+ All Categories
Home > Documents > Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that...

Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that...

Date post: 16-Jul-2020
Category:
Upload: others
View: 8 times
Download: 0 times
Share this document with a friend
207
1 Probing and Manipulating the Interlayer Coupling in 2D Structures by Zachariah Boston Hennighausen Bachelor of Science in Electrical Engineering, United States Coast Guard Academy A dissertation submitted to The Faculty of the College of Science of Northeastern University in partial fulfillment of the requirements for the degree of Doctor of Philosophy May 02, 2019 Dissertation directed by Dr. Professor Swastik Kar Associate Professor of Physics
Transcript
Page 1: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

1

Probing and Manipulating the Interlayer Coupling in 2D Structures

by Zachariah Boston Hennighausen

Bachelor of Science in Electrical Engineering, United States Coast Guard Academy

A dissertation submitted to

The Faculty of

the College of Science of

Northeastern University

in partial fulfillment of the requirements

for the degree of Doctor of Philosophy

May 02, 2019

Dissertation directed by

Dr. Professor Swastik Kar

Associate Professor of Physics

Page 2: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

2

Acknowledgments

The list of people to thank and acknowledge who have made this dissertation and work possible

is extensive to say the least. The indirect contributions have been equally as influential as the

direct contributions, and often small acts have had an outsized influence. This is in part due to

the nature of science, where progress is often non-linear and unpredictable, and unexpected

sources can stimulate significant discoveries. However, it is also due to the fact that our work

lives are not decoupled or independent from our personal lives. The people who have developed

or assisted me with non-science endeavors, both during and before starting graduate school, has

arguably been more influential than those related to science. The list is surely extensive and I am

appreciative of everyone. Although I want to thank every person in this section, an important

component of science writing is the ability to compress a large volume of data into a concise piece,

and I will attempt apply this here.

Friends and shipmates are constantly challenging me with new adventures that inspire creativity

and allow for reflection periods. Additionally, they have generously inquired into my work with

difficult and probing questions that have improved not only my own understanding, but also

how to convey and explain the work. Dr. Albert Einstein is credited with saying,

“If you can’t explain it simply, you don’t understand it well.”

An interested audience offers scientists an opportunity to explain and teach their research, a

process that can improve the scientist’s own understanding. One such notable occasion was in

Tahoe, CA in 2018 where very late into the night after an extreme day, some friends asked to hear

my APS March Meeting presentation on my work, and honored me by actively engaging with

my presentation. These friends and shipmates include: Travis Murray for teaching me of the

wealth of opportunities that exist globally; Emily Theresa Sonner Kehrt for engaging discussions

that probe the soul and foundation of my research; Kip Hutchinson for inspiring creativity using

unlikely sources; and Kevin Sorrell for highly technical discussions on a spectrum of topics. I

Page 3: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

3

would also like to thank Luke Bruns, Jason Acuna, Nicole Bredariol, Tommy Adams, Nicholas

Custer, and, to a very indirect extent, Rossi.

Both my extended and immediate family have provided guidance, mentorship, sage advice, and

counsel on a range of questions and obstacles directly related to my research, but also outside of

it. By solving problems and challenges more quickly, it freed-up energy, time, and mental stamina

for me to focus on research. There is no doubt that without them this dissertation would not be

where it is today. In particular, I would like to thank Molly Hennighausen and Hannah

Hennighausen for highly engaging discussions on a range of topics, the fruitful feedback, and

wise advice. I would also like to thank Ingrid Butler, Frank Butler, Maija Butler, and Lisa Butler

for providing feedback and smart insight on a spectrum of topics. In particular, it was Ingrid

Butler who alerted me to the history of the moiré pattern and its relation to the fabric (a fact that

shaped my writing), and Frank Butler for providing unique insight into the properties of bulk

MoS2.

My parents, Dr. Lothar Hennighausen and Dr. Prof. Priscilla Furth, deserve not only their own

paragraph acknowledging their support, but an entire dissertation. However, they likely would

advise me not to invest too much time thanking them, and to keep this paragraph practical and

short. They have built the foundation for the person I am today, and they have continued to guide

me throughout every step of my life. I have only achieved what I have due to their guidance and

hard work.

I have worked with numerous scientific groups spanning a variety of disciplines throughout my

Ph.D., and have sent samples to many laboratories located across the globe. Frequently these

collaborations led to a fruitful exchange of bold ideas, and in a few instances we were able to

produce exciting results that are currently published or will be soon. I am very thankful to all my

collaborators. In particular, I would like to thank Dr. Prof. Arun Bansil at Northeastern University

Page 4: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

4

and his group, which focuses on theoretical condensed matter physics. I believe that some of the

strongest scientific discoveries are when theory and experiment collaborate, and I feel very

fortunate I was able to collaborate with such an incredible theory group. They have been not only

very supportive of my experiments, but they also played a critical role in developing me as a

physicist. I worked mostly with Christopher Lane and Dr. Gianina Buda, and will cherish their

friendship forever.

During my Ph.D. I was fortunate enough to learn a large number of techniques and how to

operate a variety of complex scientific equipment, and this was only possible due to the hard

work of my lab mates. I would like to thank the following people for assistance with various

experimental techniques: Dr. Fangze Liu, Dr. Anthony Vargas, Dr. Ismail Bilgin, Dr. Daniel

Rubin, Colin Casey, Monika Eggenberger, Kevin Mendez, Dr. Pradeep Waduge, Dr. Wentao

Liang, and Vineet Mathur.

I would argue that I might have had one of the best Ph.D. advisors that exist, and am forever

exceptionally grateful for all his hard work, guidance, mentorship, patience, and friendship. Dr.

Prof. Swastik Kar always strived to go above and beyond, and put being a good mentor above all

else. He provided me with all the resources, freedom, and opportunities needed to be successful.

If I began to stray too far into the scientific jungle and was in danger of getting lost, he made sure

to guide me back.

Lastly, I would like to thank Kristi Butler (Buthausen) for being the force that made all this

possible. On a daily basis she provides sage wisdom for challenges large and small, and has gone

to impressive lengths to help me. She has been nothing less than an irreplaceable teammate and

partner, and there is no doubt most of the thanks and acknowledgment goes to her. And, finally,

our two furry roommates, Madame and Landau, deserve acknowledgment; they were always

there during both the sunny days and the colder nights.

Page 5: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

5

Abstract of Dissertation (English)

In 2004, researchers reported the synthesis of planar graphene in the free state, overturning a

decades-old prediction that 2D crystals were too thermodynamically unstable to exist in ambient.

Since then, dozens of new monolayer crystals have been demonstrated, each with their own

properties and capabilities. Very interestingly, as these monolayer materials are vertically stacked

to create 2D structures, the properties of the structure are not simply a sum of the parent crystals,

but a product of the interlayer coupling, which can redistribute charge, modify the band

structures, and induce new properties unique to either parent crystal. The term interlayer

coupling refers to the communication and interaction that exists between the parent materials,

where the absence of a coupling suggests the parent materials are electronically independent.

Understanding and controllably manipulating the interlayer coupling holds promise to not only

engineer new capabilities, but to also retain desired properties that only exist in monolayer

materials. This dissertation explores the interlayer coupling in a variety 2D structures, with an

emphasis on a particular class of 2D structures: monolayer bismuth selenide (Bi2Se3) grown on

monolayer transition metal dichalcogenides (TMDs), using vapor-phase chalcogenization. The

physical charge redistribution induced by the interlayer coupling, and its effect on the band

structure, was studied in as-grown samples. The interlayer coupling was then probed and

manipulated using both electron beams in a vacuum, and the controlled absorption and

desorption - possible intercalation and deintercalation - of oxygen. The results suggest that the

properties could be drastically and controllably altered by varying the interlayer coupling

strength.

Chapter 1 is an introduction and background to monolayer materials and 2D structures, as well

as bismuth selenide (Bi2Se3) and transition metal dichalcogenides (TMDs), the parent materials

for the primary class of 2D structures studied in this dissertation.

Chapter 2 is instrumentation and synthesis, where all the important methods and equipment are

discussed, so as to enable reproduction of the work.

Page 6: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

6

Chapter 3 demonstrates that the interlayer coupling between dissimilar 2D heterostructures can

be probed in-situ by manipulating the twist angle and structure using the focused electron beam

of a transmission electron microscope (TEM). The electron beam imparts energy that induces the

Bi2Se3 to break-up into grains, and for those grains to twist relative to the underlying monolayer

TMD.

Chapter 4 demonstrates oxygen-induced in-situ manipulation of the interlayer coupling and

exciton recombination in Bi2Se3/MoS2 2D heterostructures. This chapter studies the possible

intercalation and deintercalation of oxygen using both experimental and theoretical methods,

providing justification for the claims, as well as other possible explanations for the manipulation

of the interlayer coupling.

Chapter 5 demonstrates the tunable photoluminescence in Bi2Se3/TMD 2D heterostructures for

write-read-erase-reuse applications. Mono- and few-layer Bi2Se3 was grown on four different

TMDs – MoS2, MoSe2, WS2, and MoSe2-2xS2x – and the interlayer coupling was manipulated

possibly using the controlled intercalation and deintercalation of oxygen. Potential technologies

include ultra high-density information storage, and tunable photoluminescing pixels (PLPs).

Chapter 6 demonstrates evidence of a purely electronic two-dimensional lattice at room

temperature that resides between the parent layers of a Bi2Se3/TMD 2D heterostructure. The

results suggest that the interlayer coupling is inducing significant charge redistribution, a

surprising result considering that that interlayer bonding in their bulk counterparts is “weak”

van der Waals.

Chapter 9.1 studies and briefly discusses work related to the interlayer coupling in Non-Bi2Se3 2D

structures that include, graphene/MoS2, graphene/graphene, Sb2Se3/MoS2, and Bi2Te3/MoS2. The

results suggest that while interlayer coupling is important in other 2D structures, the Bi2Se3/TMD

2D structures have a relatively strong interlayer coupling.

