with different
Cu step, the alignment orientation of graphene along the two straight Cu step edges, Cu<110> and
Cu<211> steps are explored firstly. Figure 5.5(a) displays the atomic configurations and formation
energies of various graphene edges attaching to the two straight Cu step edges. We found that structural
match is of important for the interfaces with straight graphene edges attaching to straight Cu step edges.
Along the Cu<110> step edge (higher panel in Figure 5.5(a)), the interface with the graphene ZZ edge
shows the lowest formation energy due to the structural match between the Cu<110> step edge and
graphene ZZ edge, while the interface with the graphene AC edge is not energetically favored despite
of its straight edge structure. Tilted graphene edges distort the atomic arrangement of the Cu<110> step
edge more or less, resulting in higher formation energies. Along the Cu<211> step edge (lower panel
in Figure 5.5(a)), graphene AC edge shows obvious advantages in matching the Cu step edge and
therefore the corresponding interface has the lowest formation energy. In this case, graphene ZZ
direction is still parallel to Cu<110> direction, which is the same as the case for Cu<110> step edges.
Based on the graphene orientation along the two straight Cu step edges, the alignment of graphene
islands along an arbitrary step edge with its direction from -90 to 90ºis shown in Figure 5.5 (b), and we
can see that all islands show the same orientation along <110> segments and <211> segments because
both graphene and Cu(111) facet are 6-fold symmetric.

84
Figure 5.5 (a) Formation energies and configurations of various graphene edges attaching to the two
straight steps of Cu{111}-based low symmetric surfaces. (b) Sketch of graphene islands along a curved
step varying from -90 ~ 90º.Graphene islands along straight steps all show same orientation.

Above calculations do not consider tilted Cu step edges, but there must be a large amount of
tilted step edges between the transition of the two straight steps. Therefore, we further calculated the
formation energies of different graphene edges attaching to various tilted step edges, as shown in Table
5.1. Graphene ZZ direction (G<10>) and Cu<110> direction are chosen as the references of graphene
orientaion and Cu step edge orientations respectively, along the diagonal of the table, the relative angle
of the grpahene edge to its reference and that of Cu step edge to its reference are same, and the formation
energy is smallest along both row and column directions. It is worth noting that, except the matched
interfaces (iv) of Figure 5.3(d) where tilted graphene edges match the tilted Cu step edges, the interfaces
(iii) with graphene ZZ edge attaching to tilted Cu step edges are stable than other unmatched interfaces
with tilted garphene edges attaching to tilted Cu step edges. The corresponding atomic configurations
with lowest formation energies are shown in Figure 5.6 (a), graphene edges and Cu step edges all match
very well at the interfaces, satisfying all requirements of interface (type (iv) of Figure 5.3(d)). Moreover,
in consistent with the alignment of graphene on the straight Cu step edges, ZZ segments of graphene
𝑖×<110>
tilted edges prefer to align along <110> step segments of the SE<211> step edges, and AC segments
𝑖×<211>
of graphene tilted edges prefer to align along <211> step segments of the SE<110> step edges.

85
Table 5.1 Formation energies of graphene edges attaching to different Cu step edges of Cu{111}-based
low symmetric surfaces.

Ef (eV/Å) G<10> G<14> G<13> G<12> G<23> G<34> G<11>


SE0<110> (Cu<110>) 0.35 0.44 0.46 0.49 0.50 0.47 0.45
3×<110>
SE<211> (Cu<145>) 0.49 0.35 0.51
2×<110>
SE<211> (Cu<134>) 0.48 0.37 0.51 0.53
<110>
SE<211> (Cu<123>) 0.46 0.49 0.39 0.54 0.52
2×<211>
SE<110> (Cu<235>) 0.47 0.49 0.40 0.53
3×<211>
SE<110> (Cu<347>) 0.49 0.40 0.52
SE0<211> (Cu<112>) 0.46 0.50 0.46 0.47 0.55 0.53 0.39

Figure 5.6 (a) Matched interfaces of graphene edges attaching to different kinked Cu step edges. (b)
The intact evolution of graphene orientation (dark lines) versus different step edges. Light lines showing
kink density of the step edges.

Choosing the Cu[110] step direction in Figure 5.5(b) as the reference of 0º, the graphene ZZ
direction with the reference is used to represent the orientation of graphene island. Figure 5.6(b) shows
the calculated graphene orientation profile (solid dark line) as a function of the orientation of the Cu

86
step edges on the Cu{111}-based high-index substrates. The small hexagons in Figure 5.6(b) denote
the corresponding orientations graphene islands. On Cu substrates with the straight SE0<110> (60º× 𝑛,
n is an integer), graphene ZZ edge aligns along the step edges and the graphene orientation is 0º. On
Cu substrates with the straight SE0<211> (30º+60º× 𝑛), graphene AC edge is parallel to the step,
meanwhile, graphene ZZ direction is parallel to Cu<110> direction and therefore the graphene
orientation is also 0º. On Cu substrates with tilted step edges, as inferred in the interface (type (iv) in
Figure 5.3(d)), the kink height difference between the tilted graphene edge and the tilted Cu step edge
would induce a small misorientation angle between their segments. As shown in Figure 5.6 (b), on Cu
𝑖×<110>
substrates with SE<211> step edges, graphene ZZ direction deviates from Cu<110> direction with a
small misorientation angle varying in a range of 0º ± 0.74º (solid black line). Similarly, on Cu
𝑖×<211>
substrates with SE<110> step edges, graphene AC (ZZ) direction deviates from Cu<211> (Cu<110>)
direction with a small misangle ranging in 0º ±0.41º. Although segments have more contributions and
should be the dominant factors in determining the orientation of graphene islands, the effects of kinks
cannot be ignored especially for the step edges with a high kink density. Thus, the dotted dark lines
representing graphene orientation determined by kinks are also plotted. On Cu substrates with step
edges with a high kink density, graphene islands may have a large chance to show two orientations.
However, on all Cu{111}-based high-index low symmetric surfaces, the two orientations are quite close,
with a difference less than 1º, which is difficult be identified in experiments.

In summary, graphene orientation is not sensitive to the step edge variation on Cu{111}-based
high-index low symmetric substrate. Therefore, unidirectionally aligned graphene islands can be easily
achieved on these substrates regardless of the directions of the step edges, which means that Cu{111}-
based high-index low symmetric substrates show a very large tolerance for the synthesis of
unidirectionally aligned graphene.

5.4.2 Alignment of graphene islands on Cu{100}-based high-index low symmetric substrates

On Cu{100}-based high-index low symmetric substrates, the formation energies of various


graphene edges along the straight step edges are also compared firstly, and it is found that along both
Cu<110> and Cu<100> step edges, the interfaces with a straight graphene ZZ edge are always most
stable, as shown in Figure 5.7(a), proving the stability of the interface (type (i) of Figure 5.3(a)) along
the straight Cu step edges. For the Cu<110> step edge, the lattices of graphene ZZ edge and Cu step
edge matches very well. For Cu<100> step edge, neither ZZ nor AC edge can match it, but ZZ edge is
more energetically favorable. Based on this relationship, the alignment of graphene islands along
straight segments of a curved step edge with varying direction from -90 to 90º is illustrated in Figure

87
5.7(b). There are totally four orientations, and graphene islands change their orientations if the direction
of the dominant segments of curved step edges changes.

Figure 5.7 (a) Formation energies of various graphene edges attaching to the two straight step edges on
Cu{100}-based low symmetric surfaces. (b) Sketch of graphene islands along straight steps show 4
orientations. (c) Configurations of graphene ZZ and tilted edges attaching to tilted step edges. (d) The
evolution of graphene orientation (dark lines) versus different steps. Light lines showing kink density
of the step edges.

To further explore the graphene orientation on Cu{100}-based low symmetric substrates with
tilted step edges, we compared the stability of interface (iii) and (iv) (illustrated in Figure 5.3(c) and
(d)), and the formation energies and atomic configurations of the graphene edges (both along ZZ
𝑖×<110> 𝑖×<100>
direction and tilted direction) attaching to tilted step edges including SE<100> and SE<110> , are
shown in Figure 5.7(c). Obviously, the structure-matched interfaces that segments correspond to
segments and kinks correspond to kinks are more favorable.

Based on above results, the evolution of graphene orientation on Cu{100}-based low symmetric
substrates with different step edge directions varying from -90 to 90ºis illustrated in Figure 5.7(d), here,
the Cu[110] direction shown in Figure 5.7(b) is chosen as the orientation reference. Along straight
SE0<110> step edges, graphene orientation is 0º (orange hexagon). On Cu substrates with SE<100>
𝑖×<110>

step edges (0º ~ ±26.57º), graphene orientation deviates from the Cu<110> segment direction with a

88
misorientation angle varying in a range of 0º ±4.68º (solid black line). On Cu substrates with straight
SE0<100> step edges, graphene ZZ edge aligns along it, as a result, the orientation of graphene is rotated
by 45º, which is equivalent to -15º (blue hexagon) because of the 6-fold symmetry of graphene. On Cu
𝑖×<100>
substrates with SE<110> step edges (26.57º~ 45º), graphene orientation deviates from the Cu<100>
segment direction with a misorientation angle ranging in -15º ± 3.45º (solid red line). Obviously, the
large ranges of misorientation angles imply that graphene orientation is sensitive to the directions of
step edges with both <110> and <100> domination on Cu{100}-based low symmetric surfaces.
Therefore, on Cu{100}-based low symmetric surfaces, well-aligned graphene islands are difficult to
synthesize unless all step edges are unidirectional and parallel to each other.

5.4.3 Alignment of graphene islands on Cu{110}-based high-index low symmetric surfaces

Cu{110}-based high-index low symmetric surfaces have three types of straight step edge.
Figure 5.8(a) shows the formation energies of various graphene edges attaching to these three straight
step edges, and it turns out that the straight graphene ZZ edge attaching to these edges is always
energetically favorable, supporting interface (i) in Figure 5.3(a). Based on this, the schematics of
graphene islands along a step edge with varying direction from -90ºto 90ºis shown in Figure 5.8(b),
and the exact Cu[110] direction is still chosen as the orientation reference. Similar to that on Cu{100}-
based low symmetric surfaces, there are four graphene orientations respectively along the four straight
step edges with different directions.

For tilted Cu step edges, the stabilities of interfaces (iii) and (iv) shown in Figure 5.3(c-d) are
investigated, and we found that interfaces of type (iv) are more energetically favored. Four
representative configurations of the matched interface (iv) with different tilted step edges attached by
corresponding tilted graphene edges are shown in Figure 5.8(c), and all of them show that the ZZ
segment of graphene edge aligns along the straight segment of Cu step edges regardless the types of the
segment. Based on this, the evolution of graphene orientation with the change of Cu step edge directions
varying from -90º to 90º is calculated and shown by a liner graph (Figure 5.8(d)). On Cu{110}-based
low symmetric surfaces with straight SE0<110>step edges, graphene orientation is 0º (orange hexagon).
𝑖×<110>
On Cu{110}-based low symmetric surfaces with SE<100> step edges, graphene orientation deviates
from the Cu<110> direction with a misorientation angle of 0º ±15.37º (solid black line). On Cu{110}-
based low symmetric surfaces with the straight SE0<211> step edges, graphene orientation is rotated by
54.74º, which is equivalent to -5.26º because of the 6-fold symmetry of graphene (purple hexagon).
𝑖×<211> 𝑖×<211>
On Cu{110}-based low symmetric surfaces with SE<110> and SE<100> step edges, due to the small
difference in the kink heights of graphene tilted edges and Cu tilted step edges, the graphene
𝑖×<211>
misorientation angles are quite small, which change from 0º ~ 0.38º for the SE<110> step edges and

89
𝑖×<211>
changes from 0 ~ 0.22ºfor the SE<100> ones (solid red line). On Cu{110}-based low symmetric
surfaces with straight SE0<100> step edges, graphene orientation is 30 º (green hexagon) and, on
𝑖×<100>
Cu{110}-based low symmetric surfaces with SE<211> step edges, graphene orientation varies in a
range of 30º ± 3.34º (solid blue line). Overall, graphene orientation is not sensitive to the step edge
𝑖×<211> 𝑖×<211>
variation on Cu{110}-based low symmetric surfaces with SE<110> and SE<100> step edges, but
𝑖×<110> 𝑖×<100>
sensitive for the cases with SE<100> or SE<211> ones. Hence, on Cu{110}-based low symmetric
surfaces, well-aligned graphene islands can be synthesized under two cases: all Cu step edges are
constructed by <211> segments with kinks or all step edges are unidirectional and parallel to each other.

Figure 5.8 (a) Formation energies of various graphene edge attaching to three straight steps on
Cu{110}-based low symmetric surfaces. (b) Sketch of graphene islands along straight steps show 4
orientations. (c) Configurations of tilted graphene edges attaching to tilted step edges. (d) The evolution
of graphene orientation (dark lines) versus different steps. Light lines showing kink density of the step
edges.

5.5 Summary

In this Chapter, we for the first time present a systematic study on the alignment of graphene
on high-index low symmetric Cu surfaces, which can be classified into three categories, i.e., Cu{111}-
based, Cu{100}-based and Cu{110}-based high-index low symmetric substrates. Four types of models
for the interface structure between the substrate step edges and graphene edges are proposed. Based on

90
these models, we performed extensive DFT calculations and geometry analysis, and found that the
orientation of graphene on such substrates is highly dependent on the step edges of the substrates. On a
Cu{111}-based low symmetric surface, well-aligned graphene islands can be always achieved despite
of the directions of the step edges. On a Cu{100}-based low symmetric surface, graphene orientation
is very sensitive to the direction of the Cu step edge and well-aligned graphene islands can be achieved
only when all the step edges of the surface are parallel to each other with a constant direction. On a
Cu{110}-based low symmetric surface, if the Cu step edges are dominated by Cu<211> segments, there
is a large tolerance for the variation of step edge direction on the orientation of graphene islands, and
thus well-aligned graphene islands can be easily synthesized, otherwise well-aligned graphene islands
can be achieved only if the surface has unidirectional step edges.

91
Chapter 6 Single-crystal graphene grown on twinned Cu
substrates

6.1 Introduction

It has been widely acknowledged that the orientations of grown 2D material islands is
determined by the crystalline structure of the substrates on which epitaxial growth takes place, and
therefore, to grow multiple graphene domains with unidirectional orientation, single crystalline
substrates are intuitively required.350,351 In contrast with this conventional understanding, our theoretical
analysis shows that the growth of single crystal graphene by seamless coalescence of unidirectionally
aligned graphene islands is also possible on specific polycrystalline TM substrates. In this Chapter, we
present a theoretical study corroborated by experimental demonstrations showing that single crystalline
graphene films can be grown on properly tailored twinned Cu substrates. Inch-sized twinned Cu foils
were successfully obtained by a strain-induced annealing technique, and it is found the grown graphene
islands on two sides of the twin boundaries of the substrate show a great possibility to align
unidirectionally, and single crystalline graphene was realized by seamless stitching of these
unidirectional graphene islands. Moreover, we have proved that this method can be applied to the
synthesis of other WSSC 2D materials, paving a new way of synthesizing WSSC 2D materials.

6.2 Experimental methods and results

6.2.1 Preparation of twinned Cu foils

In this work, three different processing methods for Cu foils before annealing are designed. For
the first one, the commercial Cu foil was kept in a flat shape and supported by a quartz slab in the
furnace, as displayed in Figure 6.1(a). Then the Cu foil was annealed at 1010 °C for 1 hour in a H2
environment. After that, we oxidized the annealed foil to check its surface structure, and the optical
image in Figure 6.1(b) shows the surface orientation is very uniform without obvious surface orientation
difference. The EBSD map in Figure 6.1(c) shows a strong {001}<100> cubic texture with an average
grain size of ~ 300 μm for the annealed Cu foil, which is consistent with previous reports.352,353 It is
worth noting that a very small fraction of twined grains (the blue area) can be observed. For the second
process, the commercial Cu foil was bended by the furnace tube wall (see Figure 6.1(d)) and then
annealed. Both optical image (Figure 6.1(e)) and EBSD map (Figure 6.1(f)) of the annealed bended Cu
foil show various surface orientations but the sizes of grains are much larger than that in the annealed
flat Cu foil, indicating that extra stress can significantly promote the migration and growth of grain
boundaries during annealing.

92
Figure 6.1 (a) Photo of a flat Cu foil. (b) Optical image of the annealed flat Cu foil after oxidization.
(c) EBSD map of the annealed flat Cu foil. (d) Photo of a bended Cu foil. (e) Optical image of the
annealed bended Cu foil after oxidization. (f) EBSD map of the annealed bended Cu foil.

For the third processing method, the polycrystalline Cu foil was not only bended but also
manipulated by a microhardness indenter in the central area to induce a preferred nucleation site and
further a preferred growth of Cu grains at this site, as shown in Figure 6.2(a). The tailored Cu foil was
annealed for one hour in a H2 environment at 1010 °C. After annealing, the Cu foil was then oxidized,
and its optical image is shown in Figure 6.2(b). Obviously, there are two contrasts in the optical image,
suggesting that there are two types of Cu surfaces on the annealed Cu foil. To figure out the surface
orientation of the annealed Cu foil, EBSD was performed and the maps of three representative areas are
displayed in Figure 6.2(c). It is revealed that all dark areas are single crystalline with (116) crystalline
surface and light areas show (111) crystalline surface, and there is no in-plane rotation within the single
crystalline areas and grain boundaries only exist in the transition areas between the two single crystalline
areas. The pole and inverse pole figure in Figure 6.2(d) proved that there are no other crystalline
orientations in the annealed Cu foil, further supporting the EBSD results. In Figure 6.2(e), High
resolution transmission electron microscope (HRTEM) image of the grain boundary shows that the two
grains with (116) and (111) surfaces are mirror symmetric with respect to their grain boundary,
indicating the twin characteristic of the annealed Cu foil. Combining with the corresponding selected
area electron diffraction (SAED) pattern (insert of Figure 6.2(e)), the two grains were coaxial along one
<111> direction and had a 60°misorientation around their <111> coaxis, which is consistent with the
most popular grain boundary in FCC TMs due to its lowest formation energy, called <111>/60°twin
boundary.354,355

93
Figure 6.2 (a) Photo of a bended Cu foil with microhardness indent (shown in insert) in center area. (b)
Optical image of an oxidized twinned Cu foil. (c) EBSD maps of three representative areas of the
twinned Cu foil marked in (b), with scale bar of 500 m; (d) (111) Pole figure and inverse pole figure
of the twinned Cu foil. (e) HRTEM and SAED images of the grain boundary in the twinned Cu foil.

6.2.2 Synthesis of unidirectional graphene islands on twinned Cu foils

Using ambient pressure chemical vapor deposition (APCVD) method, graphene was
synthesized on such a twined Cu foil and the typical growth process versus the growth time is shown in
Figure 6.3(a). At nucleation stage, the coverage rate kept quite low for a certain time period, implying
the nucleation incubation process. During growth, the coverage rate increased significantly, which was
followed by a final slow growth period, indicating a self-limited growth of graphene film. After ~15
minutes growth, graphene islands merged and a continuous graphene film was formed. Three
representative optical microscopy images of graphene islands grown on the twined Cu foil were shown.
Despite that graphene islands on the two sides of a Cu twin grain boundary showed different shapes,
which were perfect hexagons on the Cu(111) surface (light part) and elongated hexagons on the Cu(116)
surface (dark part), their hexagonal shapes and the 120o inner angle of the vertices proved their single
crystalline nature. It is worth noting that all graphene islands showed the same orientation with one pair
of their ZZ edges paralleled with the Cu grain boundary, which was in sharp contrast with previous
reports of graphene grown on polycrystalline Cu foils, where graphene islands show random
orientations. Hydrogen etching at 1000ºC for 10 min was also performed to verify the crystallinity of
the obtained graphene film,356,357 and the results are shown in Figure 6.3(b). It can be seen that all etched
holes were hexagonal and unidirectionally aligned, confirming that the synthesized graphene film was
single crystalline with only point defects. Furthermore, the crystallinity of graphene grown on a 5×5

94
mm2 area Cu foil with abundant twin grain boundaries was further characterized by LEED. Figure 6.3(d)
show the diffraction patterns of the synthesized graphene at different but uniformly distributed locations
of the Cu foil in Figure 6.3(c), it turns out that the graphene film over the whole area was single
crystalline.

Figure 6.3 (a) Growth process of graphene on the twin Cu foil. (b) Optical images of graphene films
on the twin Cu foil after hydrogen etching. (c) Photo and EBSD map of the twin Cu foil. (d) LEED
patterns of 16 different areas of graphene on the twin Cu foil.

6.3 Theoretical modeling and discussions

6.3.1 Modeling of twinned Cu foils

Since we have confirmed that the grain boundaries in the annealed Cu foils are <111>/60o twin
boundaries, we established a theoretical model of such a twined bicrystal, as shown in Figure 6.4. For
the bicrystal with <111>/60o twin boundary, it requires that (i) the two grains separated by the grain
boundary are coaxial along one <111> direction, which means ∆𝜃 = |𝜃𝐵 − 𝜃𝐴 | = 0°, and (ii) the
misorientation of the two grains is 60o around the <111> coaxis, that is to say ∆𝜓 = |𝜓𝐵 − 𝜓𝐴 | = 60° ,
where θA and θB represent the angles between the normal direction of the (111) boundary plane of the
grain A and B with that of the Cu foil, 𝜓A and 𝜓B denote the rotate angle of grain A and B around their
<111> coaxis (see Figure 6.4 (a)), respectively. With fixed values of ∆𝜃 and ∆𝜓, the orientation of the
bicrystal with a <111>/60o twin boundary can be described by two degrees of freedom, 𝜃𝐴 and 𝜓𝐴 ,
which are shortened as 𝜃 and 𝜓 in the following. The atomic configuration of the (116)/(111) Cu
bicrystal is displayed in Figure 6.4 (b), and the mirror symmetry with ABC|BAC stacking is clearly
seen with respect to the grain boundary.

95
Figure 6.4 (a) The definition of the <111>/60o twin Cu foil, where θA (θB) represents the angles between
normal direction of the (111) boundary plane of the grain A (B) and that of the Cu foil, 𝜓A and 𝜓B
denote the rotate angle of grain A and B around their <111> coaxis. <111>/60o twin requires ∆𝜃 =
0° and ∆𝜓 =60°. (b) Atomic model of the twin Cu foil with a <111>/60o twin boundary.

In experiments, EBSD technique gives three Euler angles (φ1, Φ, φ2) to describe the orientations
of grains in the Cu foil. Figure 6.5(a) shows an arbitrary Cu grain in a Cu foil, the Cu grain is described
by crystalline coordination system with three axes corresponding to [100], [010] and [001] respectively,
while the Cu foil is in the coordinates of specimen, and the three axes are denoted by its transverse
direction (TD), rolling direction (RD) and normal direction (ND). Figure 6.5(b) and (c) illustrate the
three rotational operations executed by the Euler angles: rotating along the Z axis with φ1 degree, and
then rotating along the new X axis with Φ degree, and the third rotation is along the updated Z axis with
φ2, where X, Y and Z axes are corresponding to TD, RD and ND directions, respectively. The
coordinates of the Cu foil finally coincide with that of the Cu grain after these three rotations. Therefore,
the orientation of any grain in the Cu foil can be confirmed. In Table 6.2, the grain boundaries of 14
obtained bicrystals are characterized by EBSD measurements and compared with theoretical predictions,
∆𝜃 for all the 14 samples are less than 1ºand ∆𝜓 are all around 60º, further proving that all grain
boundaries are <111>/60o twin boundaries.

96
Figure 6.5 (a) Sketch of a Cu foil in coordinates of specimen and an arbitrary Cu gain in the foil in
crystal coordinates. (b) Sketch of Euler angles of a Cu grain with anticlockwise rotations are defined to
be positive. (c) Sketches of three decomposed operations through Euler angles.

6.3.2 Alignment of graphene islands on twinned Cu substrates

To investigate the orientations of graphene islands on Cu surface, the grain texture of the Cu
surfaces should be figure out first. As mentioned above, the orientation of any grain can be confirmed
by three rotational operations of Euler angles. A rotational operation with an angle of 𝜃 around a
rotational axis U= (x, y, z) where 𝑥 2 + 𝑦 2 + 𝑧 2 = 1 can be implemented by multiplying a rotate matrix
G on the coordinates of all Cu atoms:

cos 𝜃 + 𝑥 2 (1 − cos 𝜃) 𝑥𝑦(1 − cos 𝜃) − 𝑧 sin 𝜃 𝑥𝑧(1 − cos 𝜃) + 𝑦 sin 𝜃


G=[𝑥𝑦(1 − cos 𝜃) + 𝑧 sin 𝜃 cos 𝜃 + 𝑦 2 (1 − cos 𝜃) 𝑦𝑧(1 − cos 𝜃) − 𝑥 sin 𝜃] (6.1)
𝑥𝑧(1 − cos 𝜃) − 𝑦 sin 𝜃 𝑦𝑧(1 − cos 𝜃) + 𝑥 sin 𝜃 cos 𝜃 + 𝑧 2 (1 − cos 𝜃)

After knowing the orientation of the Cu grain, the surface index {hkl} of this grain can be
obtained by calculating the normal direction of the surface plane. Here, we chose three arbitrary points
of the surface plane, namely A (X1, Y1, Z1), B (X2, Y2, Z2), and C (X3, Y3, Z3). The normal direction
⃑⃑⃑⃑⃑⃑ × 𝐵𝐶
<hkl> can be obtained by 𝐴𝐵 ⃑⃑⃑⃑⃑⃑ , then we get:

(𝑌2 − 𝑌1 )(𝑍3 − 𝑍2 )
ℎ= (6.2)
(𝑍2 − 𝑍1 )(𝑌3 − 𝑌2 )

(𝑍2 − 𝑍1 )(𝑋3 − 𝑋2 )
𝑘= (6.3)
(𝑋2 − 𝑋1 )(𝑍3 − 𝑍2 )

(𝑋2 − 𝑋1 )(𝑌3 − 𝑌2 )
𝑙= (6.4)
(𝑌2 − 𝑌1 )(𝑋3 − 𝑋2 )

97
Because Cu steps play a critical role in determining the orientation of graphene grown on the
surface, as proved in Chapter 4. The step direction of an arbitrary Cu surface should also be explored.
Due to that the close-packed <110> steps or segments are energetically favored and appears more
frequently, here in this Chapter to make it sample, only <110> dominated steps are considered.

FCC Cu crystal has totally six <110> close-packed direction, from the two examples of Cu
surfaces in Figure 6.6 (a-b), we can see the lengths of the projections of the six <110> directions on a
surface are dependent on the angle between their direction and the surface plane, and the <110>
direction which has the smallest angle with respect to the surface plane acts as step segment on the
surface, as marked by blue arrows. The angles between six <110> directions with the two surfaces in
Figure 6.6 (a-b) are listed in Table 6.1, when the angle is 0º, the <110> direction is exactly the step
edge direction of surface, e.g., the step edges of Cu(611) is along Cu[01̅1]. Figure 6.6 (c) illustrates a
projection of a <110> axis (denoted by ⃑⃑⃑⃑⃑⃑
𝐵𝐴) on a Cu plane, which can be obtained by calculating the
projected points of A (a1, b1, c1) and B (a2, b2, c2).

The plane equation of the Cu foil surface is:

hx + ky + lz + d = 0 (6.5)

where 𝑑 = −(ℎ ∙ 𝑋4 + 𝑘 ∙ 𝑌4 + 𝑙 ∙ 𝑍4 ), and (X4, Y4, Z4) is a known point of the surface plan. The
projected point (Pa, Pb, Pc) can be calculated as followed:

𝑃𝑎 = 𝑎 + 𝑡 ∙ ℎ (6.6)

𝑃𝑏 = 𝑏 + 𝑡 ∙ 𝑘 (6.7)

𝑃𝑐 = 𝑐 + 𝑡 ∙ 𝑙 (6.8)

ℎ∙𝑎+𝑘∙𝑏+𝑙∙𝑐+𝑑
𝑡=− (6.9)
ℎ2 + 𝑘 2 + 𝑙 2

⃑⃑⃑⃑⃑⃑ can be obtained by:


Therefore, the projected direction of 𝐵𝐴

⃑⃑⃑⃑⃑⃑⃑⃑⃑
𝐵′ 𝐴′ = 𝐴′ (𝑃𝑎1 , 𝑃𝑏1 , 𝑃𝑐1 ) − 𝐵′ (𝑃𝑎2 , 𝑃𝑏2 , 𝑃𝑐2 ) (6.10)

And the angle (α) between a <110> axis and the Cu foil surface can be obtained by:

⃑⃑⃑⃑⃑⃑ ∙ ⃑⃑⃑⃑⃑⃑⃑⃑⃑
𝐵𝐴 𝐵′ 𝐴′
cos 𝛼 = (6.11)
⃑⃑⃑⃑⃑⃑||𝐵
|𝐵𝐴 ⃑⃑⃑⃑⃑⃑⃑⃑⃑
′ 𝐴′ |

98
Figure 6.6 (a-b) Illustrations of the relationship between 6 <110> directions and Cu steps on Cu(542)
and Cu(611) surfaces. (c) Sketch of a <110> direction projected on the surface of Cu foil.

Table 6.1 The angles (°) between <110> direction and Cu surface.

[110] [1̅10] [101] [101̅] [011] [01̅1]

(5 4 2) 39.2 12.2 47.5 18.4 71.5 6.05


(6 1 1) 53.4 35.00 53.4 35.00 13.3 0.00

After the direction of <110> segment of the step is confirmed, graphene orientation on both
sides of a twined Cu substrate can be roughly known by applying the rule that graphene ZZ segment
aligning along <110> segment of the step, which means that the misalignment angle between graphene
islands on two sides of a grain boundary can be obtained by calculating the angle between <110>
segments of the surfaces of grain A and B.

As introduced above in the modeling part, we can use two degrees of freedom to decide the
orientation of a <111>/60o Cu twin, 𝜃 and 𝜓. Because of the 6-fold symmetric of (111) plane of FCC,
𝜓 can be restricted into 60o. Therefore, by changing 𝜓 from -30º to 30º and 𝜃 from 0º to 90º, all
possible orientations of the <111>/60o Cu twin can be reached. The map in Figure 6.7 shows the
misorientation angles of graphene islands on the two sides of the <111>/60o Cu twin along all
orientations, and the misorientation angle values are denoted by colors. It can be seen that there is a
large possibility (56.92% for misorientation angle < 0.01 and 56.66% for misorientation angle < 0.001)
to get unidirectionally aligned graphene islands on <111>/60o Cu twins.

99
Figure 6.7 Map of the misangles of graphene islands on two sides of <111>/60o twined Cu foils. θ is
the angle between the normal direction of the (111) boundary plane with that of the Cu foil surface, 𝜓
is the rotate angle of the twin around their <111> coaxis.

Table 6.2 lists the experimental and theoretical results of 14 samples, where ∆𝐸𝑥𝑝 is the
graphene misorientation angles obtained by experimental measurements, and ∆𝑀𝑎𝑝 is that obtained by
theoretical calculations and marked in the map. The perfect agreement between experimental
measurements and theoretical calculations strongly supports our theory.

100
Table 6.2 Experimentally obtained misalignment angles (∆𝐸𝑥𝑝 ) and theoretically predicted ones (∆𝑀𝑎𝑝 )
of graphene islands on 14 twinned Cu foils.

Measured Euler Angles Assessment of <1 1 1>/60˚ twin boundary Misalignment Analysis
No.
𝜑1 (˚) 𝜙 (˚) 𝜑2 (˚) θ (˚) 𝜓 (˚) 𝛥θ (˚) 𝛥𝜓 (˚) ∆𝐸𝑥𝑝 (˚) ∆𝑀𝑎𝑝 (˚)
1-A 221.00 12.20 44.70 66.94 -0.07
0.02 59.79 0±0.5 0.00
1-B 283.30 52.90 41.20 66.96 -59.86
2-A 227.90 12.80 37.70 57.32 15.14
1.00 59.66 27.2 ± 0.5 27.21
2-B 15.50 37.00 37.50 58.32 -44.52
3-A 136.10 39.10 19.20 47.77 -80.08
0.51 59.17 0±0.5 0.00
3-B 284.20 9.60 90.00 48.28 -20.91
4-A 233.10 46.70 46.80 67.83 -81.76
0.36 59.71 0±0.5 0.08
4-B 144.40 14.50 75.70 67.47 -22.05
5-A 139.10 41.20 14.50 80.71 5.11
0.73 60.37 11.0±0.5 11.00
5-B 222.00 35.30 88.10 81.44 -55.26
6-A 141.10 39.30 12.50 80.28 2.82
0.13 59.72 11.0±0.5 11.28
6-B 224.90 35.50 84.80 80.41 -56.90
7-A 262.1 28.4 58.8 27.75 -14.10
0.62 61.23 1±0.5 0.00
7-B 354.2 39.9 13.6 27.13 47.13
8-A 64.5 36.6 53.7 67.05 -39.79
0.68 61.00 25±0.5 26.42
8-B 312.8 22.6 15.6 67.73 21.21
9-A 123.1 41.4 46.2 63.62 47.57
0.01 60.19 27±0.5 27.47
9-B 236.3 14 81 63.63 -12.62
10-A 71.3 39.8 80.8 82.08 -31.61
0.35 60.24 10±0.5 11.03
10-B 349.3 37.6 6 81.73 28.63
11-A 126.9 31.4 27.4 84.99 50.90
0.21 60.19 3±0.5 2.16
11-B 281 41 59.2 85.21 -9.29
12-A 51.9 44.3 87.1 33.25 -58.66
0.61 60.13 0±0.5 0.00
12-B 142 20.9 42.7 33.86 1.47
13-A 236.4 44.9 10.7 42.8 -59.12
0.15 59.86 0±0.5 0.00
13-B 336.3 12.1 42.6 42.65 0.74
14-A 261.8 36.6 45.4 18.14 -0.77
0.1 59.02 0±0.5 0.00
14-B 335.5 47 23.9 18.04 58.25
Note: i-A and i-B (i=1, 2, …, 14) represent the Cu surface on both sides of the twin boundary, measured
Euler angles are obtained from EBSD measurements.