Page 7: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

7

Abstract of Dissertation (Deutsch)

Im Jahr 2004 publizierten Forscher über die Darstellung von planarem Graphen im freien

Zustand, und darbei widerlegte eine jahrzehntealte Vorhersage, dass 2D-Kristalle nachweislich

thermodynamisch instabil sind um im Außenumgebung zu existieren. Seitdem wurden

Dutzende neuer Monoschichtkristalle mit jeweils eigenen Eigenschaften und Fähigkeiten

demonstriert. Da diese Monoschichtmaterialien zur Erzeugung von 2D-Strukturen vertikal

gestapelt sind, sind die Eigenschaften der Struktur nicht einfach eine Summe der Stammkristalle,

sondern auch ein Produkt der Zwischenschichtkopplung, die Ladung umverteilt, die

Bandstrukturen modifiziert, und neue Eigenschaften erzeugt die nicht in die Stammkristalle zu

finden sind. Der Begriff Zwischenschichtkopplung bezieht sich auf die Kommunikation und

Interaktion zwischen dis Stammkristallen, wobei das Fehlen einer Kopplung darauf hindeutet,

dass die Stammkristallen elektronischunabhängig sind. Der Zwischenschichtkopplung zu

verstehen und kontrolibar zu manipulaieren verspricht, nicht nur neue Fähigkeiten zu

entwickeln, sondern auch gewünschte Eigenschaften beizubehalten, die nur in

Monoschichtmaterialien vorhanden sind. Diese Dissertation untersucht die Kopplung zwischen

Schichten in verschiedenen 2D-Strukturen (bzw. Zwischenschichtkopplung), wobei einer

bestimmten Klasse von 2D-Strukturen liegt unter dem Brennpunkt: Monoschicht-Bismutselenid

(Bi2Se3), das auf Einschicht-Übergangsmetalldichalcogeniden (TMDs) mit Dampfphasen-

Chalcogenisierung gewachsen ist. Die physikalische Ladungsumverteilung durch die

Zwischenschichtkopplung und ihre Auswirkung auf die Bandstruktur wurden in

selbstgewachsenen Proben untersucht. Die Zwischenschichtkopplung wurde dann unter

Verwendung von beider Elektronenstrahlen im Vakuum, und der kontrollierten Absorption und

Abscheidung - mögliche Interkalation und Deinterkalation - von Sauerstoff untersucht und

manipuliert. Die Ergebnisse legen nahe, dass die Eigenschaften durch Variation der

Zwischenschichtkopplungsstärke drastisch und kontrollierbar verändert werden könnten.

Kapitel 1 ist eine Einführung und Hintergrund zu Monoschichtmaterialien und 2D-Strukturen,

sowie Monoschicht-Bismutselenid (Bi2Se3) und Übergangsmetalldichalcogeniden (TMDs), die

Page 8: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

8

Stammkristalle für die 2D-Strukturen die in dieser Dissertation hauptsächlich untersuchte

werden.

Kapitel 2 ist Instrumentierung und Synthese, in der alle wichtigen Methoden und Geräte

besprochen werden, um die Reproduktion der Arbeit zu ermöglichen.

In Kapitel 3 wird gezeigt, dass die Zwischenschichtkopplung zwischen unähnlichen 2D-

Heterostrukturen in-situ durch Manipulation des Verdrehungswinkels und der Struktur unter

Verwendung des fokussierten Elektronenstrahls eines Transmissionselektronenmikroskops

(TEM) untersucht werden kann. Der Elektronenstrahl vermittelt Energie, die das Bi2Se3 dazu

veranlasst, sich in Körner aufzubrechen, und diese Körner verdrehen sich relativ zur

darunterliegenden Monoschicht- Übergangsmetalldichalcogeniden (bzw. Monoschicht-TMD).

Kapitel 4 zeigt die sauerstoffinduzierte In-situ-Manipulation der Zwischenschichtkopplung und

der Exziton-Rekombination (bzw. Elektron-Loch-Paar-Rekombination) in Bi2Se3/MoS2-2D-

Heterostrukturen. In diesem Kapitel wird die mögliche Interkalation und Deinterkalation von

Sauerstoff anhand experimenteller und theoretischer Methoden untersucht, um die Ansprüche

zu begründen, sowie andere mögliche Erklärungen für die Manipulation der

Zwischenschichtkopplung.

Kapitel 5 demonstriert die abstimmbare Photolumineszenz in Bi2Se3/TMD-2D-Heterostrukturen

für Schreiben-Lesen-Löschen-Wiederverwenden Anwendungen. Monoschicht und

Multischichten-Bi2Se3 wurde auf vier verschiedenen Übergangsmetalldichalcogeniden (TMDs) -

MoS2, MoSe2, WS2 und MoSe2-2xS2x - gezüchtet, und die Zwischenschichtkopplung wurde

manipuliert, möglicherweise unter Verwendung der kontrollierten Interkalation und

Deintercalation von Sauerstoff. Mögliche Technologien und Anwendungen umfassen ultradichte

Datenspeicher, und abstimmbare Photolumineszierendepixeln (PLPs).

Kapitel 6 zeigt den Nachweis eines rein-elektronischen zweidimensionalen Gitter bei

Raumtemperatur, das sich zwischen den Stammschichten einer Bi2Se3/TMD-2D-Heterostruktur

befindet. Die Ergebnisse deuten darauf hin, dass die Zwischenschichtkopplung eine signifikante

Ladungsumverteilung bewirkt, ein überraschendes Ergebnis, wenn man bedenkt, dass die

Page 9: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

9

Zwischenschichtbindung in ihren Massengegenstücken "schwach" van der Waals ist. Die Arbeit

könnte sich auch mit Zweidimensionalelektronkristalle befassen.

In Kapitel 9.1 werden die Arbeiten zur Zwischenschichtkopplung in Kein-Bi2Se3-2D-Strukturen,

darunter Graphene/MoS2, Graphene/Graphene, Sb2Se3/MoS2, und Bi2Te3/MoS2, beschrieben. Die

Ergebnisse legen nahe, dass die Zwischenschichtkopplung zwar in anderen 2D-Strukturen

wichtig ist, die Bi2Se3/TMD-2D-Strukturen jedoch eine relativ starke Zwischenschichtkopplung

aufweisen.

Page 10: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

10

Table of Contents

Acknowledgments .................................................................................................................................... 2

Abstract of Dissertation (English) ......................................................................................................... 5

Abstract of Dissertation (Deutsch) ........................................................................................................ 7

Table of Contents .................................................................................................................................... 10

List of Tables ............................................................................................................................................ 15

List of Equations ..................................................................................................................................... 16

Chapter 1: Introduction ..................................................................................................................... 17

1.1 Monolayer materials: the building blocks for 2D structures ............................................. 17

1.2 2D Materials: overview of specific monolayer materials ................................................... 20

1.3 2D Structures: combining monolayer crystals to build something new .......................... 23

1.4 2D Structures: a product of the interlayer interaction ........................................................ 30

1.5 Methods to manipulate the interlayer coupling in 2D structures ..................................... 35

Chapter 2: Instrumentation and Synthesis .................................................................................... 43

2.1 Growing 2D materials using Chemical Vapor Deposition (CVD) .................................... 43

2.2 Mechanically transferring 2D materials to create 2D structures ....................................... 44

2.3 Optical setup for Raman and photoluminescence spectroscopy ...................................... 45

2.4 Manipulating interlayer coupling in Bi2Se3/TMD 2D heterostructures using laser

exposure and thermal annealing ....................................................................................................... 46

2.5 Atomic Force Microscope (AFM) ........................................................................................... 48

2.6 Scanning Electron Microscope (SEM) ................................................................................... 48

2.7 Lithography: E-Beam, Photo, and Laser ............................................................................... 49

2.8 Transmission Electron Microscope (TEM) ........................................................................... 50

2.9 Electron Beam Treatment (EBT) Studies: Manipulating the twist angle and interlayer

interface ................................................................................................................................................. 50

2.10 Computational Methods ......................................................................................................... 52

2.11 Device Fabrication.................................................................................................................... 53

Chapter 3: Probing the interlayer interaction between dissimilar 2D heterostructures by in-

situ manipulation of twist angle and structure ................................................................................. 54

3.1 Monte Carlo Simulations Reveal Statistical Significance of Histograms ......................... 81

Page 11: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

11

3.2 Material Transformation vs. Twist Angle Manipulation .................................................... 84

3.3 Negligible Role of Potential Defects on the Interlayer Interaction ................................... 85

3.4 Discussion Grain Rotation Mechanisms ............................................................................... 87

Chapter 4: Oxygen-Induced In-Situ Manipulation of the Interlayer Coupling and Exciton

Recombination in Bi2Se3/MoS2 2D Heterostructures ....................................................................... 89

4.1 Discussion of Possible exciton recombination pathways ................................................. 112

4.2 Discussion on interaction of Bi2Se3 with oxygen ............................................................... 118

Chapter 5: Tunable Photoluminescence in Bi2Se3/TMD 2D Heterostructures for Potential

Write-Read-Erase-Reuse Applications .............................................................................................. 121

5.1 Laser-Annealing in Air, and Strong Stability of As-Grown Configuration ................... 139

Chapter 6: Evidence of a purely electronic two-dimensional lattice at room temperature 142

6.1 Diophantine equation and guidance on mathematically calculating the electronic moiré

lattice commensurate unit cell. ......................................................................................................... 150

6.2 Data demonstrating that the electronic moiré lattice was observed in other 2D

heterostructures .................................................................................................................................. 151

6.3 Brief explanation of double diffraction with respect to electron diffraction ................. 157

Chapter 7: Conclusion and Discussion ......................................................................................... 159

Chapter 8: Dissemination ................................................................................................................ 164

8.1 Publications ............................................................................................................................. 164

8.2 Patents ...................................................................................................................................... 165

8.3 Presentations - Conferences .................................................................................................. 165

8.4 Presentations – Invited Talks ................................................................................................ 165

Chapter 9: Appendix ........................................................................................................................ 167

9.1 Non-Bi2Se3 2D Heterostructures........................................................................................... 167

9.1.1 Graphene/MoS2 Heterostructures ................................................................................ 167

9.1.2 Graphene/Graphene Heterostructures........................................................................ 169

9.1.3 MoS2/Sb2Se3 Heterostructures ....................................................................................... 171

9.1.4 MoS2/Bi2Te3 Heterostructures ....................................................................................... 176

9.2 Programming Code ............................................................................................................... 178

9.2.1 Fitting Lorentzian Functions to Exciton Photoluminescence Spectra .................... 178

9.2.2 Raw Search for Commensurate Unit Cells and Moiré Patterns .............................. 182

References............................................................................................................................................... 185