101
Besides, it is worth noting that the misorientation angles are exactly 0º along the dashed line
of the map, indicating perfect alignment of graphene islands on the Cu foils with those twin grain
boundaries. Figure 6.8(a) shows three atomic configurations with <111>/60o Cu twins along the dashed
line, we can see that step edges of the surfaces are all along the same Cu<110> direction. Figure 6.8(b)
shows the SEM images and atomic models of graphene on Cu twin surfaces of sample #12-14 that are
located on the dashed line. Similarly to those grown on the (116)/(111) Cu twin, graphene islands on
the two sides of the twin boundaries are highly orientated with one pair of their ZZ edges parallel with
the grain boundary, even though their shapes are either stranded hexagons or stretched hexagons along
the step edge direction. The consistency between the four samples with misaligned graphene islands
and our theoretical predictions shown in Figure 6.9 also validates our theory.

Figure 6.8 (a) Three atomic configurations of <111>/60o twin Cu foils along the point line of the map.
(b) SEM images and atomic models of well-aligned graphene islands on surface of <111>/60o Cu twin
located in the point line of the map.

102
Figure 6.9 SEM images and atomic models of misaligned graphene islands on surface of <111>/60o
Cu twin.

6.3.3 Application to other 2D materials

Since well-aligned graphene islands and single crystalline graphene films can be successfully
synthesized on the twinned Cu foils, then what is the alignment of other 2D materials with lower
symmetries on such Cu foils?

Supposing the orientation of other 2D materials is also determined by <110> segment of the
step, which has been verified for hBN and TMDCs in Chapter 4, we extend the same theory proposed
above to calculate the misorientation angle maps of C3V and C4V 2D materials on <111>/60o twinned
Cu foils (Figure 6.10). It should be noted that for C3V 2D materials, the orientation of the islands depends
on not only the step edge direction but also the stair direction (uphill or downhill). As shown in Figure
6.10(a-b), although the step directions are the same, the two C3V islands shows parallel orientation when
the stair directions are also the same but antiparallel orientations when the stair directions are opposite.
Here twinned foils with antiparallel C3V islands grown on it are marked by red in the map of Figure
6.10(c), which reduces the possibility of unidirectionally aligned C3V islands growth. For C4V 2D
materials, its orientation only depends on step direction because antiparallel C4V islands actually have
the same orientation, and the map in Figure 6.10(d) shows that well-aligned C4V islands can be obtained

103
on most of twinned FCC foils. Therefore, well-tuned twinned FCC foils open a door to grow single
crystalline 2D materials on polycrystalline substrates without nucleation control.

Figure 6.10 (a, b) Sketches of C3V 2D materials aligning on <111>/60o twinned Cu foils with parallel
prominent <110> steps but different stair directions. (c) Map of the misangles of C3V 2D materials on
two side of <111>/60o twin foils. (d) Map of the misangles of C4V 2D materials on two side of <111>/60o
twin foils.

6.3.4 Anticipation of graphene alignment on twined hexagonal close-packed (HCP) metal


substrate

Last but not least, the possibility of synthesizing unidirectionally aligned 2D materials on
twined HCP metal substrates was also explored, because HCP metals are also frequently employed as
substrates for the CVD synthesis of 2D materials. Both theoretical and experimental studies have proved
that the <0001>/30 º twin grain boundary shows the lowest formation energy and appears most
frequently in the recrystallized HCP metals.358 Here, we choose Co as an example because the lattice
constant of Co{0001} planes is closed to that of graphene (2.50Å for Co{0001} and 2.46 Å for
graphene), and suppose graphene ZZ edge prefers to align along the close-packed segment of the step
edges on Co foil surface.

Figure 6.11(a) shows the atomic model of a Co twin with the <0001>/30ºtwin boundary, and
we can see the two grains are coaxial around the <0001> direction and grain B rotates 30º around the
coaxis with respect to grain A. To figure out the step edge direction of an arbitrary Co surface, the

104
atomic arrangement of a Co crystal is sketched in Figure 6.11(b). In the {0001} plane of a HCP structure,
there are three close-packed directions, labelled by ① ~ ③. Out of the {0001} plane, there is no straight
close-packed atomic chains, instead there are six directions along which close-packed atoms are
connected in a ZZ pattern, as labeled by ④ ~ ⑨. We can see that when the step edges of a Co surface
are derived from the projection of a ZZ chain, their longest segment is composed by only two atoms.
Figure 6.11(c) shows three representative Co surface, where red point line marks a step derived from
two-atom chains and brown point line marks a step derived from the straight atomic chains, and angle
between the straight atomic chains and the surface is denoted by γ. When γ is smaller than 19.1º, the
step edges derived from straight atomic chains always show longer segments than that derived from ZZ
atomic chains, suggesting their superiority in determining the orientations of graphene islands on such
Co surfaces. When γ is larger than 19.1º, the steps are then dominated by the projections of ZZ atomic
chains. Based on this rule, the step direction of an arbitrary Co surface and further the alignment of
graphene islands on a twined Co foil can be obtained.

Figure 6.11 (a) Atomic model of HCP twin. (b) Sketch of the different directions in HCP crystal and
their coordinates in orthogonal system. (c) Three examples showing steps of different surfaces. γ is the
smallest angle between close-packed direction and the surface.

The map in Figure 6.12(a) shows the misorientation angles of graphene islands on the two sides
of <0001>/30ºtwin grain boundaries of Co foils, and two atomic models marked in the map are also

105
shown in Figure 6.12(b). We can see that unidirectionally aligned graphene islands can only be
synthesized on a very small portion of Co foils with twin grain boundaries. Therefore, HCP twinned
foils are not good substrates for the growth of unidirectionally aligned graphene islands.

Figure 6.12 (a) Map of the misangles of graphene islands on two side of Co twins with <0001>/30º
twin boundary. (b) Atomic models of graphene islands on two sides of HCP twins with different
surfaces.

6.4 Summary

In this Chapter, we propose a new method of synthesizing WSSC 2D materials on


polycrystalline TM substrates with twinned grain boundaries. Experimentally, by inducing a
microhardness indent in a bend Cu foil and annealing in a H2 atmosphere, a Cu foil with only <111>/60º
twin boundaries is successfully obtained, and we found that graphene grown on such twined Cu foils
are unidirectionally aligned in most areas, which is different from the conventional opinion that well-
aligned graphene islands can only be grown on single-crystalline Cu foils. By seamless coalescence of
such unidirectionally aligned graphene islands, WSSC graphene films were successfully obtained. To

106
reveal the alignment of graphene on twinned Cu foils, a general theoretical model is established. It is
found that there is a very large portion of all possible twinned Cu foils that can template the growth of
unidirectionally aligned graphene island. Our theoretical predictions are in perfect agreement with all
samples of our experiments, which strongly validates our theoretical model. Moreover, we extend our
theoretical models to investigate the alignment of C3V and C4V 2D materials on the twinned Cu foils and
the alignment of graphene on twinned HCP TM substrates. We predict that twinned Cu foils are
promising substrates for the growth of unidirectionally aligned C3V and C4V 2D materials and thus their
WSSCs, while it is difficult to epitaxially grown unidirectionally aligned graphene islands on twinned
HCP TM substrates.

107
Chapter 7 Alignment of hBN islands on low symmetric Cu
substrates
7.1 Introduction

Due to its large bandgap, high chemical and thermal stabilities, hBN is another promising 2D
material in a wide range of applications.91,92,95,99 As the most promising method to synthesize WSSC
2D materials at a relatively low cost, seamless coalescence of large numbers of unidirectionally aligned
2D islands on a proper substrate has drawn great attention in past several years. Similar with the
synthesis of graphene, most of experimental and theoretical studies mainly focus on the epitaxy of hBN
on high-symmetric TM substrates at the early stage. It was found that triangular hBN islands generally
show two antiparallel orientations on the Cu(111) and Cu(110) surfaces,80,113,122 and present four
different orientations on the Cu(100) surface.113,122,359 As revealed in Chapter 2, all the low-index Cu
substrates have a central inversion symmetry, while the triangular hBN lattice has no central inversion
symmetry, and as a result, hBN islands show at least two equivalent but different orientations on these
substrates. Obviously, growing WSSC hBN on low-index Cu substrates by the seamless coalescence
method is impossible.

In the year of 2019 and 2020, the synthesis of WSSC hBN has been realized on single crystalline
stepped Cu(111)360 and vicinal Cu(110)84 substrates, which significantly stimulated the experimental
efforts of growing hBN single crystals on various low symmetric Cu surfaces. In addition, the
fabrication of various high-index Cu surfaces facilities the researches of hBN growth on these low
symmetric surface.85,361 However, until now, an in-depth theoretical understanding on the mechanism
of hBN epitaxy on high-index Cu surfaces is still lacking. In this chapter, we present a systematic
theoretical study on the alignment of hBN on various high index Cu substrates.

7.2 Modeling and simulation methods

7.2.1 Modeling

The same as that in Chapter 5, the stability of the interface of a hBN edge attaching to a Cu step
edge is evaluated by its formation energy (𝐸𝑓 ). Figure 7.1(a) shows the atomic model for calculating
𝐸𝑓 . Periodic boundary condition is applied to X, Y and Z directions of the model. To avoid the imaginary
interaction of the Cu slab along Z direction, a vacuum layer with a thickness larger than 12 Å is set in
this direction. To minimize the strain along X direction induced by lattice mismatch between hBN and
Cu, the difference in the lengths of the hBN nanoribbon and the Cu slab is limited to 3.0% and, the
lattice constant of hBN is kept constant to avoid the strain in hBN nanoribbon. To minimize the

108
interaction of the other edge of the hBN nanoribbon with the terrace of the Cu substrate, the other edge
of the hBN nanoribbon is passivated by hydrogen. In addition, the length of unit cell of the model along
Y direction is chosen to be ~20 Å for all structures, so that the effect of step edges from the imaginary
neighboring cells can be neglected.

The formation energy of the interface with a hBN edge attaching to a Cu step edge is defined
as: 𝐸𝑓 = 𝜀𝑒𝑑𝑔𝑒 − 𝐸𝐵 , where 𝜀edge is the formation energy of a pristine hBN edge and 𝐸𝐵 is the binding
energy of the hBN edge attaching to the Cu step edge.

Figure 7.1 (a) Atomic model for calculating formation energy of a hBN edge attaching to a Cu step
edge of a low symmetric Cu substrate. (b) Atomic model for calculating edge energy of a pristine hBN
edge. (c) Edge energies of pristine hBN edges under different chemical potential of nitrogen, 𝜇𝑁 .

Due to the binary composition of hBN, the two edges of a hBN nanoribbon are not equivalent
except that along AC direction. Therefore, the formation energy of a hBN edge along an arbitrary
direction cannot be calculated directly by using hBN nanoribbons. Because hBN has a C3V symmetry,
an equilateral triangular hBN cluster has three equivalent edges and therefore, the edge formation
energy of a hBN edge along any directions can be calculated from two triangle hBN clusters with same
edge structures but different sizes, as shown in Figure 7.1 (b), and edge energy can be calculated by:

𝜀𝑒𝑑𝑔𝑒 = (𝐸2 − 𝐸1 − ∆𝑁𝐵 ∙ 𝜇𝐵 − ∆𝑁𝑁 ∙ 𝜇𝑁 )/3/(l2 − 𝑙1 ) (7.1)

109
where 𝐸2 and 𝐸1 represents the total energies of triangle I and II demonstrated in Figure 7.1(b),
respectively. ∆𝑁𝐵 and ∆𝑁𝑁 are difference in the numbers of B and N atoms between triangle I and
triangle II demonstrated in Figure 7.1(b), respectively. 𝜇𝐵 and 𝜇𝑁 are the chemical potentials of B and
N atoms, and satisfy 𝜇𝐵 + 𝜇𝑁 = 𝜖𝐵𝑁 , where 𝜖𝐵𝑁 is energy of a BN-pair in a perfect hBN layer and
equals to −17.74 eV. 𝜇𝑁 equals to -8.31 eV when a N2 molecule is used as the reference and equals to
-10.84 eV if α-boron is chosen as the reference (𝜇𝑁 = 𝜖𝐵𝑁 − 𝜇𝐵 = −10.84 eV). Figure 7.1(c) shows
the formation energy profile of hBN edges as a function of their orientations under different references,
where the orientation of ZZN edge is set to be 0o and thus, 30o and 60o represent AC and ZZB edges,
respectively.

The binding energy of a hBN edge attached to a Cu step edge is defined as:

𝐸𝐵 = (𝐸𝐶𝑢 + 𝐸ℎ𝐵𝑁 − 𝐸𝑇𝑜𝑡 − 𝐸𝑣𝑑𝑊 )/𝑙 (7.2)

where 𝐸Cu , 𝐸hBN , and 𝐸Tot are energies of the Cu substrate, the hBN nanoribbon, and the system of the
hBN nanoribbon on the Cu substrate. Because 𝐸𝑇𝑜𝑡 includes the vdW interaction between the hBN
nanoribbon wall and the Cu substrate (𝐸𝑣𝑑𝑊 ), 𝐸𝑣𝑑𝑊 should be deduced to calculate the binding energy
of the hBN edge to the Cu step edge. l is the unit length of the hBN nanoribbon (or Cu step edge).

7.2.2 Simulation methods

The DFT-D3 method was employed to perform all the calculations by using VASP.346,347 The
GGA approximation was used to describe the exchange-correlation effects,348 and the PAW method
was used to treat the interaction between valence electrons and ion cores.349 All the structures were
optimized by using the conjugate gradient method until the force on each atom less than 0.01 eV/Å, and
the energy difference is converged to 10-4 eV for the self-consistent iterations.

7.3 Results and discussions

Similar to of graphene, we classified hBN edges into straight edges and tilted edges. Straight
hBN edges include ZZN, ZZB and AC edges, as demonstrated in Figure 7.2(a). Tilted hBN edges can
be viewed to be constructed by either ZZ (ZZN or ZZB) segments with AC kinks or AC segments with
ZZ (ZZN or ZZB) kinks. Figure 7.2(b) shows an example of a tilted hBN edge consisting of ZZN
segments and AC kinks.

110
Figure 7.2 Various interfaces between a hBN edge and a Cu step edge. (a) Interfaces of three straight
hBN edge versus a straight step edge. (b) Interface of a tilted hBN edge versus a staight step edge. (c)
Interfaces of three straight hBN edges versus a tilted step edge. (d) Interface of a tilted hBN edge versus
a tilted step edge.

According to the structures of Cu step edges and hBN edges, there are four possible types of
interfaces between a Cu step edge and a hBN edge: (i) a straight hBN edge attached to a straight Cu
step edge (Figure 7.2 (a)); (ii) a tilted hBN edge attached to a straight Cu step edge (Figure 7.2(b)); (iii)
a straight hBN edge attached to a tilted Cu step edge (Figure 7.2 (c)); and (iv) a tilted hBN edge attached
to a tilted Cu step edge (Figure 7.2 (d)). On a Cu substrate with straight step edges, interface of type (i)
is obviously more stable than that of type (ii). For instance, our DFT calculations have shown that the
formation energy of the interface between a hBN ZZN edge and a Cu<211> step edge is lowest as
compared to those of other hBN edges attached to the Cu<211> step on Cu(110) surface.84 However,
there is only a very limited number of type (i) interfaces due to limit number of straight hBN edges and
Cu step edges. In practice, Cu substrates with tilted step edges are more common. From Figure 7.2(c),
we can know that the interfaces of type (iii) are also not stable, because a large proportion of Cu atoms
of the Cu step edge cannot be well passivated by straight hBN edges. For interfaces of type (iv), their
stabilities depend on the structural matching between tilted hBN edges and tilted Cu step edges. As
already introduced in Chapter 5, different kink heights of tilted hBN edges and tilted Cu step edges can
induce a misorientation angle, which leads to the rotation of hBN islands along tilted Cu step edges as
compared to that along straight Cu step edges.

In following parts of this Chapter, we will present a systematic study on the orientations of hBN
islands on Cu{111}-based, Cu{100}-based and Cu{110}-based high-index low symmetric substrates
based on the stabilities of the interfaces between hBN edges and Cu step edges on these Cu substrates.

111
7.3.1 Alignment of hBN islands on Cu{111}-based high-index low symmetric substrates

Figure 7.3(a) shows the six atomic configurations of straight hBN edges attached to straight Cu
step edges on Cu{111}-based low symmetric surfaces, i.e., hBN ZZN, ZZB and AC edges attached to
Cu<110> and Cu<211> steps, respectively. From atomic configurations, it can be seen that both hBN
ZZN and hBN ZZB edges show a perfect structural match with the Cu<110> step, while hBN AC edge
only matches to the Cu<211> step. The formation energies of these interfaces are calculated and shown
in in Figure 7.3(b). It should be noted that, the formation energies are functions of the chemical potential
of nitrogen (or boron) due to the unbalanced stoichiometry of hBN edges, and here the chemical potenial
of nitrogen is chosen as the reference. During CVD growth of hBN, the range of 𝜇𝑁 is usually between
the reference of N2 (𝜇𝑁 =-8.31eV) and the reference of α-boron bulk (𝜇𝑁 =-10.84 eV). From Figure
7.3(b), it can be seen that in a N-rich environment the interfaces with hBN ZZN edge attached to stragith
Cu step edges are energically perfered, while a under B-rich condition the interfaces with hBN ZZB
edge are more stable. It is noted that, on Cu substrates with straight <211> step edges (lower panel of
Figure 7.3(b)), alought the Cu<211> step edge has a good structural mach with hBN AC edge, the
interface with hBN AC edge is most stable only when 𝜇𝑁 is smaller than -9.61 eV, and the interface
with hBN ZZN edge is still ebergetically favored under a N-rich condition.

Figure 7.3 The orientation of hBN islands along straight step edges of Cu{111}-based surfaces. (a)
Configurations of three straight hBN edges attached to two straigth step edges. (b) Formation energies
of straight hBN edges attached to straigth step edges as a function of chemical potential of nitrogen,
𝜇𝑁 . (c) Illustrations of hBN islands along straight step edges at different 𝜇𝑁 ranges corresponding to
(b). Triangles denote the orientations of hBN islands with three ZZN edges and different orientations
are distinguished by colors.

112
In practice, the step edges on Cu{111}-base high-index low symmetric are not always straight
ones, and instead most of them are tilted ones. We have assumed that a tilted step edge can be view to
be constructed by stragith step segments, as shown in Figure 7.3(c). The alignment of hBN islands on
Cu substrates with tilted step edges are dependent on the orientations of the straight step segments and
also on the amibient condition. For a high 𝜇𝑁 , i.e., a N-rich condition (regime I in Figure 7.3(b) and
structure I in Figure 7.3(c)), both Cu<110> and Cu<211> step edges passivated by hBN ZZN edge are
energetically favorable, and concequently the orientation of hBN will change if the Cu step edge
changes from <110> direction to <211> direction and vice versa. If 𝜇𝑁 is in regime II (Figure 7.3(b),
and structure II on Figure 7.3(c)), where hBN ZZN edges perfer to attach to Cu<110> step edges and
hBN AC edges prefers to attach to Cu<211> step edges, two orientations of hBN islands appears along
Cu<211> step edges if a mirror refelection is operated on the structure with hBN AC edge attaching to
a Cu<211> step edge, despite one of them are equvialent with the orientaiton of hBN along Cu<110>
step edges (structure II in Figure 7.3(c)). Further decreasing 𝜇𝑁 to regime III (Figure 7.3(b)), i.e., under
a B-rich condition, Cu<110> step edges prefer to be attached by hBN ZZB edges and Cu<211> step
edges prefer hBN AC edges, and consequently hBN islands regime III show opposite orientations along
Cu<211> step edges, which is similar to that in regime II.

Based on the above predicted orientations of hBN islands on Cu{111}-base high-index low
symmetric substrates with the two straight step edges, the orientations of hBN islands on an arbitary
tilted step edge of Cu{111}-based surfaces are futher explored. As discussed in Figure 7.2(c-d), a tilted
step edge on a substrate perfers to passivated by a structural matched tilted hBN edge. Here a few rules
are assumed to construct the structural matching interface between a hBN edge and a titl Cu step edge:
(i) each kink of the tilted Cu step edge should be docked by one kink of the hBN edge; (ii) there should
also be a one-to-one docking relation for the segments of the hBN edge and those of the Cu step edge;
(iii) if there are more than one orientations of hBN islands that satisfy the above two geometry rules,
such as that a tilted Cu step edge with <211> segments can be passivated by two tilted hBN edges with
the same AC segments but different kinks, the most preferential orientation of the hBN island should
be determined by the interface between kinks with the lowest formation energy.

113
Figure 7.4 The evolution of hBN orientation along step edges of Cu{111}-based surfaces. (a) The
orientation of hBN islands versus step edge direction at 𝜇𝑁 =-8.21~ -9.61 eV. (b) Illustration of hBN
i×<110> i×<211>
islands along SE<211> and SE<110> corresponding to (a). (c) The orientation of hBN islands versus
i×<211>
step edge direction at 𝜇𝑁 =-9.61~ -9.92 eV. (d) Illustration of hBN islands along SE<110>
corresponding to (b). (e) The orientation of hBN islands versus step edge direction at 𝜇𝑁 =-9.92~ -10.84
eV. (f-h) The alignment of hBN islands along a curved step edge with direction varying from -180 to
180°at different 𝜇𝑁 ranges, corresponding to (a), (c) and (e), respectively. Pure colors at margin means
small orientation variations while gradient colors represent large orientation variations.

By using above proposed rules, we predicted the orientations of hBN islands on Cu{111}-based
high-index low symmetric substrates with step edges along different orientations within the chemical

114
potential range of −9.61 𝑒𝑉 < 𝜇𝑁 < −8.31 𝑒𝑉 (Figure 7.4(a-b) and (f)). The same as that in Chapter
5, the exact Cu[110] direction on the Cu substrate is chosen as the reference orientation for both the
hBN island and the Cu step edge. On Cu{111}-based high-index low symmetric substrates with step
i×[110] i×[110]
edges dominated by [110] segments, including SE[211̅] and SE[121] ones (SE represent step edge,

superscripts and subscripts denote the segment and kink structures, which introduced in the section 5.2),
which show orientations ranging from -19.1o to 19.1º, hBN islands show an misorientation angle
ranging from -0.36ºto 0.36ºfor −9.61 𝑒𝑉 < 𝜇𝑁 < −8.31 𝑒𝑉 and the small misorientation angle is
ascribed to the perfect match in kink heights between tilted hBN edges with ZZ segments (0.217 nm)
i×<110>
and SE<211> step edges (0.221 nm) (see Figure 7.4(b)). On Cu{111}-based high-index low symmetric
̅]
i×[211 i×[211̅]
substrates with step edges dominated by [211̅] segments, including SE[110] and SE[101̅] ones, of

which the orientation angle is from 19.1o to 41.9º,hBN islands show misorientation angles ranging from
30º−7.86ºto 30º+7.86ºat −9.61 𝑒𝑉 < 𝜇𝑁 < −8.31 𝑒𝑉 because of the large difference in kink heights
i×<211>
between tilted hBN edges with ZZ segments (0.217 nm) and SE<110> step edges (0.128 nm) (see
Figure 7.4(b)). Figure 7.4(f) demonstrates the orientations of hBN islands on the Cu{111}-based high-
index low symmetric substrates with step edge directions varying from -180ºto 180º,where pure colors
at the margin indicate small orientation variations of hBN islands and gradient colors represent large
orientation variations. Obviously, unidirectionally aligned hBN islands can only be synthesized on
i×<110>
Cu{111}-based high-index low symmetric substrates with SE<211> step edges when −9.61 𝑒𝑉 <
𝜇𝑁 < −8.31 𝑒𝑉.

Similar to above analysis, we also investigated the orientations of hBN islands on Cu{111}-
based high-index low symmetric substrates for regime II (−9.92 𝑒𝑉 < 𝜇𝑁 < −9.61 𝑒𝑉), where tilted
i×<110>
SE<211> step edges are preferred to be passivated by tilted hBN edges dominated by ZZN segments
i×<211>
and tilted SE<110> step edges prefer to be passivated by tilted hBN edges dominated by AC segments.
Due to the small kink height differences for both of these two cases, the orientation change of hBN
islands is negligible (< 0.5o). As shown in Figure 7.4(c), on Cu substrates with step edge orientations
ranging from -29.99o to 29.99ºrespect to a Cu<110> direction, the orientations of hBN islands are
almost the same. Despite antiparallel hBN islands will be formed on Cu substrates with straight
SE0<211>at 30º+60º× 𝑛, where n is an integer, there is only one orientation along SE<110>
i×<211>
step edges
because the mirror symmetry is broken by the existence of kinks (Figure 7.4(d)). Figure 7.4(g)
demonstrates the overall evolution of hBN islands with the change of Cu step edge orientation. Our
theoretical predictions are well consistent with the experimental observation on hBN growth on vicinal
Ir(111) surfaces, where hBN show only one orientation on a vicinal Ir(111) surface with <110>
dominated step edges but two orientations on a vicinal Ir(111) surface with <211> dominated step
edges.362

115
i×<110>
In regime III ( −10.84 𝑒𝑉 < 𝜇𝑁 < −9.92 𝑒𝑉 ), where SE<211> step edges prefer to be
i×<211>
passivated by tilted hBN edges with ZZB segments and SE<110> step edges prefer to be attached by
hBN edges with AC segments, hBN islands on Cu substrates show opposite orientations as compared
to regime II, as shown in Figure 5(e) and (h).

From above discussions, we can see that unidirectionaly aligned hBN agrains can be
i×<110>
synthesized on Cu{111}-based high-index low symmetric substrates with SE<211> step edges, while
antiparrallel hBN islands might be obtained on Cu{111}-based high-index low symmetric substrates
i×<211>
with SE<110> step edges. From Chapter 4, we have shown that hBN islands show antiparallel
orientations on ideal Cu(111) substrates, and therefore lowering the symmetry of the substrate is
promising in realizing the epitaxial growth of hBN.

7.3.2 Alignment of hBN islands on Cu{100}-based low symmetric substrates

Different from Cu{111}-based high-index low symmetric substrates, the straight step edges
that can exist on Cu{100}-based substrate are along Cu<110> and Cu<100> directions, respectively.
We calculated the interface formation energy profiles of straight hBN ZZN, ZZB and AC edges
attaching to the two straight Cu step edges as a function of 𝜇𝑁 , which are shown in Figure 7.5(a). Due
to the structural mismatch between the hBN AC edge and both of the two straight Cu step edges, the
interfaces of hBN AC edge attaching to the straight Cu step edges are not preferred in all 𝜇𝑁 ranges. In
regime I (𝜇𝑁 > -10.27 eV), both of the two stragith Cu step edges prefer to be passivated by the hBN
ZZN edge, and consequently hBN islands attached to the two straight steps show a 45ºmisorientation
angle (Figure 7.5(b)). In regime II (−10.63 𝑒𝑉 < 𝜇𝑁 < −10.27 𝑒𝑉), straight Cu<110> and Cu<100>
step edges prefer to be passivated by hBN ZZN and ZZB edges, respectively, and hBN islands attached
to the two stragith step edges has a included angle of 15o (Structure II in Figure 7.5(b)).

116
Figure 7.5 Orientation of hBN on Cu{100}-based low symmetric surfaces. (a) Formation energies of
straight hBN edges attached to two straigth step edges as a function of chemical potential of nitrogen, 𝜇𝑁
and the corresponding stable interfaces. (b) Alignment of hBN islands along a curved step edge
constructed by straight step edges at different 𝜇𝑁 range. (c) Algiment of hBN orientation along a curved
step edge varying from -180 to 180°at 𝜇𝑁 =-8.31~ -10.27 eV. (d) The orientation of hBN islands versus
step edge orientation at 𝜇𝑁 =-8.31~ -10.27 eV.

We further calculated the orientations of hBN islands attached to tilted step edges of Cu{100}-
based high-index low symmetric substraes. Figures 7.5(c-d) show the overall orientation of hBN islands
with the change of Cu step edge orientaion for regime I, where Cu[110] direction is still chosen as the
orientation reference. Due to the non-negligible step heigh differences (0.217 nm vs 0.256 nm for a
i×<110>
tilted hBN ZZ edge attached to a SE<100> , and 0.217 nm vs 0.181 nm for a tilted hBN ZZ edge
i×<100>
attached to a SE<110> ), the misorientation angles of hBN islands are quite large (~ ±4o).

From above analysis, it can be seen that unidirectionally aligned hBN islands can only be
realized on ideal Cu{100}-based high-index low symmetric substraes with unidirectionall and parallel
step edges, i.e., there is a very small tolerance of the orientation of hBN islands on the change of the
orientation of Cu step edges on such substrates.

117
7.3.3 Alignment of hBN islands on Cu{110}-based low symmetric surfaces

Figure 7.6 Orientation of hBN on Cu{110}-based surfaces. (a) Formation energies of straight hBN
edges attached to three straigth step edges as a function of chemical potential of nitrogen, 𝜇𝑁 and the
corresponding stable interfaces. (b) Alignment of hBN islands along a curved step edge varying from -
180 to 180°at 𝜇𝑁 =-8.31~ -10.50 eV. (c) The orientation of hBN islands versus step edge orientation
at 𝜇𝑁 =-8.31~ -10.50 eV.

Following the above analysis, we further investigated the alignment of hBN islands on
Cu{110}-based high-index low symmetric substrates. The Cu{110}-based substrates have three types
of straight step edges, namely Cu<100>, Cu<110> and Cu<211> step edges. Similar to the case of
Cu{100}-based substrates, the interfaces with straight Cu step edges passivated by hBN AC edges are
not energetically favored (Figure 7.6(a)). Figure 7.6(a) also shows that all of the three stragith step edges
prefer to be passivated by hBN ZZN edges in a dominant range of 𝜇𝑁 (𝜇𝑁 > 10.50 𝑒𝑉). Figures 7.6(b-
c) show the predicted orienations of hBN islands for 𝜇𝑁 > 10.50 𝑒𝑉 on all possible Cu{110}-based
i×<110>
high-index low symmetric substrates, where there are four types of tilted step edges (SE<211> ,
i×<211> i×<211> i×<100>
SE<110> , SE<100> and SE<211> ) constructed by the three straight step edges. Only on Cu
̅]
i×[112 ̅]
i×[112
substrate with SE[110] step edges orientated from 35.26º~54.74ºand SE[001̅] step edges orientated

from 54.74º~70.53º, the orientations of hBN islands show a very small variation with respect to the

118
orientation change of Cu step edges, suggesting that unidirectionally aligned hBN islands can be
synthesized on Cu{110}-based substrates with step edges dominated by <211> segments. In 2019,
WSSC hBN films have been synthesized on a single crystalline vicinal Cu(110) substrate with <112>
dominated tilted step edges by the coalescence of unidirectionally aligned hBN islands on the
substrate.84,309

7.4 Experimental evidence

Up to now, extensive experimental studies have been carried out on the CVD synthesis of hBN
on various Cu substrates.75,80,84 By extensvely searching published literatures, we found eight
experiemts on the measurements of hBN islands on high-index low symmetric Cu substrate. In this part,
we will compare these experimental observations with our theoretical predictions shown above.