Page 12: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

12

List of Figures

FIGURE 1.1: ARTISTIC DIAGRAM OF A GENERAL 2D MATERIAL ................................................................................. 17

FIGURE 1.2: (A) GRAPHITE AND (B) DIAMOND CRYSTAL STRUCTURES. ..................................................................... 19

FIGURE 1.3: ARTISTIC DIAGRAMS OF (A) GRAPHENE AND (B) A MONOLAYER TRANSITION METAL

DICHALCOGENIDE (TMD).................................................................................................................................. 20

FIGURE 1.4: EXPERIMENTAL AND THEORY BAND STRUCTURE FOR THE TOPOLOGICAL INSULATOR BI2SE3. ............. 22

FIGURE 1.5: ARTISTIC DIAGRAM SHOWING A REPRESENTATIVE 2D STRUCTURE. ...................................................... 24

FIGURE 1.6: GRAPHENE BAND STRUCTURE, AND DEMONSTRATION OF THE EVOLUTION OF THE BAND STRUCTURE

AS GRAPHENE LAYERS ARE STACKED. ................................................................................................................ 25

FIGURE 1.7: DEMONSTRATION HOW THE PL IS MODIFIED BY SIMPLY INCREASING THE LAYER COUNT OF MOS2 .... 27

FIGURE 1.8: ANGLE RESOLVES PHOTOEMISSION SPECTROSCOPY (ARPES) OF BI2SE3 ............................................... 29

FIGURE 1.9: ARTISTIC DIAGRAM ILLUSTRATING THE POSSIBILITIES OF DESIGNING VERTICALLY STACKED 2D

STRUCTURES ........................................................................................................................................................ 30

FIGURE 1.10: DENSITY FUNCTIONAL THEORY (DFT) CALCULATED BI2SE3/MOSE2 2D HETEROSTRUCTURE ............ 32

FIGURE 1.11: DFT CALCULATED 2D STRUCTURE WITH ONE WS2 LAYER ON THE BOTTOM AND ONE BI2SE3 LAYER

ON THE TOP ......................................................................................................................................................... 34

FIGURE 1.12: TUNING THE INTERLAYER INTERACTION AND COUPLING USING THE INTERCALATION OF MOLECULES

AND ATOMS. ....................................................................................................................................................... 36

FIGURE 1.13: VARIOUS METHODS THAT MODIFY THE INTERLAYER COUPLING IN 2D STRUCTURES. ........................ 38

FIGURE 1.14: DFT CALCULATED 2D STRUCTURES WITH ONE LAYER WS2 AND ONE LAYER BI2SE3 .......................... 40

FIGURE 1.15: TWO POWERFUL METHODS TO CONSTRUCT AND MANIPULATE 2D STRUCTURES ............................... 41

FIGURE 2.1: OPTICAL PICTURE OF BI2SE3 2D HETEROSTRUCTURES TRANSFERRED ONTO A TEM GRID .................. 51

FIGURE 3.1: ELECTRON BEAM PROBES INTERLAYER INTERACTION IN 2D HETEROSTRUCTURES. .............................. 56

FIGURE 3.2: AS-GROWN 2D HETEROSTRUCTURES PRIOR TO ELECTRON BEAM TREATMENT (EBT). ........................ 58

FIGURE 3.3: REPRESENTATIVE BI2SE3/TMD HETEROSTRUCTURES. ............................................................................ 59

FIGURE 3.4: TEM SAED PATTERNS OF FOUR 2D HETEROSTRUCTURES STUDIED, AS WELL AS BI2SE3 GROWN ON

SIO2. .................................................................................................................................................................... 60

FIGURE 3.5: TEM IMAGES OF DIFFERENT 2D HETEROSTRUCTURES. .......................................................................... 60

FIGURE 3.6 TUNING THE INTERLAYER INTERACTION THROUGH CHEMISTRY. HISTOGRAMS OF THE TWIST ANGLES

............................................................................................................................................................................ 62

FIGURE 3.7 MANIPULATING THE TWIST ANGLE IN-SITU USING AN ELECTRON BEAM. .............................................. 65

FIGURE 3.8 IN-SITU CRYSTAL MANIPULATION OF BI2SE3/MOSE2 USING AN ELECTRON BEAM. ................................. 68

FIGURE 3.9 BI2SE3/WS2 CONTROLLABLE CRYSTAL DISRUPTION, ANNEALING, AND ROTATION IN RESPONSE TO EBT.

............................................................................................................................................................................ 69

FIGURE 3.10 CONTROLLABLE CRYSTAL REARRANGEMENT IN-SITU USING AN ELECTRON BEAM. ............................ 70

FIGURE 3.11 SUMMARY OF DIFFERENT 2D HETEROSTRUCTURE’S RESPONSE TO AN EBT INTENSITY. ....................... 72

FIGURE 3.12 MANIPULATION OF THE OPTICAL AND PHOTOLUMINESCENCE PROPERTIES USING AN ELECTRON

BEAM. .................................................................................................................................................................. 74

FIGURE 3.13 MANIPULATION OF THE OPTICAL AND PHOTOLUMINESCENCE PROPERTIES USING AN ELECTRON

BEAM. .................................................................................................................................................................. 77

FIGURE 3.14 BI2SE3/MOSE2 AFFECTED WITH ELECTRON BEAM TREATMENT (EBT). .................................................. 78

FIGURE 3.15 TEXTURE AND WORK FUNCTION INDUCED CHANGES FROM EBT. ........................................................ 79

Page 13: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

13

FIGURE 3.16 MONTE CARLO SIMULATION: INSTANCE NOISE FLOOR VALUES. ........................................................ 82

FIGURE 3.17 MONTE CARLO SIMULATION DATA TO IDENTIFY THE INSTANCE NOISE FLOORS (INF) ....................... 83

FIGURE 3.18 DATA FROM THE MONTE CARLO SIMULATIONS USED TO CALCULATE THE INSTANCE NOISE FLOOR . 84

FIGURE 3.19: REPRESENTATIVE RAMAN SPECTRA OF A BI2SE3/MOS2 2D HETEROSTRUCTURE ................................. 86

FIGURE 4.1: FIGURE BRIEFLY SUMMARIZING THE CAPABILITIES DEMONSTRATED IN CHAPTER 4. ............................ 90

FIGURE 4.2: AS-GROWN BI2SE3/MOS2 VERTICAL 2D HETEROSTRUCTURES. ............................................................... 92

FIGURE 4.3: MANIPULATING THE INTERLAYER COUPLING USING OXYGEN AND ENERGY. ....................................... 93

FIGURE 4.4: ATOMIC FORCE MICROSCOPE (AFM) IMAGES OF 2D HETEROSTRUCTURE THERMALLY TREATED IN AIR.

............................................................................................................................................................................ 95

FIGURE 4.5: AFM ANALYSIS OF 2D HETEROSTRUCTURE: AS-GROWN VS. LASER EXPOSED IN AIR. ............................ 95

FIGURE 4.6: OXYGEN-INDUCED MANIPULATION OF EXCITON DYNAMICS USING LASER SETUP. ............................... 96

FIGURE 4.7: 2D HETEROSTRUCTURES THERMALLY TREATED IN AR (I.E. AN O2-FREE ENVIRONMENT) AT 240°C FOR

3 HOURS. ............................................................................................................................................................. 98

FIGURE 4.8: 2D HETEROSTRUCTURE THERMALLY TREATED IN DRY AIR (I.E. 79% N2 + 21% O2). ............................ 99

FIGURE 4.9: 2D HETEROSTRUCTURE THERMALLY TREATED IN AN N2+H2OVAPOR ENVIRONMENT. .......................... 100

FIGURE 4.10: 2D HETEROSTRUCTURES FIRST THERMALLY TREATED IN AIR, AND THEN IN N2.

PHOTOLUMINESCENCE SPECTRA AT DIFFERENT LOCATIONS. ......................................................................... 101

FIGURE 4.11: 2D HETEROSTRUCTURES THERMALLY TREATED IN AIR AND THEN IN N2. PL SPECTRA AT THE SAME

LOCATION. ........................................................................................................................................................ 102

FIGURE 4.12: TUNING THE INTERLAYER COUPLING AND PHOTOLUMINESCENCE INTENSITY. ................................ 104

FIGURE 4.13: HIGH TUNABILITY AND CONTROL OF INTERLAYER COUPLING AND PL INTENSITY. ......................... 105

FIGURE 4.14: DFT CALCULATIONS PREDICT INTERCALATED O2 DIMINISHES INTERLAYER COUPLING .................. 106

FIGURE 4.15: SITE-SELECTIVE MANIPULATION OF INTERLAYER COUPLING AND PHOTOLUMINESCENCE. ............. 107

FIGURE 4.16: RAMAN RESPONSE OF BI2SE3/MOS2 PRE AND POST LASER EXPOSURE. ............................................... 109

FIGURE 4.17: POSSIBLE NON-RADIATIVE ELECTRON-HOLE RECOMBINATION PATHWAYS LEADING TO PL

QUENCHING IN AS-GROWN 2D HETEROSTRUCTURES. ..................................................................................... 114

FIGURE 4.18: DFT CALCULATIONS COMPARING A BI2SE/MOS2 SUPERLATTICE WITH AND WITHOUT O2

INTERCALATED. ................................................................................................................................................ 116

FIGURE 4.19: DFT CALCULATIONS SHOWING THE LOCATION OF THE O2. .............................................................. 116

FIGURE 4.20: DIAGRAMS DEMONSTRATING THE CREATION OF GRAIN BOUNDARIES, FACILITATING O2

INTERCALATION INTO THE INTERLAYER REGION. ........................................................................................... 117

FIGURE 5.1: FIGURE BRIEFLY SUMMARIZING THE CAPABILITIES DEMONSTRATED IN THIS CHAPTER. ..................... 121

FIGURE 5.2: REVERSIBLE OXYGEN ABSORPTION AND ITS ROLE IN MANIPULATING THE PHOTOLUMINESCENCE

INTENSITY ......................................................................................................................................................... 123

FIGURE 5.3 AS-GROWN BI2SE3/TMD 2D HETEROSTRUCTURES ................................................................................ 124

FIGURE 5.4: REPRESENTATIVE SAED IMAGES OF AS-GROWN AND LASER-EXPOSED 2D HETEROSTRUCTURES ...... 125

FIGURE 5.5: BI2SE3/MOSE2(1-X)S2XALLOY SAED AND TEM IMAGES. ............................................................................. 126

FIGURE 5.6: MANIPULATING THE OPTICAL AND EXCITONIC PROPERTIES IN BI2SE3/TMD 2D HETEROSTRUCTURES

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

FIGURE 5.7: DEMONSTRATION THAT LASER EXPOSURE IN AIR EFFECTS ARE REVERSED THROUGH ANNEALING ... 128

FIGURE 5.8: COMPLEX PATTERNS AND GRIDS CAN BE CONTROLLABLE WRITTEN INTO BI2SE3/TMD 2D

HETEROSTRUCTURES ......................................................................................................................................... 129

FIGURE 5.9: ATOMIC FORCE MICROSCOPY (AFM) ANALYSIS OF BI2SE3/MOSE2 AND BI2SE3/WS2 ......................... 130

FIGURE 5.10: RAMAN SPECTROSCOPY STUDY OF BI2SE3/MOSE2 .............................................................................. 131

Page 14: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

14

FIGURE 5.11: RAMAN SPECTROSCOPY STUDY OF BI2SE3/MOSE2 AND BI2SE3/WS2 ................................................... 132

FIGURE 5.12: BI2SE3/WS2 SAED IMAGES AFTER LASER EXPOSURE IN AIR, AND THEN ANNEALED IN AR. ............. 133

FIGURE 5.13: TUNING THE PL INTENSITY AND PEAK POSITION ............................................................................... 133

FIGURE 5.14: MANIPULATION OF BOTH EXCITON POPULATIONS AND THEIR RECOMBINATION ENERGY ............... 134

FIGURE 5.15: REPRESENTATIVE A AND A- FITS FOR PHOTOLUMINESCENCE SPECTRA ............................................ 136

FIGURE 5.16: ANNEALING REVERSES THE CHANGES TO OPTICAL AND EXCITONIC PROPERTIES. ............................ 137

FIGURE 5.17 VERY LOW LASER POWER CAN ANNEAL BI2SE3/TMD 2D HETEROSTRUCTURES. ................................ 140

FIGURE 5.18 PROBING THE PL STABILITY OF MOIRÉ HETEROSTRUCTURES. ............................................................. 141

FIGURE 6.1 DIRECT OBSERVATION OF A NEW LATTICE IN SELECTED AREA ELECTRON DIFFRACTION OF VERTICALLY

STACKED BI2SE3/TMD 2D HETEROSTRUCTURES. ............................................................................................. 143

FIGURE 6.2 ELECTRONIC MOIRÉ LATTICE SCATTERING SITES FORMED BY THE INTERLAYER COUPLING. ................ 145

FIGURE 6.3 THE MOIRÉ SAED SPOTS ARE NOT A PRODUCT OF ONE OF THE INDIVIDUAL LATTICES. ...................... 146

FIGURE 6.4 CHARGE POOLS FORM BETWEEN NEAREST INTERLAYER NEIGHBORS. ................................................... 148

FIGURE 6.5 GUIDANCE ON CALCULATING THE MOIRÉ SUPERLATTICE COMMENSURATE UNIT CELL. ..................... 151

FIGURE 6.6 EXPERIMENTAL, MOIRÉ MODELING, AND DFT DATA FOR A BI2SE3/MOS2 2D HETEROSTRUCTURE. .... 153

FIGURE 6.7 EXPERIMENTAL, MOIRÉ MODELING, AND DFT DATA FOR A BI2SE3/MOSE2 2D HETEROSTRUCTURE. .. 154

FIGURE 6.8 EXPERIMENTAL AND MOIRÉ MODELING IMAGES FOR DIFFERENT BI2SE3/MOSE2-2XS2X 2D

HETEROSTRUCTURES AT VARIOUS TWIST ANGLES............................................................................................ 155

FIGURE 6.9 EXPERIMENTAL AND MOIRÉ MODELING IMAGES FOR A BI2SE3/WS2 2D HETEROSTRUCTURE .............. 156

FIGURE 6.10 EXPERIMENTAL AND MOIRÉ MODELING IMAGES FOR BI2SE3/WS2 2D HETEROSTRUCTURES FROM

FIGURE 2. .......................................................................................................................................................... 156

FIGURE 6.11: SIMPLIFIED DESCRIPTION OF DOUBLE DIFFRACTION DUE TO TEM ELECTRONS. ................................ 158

FIGURE 9.1 MOS2 GROWN ON TOP OF GRAPHENE TO CREATE MOS2/GRAPHENE 2D HETEROSTRUCTURES. .......... 168

FIGURE 9.2 MOS2 GROWN ON GRAPHENE USING THE SECONDARY ELECTRON SEM IMAGING MODE. .................. 168

FIGURE 9.3 ARTISTIC DIAGRAMS DEMONSTRATING MECHANICALLY TRANSFERRING ECONOMICAL GRAPHENE

PRODUCES A CONDUCTING PATH. ................................................................................................................... 170

FIGURE 9.4 IMAGES OF AS-GROWN GRAPHENE MECHANICALLY TRANSFERRED ONTO GLASS. .............................. 171

FIGURE 9.5 OPTICAL IMAGES OF MONOLAYER MOS2 PRIOR TO GROWTH OF SB2SE3. .............................................. 172

FIGURE 9.6 OPTICAL IMAGES OF HETEROSTRUCTURES WHERE SB2SE3 WAS GROWN ON TOP OF MONOLAYER MOS2.

.......................................................................................................................................................................... 173

FIGURE 9.7 PHOTOLUMINESCENCE AND RAMAN SPECTROSCOPY FROM SB2SE3/MOS2 HETEROSTRUCTURES. ....... 173

FIGURE 9.8 TEM SAED IMAGES AND TEM IMAGES OF SB2SE3/MOS2 HETEROSTRUCTURES. ................................. 175

FIGURE 9.9 OPTICAL IMAGES OF MONOLAYER MOS2 AND BI2TE3 GROWN ON MONOLAYER MOS2. ...................... 176

FIGURE 9.10 PHOTOLUMINESCENCE AND RAMAN SPECTRA OF BI2TE3/MOS2 HETEROSTRUCTURES. ..................... 177

Page 15: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

15

List of Tables

TABLE 2.1 MONOLAYER TMD GROWTH PARAMETERS USING VAPOR-PHASE CHALCOGENIZATION (VPC). .......... 44

TABLE 2.2: LOWEST LASER EXPOSURE POWERS FOR ALL BI2SE3/TMD 2D HETEROSTRUCTURES TESTED. ................. 46

TABLE 2.3: DETAILED LASER EXPOSURE PARAMETERS FOR EXPERIMENTS DESCRIBED IN CHAPTER 4. ..................... 47

Page 16: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

16

List of Equations

EQUATION 1.1: SCHRÖDINGER EQUATION ................................................................................................................ 18

EQUATION 3.1: ELECTRON BEAM DOSING DENSITY (GENERAL) ............................................................................. 76

EQUATION 3.2: ELECTRON BEAM DOSING DENSITY (VALUES) ................................................................................. 76

EQUATION 6.1: LENGTH OF MOIRE SUPERLATTICE ................................................................................................. 150

EQUATION 6.2: LATTICE RATIO OF 2D HETEROSTRUCTURES .................................................................................. 150

EQUATION 6.3: TWIST ANGLE OF 2D HETEROSTRUCTURES .................................................................................... 150

Page 17: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

17

Chapter 1: Introduction

1.1 Monolayer materials: the building blocks for 2D structures

2-Dimensional (2D) materials have demonstrated a range of impressive and unique capabilities

that have allowed them to impact a broad spectrum of fields, including transistors, DNA

sequencing, particle physics, radiation detection, superconductivity, neural regeneration,

spintronics, and body armor, just to name a few. Their diversity in properties is due not only to

the large number of monolayer materials grown and characterized, each with their own distinct

properties, but also the extreme number of 2D structures possible from combining the monolayer

materials, where each 2D structure can have properties vastly distinct from the parent crystals.

More specifically, 2D structures are not simply a sum of the parent crystals, but are a product of

their interaction as well. This incredible effect is surprising and highly unique, being confined to

nanomaterials and not observed in the 3D class. With this in mind, let us begin with the questions:

What makes 2D materials ‘2D’? And how come 2D structures are a product of the interaction,

vice simply a sum?

Figure 1.1: Artistic diagram of a general 2D material, where it is only one atom tall (i.e. in the z-axis), but is many more

in the x- and y-axes. This confines the electrons in the material to a 2D plane, thereby significantly altering the

properties.

The field of 2D materials is relatively new having only started in 2004 with the report of graphene,

where multiple definitions of what makes a material “2-Dimensional” have arisen. However, at

its core, a 2D material is ‘2D’ because the electrons have been confined to a plane (i.e. confined to

two axes), and are thereby significantly restricted from traveling along a 3rd axis. More

specifically, 2D materials are frequently only one to a few atoms high in the z-axis, but are

thousands of atoms long in the x- and y-axes (see Figure 1.1), which confines the electrons through

Page 18: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

18

boundary conditions. This may seem unusual at first glance because in the world of classical

physics, the world we mostly experience every day, materials are not expected to change

properties simply because they are made atomically thin. So why do electrons care that they are

confined, and why does changing the dimensionality significantly alter the properties? The

answer lies within quantum mechanics and the fact that electrons have wave-nature, where their

stable state needs to satisfy the Schrödinger equation (Eq. 1). Simply said, in the quantum

mechanical world dimensionality and relative size are critical in determining properties. Below

this effect is discussed in more detail.

−ħ2

2𝑚𝛻2𝜓 + 𝑉(𝑟)𝜓 = 𝐸𝜓 Equation 1.1

The properties of a material are determined not just by the type of atoms present and

stoichiometry, but also by the crystal structure. For example, both graphite and diamond are

constructed only of carbon atoms, yet their properties could not be more distinct in many ways.

Graphite is commonly found as the “lead” in pencils, where it is black, flakes easily, is breakable,

and leaves a residue when wiped on paper. Diamonds, on the other hand, are clear, diffract

colors, extremely hard, and will often scratch other materials before breaking themselves.