Wang et al. measured the alignment of hBN islands on a Cu{111}-based low symmetric
substrate, i.e., Cu(313).361 As shown in Figure 7.7(a), Cu(313) surface has striaght step edges along
Cu<110> direction, and it was found that hBN islands are parallel to the straight Cu<110> step edges,
which is consistent with our theortical prediction. The alignment of hBN islands on three Cu{100}-
based low symmetric substrates were also observed, as reported in Ref. [80] and Ref. [361] (Figures
7.7(b-d)). All of the three substrates, namely Cu(102) (Figure 7.7(b)), Cu(103) (Figure 7.7(c)), and Cu(5
0 11) (Figure 7.7(d)), show same type of step edges, which are straight Cu<100> ones. Both EBSD and
SEM characterizations have shown that one of the three ZZ edges of the triangular hBN islands grown
on these substrates is parallel to the straight Cu<110> step edges and thus all hBN islands are
unidirectionally aligned. The alignment of hBN islands on four Cu{110}-based low symmetric
substrates were also measured, in Ref. [84] and Ref. [361]. It was observed that unidrectionally aligned
hBn islands can be synthesized on the vicinal Cu(110) surfaces with straight Cu<211> step edges, where
one of the three ZZN edges of hBN islands was found to be attached to the Cu<211> step edge (Figure
7.7(e)). Figures 7.7(f-h) show the alignment of hBN islands on on Cu(14 4 15), Cu(14 1 15) and Cu(8
i×<110>
2 9) surfaces, respectively. All of these three substrates have SE<211> type titled step edges.
Theoretically, we have predicted that the misorientation angle between a ZZ direction of hBN and the
Cu<110> segments of the step edges on the three substrates are 4.02º,14.95ºand 7.91º, which are well
consistent with the experiemtnally measured values, i.e., 6±2º,15±2ºand 10±2º, respectively.

119
Figure 7.7 Comparison between experimental and theorical results. (a-c) hBN growth on Cu{100}-
based low symmetric surfaces, Cu(102), Cu(103), Cu(5 0 11).80,361 (d-g) hBN growth on Cu{110}-based
low symmetric surfaces, Cu(110), Cu(14 4 15), Cu(14 1 15), Cu(8 2 9).84,361 (h) hBN growth on a
Cu{111}-based low symmetric surface, Cu(313).361 The upper panels show experimental observations
and lower panels show the corresponding atomic models.

The perfect agreement between our theoretical predictions and experinmental obervations,
especially those about hBN grown on Cu substrates with titl step edges, strongly validates our theory
on the alignment of hBN grown on high-index low symmetri Cu substrates. Moreover, it should be
noted that our theory can be extended to the alignment of other 3-fold symmetric 2D materials grown
on an arbitrary high-index surface of various FCC TMs.

7.5 Summary

In this Chapter, we present a general theoretical model on the alignment of hBN on Cu high-
index low symmetric substrates, which have been found to be promising in synthesizing WSSC 2D

120
materials by experiments recently. We found that the alignment of hBN islands on Cu high-index low
symmetric substrates is highly determined by the structures of step edges of the substrate. Ideally,
unidirectionally aligned hBN islands can be grown on all high-index low symmetric Cu substrates with
unidirectionally aligned step edges. However, the orientations of step edges on Cu substrates always
variate more or less in practice. Therefore, the orientation variation of hBN islands with the orientation
change of step edges on high-index low symmetric Cu substrates is of critical importance on the
alignment of hBN islands. Based on our proposed theoretical model, we systematically revealed the
alignment of hBN islands on all types of high-index low symmetric Cu substrates, as summarized in
Table 7.1. Our proposed theoretical model is validated by its perfect agreement with existing
experimental observations. Moreover, our theoretical model can be extended to the growth of other 3-
fold symmetric 2D materials on an arbitrary high-index surface of FCC TMs. We believe this study will
greatly promote the controllable synthesis of various 3-fold symmetric 2D materials.

Table 7.1 Summary of the alignment of hBN on various high-index Cu surfaces.

Kink height
High- Tilted step edges Range of
μN range Edge difference (Å)
index misalignment
(eV) alignment Kink ZZN(B) AC
Surfaces Edge types angle(º)
height(Å) (2.17Å) (1.26Å)
i×<110> 2.21 0.04 -- -0.36/+0.36
(−8.31, ZZN/<110> {111}SE<211>
−9.61) ZZN/<211> {111}SE<110>
i×<211> 1.28 -0.90 -- -7.86/+7.86
i×<110> 2.21 0.04 --
Cu{111} (−9.61, ZZN/<110> {111}SE<211>
-0.36/+0.36
serial −9.92) AC/<211> {111}SE<110>
i×<211> 1.28 -- 0.02
i×<110> 2.21 0.04 --
(−9.92, ZZB/<110> {111}SE<211>
-0.36/+0.36
−10.84) AC/<211> {111}SE<110>
i×<211> 1.28 -- 0.02
i×<110> 1.81 -0.37 -- -4.21/+4.21
Cu{100} (−8.31, ZZN/<110> {100}SE<100>
serial −10.27) ZZN/<100> {100}SE<110>
i×<100> 2.56 0.38 -- -3.91/+3.91
i×<110> 3.68 1.51 -- -14.95/+14.95
{110}SE<211>
ZZN/<110> i×<211>
Cu{110} (−8.31, {110}SE<110> 2.09 -0.09 -- 0/+0.85
ZZN/<211>
serial −10.50) i×<211>
{110}SE<100> 2.09 -0.09 -- 0/+0.68
ZZN/<100>
i×<100> 2.69 0.52 -- -3.0/+3.00
{110}SE<211>

121
Chapter 8 The effect of surface roughness on the alignment of
grown 2D materials
8.1 Introduction

In practice, annealed Cu foils can not be ultra-flat at atomic scale on their whole surfaces, and
the surfaces of Cu foils show a certain degree of roughness.363-365 As examplified in Figure 8.1(a), the
Cu surface show a roughness of tens of nanometers at a micro-scale. Due to the existence of nano-
protrusions and dents on the Cu surface, step edges on the Cu are no longer along straight lines and
parallel with each other, instead, the orientations of step edges become site-dependent. Figure 8.1(b)
shows an SEM image of a Cu surface after graphene growth. It can be clearly seen that the step edges
on the Cu surface are meandering and even forming circles. Consequently, the orientations of 2D
materials grown on realistic high-index low symmetric depends on not only the crystallographic index
of the substrate substrate but also on the substrate surface roughness. Therefere, investigating the effect
of surface roughness of substrates on the alignment grown 2D materials is also highly demanded and
important for the controllable synthesis of 2D materials.

In this chapter, a theoretical model is proposed to describe the surface roughness of an arbitrary
substrate and, based on this model we further explored effect of surface roughness on the direction of
step edges of the substrate and the alignment of 2D materials on the substrate.

Figure 8.1 (a) 3D AFM image and root mean square (RMS) roughness of Cu foil.365 (b) SEM image
of a Cu surface after graphene growth.364

8.1 Modeling and discussions

Figure 8.2(a) shows the atomic structure of an ideal Cu(111)-based high-index low symmetric
Cu surface with parallel step edges along a constant direction. The heights of the atoms are coded by
different colors and the tilted angle of this surface with respect to its high-symmetric terrace plane (that
122
is Cu(111) in this example) is denoted by α. Obviously, the step density increases with the increase of
α. As discussed above, a realistic substrate surface must have a surface roughness. Figure 8.2(b) shows
an example of the possible structure of a realistic surface, where we use θ to define the angle between
the substrate surface and the tangent line of a nanoprotrusion (or nanoindent) on the surface, to describe
the sharpness of the nanoprotrusion (or nanodent). As shown in Figure 8.2(b), the step edges are no
longer aligned along a constant direction but become meandering because of the surface roughness.
Figure 8.2(e) shows the top views of three high-index Cu surfaces with a θ of 3.5º.From these top views,
we can easily find the direction change of step edges and we use ∆𝜑 to describe the maximum change
of the step edge direction from its original direction, ∆𝜑 = 𝜑𝑚𝑎𝑥 − 𝜑𝑚𝑖𝑛 .

Figure 8.2 The effects of roughness on the surface topography of a Cu foil. (a) Atomic models of an
ideal high-index Cu surface which has a tilted angle of 𝛼 with its low-index terrace. (b) Different views
of a fluctuated Cu foil surfaces with 𝜃 denoting the fluctuation degree. (c) Model of constructing a
fluctuated Cu surface. (d) Relationship between Cu foil fluctuation degree and variation of Cu step edge
direction for various high-index Cu foils. (e) Top views of various high-index Cu surfaces under a same
surface fluctuation with Height profile colored. (e) Relationship between 𝜃 and 𝛼 regard to different
step variation.

123
To quantitatively calculate ∆𝜑, a simplified cone model is used to descibe the nanoprotrusion
on the Cu surface, as shown in Figure 8.2(c). Suppose the Cu surface is in the horizon plane, the surface
of the cone can be described by:

𝑧 = −√(tan 𝜃)2 ∙ (𝑥 2 + 𝑦 2 ) (8.1)

We further assume that the angle between the high-index surface and its nearest low-index
terrace is 𝛼 and planes parallel to the low-index terrace satisfies that:

𝑧 = tan 𝛼 ∙ 𝑥 − 𝐶 (8.2)
The step edges on the cone are thus the interception lines of the cone surface and the planes of
the terraces, and they are curved lines. Moreover, the direction the projections of the tangential
directions of the interception lines are thus the directions of the step edges. Mathematically the
projection of the interception curves to the high-index surface (z = 0) satisfies:

−√(tan 𝜃)2∙ (𝑥 2 + 𝑦 2 ) = tan 𝛼 ∙ 𝑥 − 𝐶 (8.3)

From Eq. (8.3), we can get:

(tan 𝛼)2 2 − 2𝐶
tan 𝛼 𝐶2
𝑦 = ±√( − 1) ∙ 𝑥 ∙ 𝑥 + (8.4)
(tan 𝜃)2 (tan 𝜃)2 (tan 𝜃)2

𝑑𝑦 1 ((tan 𝛼)2 − (tan 𝜃)2 ) ∙ 𝑥 − 𝐶 tan 𝛼


=± (8.5)
𝑑𝑥 tan 𝜃 √((tan 𝛼)2 − (tan 𝜃)2 ) ∙ 𝑥 2 − 2𝐶 tan 𝛼 ∙ 𝑥 + 𝐶 2

The highest and lowest boundaries of the step edge direction (𝜑𝑚𝑎𝑥 and 𝜑𝑚𝑖𝑛 )can be calculated by:

𝑑2 𝑦 tan 𝜃 ∙ 𝐶 2
= ∓ =0 (8.6)
𝑑𝑥 2 [((tan 𝛼)2 − (tan 𝜃)2 ) ∙ 𝑥 2 − 2𝐶 tan 𝛼 ∙ 𝑥 + 𝐶 2 ]3/2
𝑑2 𝑦 𝑑𝑦
From Eq. (8.6), it can be seen that 𝑑𝑥 2 = 0 at 𝑥 → ±∞, suggesting that 𝑑𝑥 reaches its maximum and

minimum values at 𝑥 → ±∞, respectively, which are:

𝑑𝑦 1 ((tan 𝛼)2 − (tan 𝜃)2 ) ∙ 𝑥 tan 𝛼 2


lim =− = √( ) −1 (8.7)
𝑥→+∞ 𝑑𝑥 tan 𝜃 √((tan 𝛼)2 − (tan 𝜃)2 ) ∙ 𝑥 2 tan 𝜃

𝑑𝑦 1 ((tan 𝛼)2 − (tan 𝜃)2 ) ∙ 𝑥 tan 𝛼 2


lim = √
=− ( ) −1 (8.8)
𝑥→−∞ 𝑑𝑥 tan 𝜃 √((tan 𝛼)2 − (tan 𝜃)2 ) ∙ 𝑥 2 tan 𝜃

From Eqs. (8.7-8.8), we finally obtain:

𝑑𝑥 𝑑𝑥
∆𝜑 = 𝜑𝑚𝑎𝑥 − 𝜑𝑚𝑖𝑛 = 𝑎𝑡𝑎𝑛 ( lim ) − 𝑎𝑡𝑎𝑛 ( lim ) (8.9)
𝑥→+∞ 𝑑𝑦 𝑥→−∞ 𝑑𝑦

and equivalently we can finally have:

124
(tan 𝜃)2
2 × atan [√ ], 𝑖𝑓 𝜃 < 𝛼
∆𝜑 = (tan 𝛼)2 − (tan 𝜃)2 (8.10)
{2𝜋 , 𝑖𝑓 𝜃 ≥ 𝛼
It should be noted when 𝜃 ≥ 𝛼, ∆𝜑 will abruptly become 2𝜋 which means the appearance of
step edges along all directions because two step edges will eventually intersect with each other once
their direction difference is larger than 𝜋. Take Cu(111) surface as an example, its 𝛼 is 0°and thus any
surface roughness can lead to the appearance of step edges along all possible directions. Figure 8.2(d)
shows the ∆𝜑 curves as a function of 𝜃 for four different Cu surfaces. It can be seen that ∆𝜑 increases
significantly with 𝜃, suggesting that surface roughness has a significant effect on the step edge direction
and further the alignment of 2D materials on the substrate. In addition, from Figure 8.2(e) we can also
qualitatively reveal the effect of the density of step edges on the change of their directions under the
same 𝜃. Obviously, the variation in the direction of step edges decreases with the increase of step edge
density. Therefore, a high step edge density can effectively weaken the effect of surface roughness on
the change of the directions of step edges.

Figure 8.3(a) shows a map of largest variation of the step edge direction as a function of both
𝜃 and 𝛼. From Eq. (8.10), we can know that for high-index substrates with small 𝜃 and 𝛼, the largest
variation of the step edge direction can be estimated by:

2
(𝜃⁄𝛼 )
∆𝜑 = 2 × atan √ 2
(8.11)
1 − (𝜃⁄𝛼 )
[ ]
Figure 8.3(b) shows the ∆𝜑 curve as a function of 𝜃⁄𝛼 when both 𝜃 and 𝛼 are small. We can
see that for ∆𝜑 < 5º, 10º, 20ºand 30º, 𝜃⁄𝛼 <0.044, 0.088, 0.174 and 0.259 is required, respectively.
On a typical annealed Cu foil, 𝜃 is ~1º,67,84 and thus the critical 𝛼 of the Cu surface that satisfies the
above requirements can be calculated to be 22.72º, 11.36º, 5.75ºand 3.86º, respectively. If a substrate
is carefully polished and its 𝜃 is as low as 0.01º,366,367 unidirectionally aligned step edges on such a
substrate can exist if 𝛼 is greater than 1º.

125
Figure 8.3 (a) Variation of the step edge direction regarding to (𝜃, 𝛼). (b) Variation of n of the step
edge direction regarding to 𝜃/𝛼 with small 𝜃 and 𝛼.

Above analysis clearly shows that high-index surfaces with large 𝛼 (or equivalently large step
edge density) is superior to those with small 𝛼 in templating the synthesis of unidirectionally aligned
2D materials, because of the smaller change in the step edge direction. Take hBN grown on high-index
Cu substrates as an example, Figure 8.4 shows the most stable alignment of hBN islands on Cu(5 5 6)
and the Cu(10 10 17) surfaces, which is obtained based on the results shown in Chapter 7. Obviously,
although these two surfaces have the same 𝜃 = 3.5º, there are two orientations of hBN islands grown
on the Cu(5 5 6) surface, which has a low step edge density, and only one orientation for the hBN
islands on the Cu(10 10 17) surface with a high step edge density.

Figure 8.4 Alignments of hBN islands on Cu(5 5 6) and Cu(10 10 17) surfaces under same surface
roughness of 𝜃 = 3.5º.

126
8.3 Summary

Because the surface of a TM substrate for CVD growth of 2D materials can not be ideally flat,
the direction of a step edge on a high-index substrate surface is site-dependent, which may lead to the
change of alignment of 2D islands grown on the substrate. In this chaper, we have proposed a theoretical
model to systematically investigate the effect of surface roughness on the directions of step edges of the
substrate and further the alignment of 2D materials on the substrate. Overall, we revealed that high-
index surfaces with a large step edge density, which can be otained by cutting a single crystalline TM
along a direction with a large diviation angle with respect to the low-index surfaces, show a high
tolerance of surface roughness on the variation of step edge direction, so that the tolerance range of step
edge direction revealed in Chapters 5 and 7 can be satisfied for the realization of unidirectionally aligned
2D material islands. Therefore, we propose that high-index substrates with high step edge densities are
promising in synthesizing WSSC 2D materials by the seamless coalescence method.

127
Chapter 9 Alignment of TMDC islands on vicinal Au(111)
substrates
9.1 Introduction

Similar to hBN, TMDCs also have a binary composition and their 2H phase possesses 3-fold
symmetry (D3H). As discussed in previous Chapters, the alignment of a 2D material on a substrate is
dominated by the interplay between their symmetries. Therefore, the similarity in the symmetries
between hBN and 2H phase TMDCs may make the alignment of 2H phase TMDCs on a high-index
low symmetric substrate to be similar to that of hBN. Up to now, the substrates used for the growth of
TMDC are mostly inert ones including amorphous SiO2/Si,12,368-370 and high symmetric Al2O3(0001)371-
373
and GaN(0001) substrates,172,374,375 where the grown TMDC islands usually present multiple
orientations and thus polycrystalline TMDC films are formed.

Recent experimental studies show that Au surfaces are promising substrates for the growth of
various TMDCs, including MoS2, MoSe2, WSe2 and WS2, etc.,190,376-378 because Au can decrease the
sulfurization barrier of TM precursors and increase the growth rate.379 By reducing the nucleation
density through pretreatment of the substrate, both WSe2 and WS2 single crystals with a millimeter size
have been realized on reusable Au substrates,167,380 which is 2~3 orders larger than those grown on
insert substrates. Moreover, the synthesis of unidirectional WS2 islands has been achieved on vicinal
Au(111) substrate and WSSC WS2 films have been obtained by the seamless stitching method.179

In this chapter, we chose WS2 as an representative of the TMDC materials to explore their
alignment on Au{111}-based high-index low symmetric substrates with straight step edges.

9.2 Calculation methods

The same as all previous studies of graphene and hBN on the TM substrate surfaces, DFT-D3
calculations are carried out in this chapter to investigate the alignment of WS2 islands on Au{111}-
based low symmetric substrates. The exchange-correlation functionals are treated by GGA,348 and the
interaction between valence electrons and ion cores is described by the PAW method.349 A criteria of
force on each atom to be less than 0.01 eV/Å and an energy convergence of 10-4 eV is used for the
structural optimization.

9.3 Results and Discussions

Since the structural reconstruction of the Au(111) surface has been extensively observed with
the adsorption of S adatoms,381-383 the Au(111) surface reconstruction should be considered. Here, a

128
commonly observed Au( √3 × √3 ) R30ºhoneycomb lattice384,385 is used as the representative of
reconstructed Au(111) surfaces and the growth of WS2 on the Au(111) surface with and without S-
termination is investigated. Figure 9.1(a) shows the atomic configuration of the (√3 × √3) R30ºS-
terminated Au(111) surface.

Figure 9.1 Comparison of WS2 film on the pristine Au(111) surface and S-terminated Au(111) surface.
(a) Atomic configuration of the (√3 × √3) R30ºS-terminated Au(111) surface. (b) Formation energy
of WS2 film on the pristine Au(111) surface and on the (√3 × √3) R30ºS-terminated Au(111) surface
as a function of the chemical potential of sulfur. (c) Atomic configurations of WS2 film on the (√3 × √3)
R30ºS-terminated Au(111) surface with relative angles of 0ºand 30º.(d) Atomic configurations of WS2
film on the pristine Au(111) surface with relative angles of 0ºand 30º.

We first investigated the stabilities of a WS2 film on the Au(111) surface and on the (√3 × √3)
R30ºS-terminated Au(111) surface by calculating their formation energies, which are defined by:

𝐸𝑓 = (𝐸𝑇 − 𝜀𝐴𝑢 × 𝑁𝐴𝑢 − 𝜇𝑆 × 𝑁𝑆 − (𝜇𝑊 + 2 × 𝜇𝑆 ) × 𝑁𝑊 )/𝑁𝑊 (9.1)

where 𝐸𝑇 is the total energy of the WS2 film on a substrate, 𝜀𝐴𝑢 and 𝑁𝐴𝑢 denote the energy of a Au atom
in the Au(111) substrate and the number of Au atom in the substrate, 𝜇𝑆 and 𝑁𝑆 denote the chemical
potential of S and the number of S adatoms in the (√3 × √3) R30ºS-terminated Au(111) surface, 𝜇𝑊 +

129
2 × 𝜇𝑆 = 𝜇𝑊2 𝑆 and it denotes the energy of a WS2 unit in the WS2 film, 𝜇𝑊 and 𝑁𝑊 are the chemical
potential of W and the number of W atoms in the WS2 film. As 𝜇𝑊2 𝑆 is a constant, the formation energy
is only related to the chemical potential of S, and in Figure 9.1(b), the formation energy profiles of a
WS2 film on the Au(111) surface and on the (√3 × √3) R30ºS-terminated Au(111) surface with relative
rotation angles of 0ºand 30º(see Figure 9.1(c-d)) are given. Because 𝜇𝑆 is equal to -4.32 eV or -5.47
eV by using S bulk or W bulk as a reference, it is clear that the growth of WS2 film on the pristine
Au(111) surface is always energetically preferred in the 𝜇𝑆 range between the two values, indicating
that the Au(111) surface can remain its pristine configuration with the cover of the WS2 film. Therefore,
the pristine Au(111) surface is used as the substrate in the following exploration for the alignment of
WS2 islands.

To explore the orientation of WS2 islands on a low symmetric substrate, vicinal Au(111) surface
with <110> and <211> step edges are chosen as the substrate and a square-shape WS2 cluster with two
AC edges, one ZZW and one ZZS edge is chosen as the WS2 island, so that the interaction between
straight WS2 edges and straight Au step edges can be investigated. Different with one-atom thick 2D
materials, the edges of the three-atom thick TMDC is preferred to be self-terminated, and it has been
reported that the reconstructed edges are highly dependent of the ambient condition and various self-
terminated edges have been explored by both experimental and theoretical studies.318,339,386-388 To make
it simple, we suppose a S-rich condition and thus the reconstructed AC and ZZW edges of the WS2
cluster with S-termination are energetically preferred, as shown in Figure 9.2. The total energies of such
a WS2 cluster with its different edges attached to the Au<110> and Au<211> step edges are compared.
Because of the square shape of the WS2 cluster and the equal number of atoms for all these models, a
lower total energy implies a stronger interfacial binding energy between the WS2 edge and the Au step
edge.

Along the Au<110> step edge, ZZS and ZZW edges are more superior than the AC edge, but
there is no obvious binding energy difference for the interfaces of the Au<110> step edge attached by
ZZS and ZZW edges because both of the WS2 ZZ edges have been well passivated by S atoms.
However, the different compositions of the ZZS and such S-terminated ZZW edge must result in
different formation energies of the ZZS and such S-terminated ZZW edge, which will lead to only one
preferential orientation of the WS2 island along the Au<110> step edge. Along the Au<211> step edge,
AC edge is more superior than the two ZZ edges, as a result, two orientations of WS2 island with a
misangle of 60ºwill form because the absence of the mirror symmetry for WS2 respect to its ZZ direction,
which is similar with the cases (II) and (III) in Figure 7.3 showing the hBN AC edge attached to the
Cu<211> step edge. Although there is a large lattice mismatch between WS2 and Au(111), the results
of WS2 edges attaching to Au<110> and Au<211> step edges are quite similar to those of hBN on
Cu{111}-based low symmetric surfaces that have very close lattice constants.

130
Figure 9.2 Total energies of different edges of a WS2 cluster attaching (a) Au<110> step edge and (b)
Au<211> step edge, the unit of the total energy is eV.

It should be noted that, because of the unbalanced stoichiometry, the edge formation energies
of the WS2 edges as well as the interfacial formation energies between these edges and the Au step
edges highly depends on the ambient condition. Consequently, TMDC islands can show different
orientations under different ambient conditions, which means that achieving unidirectionally aligned
TMDC islands on a low symmetric TM substrate is possible by tuning the ambient condition. Further
explorations are required to investigate the alignment of TMDCs on low symmetric substrates under
different ambient conditions. In addition, the interfaces between various reconstructed TMDC edges
and various TM step edges also need to be studied in the future.

9.4 Summary

In this Chapter, we simply investigated the alignment of TMDC islands on low symmetric TM
substrates with straight step edges by using WS2/Au{111}-based low symmetric surfaces as an example.
The results are quite similar with those obtained in the section 7.3.1, confirming the ability of the low
symmetric Au surfaces in templating the growth of unidirectionally aligned WS2 islands. However, the
study in this Chapter only provides preliminary results and a systematical research on the alignment of

131
TMDC islands on low symmetric TM surfaces needs further study, and moreover, the exploration on
the alignment of TMDC on low symmetric inorganic substrates, such as Al2O3, is also needed as they
are also commonly used substrates in CVD growth of TMDC materials.

132
Chapter 10 Formation mechanisms of large single-crystal Cu(111)
foil by contact free annealing
10.1 Introduction

In previous Chapters, we have proved that low symmetric substrates rise the chance of the
unidirectional alignment of 2D materials of different symmetries, however, before these low symmetric
substrates are applied in the growth of 2D materials, it has been known that unidirectionally aligned
graphene islands and further WSSC graphene films can be synthesized on the Cu(111) surface, which
is also one of the earliest milestone achievements in graphene synthesis. To achieve this goal, large-
sized single crystalline Cu(111) substrates should be obtained as a prerequisite. Up to now, there have
been several methods of fabricating large-sized single crystalline Cu foils or films. Initially, single
crystalline Cu(111) films were usually fabricated by depositing and annealing of Cu thin films on
insulated substrates, such as sapphire,42,389-391 quartz,392 silicon substrates,393-395 etc., but the sizes of
obtained Cu foils are usually very limited and not cost-effective by such PVD methods.

In recent years, it was found that Cu(111) grains with a lateral size at centimeter scale can be
produced by properly annealing polycrystalline Cu foils,396,397 and strategies of reducing grain
boundaries in commercial polycrystalline Cu foil by annealing are then explored.398,399 In 2017, Xu et
al. developed a temperature-gradient-driven annealing technique and realized the production of single-
crystal Cu(111) foils with a size up to 5×50 cm2,69 which enables the first synthesis of meter-sized
single crystalline graphene. However, the control of temperature gradient in this study is difficult.

In this Chapter, we present a contact-free annealing method of achieving single-crystal Cu(111)


films with the largest size up to 32 cm2 from commercial polycrystalline Cu foils, and the size of
obtained Cu(111) films is in fact limited by the size of the tube furnaces.67 Importantly, we revealed the
formation mechanisms of single crystalline Cu(111) films from polycrystalline Cu foils by both DFT
and classical MD simulations.

10.2 Experimental setups and results

The commercial polycrystalline Cu foil was held by a quartz holder which is suspended by two
blocks. As shown in Figure 10.1, one edge of the as-received Cu foil was folded to make it held by the
quartz holder and the rest part was freely suspended without contact. The Cu foil was then annealed at
1323K for 12 hours in an atmosphere of H2/Ar (both at 10 sccm), and a single crystalline Cu(111) foil
with the largest size up to 32 cm2 was then obtained.

133
Figure 10.1 The synthesis of single-crystal Cu(111) foil by contact free annealing. (a) Diagrams of Cu
foil suspended on the quartz holder. (b) Photograph of the obtained Cu(111) foil with a size of 32 cm2.

Figure 10.2(a) shows the photography of an annealed Cu foil. It can be seen that there is no
obvious contrast in the contact-free region of 2cm × 8cm in the Cu foil, suggesting its single
crystallinity. X-ray diffraction (XRD) and EBSD were performed in different areas of the annealed Cu
foil to study its crystallographic characteristic. 2𝜃 scans shown in Figure 10.2(b) correspond to different
areas marked as P1, P2 and P3 of Figure 10.2(a) show that the Bragg reflections are all from the (111)
and (222) crystal facets, indicating its (111) crystalline surface. The EBSD maps in Figure 10.2(c) on
different areas of the Cu foil also suggest its single crystallinity. The in-plane direction of different areas
of the Cu foil are then examined, as shown in Figure 10.2(d). Clearly, all areas show same <112> rolling
direction without in-plane rotation, confirming that the annealed Cu foil is a single crystal with
{111}<112> surface in the contact-free region.

134
Figure 10.2 The characterization of the annealed single-crystal Cu(111) foil. (a) Photograph of the
obtained single-crystal Cu(111) foil. (b) XRD 2𝜃 scans of P1-P3 areas of the annealed Cu foil showing
(111) crystallographic plane. (c-d) EBSD maps and (001) pole figures of different areas of the annealed
Cu foil.

To understand the formation process of the Cu foil with {111}<112> surface during annealing,
the texture evolution with respect to the annealing time was further explored. The orientation
distribution function (ODF) from EBSD data was employed to describe the texture, and the positions
of typical textures of an FCC metal is shown in Figure 10.3(a). Before annealing, the commercial
polycrystalline Cu foils are mainly composed by {112}<111> and {110}<112> surface (Figure 10.3(b)).
After 1 hour’s annealing, small grains with {111}<112> surface started to appear (Figure 10.3(c)). It is
worth noting that only a few grains with {111}<112> surface that are located near the foil edges can
grow to large grains, indicating that the stored strain energy from mechanical processing on the edges
can accelerate grain growth. After 2 hours’ annealing, two large-sized grains with {111}<112> surface
separated by numerous small grain were obtained (Figure 10.3(d)). With an additional annealing of 5.5
hours, the grains in the middle all disappeared and a single crystalline Cu foil with {111}<112> surface
was eventually formed.

135
Figure 10.3 Texture evolution of as-received Cu foils during contact-free annealing. (a) ODF section
with 𝜑2 = 45ºshowing the positions of typical textures of FCC metal. (b) EBSD and ODF maps of an
as-received Cu foil. (c) EBSD and ODF maps of the Cu foil annealed for 1 hour. (d) Photography,
EBSD and ODF maps of the Cu foil annealed for 2 hours.

Figure 10.4 (a) SEM image of a Cu grain with {112}<111> surface, and EBSD maps of the Cu grain
before and after contact-free annealing. (b) Surface energies map of different Cu surfaces.

It is important to address that the initial {112}<111> textures of Cu foils are critical for
obtaining the final {111}<112> foils. We found that if the textures of the as-received Cu foils is not
{112}<111> dominated, the final Cu foils would show other crystallographic orientations even though
the same annealing treatment was used. Figure 10.4(a) shows the annealing process of a small Cu foil
with only {112}<111> textures (middle panel) was annealed at 1323K for 12 hours, and the obtained
Cu foil indeed showed {111}<112> surface (right panel), confirming that the final {111}<112> texture
is transformed from the {112}<111> one. From the surface energy map shown in Figure 10.4(b), it is
clear that from {112} to {111}, the surface energy keeps decreasing. It is worth noting that another
main component in the initial Cu foils, i.e., {110} surface, is a saddle point in the surface energy map,

136
indicating that there is no driving force for the rotation of grains with {110} surface. Therefore, the
transformation from {112}<111> grains to {111}<112> ones is responsible for the synthesis of final
single-crystalline Cu(111) foil.

10.3 Theoretical methods and discussions

To explore the rotation mechanism from a {112}<111> grain to a {111}<112> one, classical
MD simulations were preformed first to simulate the rotation process, and then the rotation energetics
was calculated precisely DFT method.

10.3.1 Calculation methods

Classical MD simulations were preformed to study the motion of dislocations by the LAMMPS
package. NPT ensemble was adopted with the Nosé–Hoover thermostat and barostat controlling
temperature and pressure. The pressure is maintained at 1 bar during the whole process. The velocity–
Verlet algorithm is used with a time step of 1.0 fs. Cu-Cu interaction is described by EAM potentials.
Periodic boundary condition was applied in x, y and z directions. The vacuum spacing between
neighboring images of the Cu slab is at least 12 Å to avoid imaginary interaction.

The energies related to the rotation mechanism are calculated within the framework of DFT-
D3 method as implemented in VASP. The exchange-correlation functionals are treated by GGA, and
the interaction between valence electrons and ion cores is described by the PAW method. During
structure optimization, the force on each atom is minimized to be less than 0.01 eV/Å, with an energy
convergence of 10-4 eV.

10.3.2 Modeling and discussions

Based on a series of contrast experiments, we found that there are two essential conditions to
obtain the Cu foil with {111} surface: a high annealing temperature near the Cu melting temperature
and high H2 concentration. The high annealing temperature can effectively increase the concentrations
defects in the Cu foil, such as vacancies and interstitials. By DFT calculations, the formation energy of
a vacancy in Cu bulk (1.64 eV) is much smaller than that of an interstitial in Cu bulk (2.59 eV), therefore
here we mainly considered the effect of vacancies in the MD simulations. Estimated by Arrhenius
equation, the concentration of vacancies in Cu bulk at a temperature of 1323K is ~10-6. Hydrogen can
decrease the formation energy of vacancies, for instance, the formation energies of a vacancy that
passivated by one and two H atoms are calculated to be 1.20 and 0.72 eV, respectively. As a result, the
concentration of vacancies increases drastically with a value up to ~10-3. Besides, we need to note that

137
vacancies diffuse very easily inside of the Cu foil due to their low diffusion barrier, and eventually
aggregate together.

Figure 10.5 Illustration of the movement of vacancies in Cu foil. (a) Diagram of the formation of SF in
FCC metal. (b) Atomic configurations showing stacking order of a perfect FCC metal and that of a SF.

Figure 10.5(a) illustrates that the large vacancy formed from the aggregation of a large number
of single vacancies on the plane is mechanically unstable, the adjacent lattice planes of the vacant disc
tend to collapse and become neighbors (blue atoms), forming negative Frank dislocation loops (located
at the interface between the blue and yellow atoms). As a sessile dislocation, the Frank dislocation can
move by climbing, companied with the formation of an intrinsic (111) stacking fault (SF) across through
the grain, which is due to the very low SF formation (0.326 eV/nm2). Figure 10.5(b) shows the atomic
configurations of a perfect ABC-stacking FCC structure and an FCC structure with an intrinsic (111)
SF, due to the lack of a layer of atoms, (111) SF shows AB-stacking HCP structure, which is marked
by blue atoms.

In order to investigate the movement of the vacancies in the Cu foil, classical MD simulations
were performed, as shown in Figure 10.6. Firstly, we introduced vacancies with a 4.5% concentration
in a 5.01×2.04×4.4 nm3 Cu foil with {211} surface, which was then annealed at 1000K for 10 ns.
During annealing process, lots of (111) SFs appear and then disappear, resulting in a small rotation of
~2.02ºfor the grain, which can also be confirmed by the reduction of the number of step edges on the
Cu surface. If we further introduce 5.0% vacancies in the annealed structure, further rotation happens
and the number of step edges is reduced by 2 after annealing, suggesting that the disappear of the SFs
makes the grain rotate toward to the {111} surface. These MD simulations proved that the rotation from
{112}<111> grain to {111}<112> one can happen in a stress-free environment.

138
Figure 10.6 MD simulations showing the evolution of the as-received Cu foils during contact-free
annealing. (a) Snapshots of MD simulations for annealing of a Cu grain. (b) Atomic configurations
showing orientational evolution of Cu grain during annealing.

Based on MD simulations, the formation mechanism of grain rotation is proposed, as illustrated


in Figure 10.7: (i) A large number of vacancies prefer to aggregate together to form Frank partial
dislocation (FPD). (ii) By capturing the vacancies steadily, the half atomic planes of Frank partial
dislocation become smaller and smaller and even disappear, leaving the SF only. (iii) The interaction
between the SFs and surface can activate a Shockley partial dislocation (SPD) at one side of the SF. (iv)
The SF then glides from the surface and the SF area keeps shrinking, resulting in a rotation of the grain.
It should be noted that there are two possible directions for dislocation slipping to remove the SF. One
path can form the flat terrace and another path results in two steps appearing on opposite sites
respectively (Figure 10.7(b)). Obviously, the terrace has a lower energy and is energetically more
preferred. Therefore, the SF annealing acts as the driving force for grain rotation towards to lower
surface energy, which does not need a stress field around.

139
Figure 10.7 The mechanism of grain rotation during annealing. (a) Sketch and atomic models showing
grain rotation induced by SF. (b) Relative energy of Cu gain rotation.

10.4 Summary

In this Chapter, we introduced a new method of fabricating large-size single-crystalline


Cu{111}<112> foils from commercial polycrystalline Cu foil dominated by {112}<111> grains
through a contact-free annealing. Extensive MD simulations and DFT calculations were performed to
reveal the annealing mechanism. The lowest surface energy of Cu{111} surface offers the driving force
for the formation of the final Cu{111} surface. We found that the gliding of the SFs that are formed
from the aggregation of large amounts of vacancies in the Cu foil is the main mechanism for the rotation
from dominated {112}<111> grains to Cu{111}<112> single crystal. This study has been published in
SCIENCE in 2018.67

140
Chapter 11 Rotated graphene moiré superstructures on various
transition metal surfaces
11.1 Introduction

The 2D materials – substrates interactions can not only regulate the orientation of 2D materials
during the growth, but also bring 2D materials novel structures and properties distinguished with that
in vacuum, and in this chapter, the behaviors of graphene films on various TM substrates will be
investigated. To date, graphene has been successfully synthesized on various TM surfaces, and it is
extensively reported that the characteristics of the TM substrates show a big effect on the morphology
of grown graphene films. 214,222,223,270,400-402 Most obviously, the superposition of the lattices of graphene
and TM substrate arises a kind of ordered pattern in vision, which is called graphene moirépattern, and
above graphene films generally present ups and downs to some extent respecting to the periodicity of
the moiré pattern, and here we name these periodically corrugated graphene as graphene moiré
superstructures.

Such corrugated graphene moiré superstructures on various TM surfaces attract lots of


attentions as templates for the fabrication of metal clusters and assembly of molecules,224,227,228,234,239
among them, graphene/Ru(0001) system is one of the most popular study objects ascribing to its strong
anisotropy and large height fluctuation.403-405 Previous experimental and theoretical studies mainly
concentrated on the graphene/Ru(0001) superstructures with graphene ZZ direction along the Ru<21̅1̅0>
crystalline direction due to the strong epitaxial relationship between graphene and the Ru(0001)
surface,401,406,407 however, the grown graphene which has a small rotation angle respect to the Ru(0001)
surface has also been revealed.408,409 The metal clusters synthesized on different rotated graphene moiré
superstructures on the Ru(0001) surface produced by graphene layer grows cross grain boundaries of
the substrate were found to show different sizes,410 indicting the potential of rotated/TM superstructures
in templating size-tunable metal clusters. Besides, with the maturing of the film transfer technique,411-
413
rotated graphene moirésuperstructure with a desired rotation angle is expectable.

Therefore, a study on the evolution of graphene moirésuperstructure on various TM substrates


with respect to their rotation angle is needed. In this Chapter, graphene/Ru(0001) system is chosen as
an example, and the dependence of graphene moiré superstructures as well as its application in
controllable synthesis on the rotation angle are explored first. Based on the graphene/Ru(0001) system,
a simple model for identifying the morphology of grown graphene on various TM surfaces is developed.

141
11.2 Modeling and calculation methods

11.2.1 Modeling

To build various graphene/Ru(0001) superstructures with different rotation angles, graphene


and Ru(0001) cells with different boundary direction take turns combing together, and Figures 11.1 (a)
and (b) illustrate examples of a graphene cell with ⃑⃑⃑⃑
𝑉1 and ⃑⃑⃑⃑
𝑉2 boundary directions and a Ru(0001) cell
with ⃑⃑⃑⃑
𝑉3 and ⃑⃑⃑⃑
𝑉4 boundary directions. To make sure the two boundary directions of graphene (or
Ru(0001)) maintain the included angle of 120º, the two primary vectors 𝑎⃑ and 𝑏⃑⃑ (𝑐⃑ and 𝑑⃑) with 120º
angle of graphene (or Ru(0001)) are chosen as the basic vectors, and then the two boundary directions
⃑⃑⃑⃑1 and ⃑⃑⃑⃑
of graphene (𝑉 ⃑⃑⃑⃑3 and ⃑⃑⃑⃑
𝑉2 ) and the Ru(0001) (𝑉 𝑉4 ) cells can be obtained by:

𝑉1 = 𝑚 ∙ 𝑎⃑ + 𝑛 ∙ 𝑏⃑⃑, ⃑⃑⃑⃑
⃑⃑⃑⃑ 𝑉2 = −𝑛 ∙ 𝑎⃑ + (𝑚 − 𝑛) ∙ 𝑏⃑⃑ (11.1)

𝑉3 = ℎ ∙ 𝑐⃑ + 𝑘 ∙ 𝑑⃑, ⃑⃑⃑⃑
⃑⃑⃑⃑ 𝑉4 = −𝑘 ∙ 𝑐⃑ + (ℎ − 𝑘) ∙ 𝑑⃑ (11.2)

Figure 11.1 Modeling of the graphene/Ru(0001) superstructures. (a, b) Illustrations of the construction
of supercells of graphene and the Ru(0001) surface. (c) The sizes of graphene/Ru(0001) superstructures
(blue point) and the size of the moirépattern unit cells (lines) versus the rotation angles. The sizes of
graphene/Ru(0001) superstructures chosen in DFT calculations are marked by red circles.

Defining the graphene ZZ direction and the Ru <21̅1̅0> direction as the reference direction of
graphene and the Ru(0001) surface, the rotation angles between graphene and the Ru(0001) surface can

142
be obtained by: 𝜑 = 𝜃1 − 𝜃2 , where 𝜃1 is the angle between ⃑⃑⃑⃑
𝑉1 and 𝑎⃑, 𝜃2 is the angle between ⃑⃑⃑⃑
𝑉3 and
𝑐⃑. 𝜃1 and 𝜃2 can be calculated by:

1 1
𝑚 − 2𝑛 ℎ − 2𝑘
cos 𝜃1 = , cos 𝜃2 = (11.3)
√𝑚2 + 𝑛2 − 𝑚𝑛 √ℎ2 + 𝑘2 − ℎ𝑘

For the graphene/Ru(0001) supercells, the vectors of graphene and the Ru(0001) surface obey
⃑⃑⃑⃑1 | = |𝑉
the condition of |𝑉 ⃑⃑⃑⃑3 | and a mismatching of 3.0% is allowed in modeling, the sizes of

graphene/Ru(0001) supercells with various rotation angles conforming to the rules are denotes by blue
points in Figure 11.1(c). Here, the rotation angles of graphene/Ru(0001) superstructures can be
restricted into a range of 0-30ºand the graphene/Ru(0001) supercells have a mirror symmetry regarding
to 30ºbecause both the graphene and the Ru(0001) surface are C6V symmetric.

The line in Figure 11.1(c) gives the evolution of graphene moirépattern as a function of the
rotation angle, and the derivation is as followed. The reciprocal lattice vectors of moirépattern has a
relationship of 𝑘⃑⃑ moiré =𝑘⃑⃑ Gra -𝑘⃑⃑ Ru with that of its two constituting lattices,414 therefore, the smaller
different of the two constituting lattices, the larger size of graphene moirépattern. The lattice mismatch
between graphene and the Ru(0001) surface is defined by a factor 𝛿 , which is defined as 𝛿 =
𝑎𝑅𝑢 −𝑎𝐺𝑟𝑎 415
𝑎𝑅𝑢
, and the reciprocal lattice vectors of graphene and the Ru(0001) surface then become:

2𝜋 2𝜋
𝑘⃑⃑Gra = 𝑎 (1,0) , 𝑘⃑⃑Ru = 𝑎 (cos 𝜑 , sin 𝜑) (11.4)
𝐺𝑟𝑎 𝐺𝑟𝑎 (1+𝛿)

And the moirépattern size in real space can be calculated by:

2𝜋 (1 − 𝛿)𝑎𝐺𝑟𝑎
𝜆= = (11.5)
⃑⃑⃑⃑⃑⃑
|𝑘 𝑚| √2(1 − 𝛿)(1 − cos 𝜑) + 𝛿 2

11.2.2 Calculation methods

The calculations are carried out by DFT-D3 method with VASP.346,347 Using the PAW method
to perform the interaction between valence electrons and ion core,349 and the energy cutoff for the plane-
wave function is set to 400 eV. The exchange-correlation functions are dealt with GGA.348 A force on
each atom less than 0.01 eV/Å and an energy convergence of 10-4 eV are set as the convergency
conditions for structural optimization. To exclude edge coupling, periodic boundary conditions are
applied in the in-plane direction while a vacuum spacing of 12 Å is induced in the out-plane direction.

143
11.3 Rotation-dependent graphene on the Ru(0001) surface

To explore the orientation-dependent graphene moirésuperstructures on the Ru(0001) surface,


ten graphene/Ru(0001) superstructures with an interval about 3º were optimized by dispersion‐
corrected DFT calculations.

11.3.1 Structural evolution of graphene on the Ru(0001) surface

The structural evolution of the optimized graphene/Ru(0001) superstructures regarding to


rotation angles is shown in Figure 11.2, in which the unit cells of graphene/Ru(0001) superstructures
are labelled by black rhombuses. The above graphene on the Ru(0001) surface presents a very large
height fluctuation at 0ºand the corrugated morphologies of graphene can be maintained until around
20º, while further increasing the rotation angle, the morphology of above graphene become ultra-flat.
From top views of graphene/Ru(0001) superstructures in Figure 11.2(a), three typical sites, i.e., ATOP,
FCC and HCP sites, in the unit cells of all corrugated graphene moirésuperstructures are clear, and
graphene lattice on ATOP sites usually show a larger distance with the substrate comparing with other
sites. The more intuitional heigh oscillation file in Figure 11.2(c) shows that the degree of graphene
heigh fluctuation keep decaying within 0 ~ 20ºwith the largest distance on the ATOP sites decreasing
from 3.7 Å to 2.6 Å and the lowest distance of graphene to the substrate keeping constant at ~ 2.10 Å,
implying a weak vdW interaction of graphene and Ru(0001) substrate on ATOP sites and a strong
chemical binding of graphene and other sites of Ru(0001) substrate. Once the rotation angle exceeds
20º, the ultra-flat graphene layers show a distance of ~ 3.3 Å everywhere to the substrate, which is in
the typical vdW interaction range.

144
Figure 11.2 Evolution of graphene structures on the Ru(0001) surface with different rotation angles. (a)
Top views of graphene/Ru(0001) superstructures, where ATOP, FCC and HCP regions respectively are
marked by pink Ru atoms, hexagons and triangles, and graphene/Ru(0001) unit cells are denoted by
rhombus. (b) Perspective views of a few typical graphene/Ru(0001) superstructures. (c) The height
oscillation of above graphene on the Ru(0001) surface, in which the average distances between
graphene and substrate are marked by points.

Figure 11.13 shows the distribution of C-C bond length in various graphene/Ru(0001)
superstructures, and very clearly, that the lengths of C-C bonds for corrugated graphene layers are
highly dependent of the lattice relationship between graphene and the Ru(0001) substrate. It has already
been demonstrated that in the 0ºgraphene/Ru(0001) superstructure, C atoms of graphene are sp2
hybridized on ATOP sites due to the weak interaction335 with a typical sp2 C-C bond length of ~1.42
Å.416-418 In contrast, C-C bonds on other sites present a length of 1.45 ~ 1.46 Å, indicating the strong
interaction between graphene on other sites of the Ru(0001) substrate leads C atoms here prone to sp3
hybridized to some extent. For graphene/Ru(0001) superstructures with rotation angle larger than 20º,
C-C bond length becomes ~1.42 Å everywhere, which is similar with that on ATOP sites of the
corrugated graphene/Ru(0001) superstructures. The similar distribution of C-C bond length in
corrugated graphene/Ru(0001) superstructures with that in the 0º superstructures indicates that

145
corrugated graphene layers on the Ru(0001) surface may have similar properties with the aligned
graphene on the Ru(0001) surface with 0º, while the properties of the ultra-flat graphene layers are
uniform and not sensitive to the sites and rotation angles.

Figure 11.3 C-C bond length profiles of graphene/Ru(0001) superstructures. (a) The atomic
configuration of C-C bond with a colored length profile. (b) The statistical result of C-C bond length,
with the longest, shortest and average bond length shown.

11.3.2 Electronic properties of graphene/Ru(0001) superstructures

It has been reported extensively that for graphene on the TM substrates with 0ºrotation angle,
the hybridization between π orbitals of C atoms and dz2 orbitals of TMs is responsible for the
interaction between graphene and TM substrates,385,404,405 however, the studies on the properties of
rotated graphene/TM superstructures are very limited. Here, three corrugated graphene/Ru(0001)
superstructures with rotation angles of 0º,7.9ºand 13.89ºare used to explore the effect of rotation angle
on the properties of graphene layers and the 30ºgraphene/Ru(0001) superstructure is chosen as the
representative of the ultra-flat graphene/Ru(0001) ones, as displayed in Figure 11.4. The partial density
of states (PDOSs) show that the interactions between graphene and the Ru(0001) substrate are mainly
contributed by the hybridization between C p𝑧 orbitals and Ru d𝑧2 orbital. Both PDOSs and charge
density difference (CDD) indicate that, for the corrugated graphene/Ru(0001) superstructures, the
interaction between graphene and the Ru(0001) surface at FCC and HCP sites are always much stronger
than that at ATOP site, despite the coupling at the ATOP sites becomes stronger with the increase of
the rotation angle. For the 30ºgraphene/Ru(0001) superstructure, PDOSs are similar and not site-
dependent, and there is no charge transfer shown in the CDD distribution at an isosurface level of 0.004
e/bohr3, implying an uniform property of ultra-flat graphene layers on the Ru(0001) surface.

146
Figure 11.4 Electronic properties of graphene/Ru(0001) superstructures with four different rotation
angles. (a) The PDOS of the C and Ru atoms at different sites of graphene/Ru(0001) superstructures.
(b) CDD of graphene/Ru(0001) superstructures with an isosurface level of 0.004 e/bohr3. (c) Side view
and section of the CDD of the 7.59ºgraphene/Ru(0001) superstructure.

11.3.3 Applications of graphene/Ru(0001) superstructures as templates

As one of the most popular templates for the synthesis of TM clusters, the 0ºgraphene/Ru(0001)
superstructure has already been well-explored.224,234,419 but the sizes of the synthesized TM clusters are
generally uncontrollable. Since graphene/Ru(0001) superstructures with rotation angles less than 20º
can preserve corrugated structures of graphene, these superstructures may also serves as templates and
producing a TM cluster with a desired size by choosing an appropriate graphene/Ru(0001) template is
expectable.
Here, to explore the possibility of the corrugated graphene/Ru(0001) superstructures in
templating TM clusters, we first investigate and compare the adsorption capacities of different sites of
these superstructure to a Pt atom, as shown in Figure 11.5(a). The adsorption energy of a Pt atom
attaching to different sites of a graphene/Ru(0001) superstructure, 𝐸𝐴𝑑 , is defined as:
𝐸𝐴𝑑 = 𝐸𝐺/𝑅𝑢 + 𝜇𝑃𝑡 − 𝐸𝑇𝑜𝑡 (11.6)

147
where 𝐸𝐺/𝑅𝑢 and 𝜇𝑃𝑡 represent the energy of a graphene/Ru(0001) superstructure and the energy of a
Pt atom in vacuum, respectively. 𝐸𝑇𝑜𝑡 is the total energy of a Pt atom attaching to the
graphene/Ru(0001) superstructure.
Because a prerequisite for the fabrication of dispersed TM clusters on a substrate is that the
interaction between the TM and the substrate should be larger than the cohesive energy of the TM,420
which means to grow Pt clusters on a graphene/Ru(0001) superstructure, it requires that the Pt -
graphene/Ru(0001) interaction should be larger than the Pt-Pt interaction. From Figure 11.5, we see
that both HCP and FCC sites of corrugated graphene/Ru(0001) superstructures have a stronger
adsorption energy to a Pt atom than the Pt-Pt interaction, confirming the abilities of the corrugated
graphene/Ru(0001) superstructures in templating the growth of dispersed Pt clusters. It is worth noting
that the adsorption energy of the FCC site of a graphene/Ru(0001) superstructure is larger than the HCP
site, implying the stronger graphene - Ru(0001) interaction at HCP site weakens its adsorption capacity
to Pt atoms. In contrast, the adsorption capacities of different sites of ultra-flat graphene/Ru(0001)
superstructures to a Pt atom have no different, and the adsorption energies are smaller than the Pt-Pt
interaction, which means ultra-flat graphene/Ru(0001) superstructures are not good for templating.

Figure 11.5 The adsorption energy of a Pt atom at different sites of graphene/Ru(0001) superstructures
versus the rotation angles.

Due to Pt atoms prefer FCC site of a graphene/Ru(0001) superstructure, the area of FCC site
was used as the criteria of deciding Pt cluster size, i.e., the area of 1-layer Pt cluster is similar with that
of the FCC site of the graphene/Ru(0001) structure and the reasonability of the criteria has also been
testified. As illustrated in Figure 11.6, from the formation energies of Pt clusters on the 10.89º
graphene/Ru(0001) superstructure versus the number of Pt atoms, we see that the 1-layer Pt cluster
which has a similar area with the FCC site is the one composed by 7 Pt atoms, and once Pt cluster
surpasses 7 atoms, 2-layer Pt clusters are more favorable than 1-layer ones.

148
Figure 11.6 The formation energies of Pt clusters of different size as a function of Pt atoms number.

In the following, the synthesis of 1- and 2-layer Pt clusters templated by the corrugated
graphene/Ru(0001) superstructures with different rotation angles are investigated. From Figure 11.7,
we see that Pt clusters on the 0ºgraphene/Ru(0001) template show sizes up to 45 atoms and 45+36
atoms for 1- and 2-layer structures, while that on the 19.10ºgraphene/Ru(0001) template is composed
by one Pt atom. Besides, the stabilities of 1- and 2-layer Pt clusters on these templates are compared
through their formation energies, which are listed at the bottom of the atomic configurations, clearly,
the formation energy of 2-layer Pt cluster on a graphene/Ru(0001) template is always smaller than that
of the corresponding 1-layer Pt cluster, indicting Pt clusters on the graphene/Ru(0001) templates favor
3D structures more. Here the definition of the formation energy of a Pt cluster on a graphene/Ru(0001)
template is:
EPt = (ETot/Pt − ETot − NPt × εPt )/NPt (11.7)

where ETot/Pt and ETot represent the energy of a Pt cluster attaching to a graphene/Ru(0001) template
the energy of the graphene/Ru(0001) template, εPt and NPt denote the cohesive energy of a Pt atom in
its bulk structure and the number of Pt atoms in the Pt cluster, respectively.

149
Figure 11.7 Atomic configurations of 1- and 2-layer Pt clusters on various graphene/Ru(0001)
templates, with the corresponding formation energy of Pt clusters on graphene/Ru(0001) templates (EPt )
showing at the bottom.

11.4 Formation mechanisms of graphene moirésuperstructures on the Ru(0001) surface

Based on the behaviors of graphene on the Ru(0001) surface, two models of graphene are
proposed and exhibited in Figure 11.8, which are corrugated graphene/Ru(0001) superstructures
comprised of a hump surrounded by flatland with small rotation angels and ultra-flat graphene/Ru(0001)
ones with large rotation angels. Compared with an ultra-flat graphene layer which has the weak vdW
interaction with the TM substrate, a corrugated graphene layer shows much stronger interaction which
is contributed by the strong chemical binding at the flatland area and simultaneously an extra curvature
energy due to its corrugated graphene structure, indicating that the morphology of a graphene layer on
a TM substrate is determined by the competition of binding energy and curvature energy of the
graphene/TM system. If the binding energy increase of a graphene layer changing from ultra-flat to
corrugated structure is large enough to compensate the curvature energy increase, a corrugated graphene
layer will be formed, otherwise, an ultra-flat graphene layer will be maintained.

150
Figure 11.8 Two models of graphene on the Ru(0001) surface with grey representing the graphene film
and blue representing the Ru(0001) substrate. In model I, h denotes the height oscillation degree and r
denotes the radius of the hump area.

In the following, the binding energies and curvature energies of graphene/Ru(0001) systems
are investigated first. Figure 11.9 shows the atomic configurations of graphene on the four high-
symmetric sites of the Ru(0001) surface, i.e., ATOP, FCC, HCP and Bridge as referred above. Because
of the fluctuations of graphene layers on the Ru(0001) surface with rotation angles less than 20ºare
always site-dependent, the binding energies of graphene at the four high-symmetric sites of the Ru(0001)
surface are calculated separately based on the definition:

𝐸𝐵 = (𝐸𝑇 −𝐸𝑆𝑢𝑏 − 𝐸𝐺𝑟𝑎 )/𝑁𝐶 (11.8)

where 𝐸𝑇 is the total energy of a structure with graphene lattice on a high-symmetric site of Ru(0001)
lattice, as shown in Figure 11.9(a), 𝐸𝑆𝑢𝑏 and 𝐸𝐺𝑟𝑎 are the energies of freestanding Ru(0001) and
freestanding graphene in the cell, respectively. 𝑁𝐶 denotes the number of C atoms in the graphene lattice.

151
Figure 11.9 (a) Lattice relationships of graphene on the four high-symmetric sites of the Ru(0001)
surface. (b) Binding energies of graphene with the four sites of the Ru(0001) surface as a function of
the distance. (c) Illustration of a graphene/Ru(0001) common cell with different lattice relationships
distinguished by colors. (d) Linear fitting of the curvature energies of corrugated graphene moiré
superstructures. (e) Fitting of the height of graphene layers on the Ru(0001) surface versus the length
of graphene/Ru(0001) common cell (L). (f) Curvature energies of graphene layers on the Ru(0001)
surface as a function of the rotation angle and the dashed line denotes the binding energy increase of
graphene/Ru(0001) with graphene changing from an ultra-flat to a corrugated structure.

From Figure 11.9(b), we see that the binding energies of graphene on the FCC, HCP and Bridge
sites are quite strong with optimized distances of ~2.1Å, in contrast, that on the ATOP site becomes
much weaker with an optimized distance of ~3.6 Å, which are in agreement with the height oscillation
degree of the 0ºgraphene moirésuperstructure on the Ru(0001) surface, implying that the corrugated
graphene layers are cooperative phenomena by distinct interactions between graphene and different
sites of the Ru(0001) surface. Therefore, the binding energy of graphene on the Ru(0001) surface can
be estimated by the summation of that on the four high-symmetric sites of the Ru(0001) surface. In a
graphene/Ru(0001) common cell, as shown in Figure 11.9(c), areas that correspond to graphene on
different sites are marked by different colors. There are 1/3, 1/8 and 1/8 graphene lattice locating on the
ATOP, FCC and HCP sites with the rest locating on the Bridge site, based on this, the binding energy
is estimated to be ~ -208.7 eV. Using the binding energy of graphene on the ATOP site as the reference

152
of vdW interaction, the binding energy increase of the graphene/Ru(0001) superstructure with graphene
changing from an ultra-flat to a corrugated structures is ~ 136.9 eV.

The curvature energy of a graphene/Ru(0001) superstructure is mainly from hump part of the
graphene layer, and for graphene, its curvature energy ( 𝐸𝐶 ) and curvature radius (R) satisfy a
1
relationship of 𝐸𝐶 ∝ 𝑅2 .421,422 From the model I of Figure 11.8(a), the curvature radius of a corrugated

graphene layer on a TM substrate can be obtained by the formula of (𝑅 − ℎ)2 + 𝑟 2 = 𝑅 2, where h and
ℎ 2 +𝑟 2 𝑟2
r denote the height and the radius of the hump section, respectively. Due to ℎ ≪ 𝑟, 𝑅 = 2ℎ
≈ ℎ
,
ℎ 2
and therefore 𝐸𝐶 ∝ (𝑟 2 ) . By fitting curvature energies of the corrugated graphene layers on the

Ru(0001) surface in Figure 11.9(d), we get:

ℎ 2
𝐸𝐶 = 50539.98 × ( 2 ) (11.9)
𝑟

Besides, it is found that the height of graphene layer (h) on the Ru(0001) surface and the length
(L) of graphene/Ru(0001) common cell satisfy a relationship of ℎ = ℎ0 × (𝛽 ∙ 𝐿 + 𝛾 ∙ 𝐿2 ), as shown in
Figure 11.9(e). Because of ℎ = ℎ0 happens at 𝐿 = 𝐿0 , where ℎ0 and 𝐿0 correspond to the 0º
graphene/Ru(0001) superstructure, we get

1
ℎ = ℎ0 × ((𝐿 − 𝛾 ∙ 𝐿0 ) ∙ 𝐿 + 𝛾 ∙ 𝐿2 ) (11.10)
0

and 𝛾 =-1.25e-3 by fitting with the corrugated graphene/Ru(0001) superstructures.

Assuming graphene layers on the Ru(0001) are always corrugated, Figure 11.9(f) gives
curvature energies of graphene layers obtained from Eqs. (11.8) and (11.9), with the binding energy
increase value of ~136.9 is marked by a dash line. It turns out the demarcation of the corrugated and
ultra-flat graphene structures is 19.5º, which perfectly agrees with our DFT results, confirming the
proposal that graphene layer on a TM surface is determined by the competition of binding energy and
curvature energy of a graphene/Ru(0001) superstructure. Besides, the ultra-flat and corrugated
graphene/Ru(0001) superstructures with large rotation angles are compared, as displayed in Figure
11.10, it is very clear that the ultra-flat ones always show advantages in total energy when the rotation
angle is larger than 20º, further confirming the prominent roles of the corrugated graphene/Ru(0001)
superstructures at large rotation angles.

153
Figure 11.10 Comparison the total energies of the ultra-flat and corrugated graphene layers on the
Ru(0001) substrate with larger rotation angles, and the unit of the total energy here is eV.

11.5 Rotated graphene moirésuperstructures on other transition metal surfaces

Next, the proposal is tested in other graphene/TM systems, e.g., graphene/Rh(111),


graphene/Ir(111), graphene/Pt(111). Among them, the Rh(111) surface has a similar lattice constant
with the Ru(0001) surface and therefore similar sizes of graphene/Rh(111) common cells with the
graphene/Ru(0001) ones. By DFT calculations, it is found the height oscillation degree of graphene
layer without rotation on the Rh(111) surface is also similar with that on the Ru(0001) surface, as shown
in Figure 11.11(a). Based on these similarities, we assume the height change of graphene layer on the
Rh(111) surface adopt a similar tendency with that on the Ru(0001) surface, i.e., we only substitute ℎ0
and 𝐿0 with the corresponding values of graphene/Rh(111) system and keep 𝛾 unchanged in Eq. (11.10).
However, the binding energies of graphene on the four high-symmetric sites of the Ru(111) surface
shown in Figure 11.11(b) are different with that on the Ru(0001) surface, especially that on the HCP
and FCC sites are weaker than that on the Bridge sites despite of similar graphene-Rh(111) distance at
the three sites, resulting in a smaller binding energy increase of ~79.5 eV compared with the
graphene/Ru(0001) system. The competition between curvature energy and binding energy is depicted
in Figure 11.11(c), we see the demarcation of the corrugated and ultra-flat graphene structures for
graphene/Rh(111) systems is at ~15º.Besides, the results are confirmed by DFT calculations which are
shown in Figure 11.11(d) and (e).

154
Figure 11.11 (a) Top and side views of the 0ºgraphene/Rh(111) superstructure. (b) Optimized binding
energies and distance of graphene lattice on different sites of the Rh(111) surface. (c) Curvature energies
of graphene layers on the Rh(111) surface as a function of the rotation angle and the dashed line denotes
the binding energy increase of graphene/Rh(111) with graphene changing from an ultra-flat to a
corrugated structure. (d) Height oscillation of graphene layers on the Rh(111) surface regarding to the
rotation angle by DFT calculations. (e) Stereo views of graphene/Rh(111) superstructures with different
rotation angles.

In contrast with graphene/Ru(0001) and graphene/Rh(111) superstructures, graphene present


quite different behaviors in graphene/Ir(111) and graphene/Pt(111) one, as shown in Figures 11.12 and
11.13. The height oscillation degrees of graphene layer on both the Ir(111) and Pt(111) surfaces are
very week even with a rotation angle of 0º, because that interactions of graphene and substrate are all
similarly weak with values less than 100 meV/C atom on different sites and the optimized distances are
all larger than 3.3 Å. From Figures 11.12(b) and 11.13(b), it is worth noting that, the weakest interaction
between graphene and the substrate still happens at the ATOP sites whereas the strongest one appears
at the HCP site in these weak interacted systems, which are distinguished from the strong interacted
graphene/Ru(0001) and graphene/Rh(111) systems. By comparing the differences of graphene on the
Ir(111) and Pt(111) surfaces, it is found that the binding energies difference of graphene on different
sites of the Ir(111) surface is larger than that of the Pt(111) surface, resulting in a slightly larger height
oscillation of graphene layer on the Ir(111) surface than that on the Pt(111) surface. The tendency of
graphene height evolution on the Ru(0001) substrate with 𝛾 =-1.25e-3 is still be adopted, and curvature
energy curves are obtained in Figures 11.12(c) and 11.13(c). The demarcations with graphene layer

155
changing from corrugated to ultra-flat for graphene/Ir(111) and graphene/Pt(111) systems are ~15.4º
and ~14.3º,respectively, and these results are all consistent with the DFT calculations which are shown
in 11.12(d-e) and 11.13(d-e). Besides, experiments have already reported that graphene layers on the
Pt(111) surface show height oscillation of 0.05 to 0.08 nm at rotation angles of 2º, 3ºand 6ºwhereas
within 0.03 nm at 14º,19ºand 30º,208 which agrees well with our prediction.

Figure 11.12 (a) Top and side views of the 0ºgraphene/Ir(111) superstructure. (b) Optimized binding
energies and distance of graphene lattice on different sites of the Ir(111) surface. (c) Curvature energies
of graphene layers on the Ir(111) surface as a function of the rotation angle and the dashed line denotes
the binding energy increase of graphene/ Ir(111) with graphene changing from an ultra-flat to a
corrugated structure. (d) Height oscillation of graphene layers on the Ir(111) surface regarding to the
rotation angle by DFT calculations. (e) Stereo views of graphene/ Ir(111) superstructures with different
rotation angles.