Further, graphite is conductive, while diamonds are good insulators. The only different between

graphite and diamond is their crystal structure (Figure 1.2). Why is it that the crystal structure

can have such a dramatic impact on the properties? The answer lies in how the electrons perceive

the material they exist within, and then how they interact with it. By changing the crystal

structure, they will perceive the material different and respond differently.

Page 19: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

19

Figure 1.2: (a) Graphite and (b) diamond crystal structures. Reproduced from Universität Kiel.1

As stated earlier, electrons must satisfy the Schrödinger equation while they reside inside a

material. The potential term (i.e. V(r) shown in Eq. 1.1) represents objects that impart potential

energy, as well as their relative positions in space, which includes the nuclei of the atoms. Hence,

as the crystal structure changes, so does the solution to the Schrödinger equation and the electron

response, and thereby the material’s properties. More specifically, solutions to the Schrödinger

equation determine what energies (or momentums) the electrons are allowed to have. These

allowed and disallowed states are often visualized using band structure diagrams, which provide

a tremendous amount of information and can be used to predict exotic effects, such as super

conductivity, 2D electron gases, Rashba splitting, and topological insulators. Along these lines,

one important consequence of crystal structure is whether a band gap exists, and, if so, how large

it is. The band gap of a material is critical in determining the electronic and optical properties,

along with other properties. For example, graphite has a very small band gap while diamond has

a very large one, thereby bestowing conducting and insulating properties, respectively. The band

gap results because electrons are forbidden from existing at certain energy (or momentum)

ranges. It is beyond the scope of this dissertation to explore this extensive topic, which can be

found in numerous condensed matter textbooks.2 For the readers of this dissertation, the primary

takeaway is that the crystal structure is critical in determine the material’s band structure, and

thereby their properties.

Page 20: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

20

2D materials are so ubiquitous and have had such a tremendous impact on a diverse number of

fields that it is easy to forget how surprising their history is. Although numerous properties and

layer-dependent effects were surprising, one of the most surprising findings was that 2D

materials could even exist without highly specialized environments. Up until the report of

graphene in 2004, the scientific community believed that 2D materials were too unstable to exist

in ambient.3 In fact, the first report of graphene was considered so controversial and unlikely that

the manuscript was rejected twice before being accepted. Once published, the findings were

repeated by numerous groups across the globe, while expanding upon the fundamental

understanding and demonstrating applications. As a testament to the impact of graphene and the

initial research, Dr. Andre Geim and Dr. Konstantin Novoselov were awarded the 2010 Nobel

Prize in Physics, the second fastest Nobel Prize ever awarded!3

1.2 2D Materials: overview of specific monolayer materials

Up until now the dissertation has not discussed any specific materials. This section will discuss

some of the more prominent and influential 2D materials, while touching on the diversity of

properties demonstrated. This is important for discussing 2D structures later in the paper, and

understanding the influence the interlayer coupling has on the properties.

Figure 1.3: Artistic diagrams of (a) graphene and (b) a monolayer transition metal dichalcogenide (TMD). Both

materials have a hexagonal crystal structure. While graphene is only one atom tall, monolayer TMDs are three atoms

tall, where chalcogen atoms are above and below a transition metal atom. The hexagonal structure is best shown by

looking down the z-axis.

Page 21: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

21

Graphene is one layer of graphite (i.e. one layer of carbon atoms arranged in a hexagonal lattice –

see Figure 1.3). It has demonstrated the highest tensile strength recorded,4 one of the lowest room-

temperature resistivity ever recorded,5 extremely high thermal conductivity,6 Dirac-Fermions,7

and is nearly transparent.8 As far as applications, graphene is being used as optical detectors,9

biosensensors,10 massless electrodes for resonators,11 supports in transmission electron

microscopy,12 detection of explosives,13 transparent electrodes,14 flexible and wearable

electronics,15 energy storage,16 low-cost solar cells,17 single-ion detection,18 DNA sequencing,19

platform to tune the band gap through chemical doping,20 and catalysis.21 It is the most well-

known and studied of the 2D materials, and justified enormous funding initiatives, including the

European Union’s one billion euro flagship project.22 Multilayer graphene is discussed in the next

section. The opportunities and next-gen technologies possible using graphene are substantial,

where very bold and creative ventures that bridge disciplines are not as unlikely as they used to

be. For example, this dissertation posutlates that the exceptionally high tensile strength combined

with the nano-pore capabilities demonstrated by graphene, suggest that it is a good candid for

micron-sized jet injection technology, enabling localized drug deliver at micron scales.23 Neurons

with their long dendrites would be good candidates for ultra-localized drug delivery. This

dissertation also did work with graphene/MoS2 2D heterostructures, where such samples were

grown and characterized. This is discussed in Chapter 9.1.

Monolayer transition metal dichalcogenides (TMDs) are three layers of atoms arranged in a

hexagonal structure (Figure 1.3), where each layer is approximately 0.7nm in height.24 As of

writing this dissertation, dozens of monolayer TMDs have been demonstrated, each with their

own properties.25,26 Some of the properties, such as photoluminescence (PL) and being a

semiconductor, are observed in numerous materials and are considered characteristic of the

class,27–33 while other properties, such as charge density waves in NbSe2, are unique to only a few

materials.34–37 The family of monolayer TMDs is by far the largest of the monolayer and 2D

material families, and has shown promise for numerous applications across a spectrum of fields.38

This includes nano lasers,39 chemical, optical and bio sensors,31,40–43 transistors,44

Page 22: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

22

superconductivity,45 piezoelectric energy conversion,46 exciton-plasmon coupling for optical

switches and modulators,47 gating-induced Ising superconductivity,48 and nano-pores for

sequencing DNA,49 facilitating the study of signaling pathways and complex proteins such as

Stat5.50 MoS2 is discussed in the next section. This dissertation grew several MoS2-based 2D

heterostructures, which are discussed in different chapters.

Figure 1.4: Experimental and theory band structure for the topological insulator Bi2Se3. High-momentum-resolution

angle-resolved photoemission spectroscopy (ARPES) measurements along the (a) Γ-M and (b) Γ-K momentum-space

cuts. (c) The momentum distribution curves corresponding to (a). This work shows a representative band structure

from a topological insulator, including the Dirac cones. This work was reproduced with permission from 51.

Monolayer bismuth selenide (Bi2Se3) has five layers of atoms arranged in a hexagonal

rhombohedral structure, where each layer is approximately 1nm in height. A majority of this

dissertation discusses Bi2Se3 in monolayer and few layer form, although in Chapter 9.1 it also

discuss research concerning two similar compounds: bismuth telluride (Bi2Te3) and antimony

selenide (Sb2Se3). Few published works have discussed Bi2Se3 in its monolayer morphology, and

if they have, it likely involved either a support structure to stabilize it, or the Bi2Se3 samples were

nano particles.52,53 This dissertation postulates that there are two reasons that contribute to this.

The first being that Bi2Se3 is likely not very stable or unstable in its monolayer morphology in

ambient, where it requires either a surface to stabilize it or it needs to be suspended in a liquid.

Further, Bi2Se3 is well-known to be highly sensitive to air and oxygen at ambient, suggesting this

Page 23: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

23

factor exasperates and further destabilizes the compound.54–60 The second reason monolayer

Bi2Se3 likely is not researched is because the well-known topological insulating properties are not

present in the one to few layer morphology (see Figure 1.4 for more details).61 The topological

insulating properties only appear once a sufficient number of layers have been stacked to induce

band inversion, a requirement of topological insulating properties.51 Hence, without the scientific

allure to grow monolayers, which have been theoretically and experimentally shown to not have

topological insulating properties, it can be expected that such samples are less likely to be

fabricated. Similar to Bi2Se3, Bi2Te3 has been demonstrated to be a topological insulator and highly

sensitive to oxygen at ambient.51,62,63 In contrast, Sb2Se3 has a different crystallographic structure

– orthorhombic and not hexagonal - and is predicted to not have band inversion or topological

insulating properties.61 Bi2Te3 and Sb2Se3 are discussed more in Chapter 9.1. This dissertation grew

numerous Bi2Se3/TMD 2D heterostructures, which are the primary focus of the work. These are

discussed in more detail throughout.

1.3 2D Structures: combining monolayer crystals to build something new

Up until now we have only discussed monolayer materials, also known as monolayer crystals,

where a few notable examples were described (e.g. graphene, monolayer TMDs), along with a

general overview of 2D material physics. The subtle importance of the crystal structure was

explained. In summary, it determines the relative positioning of the potentials, which affect the

solutions to the Schrödinger equation (Equation 1.1) and allowed electron energies. This

contributed to the fundamental question: Why does loss of a dimension so dramatically affect the

material’s properties? The dissertation will next discuss 2D structures, materials created by

combining 2D materials. In particular, the dissertation will focus on vertically stacked 2D

structures (see Figure 1.5 for an example of a 2D structure).

Page 24: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

24

Figure 1.5: Artistic diagram showing a representative 2D structure. In this example, two different monolayer materials

(each one atom tall) were vertically stacked. The parent crystals so not need to be different for the properties to be

dramatically altered (e.g. the bright PL of monolayer MoS2 is quenched when a second MoS2 layer is placed on top).

Additionally, 2D structures can be constructed using more than two compounds, suggesting a massive number of

permutations for creating materials with application tailored properties.

Classical physics predicts that as materials are combined via vertical stacking, the new compound

will be close to a simple-summation of the properties of the parent materials, something that is

confirmed by the reader’s intuition. For example, when the body of a car is painted, the metal

does not become weaker or change its properties to a noticeable degree. The combined product

of the painted car is now a summation of both materials (e.g. it has the strength of the metal and

the shiny color of the paint). Now, why are we discussing such an example? The answer captures

one of the biggest and most important messages of this dissertation:

Vertical 2D Structures are more than a simple sum of the parent materials, but are a

product of the interlayer coupling. This effect is distinctly quantum mechanical in

nature, cannot be explained by classical physics, and is a surprise to classical intuition.