156
Figure 11.13 (a) Top and side views of the 0ºgraphene/Pt(111) superstructure. (b) Optimized binding
energies and distance of graphene lattice on different sites of the Pt(111) surface. (c) Curvature energies
of graphene layers on the Pt(111) surface as a function of the rotation angle and the dashed line denotes
the binding energy increase of graphene/Pt(111) with graphene changing from an ultra-flat to a
corrugated structure. (d) Height oscillation of graphene layers on the Pt(111) surface regarding to the
rotation angle by DFT calculations. (e) Stereo views of graphene/ Pt(111) superstructures with different
rotation angles.

All above discussions prove that the morphology of a graphene layer on a TM substrate is
determined by the competition of binding energy and curvature energy of the graphene/TM system, and
based on this, the demarcation of the corrugated and ultra-flat graphene layers can be obtained based
on the 0ºgraphene/TM superstructure and unit cells of graphene on the four high-symmetric sites of the
TM substrate instead of optimizing graphene/TM superstructures with all rotation angles, which saves
the computing resources to a large extent.

Finally, the proposal is also used to predict graphene behaviors in two other systems, i.e.,
graphene/Re(0001) and graphene/Pd(111), as shown in Figure 11.14. Graphene layer on the Re(0001)
surface presents a large height oscillation at 0º, however due to the small binding energy increase, the
corrugation of graphene layers can only be maintained to ~ 4.6º. Behaviors of graphene layers on the
Pd(111) surface are very similar with that on the Ir(111) and Pt(111) surfaces, although the height
oscillation is very weak, the very small binding energy different can only maintain the corrugation of
graphene layers to ~9.2º.

157
Figure 11.14 (a, d) Different views of the 0º graphene/Re(0001) and the 0º graphene/Pd(111)
superstructure. (b, e) Optimized binding energies and distance of graphene lattice on different sites of
the Re(0001) and the Pd(111) surface. (c) Curvature energies of graphene layers on the Re(0001) and
the Pd(111) surface as a function of the rotation angle and the dashed line denotes the binding energy
increase of systems with graphene changing from an ultra-flat to a corrugated structure.

Table 11.1 lists the lattice relationships between graphene and various TM surfaces and
summarizes the behaviors of graphene layers on these TM surfaces. On one hand, the smaller lattice
mismatch between graphene and TM lattices, the larger size of graphene/TM common cell and the
smaller curvature energy of graphene layer at a same rotation angle. On the other hand, a large binding
energy difference of graphene on different sites of a TM substrate generally implies a large graphene
height oscillation and a large binding energy increase, which means the curvature energy can be
compensate to a large extend. For example, graphene on the Ru(0001), Ir(111) and Re(0001) surfaces
all shows a very large height oscillation at 0º, moreover, graphene/Ru(0001) superstructures have the
largest common cells and strongest binding energy increase, and therefore the corrugation of graphene
on the Ru(0001) surface can be maintain the largest rotation angle range of 0~19.5º.

158
Table 11.1 Summary of lattice parameters of graphene/TM superstructures and graphene behaviors on
different TM surfaces.

Graphene/ Graphene/ Graphene/ Graphene/ Graphene/ Graphene/


Ru(0001) Rh(111) Ir(111) Pt(111) Re(0001) Pd(111)
Lattice constant
2.706 2.690 2.715 2.775 2.760 2.751
of TMs (Å)
Lattice
9.08 9.35 10.35 12.79 12.20 11.84
Mismatch (%)

L0 (Å) 27.08 26.07 23.52 18.92 19.87 20.49

h0 (Å) 1.56 1.53 0.44 0.28 1.53 0.51

Binding energy
136.90 79.50 8.30 3.20 18.50 5.50
change (eV)

Demarcation (º) 19.50 15.00 15.40 14.30 4.60 9.20

11.6 Summary

For CVD growth of graphene on a TM surface which has a large lattice mismatch with graphene,
graphene structures show two kinds of behaviors depending on the relative orientation of graphene on
the substrate. Using graphene/Ru(0001) system as an example, there are: (i) the ultra-flat graphene
layers with height variations less than 0.1 Å for rotation angles greater than 20 degree, whose structural
and electronic properties are very uniform everywhere, and (ii) the highly corrugated graphene moiré
superstructures with height variations from 0.4 to 1.5 Å for rotation angles less than 20 degree, whose
electronic properties are highly modulated by the anisotropic interaction with the substrate. Besides,
these corrugated graphene/Ru(0001) superstructures can serve as templates for the production of size-
tunable TM clusters.

The formation mechanisms of the two morphologies of graphene layers are explored, and it
revealed that the morphology of graphene on the TM substrate is determined by the competition of
graphene - substrate binding energy and the curvature energy. For an ultra-flat graphene film on a TM
surface, the interaction is contributed by the pure vdW at all sites, while for a corrugated graphene moiré
superstructure on a TM surface, the interaction is composed by the weak vdW interaction at ATOP site
and strong chemical binding at FCC, HCP and Bridge sites, therefore the corrugated graphene moiré
superstructure shows a stronger interaction with the substrate and also a higher curvature energy. By
fitting the graphene/Ru(0001) system, the evolution rules of binding energy and curvature energy are
formulated, which are also proved to be suitable for other graphene/TM systems. Based on these rules,
the critical rotation angle of graphene changing from corrugated to ultra-flat on a TM surface can be

159
estimated effectively. Our work summarizes the behaviors of graphene layers on various TM surfaces
and offers a theoretical guidance for the design of graphene/TM superstructures.

160
Chapter 12 Conclusions and further research plans
With more than 10 years’ efforts on the CVD growth of 2D materials, two routes have been
developed in synthesizing WSSC 2D materials: (i) growing from only one nucleus of a 2D material on
a substrate and (ii) seamless stitching of unidirectionally aligned 2D material islands. In Chapter 1-2,
both the current experimental work and theoretical understanding on the growth of 2D single crystals
were reviewed. Despite there is an epitaxial relationship between a 2D material and its substrate during
the nucleation stage, the crystallographic lattice of the 2D island can be maintained when it grows across
the grain boundaries of the underlying substrate due the weak interactions between the 2D material and
its substrate, and therefore synthesis of single crystalline 2D materials on polycrystalline substrates is
possible via route (i). However, a very low nucleation density (ideally only one nucleus on the whole
substrate) or a very high nucleation barrier is needed to growth WSSC 2D material, which requires a
delicate experimental setup. In addition, it usually takes hours to grow a WSSC 2D material via route
(i) because only one nucleus is allowed to be formed and grow. In contrast, the route (ii) offers a more
cost-effective method because a large amount of 2D nuclei grow simultaneously, while in this method
proper substrates that can template the growth of unidirectionally aligned 2D islands are required.

In this dissertation, a systematic theoretical study on the alignment of 2D materials on an


arbitrary TM substrate has been presented firstly. The alignment mechanisms of 2D materials on the
substrate are revealed at atomic scale. Because single crystalline Cu(111) surface is one of the most
promising substrates for the synthesis of WSSC 6-fold symmetric 2D materials, a new contact-free
annealing method of transforming commercial polycrystalline Cu foils into single-crystalline ones has
also been proposed in this dissertation. In the final part of this dissertation, we presented a systematic
study on the formation mechanism applications of graphene super moirépattern on various TM surfaces.
Because of the lattice mismatch between graphene and the substrate, graphene super moirépatterns
usually appear on most TM substrates after CVD growth, which leads to the deformation of the
graphene layer and new properties.

For the study on the alignment of 2D materials on an arbitrary TM substrate, the substrates are
classified into low-index high symmetric and high-index low symmetric ones. On a high symmetric
substrate, 2D materials usually show multi-orientations. In the Chapter 4, the symmetries of various
FCC TM surfaces and various 2D materials are introduced. It is found that a high symmetric direction
of a 2D material prefers to align along a high symmetric direction of the substrate, based on which we
proposed that unidirectionally aligned 2D islands can be synthesized on a substrate only if the symmetry
group of the substrate is a subgroup of that of the 2D material. For instance, C6V 2D material shows 1
preferential orientation on C6V FCC{111} surface and C2V FCC{110} surface, and 2 preferential

161
orientation on C4V FCC{100} surface; C3V 2D materials show 2, 4 and 2 preferential orientation on C6V
FCC{111}, C4V FCC{100} and C2V FCC{110} surface, respectively.

Different from low-index substrates, high-index substrates show much lower symmetries due
to the existence of step edges. Moreover, in principle, there are unlimited number of high-index
substrates. As have been revealed in Chapter 4, such low symmetric substrates might be more promising
in the synthesis of unidirectionally aligned 2D materials. In the Chapter 5-8, the alignment of C6V
graphene and C3V hBN on an arbitrary low symmetric Cu surface are explored. It is found that the
orientation of a 2D island on a low symmetric surface is determined by the interaction between the Cu
step edge and the edge of the 2D island, and well-aligned graphene islands can be always achieved on
the ideal low symmetric Cu surfaces which have unidirectional step edges with a constant direction.
However, in practice there are always certain variations of step edge direction because of the existence
of surface roughness. We further revealed that well-aligned graphene islands can still be maintained on
Cu{111}-based low symmetric surface and Cu{110}-based low symmetric surface with <211>
dominated meandering step edges. For the C3V hBN, its alignment on low symmetric Cu surfaces is
more complicated, and also affected by the ambient condition due to its binary composition. On
Cu{111}-based low symmetric surfaces, unidirectional hBN islands can be maintained with step edge
direction changing in a range of ± 19.10°with respect to <110> direction at 𝜇𝑁 =−8.31 ~ −9.61 eV,
while it can be maintained in a larger range with ±29.99°with respect to <110> direction at 𝜇𝑁 =−9.62
~ −10.84 eV. Cu{110}-based low symmetric surfaces with <211> dominated steps are also suitable for
hBN epitaxial growth. Moreover, we found that a high-index surface with a larger tilted angle from its
low-index terrace can tolerate higher surface roughness and keep the change of step edge direction in a
small range, and therefore is more preferred for the growth of unidrtional 2D islands, as compared to a
high-index surface with a lower tilted angle from its terrace. Except for graphene and hBN, the aligment
of WS2 islands on Au{111}-based low symmetric surfaces are also been exploerd, and the results are
similar to those of hBN on Cu{111}-based low symmetric surfaces because both WS2 and hBN are 3-
fold symmetric and have a binary composition.

For the fabrication of large-size single-crystalline Cu(111) foils, in collaboration with


experimental groups, we successfully realized single crystalline Cu{111}<112> foils with the size up
to 32 cm2 by a contact-free annealing method and explored the transformation mechanism from
polycrystalline Cu foils into single crystals at the atomic scale, as discussed in the Chapter 10. It is
revealed that the nucleation of the Cu{111}<112> starts from a Cu{112}<111> grain of the raw Cu foil,
and the driving force of the grain rotation from Cu{112}<111> to Cu{111}<112> is the decrease of
surface energy and the rotation is mediated by the gliding of the stacking faults that produced in the
high annealing temperature.

162
In Chapter 11, we systematically investigated the structures of graphene moirépatterns on
various substrates. Graphene film on a mismatched TM surfaces usually show a highly corrugated
graphene moirésuperstructure at a small rotation angle due to the non-uniform interaction between
graphene and the substrate, while the anisotropy decays and an ultra-flat graphene structure will appear
with the increase of rotation angle. DFT calculations prove that the rotation-dependent moiré
superstructures on the Ru(0001) surfaces are able to template the fabrication of the matrices of size-
tunable metal clusters. Moreover, we found that the competition of graphene - substrate binding energy
and the curvature energy in graphene is responsible for the formation of two kinds of graphene
structures, and we further formulate the evolution of the binding energy and curvature energy as a
function of rotation angle, and based on these rules, the transition from corrugated to ultra-flat graphene
film and further the morphology of graphene layers can be predicted easily. This study is very helpful
for designing an appropriate system of graphene/TM substrate for the synthesis of TM clusters with
desired structures.

To summarize, in this dissertation, we theoretically studied the alignment of graphene, hBN


and TMDCs on both high-index and low-index TM substrates, and the novel behaviors of the grown
graphene layers on various TM surfaces. In addition, we also introduced a new method of fabricating
large-size single-crystalline Cu foils and revealed the corresponding mechanism at atomic scale. All our
theoretical results are well consistent with existing experimental observations. We believe that our study
can greatly promote the controllable synthesis of 2D materials and various WSSC 2D materials will be
synthesized in the near future. It should be noted that there are still lots of scientific problems in this
area that need to be further explored. For example, the study on the alignment of TMDC on substrates
is quite preliminary and researches on the alignment of TMDC islands on insulated surfaces, such as
hBN and Al2O3, are needed. Besides, Surface reconstruction or step reconstruction is also a very
important issue for the growth of 2D materials, which also needs to be considered in the future work.
Moreover, how 2D materials growing across the step edges of a substrate also needs to be investigated,
which is also of critical importance in choosing proper substrates for the synthesis of 2D materials.

163
References
(1) Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.; Zhang, Y.; Dubonos, S. V.;
Grigorieva, I. V.; Firsov, A. A. Electric Field Effect in Atomically Thin Carbon Films. Science
2004, 306 (5696), 666.
(2) Berger, C.; Song, Z.; Li, X.; Wu, X.; Brown, N.; Naud, C.; Mayou, D.; Li, T.; Hass, J.;
Marchenkov, A. N.et al. Electronic Confinement and Coherence in Patterned Epitaxial
Graphene. Science 2006, 312 (5777), 1191.
(3) Geim, A. K.; Novoselov, K. S. The rise of graphene. Nat. Mater. 2007, 6, 183.
(4) Castro Neto, A. H.; Guinea, F.; Peres, N. M. R.; Novoselov, K. S.; Geim, A. K. The electronic
properties of graphene. Rev. Mod. Phys. 2009, 81 (1), 109.
(5) Bonaccorso, F.; Sun, Z.; Hasan, T.; Ferrari, A. C. Graphene photonics and optoelectronics. Nat.
Photonics 2010, 4, 611.
(6) Lemme, M. C.; Echtermeyer, T. J.; Baus, M.; Kurz, H. A Graphene Field-Effect Device. IEEE
Electron Device Letters 2007, 28 (4), 282.
(7) Liang, X.; Fu, Z.; Chou, S. Y. Graphene Transistors Fabricated via Transfer-Printing In Device
Active-Areas on Large Wafer. Nano Lett. 2007, 7 (12), 3840.
(8) Wang, Q. H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J. N.; Strano, M. S. Electronics and
optoelectronics of two-dimensional transition metal dichalcogenides. Nat. Nanotech. 2012, 7,
699.
(9) Butler, S. Z.; Hollen, S. M.; Cao, L.; Cui, Y.; Gupta, J. A.; Gutiérrez, H. R.; Heinz, T. F.; Hong,
S. S.; Huang, J.; Ismach, A. F.et al. Progress, Challenges, and Opportunities in Two-
Dimensional Materials Beyond Graphene. ACS Nano 2013, 7 (4), 2898.
(10) Chhowalla, M.; Shin, H. S.; Eda, G.; Li, L. J.; Loh, K. P.; Zhang, H. The chemistry of two-
dimensional layered transition metal dichalcogenide nanosheets. Nat. Chem. 2013, 5 (4), 263.
(11) Xu, M.; Liang, T.; Shi, M.; Chen, H. Graphene-Like Two-Dimensional Materials. Chem. Rev.
2013, 113 (5), 3766.
(12) Cong, C.; Shang, J.; Wu, X.; Cao, B.; Peimyoo, N.; Qiu, C.; Sun, L.; Yu, T. Synthesis and
Optical Properties of Large-Area Single-Crystalline 2D Semiconductor WS2 Monolayer from
Chemical Vapor Deposition. Advanced Optical Materials 2014, 2 (2), 131.
(13) Sun, Z.; Chang, H. Graphene and Graphene-like Two-Dimensional Materials in Photodetection:
Mechanisms and Methodology. ACS Nano 2014, 8 (5), 4133.
(14) Xia, F.; Wang, H.; Xiao, D.; Dubey, M.; Ramasubramaniam, A. Two-dimensional material
nanophotonics. Nat. Photonics 2014, 8, 899.
(15) Novoselov, K. S.; Mishchenko, A.; Carvalho, A.; Castro Neto, A. H. 2D materials and van der
Waals heterostructures. Science 2016, 353 (6298), aac9439.

164
(16) Coleman, J. N.; Lotya, M.; O’Neill, A.; Bergin, S. D.; King, P. J.; Khan, U.; Young, K.; Gaucher,
A.; De, S.; Smith, R. J.et al. Two-Dimensional Nanosheets Produced by Liquid Exfoliation of
Layered Materials. Science 2011, 331 (6017), 568.
(17) Castellanos-Gomez, A.; Barkelid, M.; Goossens, A. M.; Calado, V. E.; van der Zant, H. S. J.;
Steele, G. A. Laser-Thinning of MoS₂: On Demand Generation of a Single-Layer
Semiconductor. Nano Lett. 2012, 12 (6), 3187.
(18) Huo, C.; Yan, Z.; Song, X.; Zeng, H. 2D materials via liquid exfoliation: a review on fabrication
and applications. Sci. Bull. 2015, 60 (23), 1994.
(19) Tao, H.; Zhang, Y.; Gao, Y.; Sun, Z.; Yan, C.; Texter, J. Scalable exfoliation and dispersion of
two-dimensional materials – an update. Phys. Chem. Chem. Phys. 2017, 19 (2), 921.
(20) Smith, R. J.; King, P. J.; Lotya, M.; Wirtz, C.; Khan, U.; De, S.; O'Neill, A.; Duesberg, G. S.;
Grunlan, J. C.; Moriarty, G.et al. Large-Scale Exfoliation of Inorganic Layered Compounds in
Aqueous Surfactant Solutions. Adv. Mater. 2011, 23 (34), 3944.
(21) Zhang, Y.; Zhang, L.; Zhou, C. Review of Chemical Vapor Deposition of Graphene and Related
Applications. Acc. Chem. Res. 2013, 46 (10), 2329.
(22) Strupinski, W.; Grodecki, K.; Wysmolek, A.; Stepniewski, R.; Szkopek, T.; Gaskell, P. E.;
Grüneis, A.; Haberer, D.; Bozek, R.; Krupka, J.et al. Graphene Epitaxy by Chemical Vapor
Deposition on SiC. Nano Lett. 2011, 11 (4), 1786.
(23) Sutter, P.; Lahiri, J.; Zahl, P.; Wang, B.; Sutter, E. Scalable Synthesis of Uniform Few-Layer
Hexagonal Boron Nitride Dielectric Films. Nano Lett. 2013, 13 (1), 276.
(24) Ago, H.; Ito, Y.; Mizuta, N.; Yoshida, K.; Hu, B.; Orofeo, C. M.; Tsuji, M.; Ikeda, K.-i.; Mizuno,
S. Epitaxial Chemical Vapor Deposition Growth of Single-Layer Graphene over Cobalt Film
Crystallized on Sapphire. ACS Nano 2010, 4 (12), 7407.
(25) Li, X.; Magnuson, C. W.; Venugopal, A.; An, J.; Suk, J. W.; Han, B.; Borysiak, M.; Cai, W.;
Velamakanni, A.; Zhu, Y.et al. Graphene films with large domain size by a two-step chemical
vapor deposition process. Nano Lett. 2010, 10 (11), 4328.
(26) Mattevi, C.; Kim, H.; Chhowalla, M. A review of chemical vapour deposition of graphene on
copper. J. Mater. Chem. 2011, 21 (10), 3324.
(27) Sutter, P.; Lahiri, J.; Albrecht, P.; Sutter, E. Chemical Vapor Deposition and Etching of High-
Quality Monolayer Hexagonal Boron Nitride Films. ACS Nano 2011, 5 (9), 7303.
(28) Kim, K. K.; Hsu, A.; Jia, X.; Kim, S. M.; Shi, Y.; Hofmann, M.; Nezich, D.; Rodriguez-Nieva,
J. F.; Dresselhaus, M.; Palacios, T.et al. Synthesis of Monolayer Hexagonal Boron Nitride on
Cu Foil Using Chemical Vapor Deposition. Nano Lett. 2012, 12 (1), 161.
(29) Lee, Y.-H.; Zhang, X.-Q.; Zhang, W.; Chang, M.-T.; Lin, C.-T.; Chang, K.-D.; Yu, Y.-C.;
Wang, J. T.-W.; Chang, C.-S.; Li, L.-J.et al. Synthesis of Large-Area MoS₂ Atomic Layers with
Chemical Vapor Deposition. Adv. Mater. 2012, 24 (17), 2320.

165
(30) Wong, S. L.; Liu, H.; Chi, D. Recent progress in chemical vapor deposition growth of two-
dimensional transition metal dichalcogenides. Prog. Cryst. Growth Charact. Mater. 2016, 62
(3), 9.
(31) Narula, U.; Tan, C. M.; Lai, C. S. Copper induced synthesis of graphene using amorphous
carbon. Microelectron. Reliab. 2016, 61, 87.
(32) Glavin, N. R.; Jespersen, M. L.; Check, M. H.; Hu, J.; Hilton, A. M.; Fisher, T. S.; Voevodin,
A. A. Synthesis of few-layer, large area hexagonal-boron nitride by pulsed laser deposition.
Thin Solid Films 2014, 572, 245.
(33) Yang, Z.; Hao, J. Progress in pulsed laser deposited two-dimensional layered materials for
device applications. J. Mater. Chem. C 2016, 4 (38), 8859.
(34) Xia, J.; Li, X.-Z.; Huang, X.; Mao, N.; Zhu, D.-D.; Wang, L.; Xu, H.; Meng, X.-M. Physical
vapor deposition synthesis of two-dimensional orthorhombic SnS flakes with strong
angle/temperature-dependent Raman responses. Nanoscale 2016, 8 (4), 2063.
(35) Zhang, X.; Zhang, Y.; Yu, B.-B.; Yin, X.-L.; Jiang, W.-J.; Jiang, Y.; Hu, J.-S.; Wan, L.-J.
Physical vapor deposition of amorphous MoS2 nanosheet arrays on carbon cloth for highly
reproducible large-area electrocatalysts for the hydrogen evolution reaction. Journal of
Materials Chemistry A 2015, 3 (38), 19277.
(36) Wofford, J. M.; Oliveira, M. H.; Schumann, T.; Jenichen, B.; Ramsteiner, M.; Jahn, U.; Fölsch,
S.; Lopes, J. M. J.; Riechert, H. Molecular beam epitaxy of graphene on ultra-smooth nickel:
growth mode and substrate interactions. New J. Phys. 2014, 16 (9), 093055.
(37) Nakhaie, S.; Wofford, J. M.; Schumann, T.; Jahn, U.; Ramsteiner, M.; Hanke, M.; Lopes, J. M.
J.; Riechert, H. Synthesis of atomically thin hexagonal boron nitride films on nickel foils by
molecular beam epitaxy. Appl. Phys. Lett. 2015, 106 (21), 213108.
(38) Chen, M.-W.; Ovchinnikov, D.; Lazar, S.; Pizzochero, M.; Whitwick, M. B.; Surrente, A.;
Baranowski, M.; Sanchez, O. L.; Gillet, P.; Plochocka, P.et al. Highly Oriented Atomically
Thin Ambipolar MoSe₂ Grown by Molecular Beam Epitaxy. ACS nano 2017, 11 (6), 6355.
(39) Cheng, T. S.; Summerfield, A.; Mellor, C. J.; Davies, A.; Khlobystov, A. N.; Eaves, L.; Foxon,
C. T.; Beton, P. H.; Novikov, S. V. High-temperature molecular beam epitaxy of hexagonal
boron nitride layers. Journal of Vacuum Science & Technology B 2018, 36 (2), 02D103.
(40) He, Q.; Li, P.; Wu, Z.; Yuan, B.; Luo, Z.; Yang, W.; Liu, J.; Cao, G.; Zhang, W.; Shen, Y.et al.
Molecular Beam Epitaxy Scalable Growth of Wafer-Scale Continuous Semiconducting
Monolayer MoTe2 on Inert Amorphous Dielectrics. Adv. Mater. 2019, 31 (32), 1901578.
(41) Liu, H. J.; Jiao, L.; Xie, L.; Yang, F.; Chen, J. L.; Ho, W. K.; Gao, C. L.; Jia, J. F.; Cui, X. D.;
Xie, M. H. Molecular-beam epitaxy of monolayer and bilayer WSe 2 : a scanning tunneling
microscopy/spectroscopy study and deduction of exciton binding energy. 2D Materials 2015,
2 (3), 034004.

166
(42) Hu, B.; Ago, H.; Ito, Y.; Kawahara, K.; Tsuji, M.; Magome, E.; Sumitani, K.; Mizuta, N.; Ikeda,
K.-i.; Mizuno, S. Epitaxial growth of large-area single-layer graphene over Cu(111)/sapphire
by atmospheric pressure CVD. Carbon 2012, 50 (1), 57.
(43) Li, H.; Li, Y.; Aljarb, A.; Shi, Y.; Li, L.-J. Epitaxial Growth of Two-Dimensional Layered
Transition-Metal Dichalcogenides: Growth Mechanism, Controllability, and Scalability. Chem.
Rev. 2018, 118 (13), 6134.
(44) N'Diaye, A. T.; Coraux, J.; Plasa, T. N.; Busse, C.; Michely, T. Structure of epitaxial graphene
on Ir(111). New J. Phys. 2008, 10 (4), 16.
(45) Huang, H.; Chen, W.; Chen, S.; Wee, A. T. S. Bottom-up Growth of Epitaxial Graphene on
6H-SiC(0001). ACS Nano 2008, 2 (12), 2513.
(46) Yang, P. C.; Prater, J. T.; Liu, W.; Glass, J. T.; Davis, R. F. The formation of epitaxial hexagonal
boron nitride on nickel substrates. J. Electron. Mater. 2005, 34 (12), 1558.
(47) Jichen, D.; Feng, D. The epitaxy of 2D materials growth. Research Square 2020,
DOI:10.21203/rs.3.rs-40179/v1 10.21203/rs.3.rs-40179/v1.
(48) Markov, I.; Stoyanov, S. Mechanisms of epitaxial growth. Contemporary Physics 1987, 28 (3),
267.
(49) Pashley, D. W. The nucleation, growth, structure and epitaxy of thin surface films. Advances
in Physics 1965, 14 (55), 327.
(50) Kim, Y.-I.; Si, W.; Woodward, P. M.; Sutter, E.; Park, S.; Vogt, T. Epitaxial Thin-Film
Deposition and Dielectric Properties of the Perovskite Oxynitride BaTaO2N. Chem. Mater.
2007, 19 (3), 618.
(51) In Surface and Interface Science, 2013, DOI:10.1002/9783527680566.ch20
10.1002/9783527680566.ch20.
(52) Hull, R.; Bean, J. C. Misfit dislocations in lattice-mismatched epitaxial films. Crit. Rev. Solid
State Mater. Sci. 1992, 17 (6), 507.
(53) Chambers, S. A. Epitaxial growth and properties of thin film oxides. Surf. Sci. Rep. 2000, 39
(5), 105.
(54) Moridi, A.; Ruan, H.; Zhang, L. C.; Liu, M. Residual stresses in thin film systems: Effects of
lattice mismatch, thermal mismatch and interface dislocations. International Journal of Solids
and Structures 2013, 50 (22), 3562.
(55) Yamauchi, R.; Hamasaki, Y.; Shibuya, T.; Saito, A.; Tsuchimine, N.; Koyama, K.; Matsuda,
A.; Yoshimoto, M. Layer matching epitaxy of NiO thin films on atomically stepped sapphire
(0001) substrates. Sci. Rep. 2015, 5 (1), 14385.
(56) Deng, J.; Dong, K.; Yang, P.; Peng, Y.; Ju, G.; Hu, J.; Chow, G. M.; Chen, J. Large lattice
mismatch effects on the epitaxial growth and magnetic properties of FePt films. Journal of
Magnetism and Magnetic Materials 2018, 446, 125.

167
(57) Li, G.; Zhang, Y.-Y.; Guo, H.; Huang, L.; Lu, H.; Lin, X.; Wang, Y.-L.; Du, S.; Gao, H.-J.
Epitaxial growth and physical properties of 2D materials beyond graphene: from monatomic
materials to binary compounds. Chem. Soc. Rev. 2018, 47 (16), 6073.
(58) Kondo, D.; Hayashi, K.; Kataoka, M.; Iwai, T.; Sato, S. 2016 Compound Semiconductor Week
(CSW) [Includes 28th International Conference on Indium Phosphide & Related Materials
(IPRM) & 43rd International Symposium on Compound Semiconductors (ISCS), 2016; p 1.
(59) Davies, A.; Albar, J. D.; Summerfield, A.; Thomas, J. C.; Cheng, T. S.; Korolkov, V. V.;
Stapleton, E.; Wrigley, J.; Goodey, N. L.; Mellor, C. J.et al. Lattice-Matched Epitaxial
Graphene Grown on Boron Nitride. Nano Lett. 2018, 18 (1), 498.
(60) Yu, Q.; Lian, J.; Siriponglert, S.; Li, H.; Chen, Y. P.; Pei, S.-S. Graphene segregated on Ni
surfaces and transferred to insulators. Appl. Phys. Lett. 2008, 93 (11), 113103.
(61) Li, X.; Cai, W.; An, J.; Kim, S.; Nah, J.; Yang, D.; Piner, R.; Velamakanni, A.; Jung, I.; Tutuc,
E.et al. Large-Area Synthesis of High-Quality and Uniform Graphene Films on Copper Foils.
Science 2009, 324 (5932), 1312.
(62) Emtsev, K. V.; Bostwick, A.; Horn, K.; Jobst, J.; Kellogg, G. L.; Ley, L.; McChesney, J. L.;
Ohta, T.; Reshanov, S. A.; Röhrl, J.et al. Towards wafer-size graphene layers by atmospheric
pressure graphitization of silicon carbide. Nat. Mater. 2009, 8, 203.
(63) Reina, A.; Thiele, S.; Jia, X.; Bhaviripudi, S.; Dresselhaus, M. S.; Schaefer, J. A.; Kong, J.
Growth of large-area single- and Bi-layer graphene by controlled carbon precipitation on
polycrystalline Ni surfaces. Nano Res. 2009, 2 (6), 509.
(64) Sutter, E.; Albrecht, P.; Sutter, P. Graphene growth on polycrystalline Ru thin films. Appl. Phys.
Lett. 2009, 95 (13), 133109.
(65) Lee, S.; Lee, K.; Zhong, Z. Wafer Scale Homogeneous Bilayer Graphene Films by Chemical
Vapor Deposition. Nano Lett. 2010, 10 (11), 4702.
(66) Wu, T.; Zhang, X.; Yuan, Q.; Xue, J.; Lu, G.; Liu, Z.; Wang, H.; Wang, H.; Ding, F.; Yu, Q.et
al. Fast growth of inch-sized single-crystalline graphene from a controlled single nucleus on
Cu-Ni alloys. Nat. Mater. 2016, 15 (1), 43.
(67) Jin, S.; Huang, M.; Kwon, Y.; Zhang, L.; Li, B.-W.; Oh, S.; Dong, J.; Luo, D.; Biswal, M.;
Cunning, B. V.et al. Colossal grain growth yields single-crystal metal foils by contact-free
annealing. Science 2018, DOI:10.1126/science.aao3373 10.1126/science.aao3373.
(68) Lee, J.-H.; Lee, E. K.; Joo, W.-J.; Jang, Y.; Kim, B.-S.; Lim, J. Y.; Choi, S.-H.; Ahn, S. J.; Ahn,
J. R.; Park, M.-H.et al. Wafer-Scale Growth of Single-Crystal Monolayer Graphene on
Reusable Hydrogen-Terminated Germanium. Science 2014, 344 (6181), 286.
(69) Xu, X.; Zhang, Z.; Dong, J.; Yi, D.; Niu, J.; Wu, M.; Lin, L.; Yin, R.; Li, M.; Zhou, J.et al.
Ultrafast epitaxial growth of metre-sized single-crystal graphene on industrial Cu foil. Sci. Bull.
2017, 62 (15), 1074.

168
(70) Kumar, A.; Huei, C. Synthesis and Biomedical Applications of Graphene: Present and Future
Trends. 2013, DOI:10.5772/55728 10.5772/55728.
(71) Yan, Z.; Lin, J.; Peng, Z.; Sun, Z.; Zhu, Y.; Li, L.; Xiang, C.; Samuel, E. L.; Kittrell, C.; Tour,
J. M. Toward the Synthesis of Wafer-Scale Single-Crystal Graphene on Copper Foils. ACS
Nano 2012, 6 (10), 9110.
(72) Li, X.; Magnuson, C. W.; Venugopal, A.; Tromp, R. M.; Hannon, J. B.; Vogel, E. M.; Colombo,
L.; Ruoff, R. S. Large-area graphene single crystals grown by low-pressure chemical vapor
deposition of methane on copper. J. Am. Chem. Soc. 2011, 133 (9), 2816.
(73) Hao, Y.; Bharathi, M. S.; Wang, L.; Liu, Y.; Chen, H.; Nie, S.; Wang, X.; Chou, H.; Tan, C.;
Fallahazad, B.et al. The Role of Surface Oxygen in the Growth of Large Single-Crystal
Graphene on Copper. Science 2013, 342 (6159), 720.
(74) Braeuninger-Weimer, P.; Brennan, B.; Pollard, A. J.; Hofmann, S. Understanding and
Controlling Cu-Catalyzed Graphene Nucleation: The Role of Impurities, Roughness, and
Oxygen Scavenging. Chem. Mater. 2016, 28 (24), 8905.
(75) Vlassiouk, I. V.; Stehle, Y.; Pudasaini, P. R.; Unocic, R. R.; Rack, P. D.; Baddorf, A. P.; Ivanov,
I. N.; Lavrik, N. V.; List, F.; Gupta, N.et al. Evolutionary selection growth of two-dimensional
materials on polycrystalline substrates. Nat. Mater. 2018, 17 (4), 318.
(76) Dai, J.; Wang, D.; Zhang, M.; Niu, T.; Li, A.; Ye, M.; Qiao, S.; Ding, G.; Xie, X.; Wang, Y.et
al. How Graphene Islands Are Unidirectionally Aligned on the Ge(110) Surface. Nano Lett.
2016, 16 (5), 3160.
(77) Deng, B.; Pang, Z.; Chen, S.; Li, X.; Meng, C.; Li, J.; Liu, M.; Wu, J.; Qi, Y.; Dang, W.et al.
Wrinkle-Free Single-Crystal Graphene Wafer Grown on Strain-Engineered Substrates. ACS
Nano 2017, 11 (12), 12337.
(78) Nguyen, V. L.; Shin, B. G.; Duong, D. L.; Kim, S. T.; Perello, D.; Lim, Y. J.; Yuan, Q. H.;
Ding, F.; Jeong, H. Y.; Shin, H. S.et al. Seamless stitching of graphene domains on polished
copper (111) foil. Adv. Mater. 2015, 27 (8), 1376.
(79) Wang, Z.-J.; Dong, J.; Li, L.; Dong, G.; Cui, Y.; Yang, Y.; Wei, W.; Blume, R.; Li, Q.; Wang,
L.et al. The Coalescence Behavior of Two-Dimensional Materials Revealed by Multiscale In
Situ Imaging during Chemical Vapor Deposition Growth. ACS Nano 2020,
DOI:10.1021/acsnano.9b08221 10.1021/acsnano.9b08221.
(80) Li, J.; Li, Y.; Yin, J.; Ren, X.; Liu, X.; Jin, C.; Guo, W. Growth of Polar Hexagonal Boron
Nitride Monolayer on Nonpolar Copper with Unique Orientation. Small 2016, 12 (27), 3645.
(81) Li, P.; Wei, W.; Zhang, M.; Mei, Y.; Chu, P. K.; Xie, X.; Yuan, Q.; Di, Z. Wafer-scale growth
of single-crystal graphene on vicinal Ge(001) substrate. Nano Today 2020, 34, 100908.

169
(82) Hou, Y.; Wang, B.; Zhan, L.; Qing, F.; Wang, X.; Niu, X.; Li, X. Surface crystallographic
structure insensitive growth of oriented graphene domains on Cu substrates. Mater. Today 2020,
36, 10.
(83) Xia, K.; Artyukhov, V. I.; Sun, L.; Zheng, J.; Jiao, L.; Yakobson, B. I.; Zhang, Y. Growth of
large-area aligned pentagonal graphene domains on high-index copper surfaces. Nano Res.
2016, 9 (7), 2182.
(84) Wang, L.; Xu, X.; Zhang, L.; Qiao, R.; Wu, M.; Wang, Z.; Zhang, S.; Liang, J.; Zhang, Z.;
Zhang, Z.et al. Epitaxial growth of a 100-square-centimetre single-crystal hexagonal boron
nitride monolayer on copper. Nature 2019, 570 (7759), 91.
(85) Wu, M.; Zhang, Z.; Xu, X.; Zhang, Z.; Duan, Y.; Dong, J.; Qiao, R.; You, S.; Wang, L.; Qi,
J.et al. Seeded growth of large single-crystal copper foils with high-index facets. Nature 2020,
581 (7809), 406.
(86) Li, Y.; Sun, L.; Chang, Z.; Liu, H.; Wang, Y.; Liang, Y.; Chen, B.; Ding, Q.; Zhao, Z.; Wang,
R.et al. Large Single-Crystal Cu Foils with High-Index Facets by Strain-Engineered
Anomalous Grain Growth. Adv. Mater. 2020, 32 (29), e2002034.
(87) Geng, D.; Wu, B.; Guo, Y.; Huang, L.; Xue, Y.; Chen, J.; Yu, G.; Jiang, L.; Hu, W.; Liu, Y.
Uniform hexagonal graphene flakes and films grown on liquid copper surface. Proc. Natl. Acad.
Sci. U. S. A. 2012, 109 (21), 7992.
(88) Wu, Y. A.; Fan, Y.; Speller, S.; Creeth, G. L.; Sadowski, J. T.; He, K.; Robertson, A. W.; Allen,
C. S.; Warner, J. H. Large Single Crystals of Graphene on Melted Copper Using Chemical
Vapor Deposition. ACS Nano 2012, 6 (6), 5010.
(89) Zeng, M.; Wang, L.; Liu, J.; Zhang, T.; Xue, H.; Xiao, Y.; Qin, Z.; Fu, L. Self-Assembly of
Graphene Single Crystals with Uniform Size and Orientation: The First 2D Super-Ordered
Structure. J. Am. Chem. Soc. 2016, 138 (25), 7812.
(90) Dong, J.; Geng, D.; Liu, F.; Ding, F. Formation of Twinned Graphene Polycrystals. Angew.
Chem. Int. Ed. 2019, 58 (23), 7723.
(91) Watanabe, K.; Taniguchi, T.; Kanda, H. Direct-bandgap properties and evidence for ultraviolet
lasing of hexagonal boron nitride single crystal. Nat. Mater. 2004, 3, 404.
(92) Song, L.; Ci, L.; Lu, H.; Sorokin, P. B.; Jin, C.; Ni, J.; Kvashnin, A. G.; Kvashnin, D. G.; Lou,
J.; Yakobson, B. I.et al. Large Scale Growth and Characterization of Atomic Hexagonal Boron
Nitride Layers. Nano Lett. 2010, 10 (8), 3209.
(93) Yu, B.; Xing, W.; Guo, W.; Qiu, S.; Wang, X.; Lo, S.; Hu, Y. Thermal exfoliation of hexagonal
boron nitride for effective enhancements on thermal stability, flame retardancy and smoke
suppression of epoxy resin nanocomposites via sol–gel process. Journal of Materials Chemistry
A 2016, 4 (19), 7330.

170
(94) Liu, Z.; Gong, Y.; Zhou, W.; Ma, L.; Yu, J.; Idrobo, J. C.; Jung, J.; MacDonald, A. H.; Vajtai,
R.; Lou, J.et al. Ultrathin high-temperature oxidation-resistant coatings of hexagonal boron
nitride. Nat. Commu. 2013, 4 (1), 2541.
(95) Hui, F.; Pan, C.; Shi, Y.; Ji, Y.; Grustan-Gutierrez, E.; Lanza, M. On the use of two dimensional
hexagonal boron nitride as dielectric. Microelectron. Eng. 2016, 163, 119.
(96) Li, L. H.; Santos, E. J. G.; Xing, T.; Cappelluti, E.; Roldán, R.; Chen, Y.; Watanabe, K.;
Taniguchi, T. Dielectric Screening in Atomically Thin Boron Nitride Nanosheets. Nano Lett.
2015, 15 (1), 218.
(97) Khan, A. F.; Brownson, D. A. C.; Randviir, E. P.; Smith, G. C.; Banks, C. E. 2D Hexagonal
Boron Nitride (2D-hBN) Explored for the Electrochemical Sensing of Dopamine. Anal. Chem.
2016, 88 (19), 9729.
(98) Dauber, J.; Sagade, A. A.; Oellers, M.; Watanabe, K.; Taniguchi, T.; Neumaier, D.; Stampfer,
C. Ultra-sensitive Hall sensors based on graphene encapsulated in hexagonal boron nitride.
Appl. Phys. Lett. 2015, 106 (19), 193501.
(99) Kubota, Y.; Watanabe, K.; Tsuda, O.; Taniguchi, T. Deep Ultraviolet Light-Emitting
Hexagonal Boron Nitride Synthesized at Atmospheric Pressure. Science 2007, 317 (5840), 932.
(100) Chae, S. H.; Hone, J.; Seo, D.; Kwon, J.; Lee, G.; Choi, H.; Cao, Q.; Hua, X.; Herman, I. P.;
Shih, E.et al. 2019 Conference on Lasers and Electro-Optics (CLEO), 2019; p 1.
(101) Hu, S.; Lozada-Hidalgo, M.; Wang, F. C.; Mishchenko, A.; Schedin, F.; Nair, R. R.; Hill, E.
W.; Boukhvalov, D. W.; Katsnelson, M. I.; Dryfe, R. A. W.et al. Proton transport through one-
atom-thick crystals. Nature 2014, 516, 227.
(102) Mogg, L.; Zhang, S.; Hao, G. P.; Gopinadhan, K.; Barry, D.; Liu, B. L.; Cheng, H. M.; Geim,
A. K.; Lozada-Hidalgo, M. Perfect proton selectivity in ion transport through two-dimensional
crystals. Nat. Commun. 2019, 10 (1), 4243.
(103) Levendorf, M. P.; Kim, C. J.; Brown, L.; Huang, P. Y.; Havener, R. W.; Muller, D. A.; Park, J.
Graphene and boron nitride lateral heterostructures for atomically thin circuitry. Nature 2012,
488 (7413), 627.
(104) Geim, A. K.; Grigorieva, I. V. Van der Waals heterostructures. Nature 2013, 499, 419.
(105) Hunt, B.; Sanchez-Yamagishi, J. D.; Young, A. F.; Yankowitz, M.; LeRoy, B. J.; Watanabe,
K.; Taniguchi, T.; Moon, P.; Koshino, M.; Jarillo-Herrero, P.et al. Massive Dirac Fermions and
Hofstadter Butterfly in a van der Waals Heterostructure. Science 2013, 340 (6139), 1427.
(106) Lee, G.-H.; Yu, Y.-J.; Cui, X.; Petrone, N.; Lee, C.-H.; Choi, M. S.; Lee, D.-Y.; Lee, C.; Yoo,
W. J.; Watanabe, K.et al. Flexible and Transparent MoS2 Field-Effect Transistors on
Hexagonal Boron Nitride-Graphene Heterostructures. ACS Nano 2013, 7 (9), 7931.

171
(107) Wang, L.; Wu, B.; Chen, J.; Liu, H.; Hu, P.; Liu, Y. Monolayer Hexagonal Boron Nitride Films
with Large Domain Size and Clean Interface for Enhancing the Mobility of Graphene-Based
Field-Effect Transistors. Adv. Mater. 2014, 26 (10), 1559.
(108) Lu, J.; Yeo, P. S. E.; Zheng, Y.; Xu, H.; Gan, C. K.; Sullivan, M. B.; Castro Neto, A. H.; Loh,
K. P. Step Flow Versus Mosaic Film Growth in Hexagonal Boron Nitride. J. Am. Chem. Soc.
2013, 135 (6), 2368.
(109) Zhu, H.; Zhao, X.; Li, H.; Zhao, R. Revealing stable geometries and magic clusters of hexagonal
boron nitride in the nucleation of chemical vapor deposition growth on Ni(111)/Cu(111)
surfaces: a theoretical study. Phys. Chem. Chem. Phys. 2020, 22 (7), 4023.
(110) Chang, R.-J.; Wang, X.; Wang, S.; Sheng, Y.; Porter, B.; Bhaskaran, H.; Warner, J. H. Growth
of Large Single-Crystalline Monolayer Hexagonal Boron Nitride by Oxide-Assisted Chemical
Vapor Deposition. Chem. Mater. 2017, 29 (15), 6252.
(111) Wu, Q.; Park, J.-H.; Park, S.; Jung, S. J.; Suh, H.; Park, N.; Wongwiriyapan, W.; Lee, S.; Lee,
Y. H.; Song, Y. J. Single Crystalline Film of Hexagonal Boron Nitride Atomic Monolayer by
Controlling Nucleation Seeds and Domains. Sci. Rep. 2015, 5 (1), 16159.
(112) Zhu, T.; Liang, Y.; Zhang, C.; Wang, Z.; Dong, M.; Wang, C.; Yang, M.; Goto, T.; Tu, R.;
Zhang, S. A high-throughput synthesis of large-sized single-crystal hexagonal boron nitride on
a Cu–Ni gradient enclosure. RSC Advances 2020, 10 (27), 16088.
(113) Song, X.; Gao, J.; Nie, Y.; Gao, T.; Sun, J.; Ma, D.; Li, Q.; Chen, Y.; Jin, C.; Bachmatiuk, A.et
al. Chemical vapor deposition growth of large-scale hexagonal boron nitride with controllable
orientation. Nano Res. 2015, 8 (10), 3164.
(114) Behura, S.; Nguyen, P.; Che, S.; Debbarma, R.; Berry, V. Large-Area, Transfer-Free, Oxide-
Assisted Synthesis of Hexagonal Boron Nitride Films and Their Heterostructures with MoS2
and WS2. J. Am. Chem. Soc. 2015, 137 (40), 13060.
(115) Li, X.; Li, Y.; Wang, Q.; Yin, J.; Li, J.; Yu, J.; Guo, W. Oxygen-suppressed selective growth
of monolayer hexagonal boron nitride on copper twin crystals. Nano Res. 2017, 10 (3), 826.
(116) Lu, G.; Wu, T.; Yuan, Q.; Wang, H.; Wang, H.; Ding, F.; Xie, X.; Jiang, M. Synthesis of large
single-crystal hexagonal boron nitride grains on Cu-Ni alloy. Nat. Commun. 2015, 6, 6160.
(117) Ji, Y.; Calderon, B.; Han, Y.; Cueva, P.; Jungwirth, N. R.; Alsalman, H. A.; Hwang, J.; Fuchs,
G. D.; Muller, D. A.; Spencer, M. G. Chemical Vapor Deposition Growth of Large Single-
Crystal Mono-, Bi-, Tri-Layer Hexagonal Boron Nitride and Their Interlayer Stacking. ACS
Nano 2017, 11 (12), 12057.
(118) Meng, J.; Zhang, X.; Wang, Y.; Yin, Z.; Liu, H.; Xia, J.; Wang, H.; You, J.; Jin, P.; Wang, D.et
al. Aligned Growth of Millimeter-Size Hexagonal Boron Nitride Single-Crystal Domains on
Epitaxial Nickel Thin Film. Small 2017, 13 (18).

172
(119) Stehle, Y.; Meyer, H. M.; Unocic, R. R.; Kidder, M.; Polizos, G.; Datskos, P. G.; Jackson, R.;
Smirnov, S. N.; Vlassiouk, I. V. Synthesis of Hexagonal Boron Nitride Monolayer: Control of
Nucleation and Crystal Morphology. Chem. Mater. 2015, 27 (23), 8041.
(120) Ren, X.; Dong, J.; Yang, P.; Li, J.; Lu, G.; Wu, T.; Wang, H.; Guo, W.; Zhang, Z.; Ding, F.et
al. Grain boundaries in chemical-vapor-deposited atomically thin hexagonal boron nitride.
Physical Review Materials 2019, 3 (1), 014004.
(121) Tay, R. Y.; Park, H. J.; Ryu, G. H.; Tan, D.; Tsang, S. H.; Li, H.; Liu, W.; Teo, E. H. T.; Lee,
Z.; Lifshitz, Y.et al. Synthesis of aligned symmetrical multifaceted monolayer hexagonal boron
nitride single crystals on resolidified copper. Nanoscale 2016, 8 (4), 2434.
(122) Wood, G. E.; Marsden, A. J.; Mudd, J. J.; Walker, M.; Asensio, M.; Avila, J.; Chen, K.; Bell,
G. R.; Wilson, N. R. van der Waals epitaxy of monolayer hexagonal boron nitride on copper
foil: growth, crystallography and electronic band structure. 2D Materials 2015, 2 (2), 025003.
(123) Wang, S.; Dearle, A. E.; Maruyama, M.; Ogawa, Y.; Okada, S.; Hibino, H.; Taniyasu, Y.
Catalyst-Selective Growth of Single-Orientation Hexagonal Boron Nitride toward High-
Performance Atomically Thin Electric Barriers. Adv. Mater. 2019, 31 (24), e1900880.
(124) Chen, T.-A.; Chuu, C.-P.; Tseng, C.-C.; Wen, C.-K.; Wong, H. S. P.; Pan, S.; Li, R.; Chao, T.-
A.; Chueh, W.-C.; Zhang, Y.et al. Wafer-scale single-crystal hexagonal boron nitride
monolayers on Cu (111). Nature 2020, 579 (7798), 219.
(125) Lynch, R. W.; Drickamer, H. G. Effect of High Pressure on the Lattice Parameters of Diamond,
Graphite, and Hexagonal Boron Nitride. The Journal of Chemical Physics 1966, 44 (1), 181.
(126) Lee, J. S.; Choi, S. H.; Yun, S. J.; Kim, Y. I.; Boandoh, S.; Park, J.-H.; Shin, B. G.; Ko, H.; Lee,
S. H.; Kim, Y.-M.et al. Wafer-scale single-crystal hexagonal boron nitride film via self-
collimated grain formation. Science 2018, 362 (6416), 817.
(127) Liu, Y.; Zou, X.; Yakobson, B. I. Dislocations and Grain Boundaries in Two-Dimensional
Boron Nitride. ACS Nano 2012, 6 (8), 7053.
(128) Abadi, R.; Uma, R. P.; Izadifar, M.; Rabczuk, T. The effect of temperature and topological
defects on fracture strength of grain boundaries in single-layer polycrystalline boron-nitride
nanosheet. Computational Materials Science 2016, 123, 277.
(129) Geng, D.; Meng, L.; Chen, B.; Gao, E.; Yan, W.; Yan, H.; Luo, B.; Xu, J.; Wang, H.; Mao, Z.et
al. Controlled Growth of Single-Crystal Twelve-Pointed Graphene Grains on a Liquid Cu
Surface. Adv. Mater. 2014, 26 (37), 6423.
(130) Wu, B.; Geng, D.; Xu, Z.; Guo, Y.; Huang, L.; Xue, Y.; Chen, J.; Yu, G.; Liu, Y. Self-organized
graphene crystal patterns. NPG Asia Materials 2013, 5 (2), e36.
(131) Tan, C.; Zhang, H. Two-dimensional transition metal dichalcogenide nanosheet-based
composites. Chem. Soc. Rev. 2015, 44 (9), 2713.

173
(132) Huang, X.; Zeng, Z.; Zhang, H. Metal dichalcogenide nanosheets: preparation, properties and
applications. Chem. Soc. Rev. 2013, 42 (5), 1934.
(133) Dong, R.; Kuljanishvili, I. Review Article: Progress in fabrication of transition metal
dichalcogenides heterostructure systems. J Vac. Sci. Technol. B Nanotechnol. Microelectron.
2017, 35 (3), 030803.
(134) Xie, L. M. Two-dimensional transition metal dichalcogenide alloys: preparation,
characterization and applications. Nanoscale 2015, 7 (44), 18392.
(135) Hemmat, Z.; Cavin, J.; Ahmadiparidari, A.; Ruckel, A.; Rastegar, S.; Misal, S. N.; Majidi, L.;
Kumar, K.; Wang, S.; Guo, J.et al. Quasi-Binary Transition Metal Dichalcogenide Alloys:
Thermodynamic Stability Prediction, Scalable Synthesis, and Application. Adv. Mater. 2020,
32 (26), 1907041.
(136) Duan, X.; Wang, C.; Fan, Z.; Hao, G.; Kou, L.; Halim, U.; Li, H.; Wu, X.; Wang, Y.; Jiang,
J.et al. Synthesis of WS2xSe2–2x Alloy Nanosheets with Composition-Tunable Electronic
Properties. Nano Lett. 2016, 16 (1), 264.
(137) Wang, G.; Robert, C.; Suslu, A.; Chen, B.; Yang, S.; Alamdari, S.; Gerber, I. C.; Amand, T.;
Marie, X.; Tongay, S.et al. Spin-orbit engineering in transition metal dichalcogenide alloy
monolayers. Nat. Commun. 2015, 6, 10110.
(138) Voiry, D.; Mohite, A.; Chhowalla, M. Phase engineering of transition metal dichalcogenides.
Chem. Soc. Rev. 2015, 44 (9), 2702.
(139) Zhou, J.; Lin, J.; Huang, X.; Zhou, Y.; Chen, Y.; Xia, J.; Wang, H.; Xie, Y.; Yu, H.; Lei, J.et
al. A library of atomically thin metal chalcogenides. Nature 2018, 556 (7701), 355.
(140) Xiao, Y.; Zhou, M.; Liu, J.; Xu, J.; Fu, L. Phase engineering of two-dimensional transition
metal dichalcogenides. Sci. China Mater. 2019, 62 (6), 759.
(141) Li, Y.; Duerloo, K.-A. N.; Wauson, K.; Reed, E. J. Structural semiconductor-to-semimetal
phase transition in two-dimensional materials induced by electrostatic gating. Nat. Commun.
2016, 7, 10671.
(142) Huang, H. H.; Fan, X.; Singh, D. J.; Zheng, W. T. Recent progress of TMD nanomaterials:
phase transitions and applications. Nanoscale 2020, 12 (3), 1247.
(143) Heising, J.; Kanatzidis, M. G. Structure of Restacked MoS2 and WS2 Elucidated by Electron
Crystallography. J. Am. Chem. Soc. 1999, 121 (4), 638.
(144) Toh, R. J.; Sofer, Z.; Luxa, J.; Sedmidubský, D.; Pumera, M. 3R phase of MoS2 and WS2
outperforms the corresponding 2H phase for hydrogen evolution. Chem. Commun. 2017, 53
(21), 3054.
(145) Zhao, W.; Ding, F. Energetics and kinetics of phase transition between a 2H and a 1T MoS2
monolayer-a theoretical study. Nanoscale 2017, 9 (6), 2301.

174
(146) Tang, S.; Zhang, C.; Jia, C.; Ryu, H.; Hwang, C.; Hashimoto, M.; Lu, D.; Liu, Z.; Devereaux,
T. P.; Shen, Z.-X.et al. Electronic structure of monolayer 1T′-MoTe2 grown by molecular beam
epitaxy. APL Materials 2018, 6 (2), 026601.
(147) Zhang, Y.-J.; Wang, R.-N.; Dong, G.-Y.; Wang, S.-F.; Fu, G.-S.; Wang, J.-L. Mechanical
properties of 1T-, 1T′-, and 1H-MX2 monolayers and their 1H/1T′-MX2 (M = Mo, W and X =
S, Se, Te) heterostructures. AIP Adv. 2019, 9 (12), 125208.
(148) Ambrosi, A.; Sofer, Z.; Pumera, M. 2H → 1T phase transition and hydrogen evolution activity
of MoS₂, MoSe₂, WS₂ and WSe₂ strongly depends on the MX₂ composition. Chem. Commun.
2015, 51 (40), 8450.
(149) Py, M. A.; Haering, R. R. Structural destabilization induced by lithium intercalation in MoS2
and related compounds. Can. J. Phys. 1983, 61 (1), 76.
(150) Mak, K. F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T. F. Atomically Thin MoS₂: A New Direct-Gap
Semiconductor. Phys. Rev. Lett. 2010, 105 (13), 136805.
(151) Ellis, J. K.; Lucero, M. J.; Scuseria, G. E. The indirect to direct band gap transition in
multilayered MoS₂ as predicted by screened hybrid density functional theory. Appl. Phys. Lett.
2011, 99 (26), 261908.
(152) Choi, W.; Choudhary, N.; Han, G. H.; Park, J.; Akinwande, D.; Lee, Y. H. Recent development
of two-dimensional transition metal dichalcogenides and their applications. Mater. Today 2017,
20 (3), 116.
(153) Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-layer MoS₂
transistors. Nat. Nanotech. 2011, 6, 147.
(154) Liu, G.-B.; Xiao, D.; Yao, Y.; Xu, X.; Yao, W. Electronic structures and theoretical modelling
of two-dimensional group-VIB transition metal dichalcogenides. Chem. Soc. Rev. 2015, 44 (9),
2643.
(155) Xu, X.; Yao, W.; Xiao, D.; Heinz, T. F. Spin and pseudospins in layered transition metal
dichalcogenides. Nat. Phys. 2014, 10, 343.
(156) Jariwala, D.; Sangwan, V. K.; Lauhon, L. J.; Marks, T. J.; Hersam, M. C. Emerging Device
Applications for Semiconducting Two-Dimensional Transition Metal Dichalcogenides. ACS
Nano 2014, 8 (2), 1102.
(157) Hu, Z.; Wu, Z.; Han, C.; He, J.; Ni, Z.; Chen, W. Two-dimensional transition metal
dichalcogenides: interface and defect engineering. Chem. Soc. Rev. 2018, 47 (9), 3100.
(158) Thakar, K.; Lodha, S. Optoelectronic and photonic devices based on transition metal
dichalcogenides. Materials Research Express 2020, 7 (1), 014002.
(159) Kuc, A.; Heine, T.; Kis, A. Electronic properties of transition-metal dichalcogenides. MRS Bull.
2015, 40 (7), 577.

175
(160) Ji, Q.; Zhang, Y.; Gao, T.; Zhang, Y.; Ma, D.; Liu, M.; Chen, Y.; Qiao, X.; Tan, P.-H.; Kan,
M.et al. Epitaxial Monolayer MoS₂ on Mica with Novel Photoluminescence. Nano Lett. 2013,
13 (8), 3870.
(161) Xu, H.; Zhou, W.; Zheng, X.; Huang, J.; Feng, X.; Ye, L.; Xu, G.; Lin, F. Control of the
Nucleation Density of Molybdenum Disulfide in Large-Scale Synthesis Using Chemical Vapor
Deposition. Materials 2018, 11 (6).
(162) Kim, H.; Ovchinnikov, D.; Deiana, D.; Unuchek, D.; Kis, A. Suppressing Nucleation in Metal–
Organic Chemical Vapor Deposition of MoS₂ Monolayers by Alkali Metal Halides. Nano Lett.
2017, 17 (8), 5056.
(163) van der Zande, A. M.; Huang, P. Y.; Chenet, D. A.; Berkelbach, T. C.; You, Y.; Lee, G. H.;
Heinz, T. F.; Reichman, D. R.; Muller, D. A.; Hone, J. C. Grains and grain boundaries in highly
crystalline monolayer molybdenum disulphide. Nat. Mater. 2013, 12 (6), 554.
(164) Najmaei, S.; Liu, Z.; Zhou, W.; Zou, X.; Shi, G.; Lei, S.; Yakobson, B. I.; Idrobo, J. C.; Ajayan,
P. M.; Lou, J. Vapour phase growth and grain boundary structure of molybdenum disulphide
atomic layers. Nat. Mater. 2013, 12 (8), 754.
(165) Liu, B.; Fathi, M.; Chen, L.; Abbas, A.; Ma, Y.; Zhou, C. Chemical Vapor Deposition Growth
of Monolayer WSe₂ with Tunable Device Characteristics and Growth Mechanism Study. ACS
Nano 2015, 9 (6), 6119.
(166) Krane, N.; Lotze, C.; Franke, K. J. Moiré structure of MoS₂ on Au(111): Local structural and
electronic properties. Surf. Sci. 2018, DOI:10.1016/j.susc.2018.03.015
10.1016/j.susc.2018.03.015.
(167) Gao, Y.; Liu, Z.; Sun, D.-M.; Huang, L.; Ma, L.-P.; Yin, L.-C.; Ma, T.; Zhang, Z.; Ma, X.-L.;
Peng, L.-M.et al. Large-area synthesis of high-quality and uniform monolayer WS₂ on reusable
Au foils. Nat. Commun. 2015, 6, 8569.
(168) Gronborg, S. S.; Ulstrup, S.; Bianchi, M.; Dendzik, M.; Sanders, C. E.; Lauritsen, J. V.;
Hofmann, P.; Miwa, J. A. Synthesis of Epitaxial Single-Layer MoS₂ on Au(111). Langmuir
2015, 31 (35), 9700.
(169) Zhang, X.; Choudhury, T. H.; Chubarov, M.; Xiang, Y.; Jariwala, B.; Zhang, F.; Alem, N.;
Wang, G. C.; Robinson, J. A.; Redwing, J. M. Diffusion-Controlled Epitaxy of Large Area
Coalesced WSe2 Monolayers on Sapphire. Nano Lett. 2018, 18 (2), 1049.
(170) Aljarb, A.; Cao, Z.; Tang, H.-L.; Huang, J.-K.; Li, M.; Hu, W.; Cavallo, L.; Li, L.-J. Substrate
Lattice-Guided Seed Formation Controls the Orientation of 2D Transition-Metal
Dichalcogenides. ACS Nano 2017, 11 (9), 9215.
(171) Dumcenco, D.; Ovchinnikov, D.; Marinov, K.; Lazić, P.; Gibertini, M.; Marzari, N.; Sanchez,
O. L.; Kung, Y.-C.; Krasnozhon, D.; Chen, M.-W.et al. Large-Area Epitaxial Monolayer MoS₂.
ACS Nano 2015, 9 (4), 4611.

176
(172) Ruzmetov, D.; Zhang, K.; Stan, G.; Kalanyan, B.; Bhimanapati, G. R.; Eichfeld, S. M.; Burke,
R. A.; Shah, P. B.; O’Regan, T. P.; Crowne, F. J.et al. Vertical 2D/3D Semiconductor
Heterostructures Based on Epitaxial Molybdenum Disulfide and Gallium Nitride. ACS Nano
2016, 10 (3), 3580.
(173) Chen, Z.; Liu, H.; Chen, X.; Chu, G.; Chu, S.; Zhang, H. Wafer-Size and Single-Crystal MoSe₂
Atomically Thin Films Grown on GaN Substrate for Light Emission and Harvesting. ACS Appl.
Mater. Interfaces 2016, 8 (31), 20267.
(174) Zhang, F.; Wang, Y.; Erb, C.; Wang, K.; Moradifar, P.; Crespi, V. H.; Alem, N. Full orientation
control of epitaxial MoS₂ on hBN assisted by substrate defects. Phys. Rev. B 2019, 99 (15),
155430.
(175) Okada, M.; Sawazaki, T.; Watanabe, K.; Taniguch, T.; Hibino, H.; Shinohara, H.; Kitaura, R.
Direct Chemical Vapor Deposition Growth of WS2 Atomic Layers on Hexagonal Boron Nitride.
ACS Nano 2014, 8 (8), 8273.
(176) Wu, L.; Yang, W.; Wang, G. Mechanism of substrate-induced anisotropic growth of monolayer
WS2 by kinetic Monte Carlo simulations. npj 2D Materials and Applications 2019, 3 (1), 6.
(177) Hwang, Y.; Shin, N. Hydrogen-assisted step-edge nucleation of MoSe2 monolayers on sapphire
substrates. Nanoscale 2019, 11 (16), 7701.
(178) Chen, L.; Liu, B.; Ge, M.; Ma, Y.; Abbas, A. N.; Zhou, C. Step-Edge-Guided Nucleation and
Growth of Aligned WSe₂ on Sapphire via a Layer-over-Layer Growth Mode. ACS Nano 2015,
9 (8), 8368.
(179) Yang, P.; Zhang, S.; Pan, S.; Tang, B.; Liang, Y.; Zhao, X.; Zhang, Z.; Shi, J.; Huan, Y.; Shi,
Y.et al. Epitaxial Growth of Centimeter-Scale Single-Crystal MoS2 Monolayer on Au(111).
ACS Nano 2020, DOI:10.1021/acsnano.0c01478 10.1021/acsnano.0c01478.
(180) Griep, M. H.; Sandoz-Rosado, E.; Tumlin, T. M.; Wetzel, E. Enhanced Graphene Mechanical
Properties through Ultrasmooth Copper Growth Substrates. Nano Lett. 2016, 16 (3), 1657.
(181) Sinterhauf, A.; Traeger, G. A.; Momeni Pakdehi, D.; Schädlich, P.; Willke, P.; Speck, F.;
Seyller, T.; Tegenkamp, C.; Pierz, K.; Schumacher, H. W.et al. Substrate induced nanoscale
resistance variation in epitaxial graphene. Nat. Commun. 2020, 11 (1), 555.
(182) Zhang, K.; Feng, Y.; Wang, F.; Yang, Z.; Wang, J. Two dimensional hexagonal boron nitride
(2D-hBN): synthesis, properties and applications. J. Mater. Chem. C 2017, 5 (46), 11992.
(183) Joshi, N.; Ghosh, P. Substrate-induced changes in the magnetic and electronic properties of
hexagonal boron nitride. Phys. Rev. B 2013, 87 (23), 235440.
(184) Chae, W. H.; Cain, J. D.; Hanson, E. D.; Murthy, A. A.; Dravid, V. P. Substrate-induced strain
and charge doping in CVD-grown monolayer MoS2. Appl. Phys. Lett. 2017, 111 (14), 143106.