It should be noted that when we discuss “vertical stacking”, mixing or the creation of alloys is

not included in the definition. 2D Structures have distinct layers where the crystal structure and

chemistry are being maintained - atoms are not being exchanged between the layers to create

alloys, and their structure is not evolving to a new phase. Additionally, it is documented that the

interface between two 3D materials placed in contact can be changed due to quantum mechanical

effects; however, in these cases only the interface, and the not the entire 3D material itself, is being

altered. In the case of 2D structures, the properties of the individual layers can be changed and

Page 25: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

25

dramatically altered from their original state, due to the interlayer coupling. Before discussing

the impact and physics of the interlayer coupling, this dissertation will review the applications

and capabilities of 2D structures. The enormous diversity and remarkability of the effects

demonstrated in previous work will serve as foundation and guide for understanding the impact

and influence of the interlayer interaction. Below a variety of examples of 2D structures with

accompanying figures, where the information is a collection from multiple published works and

serves as a short review of the 2D structure work thus far.

Figure 1.6: Graphene band structure, and demonstration of the evolution of the band structure as graphene layers are

stacked. (a) Graphene band structure with an inset showing the Dirac cone. (b) Same graphene Dirac cone band

structure, but plotted differently. Note how the slope is approximately linear, suggesting that the electrons in graphene

have a linear energy-momentum relationship, an indication that Dirac fermions exist within the material. (c) Band

structure from bilayer graphene with an AB stacking. Bilayer graphene is two layers vertically stacked on top of each

other. Note, very interestingly, how the band structure is evolving from the monolayer morphology. The Dirac cone is

no longer sharp at the interface, but more rounded, due to the interaction of the two layers. If the layers were

electronically independent and behaved monolayer, they would each have the band structure in (a) and (b). The plots

in (b) and (c) were reprinted with permission from 64. (d) Band structure from bilayer graphene stacked in the AB

configuration, and (c) band structure of bilayer graphene stacked in the AA configuration. Very interestingly, simply

changing their stacking – how the layers are interacting – dramatically changes the band structure and opens up a

sizeable band gap. These plots were reprinted with permission from 65. Together this figure demonstrates how

significantly the band structure, and thereby properties, change when 2D materials are stacked.

Graphene is the most widely researched and well-known 2D material. This is not only due to its

incredible properties (discussed in the previous section), but also in part because it was the first

2D material identified, where it was first reported in 2004. After studying graphene in its

monolayer form, a natural next step was to start stacking it. After all, the properties of graphene

and graphite are drastically different, so the following questions naturally arise: Why are their

properties so different? Where do the transitions occur between monolayer and bulk? And is

Page 26: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

26

there a natural “linear” progression in properties from monolayer layer to bulk? Or are there

intermittent configurations with different properties from either monolayer or bulk?

This dissertation attempted to answer all those questions thoroughly; however, short answers are

as follows: (1) The properties are different due to the interlayer interaction. (2) It is not always

clear when the transitions between monolayer and bulk occur, and the transition is different for

each material, but usually it is around ten layers that the properties are close to bulk than

monolayer. (3) and (4) Very interestingly, the transition from monolayer to bulk is not linear. New

properties unique to either bulk or monolayer emerge as the layer count is changed and as the

layers are twisted relative to each other. The most prominent example is how bilayer graphene

twisted to magic-angles will be superconducting.66,67

As shown in Figure 1.6, the band structure of graphene is altered as the layer count increases.

This modifies the properties, both inducing new capabilities while removing others. Figure 1.6a

shows the band structure for graphene, where the well-known Dirac cones are visible. The fact

that six Dirac cones are shown is due to the six-fold symmetry of the crystal structure. The inset

shows a close-up of a Dirac cone, where the linear behavior of the sides is better shown. This

linear structure is very important and unique to graphene because it leads to the formation of

Dirac fermions in graphene, particles with a linear dependence on their energy, an interesting

phenotype that also photons have. Figure 1.6b-c show the band structures of monolayer and

bilayer graphene. The linear behavior and Dirac cone is clearly visible in Figure 1.6b, while Figure

1.6c has lost the sharp structure of the cone with more rounded ends.64 Further, it appears as

though a band gap is opening up. It is important to note that Figure 1.6c specifies AB stacking

because the stacking is critical in determining the band structure, as shown in Figure 1.6d-e,

where both AB and AA are shown. AA stacking has been shown to open up a band gap, while

AB retains more of the monolayer structure.65 Together this data demonstrates that stacking

graphene changes the band structure significantly, underlining the importance of the interlayer

interaction. Further, how they are stacked determines the band structure.

As a result of the interlayer interaction and layer-dependent properties, bilayer and multilayer

graphene have demonstrated the following applications: high conductivity,68 large tensile

Page 27: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

27

strength and flexibility,69 tunable infrared properties for low-cost high-performance detectors,70

platform to study phono-exciton coupled particles,70 high thermal conductivity,71 high photon

transparency,72 the impermeability to gas species,73 sensitivity to tune the properties through

chemical functionalization,74 transparent electrodes,14 terahertz electronics,75,76 thermoelectric

power harvesting,77 tunable plasmonic devices,78 high sensitivity energy detector (i.e. a

bolometer),79 battery applications with superdense lithium storage80,81, ultrafast lithium diffusion

into and out of the structure,82 reinforcing nanocomposites,83 humidity sensors,84 platform to

study van Hove singularities,85 system to probe Diract fermions and Hofstadter butterflies,86,87

platform to study superconductivity,66,67 topological insulating and transport properties,88

coherent commensurate electronic states,89 biosensors,90 and low cost solar cells,17 The following

reference reviews contain more information.91,92

Figure 1.7: Demonstration how the PL is modified by simply increasing the layer count of MoS2 , a conclusion that can

be applied to the other monolayer materials. (a)-(d) Very interestingly, as the layer count is increased, the band

structure undergoes major modifications. For example, in the case of MoS2, as well as many other TMDs, the band

structure shifts from being (a) direct band gap in the monolayer morphology to an indirect band gap as the layer count

increases – see (b)-(d). The data is reproduced with permission from 93. (e) Demonstrates this effect best using PL

mapping, where the bilayer 2D structure has significantly weaker PL. Data is reproduced with permission from 94. (f)

Shows representative PL spectra from monolayer and bilayer MoS2, including a PL with the corresponding optical

image. The large change in PL magnitude is clearly seen, a reflection of the changing band structure due to the second

layer. This work is reproduced with permission from 95. (g) Artistic diagram of monolayer MoS2 with a neutral exciton.

Page 28: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

28

Charged excitons have also been reported within the monolayer TMDs. It is believed that the excitons are very tightly

bound in large part because the electric field lines are able to extend out of the material, suggesting the dielectric

environment surrounding the monolayer TMDs affects the exciton dynamics.

One of the most striking and notorious examples of a change in properties due to layer count is

the quenching of the PL from monolayer to bilayer in certain TMDs. Certain TMDs are direct

band gap and contain a very bright PL. This includes MoS2, MoSe3, WS2 and their alloys, materials

that will be discussed in depth later in the dissertation. The bright PL is due to very tightly bound

excitons that exist within the material.30,33 These excitons have some of the highest binding

energies ever recorded, due to the monolayer nature.96,97 More specifically, as the electron-hole

pair is formed, the electric field lines will reach out of the material (see Figure 1.7g). Very

interestingly, as the layer count in increased, the band structure will begin to evolve. Numerous

changes are made; however, one of the most important ones is that the band structure will become

indirect band gap, as shown in Figure 1.7a-d. A consequence of the band structure evolution is

that the PL will quench. This is best shown in the PL mapping in Figure 1.7e-f, where the bilayer

regions have a significantly reduced PL intensity. Other changes to the PL result from the change

in layer count, including the peak position shifts and the relative ratio of the excitons changes;95

however, these changes are beyond the scope of this dissertation.

As a result of the interlayer interaction and layer-dependent properties, bilayer and multilayer

TMDs-based 2D structures have demonstrated the following applications: photovoltaic energy

harvesting,98 photoelectrochemical solar energy conversion,99,100 coupling magnified spin-orbit

effect,101 photodetectors,102 ultra high-density information storage,103 advanced oxygen sensors,104

advanced p-n diodes,105 platform to study plasmonic and phono-exciton physics,106 lithium

intercalation and storage for batteries,107 sensitive photon calorimeters,104 high-performance

spintronics,108 DNA sequencing,109 tunable Schottky device contacts,110 gate-tunable

ferromagnetism for memory and spin manipulation,111 biosensors,42 tunable magnetism and

spin,112 ultrahigh-q room temperature spin filters,113 ultrahigh “one million percent”

magnetoresistance,114 platform to study magnetic Skyrmions,115 and energy efficient artificial

synapsis.116

Page 29: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

29

Figure 1.8: Angle resolves photoemission spectroscopy (ARPES) of Bi2Se3 at different layer numbers along the Γ-K

direction. As the layer count is increased, the band gap closes and band inversion appears that allows for topologically

protected states. This experiment agrees strongly with theory and demonstrates the impact of layer count. The band

inversion and topologically protected states are not present in bulk, demonstrating that this effect is only present in

samples with the correct layer count. This also further demonstrates that the evolution of properties from monolayer

to bulk is not “linear”, where a progression from one type to another is observed, but that unique states and properties

exist in between. The following figure was reprinted with permission from 117.

Figure 1.8 shows angle resolved photoemission data (ARPES) from several different Bi2Se3

samples with different layer counts. The samples were grown using molecular beam epitaxy

(MBE), suggesting that they have low doping and defects. It can be seen that as the layer count

increases, the band gap diminishes and band inversion starts to appear, suggesting topological

states. These experimental results are in strong agreement with DFT theory calculations and

theory.51,118 This band inversion and the topological insulating properties vanish as the layer count

increases, and is only present when the correct number of layers is present. Interestingly, in the

samples with lower layer counts, where a band gap is present, sizeable Rashba-type spin–orbit

splitting is observed. This is likely due to the substrate-induced potential difference between the

two surfaces,117 further demonstrating the powerful impact of the interlayer coupling, and its

ability to influence not just the neighboring atoms, but also atoms several angstroms away.

Page 30: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

30

Figure 1.9: Artistic diagram illustrating the possibilities of designing vertically stacked 2D structures , and how the

careful control of not just the parent materials, but also the stacking order, stacking configuration, and twist angle could

lead to the engineering of materials with application tailored properties. The figure was reprinted with permission

from 38.