177
(185) Giovannetti, G.; Khomyakov, P. A.; Brocks, G.; Kelly, P. J.; van den Brink, J. Substrate-
induced band gap in graphene on hexagonal boron nitride: Ab initio density functional
calculations. Phys. Rev. B 2007, 76 (7), 073103.
(186) Zeller, P.; Günther, S. What are the possible moirépatterns of graphene on hexagonally packed
surfaces? Universal solution for hexagonal coincidence lattices, derived by a geometric
construction. New J. Phys. 2014, 16 (8), 083028.
(187) Dedkov, Y.; Voloshina, E. Multichannel scanning probe microscopy and spectroscopy of
graphene moiréstructures. Phys. Chem. Chem. Phys. 2014, 16 (9), 3894.
(188) Auwärter, W. Hexagonal boron nitride monolayers on metal supports: Versatile templates for
atoms, molecules and nanostructures. Surf. Sci. Rep. 2019, 74 (1), 1.
(189) Summerfield, A.; Kozikov, A.; Cheng, T. S.; Davies, A.; Cho, Y.-J.; Khlobystov, A. N.; Mellor,
C. J.; Foxon, C. T.; Watanabe, K.; Taniguchi, T.et al. Moiré-Modulated Conductance of
Hexagonal Boron Nitride Tunnel Barriers. Nano Lett. 2018, 18 (7), 4241.
(190) Yasuda, S.; Takahashi, R.; Osaka, R.; Kumagai, R.; Miyata, Y.; Okada, S.; Hayamizu, Y.;
Murakoshi, K. Out-of-Plane Strain Induced in a MoiréSuperstructure of Monolayer MoS2 and
MoSe2 on Au(111). Small 2017, 13 (31), 1700748.
(191) Yasuda, S.; Takahashi, R.; Osaka, R.; Kumagai, R.; Miyata, Y.; Okada, S.; Hayamizu, Y.;
Murakoshi, K. Out-of-Plane Strain Induced in a MoiréSuperstructure of Monolayer MoS(2)
and MoSe(2) on Au(111). Small (Weinheim an der Bergstrasse, Germany) 2017, 13 (31).
(192) Wintterlin, J.; Bocquet, M. L. Graphene on metal surfaces. Surf. Sci. 2009, 603 (10), 1841.
(193) Rasool, H. I.; Song, E. B.; Mecklenburg, M.; Regan, B. C.; Wang, K. L.; Weiller, B. H.;
Gimzewski, J. K. Atomic-scale characterization of graphene grown on copper (100) single
crystals. J. Am. Chem. Soc. 2011, 133 (32), 12536.
(194) Cho, J.; Gao, L.; Tian, J.; Cao, H.; Wu, W.; Yu, Q.; Yitamben, E. N.; Fisher, B.; Guest, J. R.;
Chen, Y. P.et al. Atomic-Scale Investigation of Graphene Grown on Cu Foil and the Effects of
Thermal Annealing. ACS Nano 2011, 5 (5), 3607.
(195) Zou, Z.; Carnevali, V.; Jugovac, M.; Patera, L. L.; Sala, A.; Panighel, M.; Cepek, C.; Soldano,
G.; Mariscal, M. M.; Peressi, M.et al. Graphene on nickel (100) micrograins: Modulating the
interface interaction by extended moirésuperstructures. Carbon 2018, 130, 441.
(196) Murata, Y.; Petrova, V.; Kappes, B. B.; Ebnonnasir, A.; Petrov, I.; Xie, Y.-H.; Ciobanu, C. V.;
Kodambaka, S. MoiréSuperstructures of Graphene on Faceted Nickel Islands. ACS Nano 2010,
4 (11), 6509.
(197) Dedkov, Y.; Voloshina, E.; Fonin, M. Scanning probe microscopy and spectroscopy of
graphene on metals. Phys. Status Solidi B 2015, 252 (3), 451.

178
(198) Busse, C.; Lazic, P.; Djemour, R.; Coraux, J.; Gerber, T.; Atodiresei, N.; Caciuc, V.; Brako, R.;
N'Diaye, A. T.; Blugel, S.et al. Graphene on Ir(111): physisorption with chemical modulation.
Phys. Rev. Lett. 2011, 107 (3), 036101.
(199) Sutter, P.; Sadowski, J. T.; Sutter, E. Graphene on Pt(111): Growth and substrate interaction.
Phys. Rev. B 2009, 80 (24).
(200) Voloshina, E.; Dedkov, Y. Atomic force spectroscopy and density-functional study of graphene
corrugation on Ru(0001). Phys. Rev. B 2016, 93 (23), 235418.
(201) Marchini, S.; Günther, S.; Wintterlin, J. Scanning tunneling microscopy of graphene on
Ru(0001). Phys. Rev. B 2007, 76 (7), 075429.
(202) Gao, L.; Liu, Y.; Shi, R.; Ma, T.; Hu, Y.; Luo, J. Influence of interface interaction on the moiré
superstructures of graphene on transition-metal substrates. RSC Advances 2017, 7 (20), 12179.
(203) Pan, Y.; Zhang, H.; Shi, D.; Sun, J.; Du, S.; Liu, F.; Gao, H.-j. Highly Ordered, Millimeter-
Scale, Continuous, Single-Crystalline Graphene Monolayer Formed on Ru (0001). Adv. Mater.
2009, 21 (27), 2777.
(204) Voloshina, E.; Berdunov, N.; Dedkov, Y. Restoring a nearly free-standing character of
graphene on Ru(0001) by oxygen intercalation. Sci. Rep. 2016, 6, 20285.
(205) Stojanov, P.; Voloshina, E.; Dedkov, Y.; Schmitt, S.; Haenke, T.; Thissen, A. Graphene on
Rh(111): Combined DFT, STM, and NC-AFM Studies. Procedia Eng. 2014, 93, 8.
(206) Voloshina, E. N.; Dedkov, Y. S.; Torbrügge, S.; Thissen, A.; Fonin, M. Graphene on Rh(111):
Scanning tunneling and atomic force microscopies studies. Appl. Phys. Lett. 2012, 100 (24),
241606.
(207) Kwon, S.-Y.; Ciobanu, C. V.; Petrova, V.; Shenoy, V. B.; Bareño, J.; Gambin, V.; Petrov, I.;
Kodambaka, S. Growth of Semiconducting Graphene on Palladium. Nano Lett. 2009, 9 (12),
3985.
(208) Gao, M.; Pan, Y.; Huang, L.; Hu, H.; Zhang, L. Z.; Guo, H. M.; Du, S. X.; Gao, H. J. Epitaxial
growth and structural property of graphene on Pt(111). Appl. Phys. Lett. 2011, 98 (3), 033101.
(209) Merino, P.; Švec, M.; Pinardi, A. L.; Otero, G.; Martín-Gago, J. A. Strain-Driven Moiré
Superstructures of Epitaxial Graphene on Transition Metal Surfaces. ACS Nano 2011, 5 (7),
5627.
(210) Coraux, J.; N‘Diaye, A. T.; Busse, C.; Michely, T. Structural Coherency of Graphene on Ir(111).
Nano Lett. 2008, 8 (2), 565.
(211) Voloshina, E. N.; Fertitta, E.; Garhofer, A.; Mittendorfer, F.; Fonin, M.; Thissen, A.; Dedkov,
Y. S. Electronic structure and imaging contrast of graphene moire on metals. Sci. Rep. 2013, 3,
1072.
(212) Wang, B.; Bocquet, M. L. Interfacial coupling in rotational monolayer and bilayer graphene on
Ru(0001) from first principles. Nanoscale 2012, 4 (15), 4687.

179
(213) Rogge, P. C.; Thürmer, K.; Foster, M. E.; McCarty, K. F.; Dubon, O. D.; Bartelt, N. C. Real-
time observation of epitaxial graphene domain reorientation. Nat. Commu. 2015, 6 (1), 6880.
(214) Sun, B.; Ouyang, W.; Gu, J.; Wang, C.; Wang, J.; Mi, L. Formation of Moirésuperstructure of
epitaxial graphene on Pt(111): A molecular dynamic simulation investigation. Materials
Chemistry and Physics 2020, 253, 123126.
(215) Gao, L.; Liu, Y.; Ma, T.; Shi, R.; Hu, Y.; Luo, J. Effects of interfacial alignments on the stability
of graphene on Ru(0001) substrate. Appl. Phys. Lett. 2016, 108 (26), 261601.
(216) Meng, L.; Wu, R.; Zhang, L.; Li, L.; Du, S.; Wang, Y.; Gao, H. J. Multi-oriented moire
superstructures of graphene on Ir(111): experimental observations and theoretical models. J.
Phys. Condens. Matter 2012, 24 (31), 314214.
(217) Cao, Y.; Fatemi, V.; Fang, S.; Watanabe, K.; Taniguchi, T.; Kaxiras, E.; Jarillo-Herrero, P.
Unconventional superconductivity in magic-angle graphene superlattices. Nature 2018, 556
(7699), 43.
(218) Zhao, M.; Zhuang, J.; Cheng, Q.; Hao, W.; Du, Y. Moiré-Potential-Induced Band Structure
Engineering in Graphene and Silicene. Small (Weinheim an der Bergstrasse, Germany) 2019,
n/a (n/a), 1903769.
(219) Preobrajenski, A. B.; Ng, M. L.; Vinogradov, A. S.; Mårtensson, N. Controlling graphene
corrugation on lattice-mismatched substrates. Phys. Rev. B 2008, 78 (7).
(220) Zhang, H. G.; Hu, H.; Pan, Y.; Mao, J. H.; Gao, M.; Guo, H. M.; Du, S. X.; Greber, T.; Gao,
H. J. Graphene based quantum dots. Journal of Physics: Condensed Matter 2010, 22 (30),
302001.
(221) Brugger, T.; Günther, S.; Wang, B.; Dil, J. H.; Bocquet, M.-L.; Osterwalder, J.; Wintterlin, J.;
Greber, T. Comparison of electronic structure and template function of single-layer graphene
and a hexagonal boron nitride nanomesh on Ru(0001). Phys. Rev. B 2009, 79 (4).
(222) N'Diaye, A. T.; Bleikamp, S.; Feibelman, P. J.; Michely, T. Two-dimensional Ir cluster lattice
on a graphene moire on Ir(111). Phys. Rev. Lett. 2006, 97 (21), 215501.
(223) Sicot, M.; Bouvron, S.; Zander, O.; Rüdiger, U.; Dedkov, Y. S.; Fonin, M. Nucleation and
growth of nickel nanoclusters on graphene Moiréon Rh(111). Appl. Phys. Lett. 2010, 96 (9),
093115.
(224) Pan, Y.; Zhang, H.; Shi, D.; Sun, J.; Du, S.; Liu, F.; Gao, H.-j. Highly Ordered, Millimeter-
Scale, Continuous, Single-Crystalline Graphene Monolayer Formed on Ru (0001). Adv. Mater.
2009, 21 (27).
(225) Sutter, E.; Wang, B.; Albrecht, P.; Lahiri, J.; Bocquet, M. L.; Sutter, P. Templating of arrays of
Ru nanoclusters by monolayer graphene/Ru Moires with different periodicities. J. Phys.
Condens. Matter 2012, 24 (31), 314201.

180
(226) Martínez-Galera, A. J.; Gómez-Rodríguez, J. M. Surface Diffusion of Simple Organic
Molecules on Graphene on Pt(111). J. Phys. Chem. C 2011, 115 (46), 23036.
(227) Soy, E.; Liang, Z.; Trenary, M. Formation of Pt and Rh Nanoclusters on a Graphene Moiré
Pattern on Cu(111). J. Phys. Chem. C 2015, 119 (44), 24796.
(228) Wang, B.; Yoon, B.; Konig, M.; Fukamori, Y.; Esch, F.; Heiz, U.; Landman, U. Size-selected
monodisperse nanoclusters on supported graphene: bonding, isomerism, and mobility. Nano
Lett. 2012, 12 (11), 5907.
(229) Li, G.; Huang, L.; Xu, W.; Que, Y.; Zhang, Y.; Lu, J.; Du, S.; Liu, Y.; Gao, H. J. Constructing
molecular structures on periodic superstructure of graphene/Ru(0001). Philos. Trans. A Math.
Phys. Eng. Sci. 2014, 372 (2013), 20130015.
(230) Zhang, H. G.; Sun, J. T.; Low, T.; Zhang, L. Z.; Pan, Y.; Liu, Q.; Mao, J. H.; Zhou, H. T.; Guo,
H. M.; Du, S. X.et al. Assembly of iron phthalocyanine and pentacene molecules on a graphene
monolayer grown on Ru(0001). Phys. Rev. B 2011, 84 (24), 245436.
(231) Zhang, L.; Dong, J.; Guan, Z.; Zhang, X.; Ding, F. The alignment-dependent properties and
applications of graphene moirésuperstructures on the Ru(0001) surface. Nanoscale 2020, 12
(24), 12831.
(232) Teng, D.; Vilhelmsen, L. B.; Sholl, D. S. Investigating energetics of Au8 on graphene/Ru(0001)
using a genetic algorithm and density functional theory. Surf. Sci. 2014, 628, 98.
(233) Fan, L.; Wang, K.; Wei, J.; Zhong, M.; Wu, D.; Zhu, H. Correlation between nanoparticle
location and graphene nucleation in chemical vapour deposition of graphene. Journal of
Materials Chemistry A 2014, 2 (32), 13123.
(234) Zhang, L. Z.; Du, S. X.; Sun, J. T.; Huang, L.; Meng, L.; Xu, W. Y.; Pan, L. D.; Pan, Y.; Wang,
Y. L.; Hofer, W. A.et al. Growth Mechanism of Metal Clusters on a Graphene/Ru(0001)
Template. Adv. Mater. Interfaces 2014, 1 (3), 1300104.
(235) Zhou, Z.; Gao, F.; Goodman, D. W. Deposition of metal clusters on single-layer
graphene/Ru(0001): Factors that govern cluster growth. Surf. Sci. 2010, 604 (13-14), L31.
(236) Yang, K.; Xiao, W. D.; Jiang, Y. H.; Zhang, H. G.; Liu, L. W.; Mao, J. H.; Zhou, H. T.; Du, S.
X.; Gao, H. J. Molecule–Substrate Coupling between Metal Phthalocyanines and Epitaxial
Graphene Grown on Ru(0001) and Pt(111). J. Phys. Chem. C 2012, 116 (26), 14052.
(237) MacLeod, J. M.; Rosei, F. Molecular Self-Assembly on Graphene. Small (Weinheim an der
Bergstrasse, Germany) 2014, 10 (6), 1038.
(238) Banerjee, K.; Kumar, A.; Canova, F. F.; Kezilebieke, S.; Foster, A. S.; Liljeroth, P. Flexible
Self-Assembled Molecular Templates on Graphene. J. Phys. Chem. C 2016, 120 (16), 8772.
(239) Néel, N.; Kröger, J. Template Effect of the Graphene Moiré Lattice on Phthalocyanine
Assembly. Molecules 2017, 22 (5), 731.

181
(240) Mao, J.; Zhang, H.; Jiang, Y.; Pan, Y.; Gao, M.; Xiao, W.; Gao, H. J. Tunability of
Supramolecular Kagome Lattices of Magnetic Phthalocyanines Using Graphene-Based Moiré
Patterns as Templates. J. Am. Chem. Soc. 2009, 131 (40), 14136.
(241) Hämäläinen, S. K.; Stepanova, M.; Drost, R.; Liljeroth, P.; Lahtinen, J.; Sainio, J. Self-
Assembly of Cobalt-Phthalocyanine Molecules on Epitaxial Graphene on Ir(111). J. Phys.
Chem. C 2012, 116 (38), 20433.
(242) Shi, Q.; Tokarska, K.; Ta, H. Q.; Yang, X.; Liu, Y.; Ullah, S.; Liu, L.; Trzebicka, B.;
Bachmatiuk, A.; Sun, J.et al. Substrate Developments for the Chemical Vapor Deposition
Synthesis of Graphene. Advanced Materials Interfaces 2020, 7 (7), 1902024.
(243) Jeong, H.; Kim, D. Y.; Kim, J.; Moon, S.; Han, N.; Lee, S. H.; Okello, O. F. N.; Song, K.; Choi,
S.-Y.; Kim, J. K. Wafer-scale and selective-area growth of high-quality hexagonal boron nitride
on Ni(111) by metal-organic chemical vapor deposition. Sci. Rep. 2019, 9 (1), 5736.
(244) Gao, J.; Yuan, Q.; Hu, H.; Zhao, J.; Ding, F. Formation of Carbon Clusters in the Initial Stage
of Chemical Vapor Deposition Graphene Growth on Ni(111) Surface. The Journal of Physical
Chemistry C 2011, 115 (36), 17695.
(245) Gao, J.; Yip, J.; Zhao, J.; Yakobson, B. I.; Ding, F. Graphene nucleation on transition metal
surface: structure transformation and role of the metal step edge. J. Am. Chem. Soc. 2011, 133
(13), 5009.
(246) Yuan, Q.; Gao, J.; Shu, H.; Zhao, J.; Chen, X.; Ding, F. Magic carbon clusters in the chemical
vapor deposition growth of graphene. J. Am. Chem. Soc. 2012, 134 (6), 2970.
(247) Gao, J.; Ding, F. The structure and stability of magic carbon clusters observed in graphene
chemical vapor deposition growth on Ru(0001) and Rh(111) surfaces. Angew. Chem. Int. Ed.
Engl. 2014, 53 (51), 14031.
(248) Meng, L.; Sun, Q.; Wang, J.; Ding, F. Molecular Dynamics Simulation of Chemical Vapor
Deposition Graphene Growth on Ni (111) Surface. The Journal of Physical Chemistry C 2012,
116 (10), 6097.
(249) Wang, Y.; Page, A. J.; Nishimoto, Y.; Qian, H. J.; Morokuma, K.; Irle, S. Template effect in
the competition between Haeckelite and graphene growth on Ni(111): quantum chemical
molecular dynamics simulations. J. Am. Chem. Soc. 2011, 133 (46), 18837.
(250) Arifin, R.; Shibuta, Y.; Shimamura, K.; Shimojo, F. First principles molecular dynamics
simulation of graphene growth on Nickel (111) surface. IOP Conference Series: Materials
Science and Engineering 2016, 128, 012032.
(251) Han, G. H.; Güneş, F.; Bae, J. J.; Kim, E. S.; Chae, S. J.; Shin, H.-J.; Choi, J.-Y.; Pribat, D.;
Lee, Y. H. Influence of Copper Morphology in Forming Nucleation Seeds for Graphene Growth.
Nano Lett. 2011, 11 (10), 4144.

182
(252) Shu, H.; Chen, X.; Tao, X.; Ding, F. Edge Structural Stability and Kinetics of Graphene
Chemical Vapor Deposition Growth. ACS Nano 2012, 6 (4), 3243.
(253) Dong, J.; Zhang, L.; Ding, F. Kinetics of Graphene and 2D Materials Growth. Adv. Mater. 2019,
31 (9), 1801583.
(254) Shu, H.; Chen, X.; Ding, F. The edge termination controlled kinetics in graphene chemical
vapor deposition growth. Chem. Sci. 2014, 5 (12), 4639.
(255) Liu, Z.-Q.; Dong, J.; Ding, F. The geometry of hexagonal boron nitride clusters in the initial
stages of chemical vapor deposition growth on a Cu(111) surface. Nanoscale 2019, 11 (28),
13366.
(256) Liu, S.; van Duin, A. C. T.; van Duin, D. M.; Liu, B.; Edgar, J. H. Atomistic Insights into
Nucleation and Formation of Hexagonal Boron Nitride on Nickel from First-Principles-Based
Reactive Molecular Dynamics Simulations. ACS Nano 2017, 11 (4), 3585.
(257) Liu, S.; Comer, J.; van Duin, A. C. T.; van Duin, D. M.; Liu, B.; Edgar, J. H. Predicting the
preferred morphology of hexagonal boron nitride domain structure on nickel from ReaxFF-
based molecular dynamics simulations. Nanoscale 2019, 11 (12), 5607.
(258) Felter, J.; Raths, M.; Franke, M.; Kumpf, C. In situ study of two-dimensional dendritic growth
of hexagonal boron nitride. 2D Materials 2019, 6 (4), 045005.
(259) Zhang, Z.; Liu, Y.; Yang, Y.; Yakobson, B. I. Growth Mechanism and Morphology of
Hexagonal Boron Nitride. Nano Lett. 2016, 16 (2), 1398.
(260) Arias, P.; Ebnonnasir, A.; Ciobanu, C. V. Growth Kinetics of Two-Dimensional Hexagonal
Boron Nitride Layers on Pd(111). 2020, 20 (4), 2886.
(261) Barreto, P. R. R.; Kull, A. E.; Cappelli, M. A. Kinetic and Surface Mechanisms to Growth of
Hexagonal Boron Nitride. MRS Proceedings 2011, 750, Y5.13.
(262) Zhu, D.; Shu, H.; Jiang, F.; Lv, D.; Asokan, V.; Omar, O.; Yuan, J.; Zhang, Z.; Jin, C. Capture
the growth kinetics of CVD growth of two-dimensional MoS2. npj 2D Materials and
Applications 2017, 1 (1), 8.
(263) Taheri, P.; Wang, J.; Xing, H.; Destino, J. F.; Arik, M. M.; Zhao, C.; Kang, K.; Blizzard, B.;
Zhang, L.; Zhao, P.et al. Growth mechanism of largescale MoS2 monolayer by sulfurization of
MoO3 film. Materials Research Express 2016, 3 (7), 075009.
(264) Yuan, Q.; Hu, H.; Gao, J.; Ding, F.; Liu, Z.; Yakobson, B. I. Upright standing graphene
formation on substrates. J. Am. Chem. Soc. 2011, 133 (40), 16072.
(265) Zhang, X.; Xu, Z.; Hui, L.; Xin, J.; Ding, F. How the Orientation of Graphene Is Determined
during Chemical Vapor Deposition Growth. J. Phys. Chem. Lett. 2012, 3 (19), 2822.
(266) Dong, J.; Zhang, L.; Zhang, K.; Ding, F. How graphene crosses a grain boundary on the catalyst
surface during chemical vapour deposition growth. Nanoscale 2018, 10 (15), 6878.

183
(267) Yu, Q.; Jauregui, L. A.; Wu, W.; Colby, R.; Tian, J.; Su, Z.; Cao, H.; Liu, Z.; Pandey, D.; Wei,
D.et al. Control and characterization of individual grains and grain boundaries in graphene
grown by chemical vapour deposition. Nat. Mater. 2011, 10 (6), 443.
(268) Rasool, H. I.; Song, E. B.; Allen, M. J.; Wassei, J. K.; Kaner, R. B.; Wang, K. L.; Weiller, B.
H.; Gimzewski, J. K. Continuity of graphene on polycrystalline copper. Nano Lett. 2011, 11
(1), 251.
(269) Wang, Y.; Cheng, Y.; Wang, Y.; Zhang, S.; Xu, C.; Zhang, X.; Wang, M.; Xia, Y.; Li, Q.; Zhao,
P.et al. Chemical Vapor Deposition Growth of Graphene Domains Across the Cu Grain
Boundaries. Nano 2018, 13 (08), 1850088.
(270) Gao, L.; Guest, J. R.; Guisinger, N. P. Epitaxial graphene on Cu(111). Nano Lett. 2010, 10 (9),
3512.
(271) Yu, Q.; Jauregui, L. A.; Wu, W.; Colby, R.; Tian, J.; Su, Z.; Cao, H.; Liu, Z.; Pandey, D.; Wei,
D.et al. Control and characterization of individual grains and grain boundaries in graphene
grown by chemical vapour deposition. Nat. Mater. 2011, 10 (6), 443.
(272) Li, J.; Zhuang, J.; Shen, C.; Tian, Y.; Que, Y.; Ma, R.; Pan, J.; Zhang, Y.; Wang, Y.; Du, S.et
al. Impurity-induced formation of bilayered graphene on copper by chemical vapor deposition.
Nano Res. 2016, 9 (9), 2803.
(273) Shi, J.; Zhang, X.; Ma, D.; Zhu, J.; Zhang, Y.; Guo, Z.; Yao, Y.; Ji, Q.; Song, X.; Zhang, Y.et
al. Substrate Facet Effect on the Growth of Monolayer MoS₂ on Au Foils. ACS Nano 2015, 9
(4), 4017.
(274) Huang, M.; Biswal, M.; Park, H. J.; Jin, S.; Qu, D.; Hong, S.; Zhu, Z.; Qiu, L.; Luo, D.; Liu,
X.et al. Highly Oriented Monolayer Graphene Grown on a Cu/Ni(111) Alloy Foil. ACS Nano
2018, 12 (6), 6117.
(275) Murdock, A. T.; Koos, A.; Britton, T. B.; Houben, L.; Batten, T.; Zhang, T.; Wilkinson, A. J.;
Dunin-Borkowski, R. E.; Lekka, C. E.; Grobert, N. Controlling the Orientation, Edge Geometry,
and Thickness of Chemical Vapor Deposition Graphene. ACS Nano 2013, 7 (2), 1351.
(276) Fan, L.; Li, Z.; Xu, Z.; Wang, K.; Wei, J.; Li, X.; Zou, J.; Wu, D.; Zhu, H. Step driven
competitive epitaxial and self-limited growth of graphene on copper surface. AIP Adv. 2011, 1
(3), 032145.
(277) Hayashi, K.; Sato, S.; Yokoyama, N. Anisotropic graphene growth accompanied by step
bunching on a dynamic copper surface. Nanotechnology 2012, 24 (2), 025603.
(278) Zou, Z.; Carnevali, V.; Patera, L. L.; Jugovac, M.; Cepek, C.; Peressi, M.; Comelli, G.; Africh,
C. Operando atomic-scale study of graphene CVD growth at steps of polycrystalline nickel.
Carbon 2020, 161, 528.

184
(279) Yi, D.; Luo, D.; Wang, Z.-J.; Dong, J.; Zhang, X.; Willinger, M.-G.; Ruoff, R. S.; Ding, F.
What Drives Metal-Surface Step Bunching in Graphene Chemical Vapor Deposition? Phys.
Rev. Lett. 2018, 120 (24), 246101.
(280) Zhang, X.; Stradi, D.; Liu, L.; Luo, H.; Brandbyge, M.; Gu, G. Tunneling spectra of graphene
on copper unraveled. Phys. Chem. Chem. Phys. 2016, 18 (25), 17081.
(281) Didar, B. R.; Khosravian, H.; Balbuena, P. B. Temperature effect on the nucleation of graphene
on Cu (111). RSC Advances 2018, 8 (49), 27825.
(282) Yuan, Q.; Yakobson, B. I.; Ding, F. Edge-Catalyst Wetting and Orientation Control of
Graphene Growth by Chemical Vapor Deposition Growth. J. Phys. Chem. Lett. 2014, 5 (18),
3093.
(283) Yuan, Q.; Yakobson, B. I.; Ding, F. Edge-Catalyst Wetting and Orientation Control of
Graphene Growth by Chemical Vapor Deposition Growth. J Phys. Chem. Lett. 2014, 5 (18),
3093.
(284) Ogawa, Y.; Hu, B.; Orofeo, C. M.; Tsuji, M.; Ikeda, K.-i.; Mizuno, S.; Hibino, H.; Ago, H.
Domain Structure and Boundary in Single-Layer Graphene Grown on Cu(111) and Cu(100)
Films. J. Phys. Chem. Lett. 2012, 3 (2), 219.
(285) Zhang, B.; Lee, W. H.; Piner, R.; Kholmanov, I.; Wu, Y.; Li, H.; Ji, H.; Ruoff, R. S. Low-
Temperature Chemical Vapor Deposition Growth of Graphene from Toluene on
Electropolished Copper Foils. ACS Nano 2012, 6 (3), 2471.
(286) Wu, R.; Ding, Y.; Yu, K. M.; Zhou, K.; Zhu, Z.; Ou, X.; Zhang, Q.; Zhuang, M.; Li, W.-D.;
Xu, Z.et al. Edge-Epitaxial Growth of Graphene on Cu with a Hydrogen-Free Approach. Chem.
Mater. 2019, 31 (7), 2555.
(287) Wofford, J. M.; Nie, S.; Thürmer, K.; McCarty, K. F.; Dubon, O. D. Influence of lattice
orientation on growth and structure of graphene on Cu(001). Carbon 2015, 90, 284.
(288) Lahiri, J.; Lin, Y.; Bozkurt, P.; Oleynik, II; Batzill, M. An extended defect in graphene as a
metallic wire. Nat. Nanotechnol. 2010, 5 (5), 326.
(289) Appelhans, D. J.; Carr, L. D.; Lusk, M. T. Embedded ribbons of graphene allotropes: an
extended defect perspective. New J. Phys. 2010, 12 (12), 125006.
(290) Tay, R. Y.; Griep, M. H.; Mallick, G.; Tsang, S. H.; Singh, R. S.; Tumlin, T.; Teo, E. H.; Karna,
S. P. Growth of large single-crystalline two-dimensional boron nitride hexagons on
electropolished copper. Nano Lett. 2014, 14 (2), 839.
(291) Chatterjee, S.; Luo, Z.; Acerce, M.; Yates, D. M.; Johnson, A. T. C.; Sneddon, L. G. Chemical
Vapor Deposition of Boron Nitride Nanosheets on Metallic Substrates via
Decaborane/Ammonia Reactions. Chem. Mater. 2011, 23 (20), 4414.

185
(292) Ćavar, E.; Westerström, R.; Mikkelsen, A.; Lundgren, E.; Vinogradov, A. S.; Ng, M. L.;
Preobrajenski, A. B.; Zakharov, A. A.; Mårtensson, N. A single h-BN layer on Pt(111). Surf.
Sci. 2008, 602 (9), 1722.
(293) Khan, M. H.; Huang, Z.; Xiao, F.; Casillas, G.; Chen, Z.; Molino, P. J.; Liu, H. K. Synthesis of
Large and Few Atomic Layers of Hexagonal Boron Nitride on Melted Copper. Sci. Rep. 2015,
5 (1), 7743.
(294) Yin, J.; Yu, J.; Li, X.; Li, J.; Zhou, J.; Zhang, Z.; Guo, W. Large Single-Crystal Hexagonal
Boron Nitride Monolayer Domains with Controlled Morphology and Straight Merging
Boundaries. Small 2015, 11 (35), 4497.
(295) Taslim, A. B.; Nakajima, H.; Lin, Y.-C.; Uchida, Y.; Kawahara, K.; Okazaki, T.; Suenaga, K.;
Hibino, H.; Ago, H. Synthesis of sub-millimeter single-crystal grains of aligned hexagonal
boron nitride on an epitaxial Ni film. Nanoscale 2019, 11 (31), 14668.
(296) Zhao, R.; Gao, J.; Liu, Z.; Ding, F. The reconstructed edges of the hexagonal BN. Nanoscale
2015, 7 (21), 9723.
(297) Mukherjee, R.; Bhowmick, S. Edge Stabilities of Hexagonal Boron Nitride Nanoribbons: A
First-Principles Study. Journal of Chemical Theory and Computation 2011, 7 (3), 720.
(298) Yamijala, S. S. R. K. C.; Pati, S. K. Effects of edge passivations on the electronic and magnetic
properties of zigzag boron-nitride nanoribbons with even and odd-line stone–wales (5–7 pair)
defects. Indian Journal of Physics 2014, 88 (9), 931.
(299) Nemnes, G. A.; Visan, C.; Anghel, D. V.; Manolescu, A. Molecular dynamics of halogenated
graphene - hexagonal boron nitride nanoribbons. Journal of Physics: Conference Series 2016,
738, 012027.
(300) Voznyy, O.; Güçlü, A. D.; Potasz, P.; Hawrylak, P. Effect of edge reconstruction and
passivation on zero-energy states and magnetism in triangular graphene quantum dots with
zigzag edges. Phys. Rev. B 2011, 83 (16), 165417.
(301) Zhao, R.; Li, F.; Liu, Z.; Liu, Z.; Ding, F. The transition metal surface passivated edges of
hexagonal boron nitride (h-BN) and the mechanism of h-BN's chemical vapor deposition (CVD)
growth. Phys. Chem. Chem. Phys. 2015, 17 (43), 29327.
(302) Caneva, S.; Weatherup, R. S.; Bayer, B. C.; Brennan, B.; Spencer, S. J.; Mingard, K.; Cabrero-
Vilatela, A.; Baehtz, C.; Pollard, A. J.; Hofmann, S. Nucleation control for large, single
crystalline domains of monolayer hexagonal boron nitride via Si-doped Fe catalysts. Nano Lett.
2015, 15 (3), 1867.
(303) Grad, G. B.; Blaha, P.; Schwarz, K.; Auwärter, W.; Greber, T. Density functional theory
investigation of the geometric and spintronic structure of h-BN/Ni(111) in view of
photoemission and STM experiments. Phys. Rev. B 2003, 68 (8), 085404.