As shown above, 2D materials in their monolayer morphology have demonstrated incredible

properties. When these same materials are then stacked to from multilayer structures and

heterostructures, the properties and capabilities change. Combining 2D materials does not result

in a simple summation of their properties. It was shown that the research community has used

the 2D structures to both study fundamental science and design application-focused materials

and devices. The extensive volume of published results and patents is a testament to the impact

of the field and potential of this technology to solve problems. Figure 1.9 is an artistic diagram of

how the knowledge and research can be used to construct technology. By vertically stacking 2D

materials in different configurations, custom materials can be constructed. The extreme number

of permutations possible hints at the opportunities, and is a key driver and justification for further

exploration and research.

1.4 2D Structures: a product of the interlayer interaction

Page 31: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

31

The interlayer interaction is critical to understanding the properties of 2D structures. It is a very

broad term that attempts to capture the physical changes that lead to changes in properties. At a

fundamental level, the interlayer interaction describes the type and magnitude of

“communication” between the different layers. The more communication there is, the more

opportunity and probability that new properties will manifest, while, on the other hand, if there

was no communication, the materials would act monolayer and independent. With this

framework in mind we are presented with the following questions: What does “communication”

mean? What are the different types of communication”? What factors determine whether

materials are communicating? How come this effect is not observed when 3D materials are made

to communicate with one another?

The extraordinary impact of the interlayer interaction and coupling is best observed using

vertically stacked 2D materials. Although the study of interlayer physics is decade’s old and

delivered tremendous findings, the measurements were confined to studying the interlayer

region, and not the material as a whole. One of the most well-known examples of this is a highly

cited work that studied the metal-insulator transition in the interface in a silicon-based system.119

In this work bulk silicon pieces were placed in contact, and a two dimensional electron gas formed

at the interface, suggesting the materials were communicating and interacting. Despite this strong

interaction, the properties of the silicon pieces themselves were only negligibly impacted,

demonstrating that 3D materials are not significantly affected by inter-material interactions. This

effect is in agreement with our intuition, and something we observe frequently in the everyday,

classical world. For example, when a book is placed on a desk, neither the book nor the desk

change their properties. We also used the example of a painted car earlier, where the paint does

not affect the strength of the metal, allowing the combination to yield strong vehicles with a shiny

look. Now that we have established that the coupling formed between 3D materials generally

does not affect the properties of the whole material, the question is: Why does it affect the

properties of 2D materials? The answer is due to the fact that the 2D material is exceptionally

thin, so changes in the environment as able to propagate through the entire material, and that the

Page 32: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

32

interlayer separation is on the same order as the 2D material thickness, suggesting its influence is

on the same order as well.

Figure 1.10: Density functional theory (DFT) calculated Bi2Se3/MoSe2 2D heterostructure showing the side-view and z-

axis view on the left and right side, respectively. The calculated structure is one layer of MoSe2 on the bottom with one

layer of Bi2Se3 on the top, where the twist angle is 0°. Note the large interlayer separation, which is 24.9% of the total

height, indicating that it will influence the properties. For example, electrons traveling through the material will likely

interact with the massive volume of the interlayer region, and modify their behavior. To form the unit cell of the above

2D structure, four unit cells of Bi2Se3 matched closely with five unit cells of MoSe2, which is the structure used and can

be seen in the z-axis view on the right side. The formation energy density predicts that this 2D structure is significantly

more stable than the two layers individually, and a predictor that strong interlayer coupling is formed. The above

structure was calculated by students in Prof. Arun Bansil’s group and reproduced with their permission.

Figure 1.10 shows a representative 2D structure relaxed and calculated using density function

theory (DFT), suggesting the dimensions are approximately to scale and closely match

experiment. The 2D structure is one tri-layer of MoSe2 on the bottom with one quintuple-layer of

Bi2Se3 on top, where four unit cells of Bi2Se3 match with high precision five unit cells of MoSe2 (see

view along z-axis on the right side of Figure 1.10), and the twist angle was made to be 0°.

Throughout the dissertation the terms “tri-layer” and “quintuple-layer” may be shortened to just

the word “layer”, a common abbreviation in the community. The above structure was relaxed

and calculated using DFT by Dr. Gianina Buda and Dr. Christopher Lane from the Prof. Arun

Bansil group at Northeastern University. The group uses highly advanced theoretical techniques

such as strongly constrained and appropriately normed (SCAN), where their work has been

Page 33: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

33

found to be in strong agreement with experiment for bulk and 2D materials.51,120–122 They were

and are important collaborators for the DFT work discussed in the dissertation. The DFT work

discussed in the dissertation was done by this group as a part of a collaboration. As shown in

Figure 1.10, the combined structure is predicted to be 13.52Å tall, where the interlayer region is

predicted to be a massive 24.9% of the height, more than the height of the MoSe2 layer!

Considering that the interlayer region does not contain any atoms, this is a massive distance. The

large percentage suggests that it would have a large influence on the properties. As will be shown

later in the disseration, the experimental and theoretical data suggests that charge pools develop

in the interlayer region that act as scattering sites. It is reasonable to assume that these potentials

will not only affect the solutions to the Schrödinger equation and band structure, but might have

their own band structure as well. With this in mind, the large volume of the interlayer region

indicates it will have a significant influence

As will be discussed later in the dissertation, DFT calculations completed by Prof. Bansil’s group

predict numerous other changes, including a change in lattice constants for the parent materials.

This is a result of the influence of the quantum mechanics, which induce a redistribution of charge

and electrons, thereby affecting atoms several angstroms from the interlayer region. The fact that

the interlayer region induces such a large influence angstroms away is in agreement with the

claim that the properties of the parent materials are changed by the interlayer coupling. The

formation energy density on the bottom predicts that the following structure is highly stable and

preferred over two individual layers. The formation energy also implies that there is a strong

coupling because the materials are interacting - and communicating - to produce the electron

bonding required to decrease the energy.

Figure 1.10 and the above discussion demonstrated that the interlayer separation is predicted to

occupy a significant volume of the 2D structure, suggesting it should have a sizeable influence.

Further, the fact that the 2D parent crystals change lattice parameters suggests that the interlayer

interaction affects more than the nearest neighbor atoms, but also affects the atoms several

angstroms away. The far-reaching effect of the interlayer interaction indicates that its influence

extends throughout the entire 2D parent materials, or at a minimum, most of it, offering an

Page 34: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

34

explanation why 2D parent materials change their properties, but not their bulk counterparts,

when coupled together. Next we will show that the interlayer interaction induces a charge

redistribution into the interlayer region into charge pools between nearest neighbors, suggesting

the formation of a charge-based lattice. Our research suggests that the formation of this charge

lattice is central to the manifestation of new properties.

Figure 1.11: DFT calculated 2D structure with one WS2 layer on the bottom and one Bi2Se3 layer on the top , where twist

angle is 19.1°. In contrast to the 2D structure shown in Figure 1.10, the construction of this supercell is more complicated

and not a simple ratio of 4:5 of Bi2Se3:TMD (see Chapter 6 for amplifying information). This is due to the twist angle

and moiré pattern formed, which requires two lattice vectors per parent material, vice two. The left part is a view along

the z-axis, while the right image is an off-axis view to show the interlayer charge redistribution. The left image shows

the spatially varying charge density, where red corresponds to higher densities. It can be seen from both images that

charge pools form between nearest neighbors.

Figure 1.11 shows a DFT calculated 2D structure with one layer of WS2 and one layer of Bi2Se3

twisted to the angle of 19.1°. Very interestingly, the interlayer interaction induces the formation

of charge pools in the interlayer region that exist between nearest neighbors. These charge pools

are due to significant charge redistribution from the parent crystals, suggesting that the electrons

have undergone a major migration. This migration is a physical change due to the interlayer

interaction, and suggests the materials form a strong interlayer coupling. Further, the charge

distribution suggests that bonding has formed between the layers, which facilitates electron

transport, electron communication, and lattice vibrational communication. As discussed later in

the dissertation, these charge pools are often strong enough to diffract high-energy electrons, a

Page 35: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

35

very surprising finding considering that it is, as far as we know, the first instance of select area

electron diffraction (SAED) being used to reveal charge redistribution.

The physical changes due to the interlayer interaction are not well-studied, but their existence is

supported by other works. For example, the Raman vibrational modes have been shown to be

dependent on the interlayer interaction, suggesting the charge redistribution affects the structure,

k-constant of the vibrational modes, and the phonons.123–126 Some of those works have even begun

to use Raman as a fingerprint or identifier to quantify the interlayer coupling strength and

characteristics. Other works have used scanning tunneling microscopy (STM) to study the

interlayer bonding and importance of nearest neighbors.127,128 Although an STM cannot probe the

charge redistribution, it is able to demonstrate a change in band gap and tunneling energy. This

in itself demonstrates that the layers are interacting and suggests charge redistribution, since a

change in properties requires a physical change at some level. Said another way, if there were no

physical changes induced in either parent material, the interlayer region would not have charge

redistribution, there would be no new potential sites in the interlayer region, and the parent

materials would be in indistinguishable from their monolayer morphology. Hence, the materials

would be effectively independent and the interlayer coupling would be effectively zero. The

concept of interlayer coupling strength and material independence – including electronic and

vibration independence - will be discussed next and throughout the dissertation, including

methods to control the interlayer coupling magnitude.

1.5 Methods to manipulate the interlayer coupling in 2D structures

This dissertation has so far established that 2D materials can have significantly different

properties than their bulk counterparts, and has established that vertically stacking and

combining monolayer materials produces 2D structures whose properties are not a simple sum

of the parent materials. It then showed that the interaction and communication between the

parent 2D materials is what induces the new properties and changes observed. Further, it was

shown that the interlayer region is a significant volume of the overall 2D structure, and that the

interlayer interaction is able to induced significant charge redistribution, as well as changes to the

Page 36: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

36

lattice constants of the parent materials. The changes induced allow the strength or magnitude of

the interlayer interaction can be quantified and discussed as a range of values, vice a binary choice

of “on/off”. Higher values lead to more physical changes and changes to the properties, while

lesser values lead to less change overall. Taken together, these findings suggest that if the

interlayer interaction could be manipulated, so could the properties, providing a mechanism and

parameter to tune, engineer, and build new materials. Next this dissertation will demonstrate

methods that have been used to manipulate and tune the interlayer interaction.

Figure 1.12: Tuning the interlayer interaction and coupling using the intercalation of molecules and atoms. (a) and (b)

are data from a recently published highly-viewed publication where bulk, 3D materials were intercalated with various

molecules, and they showed the interlayer coupling could be diminished. More specifically, when molecules

intercalated between the layers, the interlayer coupling diminished, allowing each layer to behave more independent

and exhibit properties similar to their monolayer morphology. This work demonstrated it using several layered, van

der Waals materials, including, MoS2, phosphorous, and graphite. (a) is pre intercalation and (b) is post intercalation.