186
(304) Laskowski, R.; Blaha, P.; Schwarz, K. Bonding of hexagonal BN to transition metal surfaces:
An ab initio density-functional theory study. Phys. Rev. B 2008, 78 (4), 045409.
(305) Auwärter, W.; Muntwiler, M.; Osterwalder, J.; Greber, T. Defect lines and two-domain
structure of hexagonal boron nitride films on Ni(111). Surf. Sci. 2003, 545 (1), L735.
(306) Zhao, R.; Zhao, X.; Liu, Z.; Ding, F.; Liu, Z. Controlling the orientations of h-BN during growth
on transition metals by chemical vapor deposition. Nanoscale 2017, 9 (10), 3561.
(307) Späth, F.; Gebhardt, J.; Düll, F.; Bauer, U.; Bachmann, P.; Gleichweit, C.; Görling, A.;
Steinrück, H. P.; Papp, C. Hydrogenation and hydrogen intercalation of hexagonal boron nitride
on Ni(1 1 1): reactivity and electronic structure. 2D Materials 2017, 4 (3), 035026.
(308) Bokdam, M.; Brocks, G.; Katsnelson, M. I.; Kelly, P. J. Schottky barriers at hexagonal boron
nitride/metal interfaces: A first-principles study. Phys. Rev. B 2014, 90 (8), 085415.
(309) Bets, K. V.; Gupta, N.; Yakobson, B. I. How the Complementarity at Vicinal Steps Enables
Growth of 2D Monocrystals. Nano Lett. 2019, 19 (3), 2027.
(310) Jeong, H.-C.; Williams, E. D. Steps on surfaces: experiment and theory. Surf. Sci. Rep. 1999,
34 (6), 171.
(311) Driver, S. M.; Toomes, R. L.; Woodruff, D. P. A scanning tunnelling microscopy study of C
and N adsorption phases on the vicinal Ni(100) surfaces Ni(810) and Ni(911). Surf. Sci. 2016,
646, 114.
(312) Andryushechkin, B. V.; Cherkez, V. V.; Pavlova, T. V.; Zhidomirov, G. M.; Eltsov, K. N.
Structural transformations of Cu(110) surface induced by adsorption of molecular chlorine. Surf.
Sci. 2013, 608, 135.
(313) Eda, G.; Fujita, T.; Yamaguchi, H.; Voiry, D.; Chen, M.; Chhowalla, M. Coherent Atomic and
Electronic Heterostructures of Single-Layer MoS2. ACS Nano 2012, 6 (8), 7311.
(314) Lin, Y.-C.; Dumcenco, D. O.; Huang, Y.-S.; Suenaga, K. Atomic mechanism of the
semiconducting-to-metallic phase transition in single-layered MoS2. Nat. Nanotech. 2014, 9
(5), 391.
(315) Zan, W.; Hu, Z.; Zhang, Z.; Yakobson, B. I. Phase crossover in transition metal dichalcogenide
nanoclusters. Nanoscale 2016, 8 (45), 19154.
(316) Jin, Q.; Liu, N.; Chen, B.; Mei, D. Mechanisms of Semiconducting 2H to Metallic 1T Phase
Transition in Two-dimensional MoS2 Nanosheets. The Journal of Physical Chemistry C 2018,
122 (49), 28215.
(317) Huang, H. H.; Fan, X.; Singh, D. J.; Zheng, W. T. First principles study on 2H–1T′ transition
in MoS2 with copper. Phys. Chem. Chem. Phys. 2018, 20 (42), 26986.
(318) Sang, X.; Li, X.; Zhao, W.; Dong, J.; Rouleau, C. M.; Geohegan, D. B.; Ding, F.; Xiao, K.;
Unocic, R. R. In situ edge engineering in two-dimensional transition metal dichalcogenides.
Nat. Commun. 2018, 9 (1), 2051.

187
(319) Chang, Y.-H.; Zhang, W.; Zhu, Y.; Han, Y.; Pu, J.; Chang, J.-K.; Hsu, W.-T.; Huang, J.-K.;
Hsu, C.-L.; Chiu, M.-H.et al. Monolayer MoSe₂ Grown by Chemical Vapor Deposition for Fast
Photodetection. ACS Nano 2014, 8 (8), 8582.
(320) Huang, J.-K.; Pu, J.; Hsu, C.-L.; Chiu, M.-H.; Juang, Z.-Y.; Chang, Y.-H.; Chang, W.-H.; Iwasa,
Y.; Takenobu, T.; Li, L.-J. Large-Area Synthesis of Highly Crystalline WSe₂ Monolayers and
Device Applications. ACS Nano 2014, 8 (1), 923.
(321) Ji, Q.; Kan, M.; Zhang, Y.; Guo, Y.; Ma, D.; Shi, J.; Sun, Q.; Chen, Q.; Zhang, Y.; Liu, Z.
Unravelling Orientation Distribution and Merging Behavior of Monolayer MoS2 Domains on
Sapphire. Nano Lett. 2015, 15 (1), 198.
(322) Ji, H. G.; Maruyama, M.; Aji, A. S.; Okada, S.; Matsuda, K.; Ago, H. van der Waals interaction-
induced photoluminescence weakening and multilayer growth in epitaxially aligned WS2. Phys.
Chem. Chem. Phys. 2018, 20 (47), 29790.
(323) Born, M.; Oppenheimer, R. Zur Quantentheorie der Molekeln. Annalen der Physik 1927, 389
(20), 457.
(324) Schrödinger, E. An Undulatory Theory of the Mechanics of Atoms and Molecules. Phys. Rev.
1926, 28 (6), 1049.
(325) Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136 (3B), B864.
(326) Kohn, W.; Sham, L. J. Self-Consistent Equations Including Exchange and Correlation Effects.
Phys. Rev. 1965, 140 (4A), A1133.
(327) Barth, U. v.; Hedin, L. A local exchange-correlation potential for the spin polarized case. i.
Journal of Physics C: Solid State Physics 1972, 5 (13), 1629.
(328) Vosko, S. H.; Wilk, L.; Nusair, M. Accurate spin-dependent electron liquid correlation energies
for local spin density calculations: a critical analysis. Can. J. Phys. 1980, 58 (8), 1200.
(329) Alder, B. J.; Wainwright, T. E. Phase Transition for a Hard Sphere System. The Journal of
Chemical Physics 1957, 27 (5), 1208.
(330) Zhou, X. W.; Wadley, H. N. G.; Johnson, R. A.; Larson, D. J.; Tabat, N.; Cerezo, A.; Petford-
Long, A. K.; Smith, G. D. W.; Clifton, P. H.; Martens, R. L.et al. Atomic scale structure of
sputtered metal multilayers. Acta Mater. 2001, 49 (19), 4005.
(331) Verlet, L. Computer "Experiments" on Classical Fluids. I. Thermodynamical Properties of
Lennard-Jones Molecules. Phys. Rev. 1967, 159 (1), 98.
(332) Swope, W. C.; Andersen, H. C.; Berens, P. H.; Wilson, K. R. A computer simulation method
for the calculation of equilibrium constants for the formation of physical clusters of molecules:
Application to small water clusters. The Journal of Chemical Physics 1982, 76 (1), 637.
(333) Dong, J.; Zhang, L.; Dai, X.; Ding, F. The epitaxy of 2D materials growth. Nat. Commun. 2020,
11 (1), 5862.

188
(334) Ma, Y.; Gao, L.; Yan, Y.; Su, Y.; Qiao, L. Tune the chemical activity of graphene via the
transition metal substrate. RSC Advances 2018, 8 (21), 11807.
(335) Zhu, L.; Ding, F. How the moiré superstructure determines the formation of highly stable
graphene quantum dots on Ru(0001) surface. Nanoscale Horizons 2019, 4 (3), 625.
(336) Alem, N.; Erni, R.; Kisielowski, C.; Rossell, M. D.; Gannett, W.; Zettl, A. Atomically thin
hexagonal boron nitride probed by ultrahigh-resolution transmission electron microscopy. Phys.
Rev. B 2009, 80 (15), 155425.
(337) Liu, P.; Tian, H.; Windl, W.; Gu, G.; Duscher, G.; Wu, Y.; Zhao, M.; Guo, J.; Xu, B.; Liu, L.
Direct imaging of the nitrogen-rich edge in monolayer hexagonal boron nitride and its band
structure tuning. Nanoscale 2019, 11 (43), 20676.
(338) Auwärter, W.; Suter, H. U.; Sachdev, H.; Greber, T. Synthesis of One Monolayer of Hexagonal
Boron Nitride on Ni(111) from B-Trichloroborazine (ClBNH)3. Chem. Mater. 2004, 16 (2),
343.
(339) Li, D.; Ding, F. Environment-dependent edge reconstruction of transition metal
dichalcogenides: a global search. Materials Today Advances 2020, 8, 100079.
(340) Cheng, Y.; Bi, H.; Che, X.; Zhao, W.; Li, D.; Huang, F. Rapid growth of large-area single-
crystal graphene film by seamless stitching using resolidified copper foil on a molybdenum
substrate. Journal of Materials Chemistry A 2019, 7 (31), 18373.
(341) Wofford, J. M.; Nie, S.; McCarty, K. F.; Bartelt, N. C.; Dubon, O. D. Graphene Islands on Cu
foils: the interplay between shape, orientation, and defects. Nano Lett. 2010, 10 (12), 4890.
(342) Yuan, Q.; Song, G.; Sun, D.; Ding, F. Formation of Graphene Grain Boundaries on Cu(100)
Surface and a Route Towards Their Elimination in Chemical Vapor Deposition Growth. Sci.
Rep. 2014, 4, 6541.
(343) Wilson, N. R.; Marsden, A. J.; Saghir, M.; Bromley, C. J.; Schaub, R.; Costantini, G.; White,
T. W.; Partridge, C.; Barinov, A.; Dudin, P.et al. Weak mismatch epitaxy and structural
Feedback in graphene growth on copper foil. Nano Res. 2013, 6 (2), 99.
(344) Jacobberger, R. M.; Murray, E. A.; Fortin-Deschênes, M.; Göltl, F.; Behn, W. A.; Krebs, Z. J.;
Levesque, P. L.; Savage, D. E.; Smoot, C.; Lagally, M. G.et al. Alignment of semiconducting
graphene nanoribbons on vicinal Ge(001). Nanoscale 2019, 11 (11), 4864.
(345) Deng, B.; Wu, J.; Zhang, S.; Qi, Y.; Zheng, L.; Yang, H.; Tang, J.; Tong, L.; Zhang, J.; Liu,
Z.et al. Anisotropic Strain Relaxation of Graphene by Corrugation on Copper Crystal Surfaces.
Small 2018, 14 (22), 1800725.
(346) Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and
semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6 (1), 15.
(347) Kresse, G.; Hafner, J. Ab initio molecular dynamics for open-shell transition metals. Phys. Rev.
B 1993, 48 (17), 13115.

189
(348) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple.
Phys. Rev. Lett. 1996, 77 (18), 3865.
(349) Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave
method. Phys. Rev. B 1999, 59 (3), 1758.
(350) Lee, J. H.; Lee, E. K.; Joo, W. J.; Jang, Y.; Kim, B. S.; Lim, J. Y.; Choi, S. H.; Ahn, S. J.; Ahn,
J. R.; Park, M. H. Wafer-scale growth of single-crystal monolayer graphene on reusable
hydrogen-terminated germanium. Science 2014, 344 (6181), 286.
(351) Sutter, P. W.; Flege, J. I.; Sutter, E. A. Epitaxial graphene on ruthenium. Nature Materials 2008,
7 (5), 406.
(352) Polkowski, W.; Jóźwik, P.; Polański, M.; Bojar, Z. Microstructure and texture evolution of
copper processed by differential speed rolling with various speed asymmetry coefficient.
Materials Science and Engineering: A 2013, 564, 289.
(353) Wang, X.; Liu, X.; Xie, J. Mechanism of surface texture evolution in pure copper strips
subjected to double rolling. Progress in Natural Science: Materials International 2014, 24 (1),
75.
(354) Wolf, D. Structure-energy correlation for grain boundaries in F.C.C. metals—III. Symmetrical
tilt boundaries. Acta Metallurgica Et Materialia 1989, 38 (5), 791.
(355) Bulatov, V. V.; Reed, B. W.; Kumar, M. Grain boundary energy function for fcc metals. Acta
Materialia 2014, 65 (65), 161.
(356) Reina, A.; Jia, X.; Ho, J.; Nezich, D.; Son, H.; Bulovic, V.; Dresselhaus, M. S.; Kong, J. Large
area, few-layer graphene films on arbitrary substrates by chemical vapor deposition. Nano
Letters 2009, 9 (1), 30.
(357) Vlassiouk, I. V.; Stehle, Y.; Pudasaini, P. R.; Unocic, R. R.; Rack, P. D.; Baddorf, A. P.; Ivanov,
I. N.; Lavrik, N. V.; List, F.; Gupta, N. Evolutionary selection growth of two-dimensional
materials on polycrystalline substrates. Nature Materials 2018.
(358) Liu, X.; Wang, J. Low-energy, Mobile Grain Boundaries in Magnesium. Sci. Rep. 2016, 6 (1),
21393.
(359) Liu, L.; Siegel, D. A.; Chen, W.; Liu, P.; Guo, J.; Duscher, G.; Zhao, C.; Wang, H.; Wang, W.;
Bai, X.et al. Unusual role of epilayer–substrate interactions in determining orientational
relations in van der Waals epitaxy. Proc. Natl. Acad. Sci. U. S. A. 2014, 111 (47), 16670.
(360) Chen, T. A.; Chuu, C. P.; Tseng, C. C.; Wen, C. K.; Wong, H. P.; Pan, S.; Li, R.; Chao, T. A.;
Chueh, W. C.; Zhang, Y.et al. Wafer-scale single-crystal hexagonal boron nitride monolayers
on Cu (111). Nature 2020, 579 (7798), 219.
(361) Wang, S.; Dearle, A. E.; Maruyama, M.; Ogawa, Y.; Okada, S.; Hibino, H.; Taniyasu, Y.
Catalyst-Selective Growth of Single-Orientation Hexagonal Boron Nitride toward High-
Performance Atomically Thin Electric Barriers. Adv. Mater. 2019, 31 (24), 1900880.

190
(362) Poelsema, B.; Zandvliet, H. J. W.; van Houselt, A. Stoichiometric edges during the intrinsic
growth of hexagonal boron nitride on Ir(111). New J. Phys. 2019, 21 (9), 092001.
(363) Gao, L.; Guest, J. R.; Guisinger, N. P. Epitaxial Graphene on Cu(111). Nano Lett. 2010, 10 (9),
3512.
(364) Jeong, H.; Hwang, W.-T.; Song, Y.; Kim, J.-K.; Kim, Y.; Hihath, J.; Chung, S.; Lee, T. Highly
uniform monolayer graphene synthesis via a facile pretreatment of copper catalyst substrates
using an ammonium persulfate solution. RSC Advances 2019, 9 (36), 20871.
(365) Tian, J.; Hu, B.; Wei, Z.; Jin, Y.; Luo, Z.; Xia, M.; Pan, Q.; Liu, Y. Surface structure deduced
differences of copper foil and film for graphene CVD growth. Appl. Surf. Sci. 2014, 300, 73.
(366) Nutsch, A.; Pfitzner, L. Chemical Mechanical Planarization (CMP) Metrology for 45/32 nm
Technology Generations. AIP Conference Proceedings 2007, 931 (1), 173.
(367) Dunn, T.; Lee, C.; Tronolone, M.; Shorey, A. 2012 IEEE 62nd Electronic Components and
Technology Conference, 2012; p 1239.
(368) Elías, A. L.; Perea-López, N.; Castro-Beltrán, A.; Berkdemir, A.; Lv, R.; Feng, S.; Long, A. D.;
Hayashi, T.; Kim, Y. A.; Endo, M.et al. Controlled Synthesis and Transfer of Large-Area WS2
Sheets: From Single Layer to Few Layers. ACS Nano 2013, 7 (6), 5235.
(369) Chen, J.; Wen, Y.; Guo, Y.; Wu, B.; Huang, L.; Xue, Y.; Geng, D.; Wang, D.; Yu, G.; Liu, Y.
Oxygen-Aided Synthesis of Polycrystalline Graphene on Silicon Dioxide Substrates. J. Am.
Chem. Soc. 2011, 133 (44), 17548.
(370) Kang, K.; Xie, S.; Huang, L.; Han, Y.; Huang, P. Y.; Mak, K. F.; Kim, C.-J.; Muller, D.; Park,
J. High-mobility three-atom-thick semiconducting films with wafer-scale homogeneity. Nature
2015, 520 (7549), 656.
(371) Zhang, Y.; Zhang, Y.; Ji, Q.; Ju, J.; Yuan, H.; Shi, J.; Gao, T.; Ma, D.; Liu, M.; Chen, Y.et al.
Controlled Growth of High-Quality Monolayer WS2 Layers on Sapphire and Imaging Its Grain
Boundary. ACS Nano 2013, 7 (10), 8963.
(372) Lim, Y.-F.; Priyadarshi, K.; Bussolotti, F.; Gogoi, P. K.; Cui, X.; Yang, M.; Pan, J.; Tong, S.
W.; Wang, S.; Pennycook, S. J.et al. Modification of Vapor Phase Concentrations in MoS2
Growth Using a NiO Foam Barrier. ACS Nano 2018, 12 (2), 1339.
(373) Yu, H.; Liao, M.; Zhao, W.; Liu, G.; Zhou, X. J.; Wei, Z.; Xu, X.; Liu, K.; Hu, Z.; Deng, K.et
al. Wafer-Scale Growth and Transfer of Highly-Oriented Monolayer MoS2 Continuous Films.
ACS Nano 2017, 11 (12), 12001.
(374) Yan, P.; Tian, Q.; Yang, G.; Weng, Y.; Zhang, Y.; Wang, J.; Xie, F.; Lu, N. Epitaxial growth
and interfacial property of monolayer MoS2 on gallium nitride. RSC Advances 2018, 8 (58),
33193.

191
(375) Chen, Z.; Liu, H.; Chen, X.; Chu, G.; Chu, S.; Zhang, H. Wafer-Size and Single-Crystal MoSe2
Atomically Thin Films Grown on GaN Substrate for Light Emission and Harvesting. ACS Appl.
Mater. Interfaces 2016, 8 (31), 20267.
(376) Shi, J.; Yang, Y.; Zhang, Y.; Ma, D.; Wei, W.; Ji, Q.; Zhang, Y.; Song, X.; Gao, T.; Li, C.et al.
Monolayer MoS2 Growth on Au Foils and On-Site Domain Boundary Imaging. Adv. Funct.
Mater. 2015, 25 (6), 842.
(377) Yun, S. J.; Chae, S. H.; Kim, H.; Park, J. C.; Park, J.-H.; Han, G. H.; Lee, J. S.; Kim, S. M.; Oh,
H. M.; Seok, J.et al. Synthesis of Centimeter-Scale Monolayer Tungsten Disulfide Film on Gold
Foils. ACS Nano 2015, 9 (5), 5510.
(378) Zhang, Y.; Yao, Y.; Sendeku, M. G.; Yin, L.; Zhan, X.; Wang, F.; Wang, Z.; He, J. Recent
Progress in CVD Growth of 2D Transition Metal Dichalcogenides and Related Heterostructures.
Adv. Mater. 2019, 31 (41), 1901694.
(379) Chen, M.; Zhang, A.; Liu, Y.; Cui, D.; Li, Z.; Chung, Y.-H.; Mutyala, S. P.; Mecklenburg, M.;
Nie, X.; Xu, C.et al. Gold-vapor-assisted chemical vapor deposition of aligned monolayer
WSe2 with large domain size and fast growth rate. Nano Res. 2020, 13 (10), 2625.
(380) Gao, Y.; Hong, Y.-L.; Yin, L.-C.; Wu, Z.; Yang, Z.; Chen, M.-L.; Liu, Z.; Ma, T.; Sun, D.-M.;
Ni, Z.et al. Ultrafast Growth of High-Quality Monolayer WSe2 on Au. Adv. Mater. 2017, 29
(29), 1700990.
(381) Gao, X.; Zhang, Y.; Weaver, M. J. Observing surface chemical transformations by atomic-
resolution scanning tunneling microscopy: sulfide electrooxidation on gold(111). The Journal
of Physical Chemistry 1992, 96 (11), 4156.
(382) Walen, H.; Liu, D.-J.; Oh, J.; Lim, H.; Evans, J. W.; Kim, Y.; Thiel, P. A. Self-organization of
S adatoms on Au(111): √3R30° rows at low coverage. J. Chem. Phys. 2015, 143 (1), 014704.
(383) Yu, M.; Ascolani, H.; Zampieri, G.; Woodruff, D. P.; Satterley, C. J.; Jones, R. G.; Dhanak, V.
R. The Structure of Atomic Sulfur Phases on Au(111). J. Phys. Chem. C 2007, 111 (29), 10904.
(384) Carro, P.; Andreasen, G.; Vericat, C.; Vela, M. E.; Salvarezza, R. C. New aspects of the surface
chemistry of sulfur on Au(111): Surface structures formed by gold-sulfur complexes. Applied
Surface Science 2019, 487, 848.
(385) Carro, P.; Torrelles, X.; Salvarezza, R. C. A novel model for the (√3 × √3)R30° alkanethiolate–
Au(111) phase based on alkanethiolate–Au adatom complexes. Phys. Chem. Chem. Phys. 2014,
16 (35), 19017.
(386) Zhao, X.; Fu, D.; Ding, Z.; Zhang, Y.-Y.; Wan, D.; Tan, S. J. R.; Chen, Z.; Leng, K.; Dan, J.;
Fu, W.et al. Mo-Terminated Edge Reconstructions in Nanoporous Molybdenum Disulfide Film.
Nano Lett. 2018, 18 (1), 482.

192
(387) Hu, G.; Fung, V.; Sang, X.; Unocic, R. R.; Ganesh, P. Superior electrocatalytic hydrogen
evolution at engineered non-stoichiometric two-dimensional transition metal dichalcogenide
edges. Journal of Materials Chemistry A 2019, 7 (31), 18357.
(388) Hu, G.; Fung, V.; Sang, X.; Unocic, R. R.; Ganesh, P. Predicting synthesizable multi-functional
edge reconstructions in two-dimensional transition metal dichalcogenides. npj Computational
Materials 2020, 6 (1), 44.
(389) Verguts, K.; Vermeulen, B.; Vrancken, N.; Schouteden, K.; Van Haesendonck, C.;
Huyghebaert, C.; Heyns, M.; De Gendt, S.; Brems, S. Epitaxial Al2O3(0001)/Cu(111)
Template Development for CVD Graphene Growth. The Journal of Physical Chemistry C 2016,
120 (1), 297.
(390) Verguts, K.; Vrancken, N.; Vermeulen, B. F.; Huyghebaert, C.; Terryn, H.; Brems, S.; De Gendt,
S. Single-Layer Graphene Synthesis on a Al2O3(0001)/Cu(111) Template Using Chemical
Vapor Deposition. ECS Journal of Solid State Science and Technology 2016, 5 (11), Q3060.
(391) Reddy, K. M.; Gledhill, A. D.; Chen, C.-H.; Drexler, J. M.; Padture, N. P. High quality,
transferrable graphene grown on single crystal Cu(111) thin films on basal-plane sapphire. Appl.
Phys. Lett. 2011, 98 (11), 113117.
(392) Tao, L.; Lee, J.; Holt, M.; Chou, H.; McDonnell, S. J.; Ferrer, D. A.; Babenco, M. G.; Wallace,
R. M.; Banerjee, S. K.; Ruoff, R. S.et al. Uniform Wafer-Scale Chemical Vapor Deposition of
Graphene on Evaporated Cu (111) Film with Quality Comparable to Exfoliated Monolayer.
The Journal of Physical Chemistry C 2012, 116 (45), 24068.
(393) Jiang, H.; Klemmer, T. J.; Barnard, J. A.; Doyle, W. D.; Payzant, E. A. Epitaxial growth of
Cu(111) films on Si(110) by magnetron sputtering: orientation and twin growth. Thin Solid
Films 1998, 315 (1), 13.
(394) Jiang, H.; Klemmer, T. J.; Barnard, J. A.; Payzant, E. A. Epitaxial growth of Cu on Si by
magnetron sputtering. Journal of Vacuum Science & Technology A 1998, 16 (6), 3376.
(395) Demczyk, B. G.; Naik, R.; Auner, G.; Kota, C.; Rao, U. Growth of Cu films on hydrogen
terminated Si(100) and Si(111) surfaces. J. Appl. Phys. 1994, 75 (4), 1956.
(396) Reckinger, N.; Tang, X.; Joucken, F.; Lajaunie, L.; Arenal, R.; Dubois, E.; Hackens, B.;
Henrard, L.; Colomer, J.-F. Oxidation-assisted graphene heteroepitaxy on copper foil.
Nanoscale 2016, 8 (44), 18751.
(397) Brown, L.; Lochocki, E. B.; Avila, J.; Kim, C.-J.; Ogawa, Y.; Havener, R. W.; Kim, D.-K.;
Monkman, E. J.; Shai, D. E.; Wei, H. I.et al. Polycrystalline Graphene with Single Crystalline
Electronic Structure. Nano Lett. 2014, 14 (10), 5706.
(398) Sharma, K. P.; Shinde, S. M.; Rosmi, M. S.; Sharma, S.; Kalita, G.; Tanemura, M. Effect of
copper foil annealing process on large graphene domain growth by solid source-based chemical
vapor deposition. Journal of Materials Science 2016, 51 (15), 7220.

193
(399) Huet, B.; Raskin, J.-P. Role of Cu foil in-situ annealing in controlling the size and thickness of
CVD graphene domains. Carbon 2018, 129, 270.
(400) Bianchini, F.; Patera, L. L.; Peressi, M.; Africh, C.; Comelli, G. Atomic Scale Identification of
Coexisting Graphene Structures on Ni(111). J Phys Chem Lett 2014, 5 (3), 467.
(401) Wang, B.; Bocquet, M. L.; Marchini, S.; Gunther, S.; Wintterlin, J. Chemical origin of a
graphene moire overlayer on Ru(0001). Phys Chem Chem Phys 2008, 10 (24), 3530.
(402) Voloshina, E.; Dedkov, Y. Graphene on metallic surfaces: problems and perspectives. Phys.
Chem. Chem. Phys. 2012, 14 (39), 13502.
(403) Sutter, P. W.; Flege, J. I.; Sutter, E. A. Epitaxial graphene on ruthenium. Nat. Mater. 2008, 7
(5), 406.
(404) Marchini, S.; Günther, S.; Wintterlin, J. Scanning tunneling microscopy of graphene on
Ru(0001). Phys. Rev. B 2007, 76 (7).
(405) Donner, K.; Jakob, P. Structural properties and site specific interactions of Pt with the
graphene/Ru(0001) moire overlayer. J. Chem. Phys. 2009, 131 (16), 164701.
(406) Jiang, D.-e.; Du, M.-H.; Dai, S. First principles study of the graphene/Ru(0001) interface. J.
Chem. Phys. 2009, 130 (7), 074705.
(407) Silva, C. C.; Iannuzzi, M.; Duncan, D. A.; Ryan, P. T. P.; Clarke, K. T.; Küchle, J. T.; Cai, J.;
Jolie, W.; Schlueter, C.; Lee, T.-L.et al. Valleys and Hills of Graphene on Ru(0001). J. Phys.
Chem. C 2018, 122 (32), 18554.
(408) Man, K. L.; Altman, M. S. Small-angle lattice rotations in graphene on Ru(0001). Phys. Rev. B
2011, 84 (23), 235415.
(409) Borca, B.; Barja, S.; Garnica, M.; Minniti, M.; Politano, A.; Rodriguez-Garcí
a, J. M.; Hinarejos,
J. J.; Farías, D.; Parga, A. L. V. d.; Miranda, R. Electronic and geometric corrugation of
periodically rippled, self-nanostructured graphene epitaxially grown on Ru(0001). New J. Phys.
2010, 12 (9), 093018.
(410) Sutter, E.; Albrecht, P.; Wang, B.; Bocquet, M.-L.; Wu, L.; Zhu, Y.; Sutter, P. Arrays of Ru
nanoclusters with narrow size distribution templated by monolayer graphene on Ru. Surf. Sci.
2011, 605 (17), 1676.
(411) Leong, W. S.; Wang, H.; Yeo, J.; Martin-Martinez, F. J.; Zubair, A.; Shen, P.-C.; Mao, Y.;
Palacios, T.; Buehler, M. J.; Hong, J.-Y.et al. Paraffin-enabled graphene transfer. Nat. Commu.
2019, 10 (1), 867.
(412) Suk, J. W.; Kitt, A.; Magnuson, C. W.; Hao, Y.; Ahmed, S.; An, J.; Swan, A. K.; Goldberg, B.
B.; Ruoff, R. S. Transfer of CVD-Grown Monolayer Graphene onto Arbitrary Substrates. ACS
Nano 2011, 5 (9), 6916.
(413) Kang, J.; Shin, D.; Bae, S.; Hong, B. H. Graphene transfer: key for applications. Nanoscale
2012, 4 (18), 5527.

194
(414) N'Diaye, A. T.; Coraux, J.; Plasa, T. N.; Busse, C.; Michely, T. Structure of epitaxial graphene
on Ir(111). New J. Phys. 2008, 10 (4), 043033.
(415) Tang, S.; Wang, H.; Zhang, Y.; Li, A.; Xie, H.; Liu, X.; Liu, L.; Li, T.; Huang, F.; Xie, X.et al.
Precisely aligned graphene grown on hexagonal boron nitride by catalyst free chemical vapor
deposition. Sci. Rep. 2013, 3, 2666.
(416) Pauling, L. The nature of the chemical bond. J. Chem. Educ. 1992, 69 (7), 519.
(417) Yang, G.; Li, L.; Lee, W. B.; Ng, M. C. Structure of graphene and its disorders: a review. Sci
Technol Adv Mater 2018, 19 (1), 613.
(418) Feng, Z.; Dong, H.; Ju, S.; Wen, B.; Zhang, Y.; Melnik, R. High bond difference parameter-
induced low thermal transmission in carbon allotropes with sp2 and sp3 hybridization. Phys.
Chem. Chem. Phys. 2019, 21 (23), 12611.
(419) Díez-Albar, J.; Jiménez-Sánchez, M. D.; Gómez-Rodríguez, J. M. Nanowriting with Clusters
on Graphene on Ru(0001). J. Phys. Chem. C 2019, 123 (9), 5525.
(420) Wang, B.; Bocquet, M.-L. Monolayer Graphene and h-BN on Metal Substrates as Versatile
Templates for Metallic Nanoclusters. J. Phys. Chem. Lett. 2011, 2 (18), 2341.
(421) Yue, K.; Gao, W.; Huang, R.; Liechti, K. M. Analytical methods for the mechanics of graphene
bubbles. J. Appl. Phys. 2012, 112 (8), 083512.
(422) Svatek, S. A.; Scott, O. R.; Rivett, J. P.; Wright, K.; Baldoni, M.; Bichoutskaia, E.; Taniguchi,
T.; Watanabe, K.; Marsden, A. J.; Wilson, N. R.et al. Adsorbate-induced curvature and
stiffening of graphene. Nano Lett. 2015, 15 (1), 159.

195
Acknowledgement
First of all, I would like to thank my supervisor, Prof. Feng Ding, for his careful guidance on
this dissertation. Prof. Feng Ding is a model for me in my current and future work and life. His broad
academic horizon, rigorous academic attitude and open mode of thinking deeply impress me and
encourage me to keep moving forward. During the four years of my Ph.D. career, Prof. Feng Ding has
given me the most selfless encouragement and support in my study, especially in the topic selection,
research process and thesis writing. Here I would like to extend my highest respect and sincere thanks
to my supervisor and wish Prof. Feng Ding good health and good luck in everything.

I would like to thank all supports of Ulsan National Institute of Science and Technology
(UNIST) and Center for Multidimensional Carbon Materials, Institute for Basic Science (IBS-CMCM)
for creating a good condition for learning and scientific research. Thanks to all my friends and
colleagues of Prof. Feng Ding’s group for their encouragement and suggestions on my study and life,
their persistent pursuit on science inspires me to keep moving forward.

Thanks to all my collaborators, including Prof. Haofei Shi and Dr. Xin Li from Chinese
Academy of Sciences, Prof. Guilin Wu from Chongqing University, Prof. Kaihui Liu and Dr. Li Wang
from Peking University, Prof. Rodney Ruoff and Dr. Sunghwan Jin from IBS-CMCM, for their
experimental support on my theoretical research.

Thanks to my parents and family members, their endless love for me is the power of my
progress, I will use my whole lifetime to repay them. Special thanks to my husband Dr. Jichen Dong
for his company and support in my study, life and all other aspects.

Finally, I would like to express my sincere thanks and blessings to all the professors of the
thesis committee for their review and evaluation on this dissertation and participation in my defense.

196
197

You might also like