This work was reprinted with permission from 129. (c)-(g) are from a paper where the interlayer coupling in bilayer

graphene was modified using intercalated lithium. (c)-(d) show the experiment how the lithium was intercalated using

artistic representation, where (c) is pre-lithiation and (d) is post lithiation. (e)-(g) are density functional theory (DFT)

calculated results for bilayer graphene intercalated with different amounts of lithium. It can be seen that as the lithium

intercalates, it modifies the charge distribution, thereby influencing the coupling and communication between the

Page 37: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

37

layers, as well as the properties of the individual graphene layers themselves. This work was reprinted with permission

from 81.

One of the most effective methods to tune the interlayer coupling is by intercalating atoms or

molecules. The intercalation of atoms and molecules has been studied for decades, and has found

applications in chemical storage, catalysis, energy storage, material doping, and ionic

transport.130–132 The strong interlayer interaction in 2D structures as provided intercalation

research with a new platform and system to study the effects. Further, the experiments can be

repeated between 2D and 3D systems to study changes in the interlayer interaction. One of the

most striking examples is the differences in lithium intercalation between bilayer graphene 2D

structures and bulk graphite. As shown in Figure 1.12c-g, bilayer graphene is able to store lithium

in super dense configurations that have not been demonstrated in bulk graphite.81,82 This is likely

due to the changes in the interlayer coupling. Additionally, once the lithium atoms have started

intercalating, the interlayer interaction and communication will begin to be disrupted. Hence,

this method is one way to tune the interlayer coupling and effects in bilayer graphene 2D

structures. A more global and universal method was published in the summer of 2018 showing

how bulk, 3D van der Waals layered and stacked materials are able to have the interlayer

coupling manipulated by intercalating and deintercalating molecules (see Figure 1.12a-b).129 The

molecules are often specific to certain compounds. For example, MoS2 can be intercalated with

quaternary ammonium molecules with variable carbon chain lengths, and phosphorous can be

intercalated with cetyl-trimethylammonium bromide, where in each case the interlayer coupling

is disrupted the bulk crystal’s layers behave more independent and exhibit properties of their

monolayer morphology.

One of the most studied topics has been the intercalation of oxygen between graphene and its

metal substrate, where it has been shown that oxygen decouples the materials and diminishes

their communication, inducing the graphene to behave more freestanding and independent.127,133–

142 Further, it was shown that the intercalated oxygen makes the graphene domains easier to

rotate, suggesting the bonding (potentially between nearest neighbors) has been diminished and

the interlayer coupling induced charge distribution has been reversed.134 The work also showed

Page 38: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

38

that the intercalation of oxygen and other materials is an edge limited effect, where the

intercalation is at boundaries and not through the material.82,135,139 As shown later in the

dissertation, we are able to intercalate oxygen between the Bi2Se3 and the monolayer TMD,

diminishing the interlayer coupling and inducing the materials to behave more independent.104

Previous work has used the intercalation of ionic compounds to modify the interlayer coupling

as well; however, the stability is not as high, and the ionic compounds tend to induce changes to

other properties, such as charge transfer or doping.107,143

Figure 1.13: Various methods that modify the interlayer coupling in 2D structures. (a) A tool to apply extremely high

pressures in the vertical direction was used to manipulate the interlayer separation distance, where it was found that

materials pressed together were more coupled. This fits intuition in that the atoms will be close, insinuating that they

will communicate more. This work was used to show that the magic-angle induced superconductivity in bilayer

graphene can be tuned. The work was reprinted with permission from 144. (b) This is a highly original work where an

atomic force microscope (AFM) was used to remove contaminants and press materials together, which improved the

coupling. This work was reprinted with permission from 145. (c) This is a similar setup to (a) where a high-pressure

setup is used to press materials together and decrease the interlayer separation. In this study a 2D structure of different

TMDs (i.e. MoS2 and WS2) was used, and it was found that not only did the photoluminescence change, but the Raman

response and vibrational modes changed as well. This is important because it shows that not only are the electronic

properties affected, but the lattice and phonon modes are affected as well. This work was reprinted with permission

from 146. (d) Annealing has been shown to improve and strengthen the interlayer coupling in 2D structures. The exact

mechanisms are not fully understood, but it is believed that possibly contaminants are desorbing, or the materials are

gently perturbed with heat into more stable configurations. For example , it has been shown that annealing graphene

on hBN surfaces causes the graphene domains to rotate into more stable moiré patterns.147 It can be seen that the

photoluminescence peaks are modified due to annealing sessions, and that the data suggests the interlayer coupling

was modified. This work was reprinted with permission from 148.

Page 39: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

39

Figure 1.13 demonstrates numerous other methods to modify the interlayer coupling using non-

intercalation methods. Both Figure 1.13a and Figure 1.13c demonstrate that decreasing the

interlayer separation strengthens the interlayer coupling, which fits intuition. This is because the

decrease distance allows for more atomic interaction and strong interlayer bonds.144,146 The results

suggest that the interlayer separation affects both the electronic coupling and the structural

coupling, where changes to the gating behavior and the Raman modes were observed,

respectively. Figure 1.13b shows how an atomic force microscope (AFM) can be used to remove

contaminants and press materials together, thereby strengthening the coupling.145 Figure 1.13d

demonstrates that annealing is able to strengthen the coupling in 2D structures. The exact

mechanism for why this happens is unclear, but it is believed either contaminants desorb, or the

layers are gently perturbed by the heat to twist and stretch/contract into more stable

configurations and moiré patterns.147–150 The fact that entire grains more than thousands of atoms

could rotate and twist due to gentle thermal annealing is an indication of the extreme influence

and forces the interlayer coupling has. A unique method to tune the interlayer coupling was using

strain, where the 2D structure would be strained in the lateral direction, which would place force

on the in-plane direction.151 Although this method shows signs of controllably modifying the

strain, it suffers the hindrance that the strain effects are present simultaneously, clouding the

experimental results because the interlayer coupling is no longer an independent parameter. The

previous two figures methods were shown that manipulate the interlayer coupling and

interaction strength. Next the dissertation will discuss how manipulating the twist angle can

induce large changes to the band structure. For the interlayer interaction to have an effect and

produce the desired properties, both the interlayer coupling magnitude must be large enough,

and the twist angle must be correctly set to induce the correct band structure. Said another way,

the interlayer coupling facilitates electron movement and communication across the interlayer

region, while the twist angle helps determine the moiré pattern and new band structure created.

Page 40: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

40

Figure 1.14: DFT calculated 2D structures with one layer WS2 and one layer Bi2Se3 , where the twist angle has been

altered. This figure demonstrates that the twist angle and resulting moiré pattern significantly influence the stability

and crystal structure. Later the dissertation shows that the twist angle and moiré pattern also have a significant

influence on the band structure and properties. Both a side-view and z-axis view are shown of the 2D structures. (a)

Twist angle of 0° was used, where three Bi2Se3 unit cells equaled four WS2 unit cells. The side-view shows the significant

interlayer separation distance, which is, very interestingly, 0.3Å larger than the height of the WS2. Both the WS2 and

Bi2Se3 lattice constants changed as a result of the interlayer interaction. The formation energy density demonstrates

that the combined 2D structure is more stable than the layers individually. (b) The same 2D structure except twisted to

the angle of 5.1°, where interlayer separation and WS2 height have increased, while the Bi2Se3 height has decreased.

Additionally, the formation energy density is significantly diminished, in agreement with experiment that 2D

structures grown at this angle are observed significantly less frequently, suggesting this twist angle is less stable than

0°. The above structure was calculated by students in Prof. Arun Bansil’s group and reproduced with their permission.

The twist angle of 2D structures has been demonstrated to be critical in determining its properties.

This is because the twist angle determines the moiré pattern and supercell created. Although this

topic is discussed in more detail throughout the dissertation, a short explanation is provided here.

When two periodic lattices or arrangements are overlaid – or effectively multiplied – they will

form a moiré pattern.152 Said another way, moiré patterns, or moiré supercells, are a new periodic

arrangement or ordering that is a manifestation of the interaction of the parent lattices or periodic

arrangements. This is clearly an interesting mathematical artifact, but the question then emerges:

How does a new larger periodic ordering affect the properties? The answer lies in the explanation

for what determines a material’s band structure.

As discussed earlier in this dissertation, electrons exhibit wave properties, and these waves

interact with the periodically spaced potentials of a crystal. The stable states are described by the

solutions to the Schrödinger equation (see Equation 1.1). If the potential term in the equation (i.e.

the V(r) term) is changed by the change in crystal structure and the addition of new atoms, then

the solutions and band structure will evolve, and the corresponding properties.153,154 The concept

Page 41: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

41

of the band structure radically changing in response to crystallographic changes was also

discussed earlier, where the example of graphite and diamond was compared, two materials

composed solely of carbon atoms, but arranged in different configurations. A similar effect takes

place when the moiré pattern and super cell is changed, which is effectively changing the periodic

unit cell the electrons interact with.12,103,155–159 Now that we have established the significant

influence of the twist angle and moiré pattern, the question becomes how this parameter can be

manipulated to control the properties? In the next section, methods to manipulate the twist angle

are discussed. Additionally, Chapter 2 of the dissertation discusses a method using a transmission

electron microscope (TEM) to manipulate the twist angle and probe the interlayer interaction.

Figure 1.15: Two powerful methods to construct and manipulate 2D structures are to (a)-(d) mechanically transfer 2D

materials on top of each other, and (e)-(h) rotate 2D materials using an atomic force microscope (AFM). (a)-(d)

Demonstration of magic-angle induced superconductivity in bilayer graphene. When the 2D structure bilayer graphene

is twisted to magic-angles, a moiré pattern and supercell is formed that has band flattening and superconductivity. The

quality of the 2D structure to measure such sensitive effects, suggests that mechanically transferring 2D structures

produces a sufficiently strong interlayer coupling. (a) The moiré pattern formed at the magic-angle of 1.08° for a bilayer

graphene 2D structure. This is reprinted with permission from 67. (b) Artistic diagram of a device of a bilayer graphene

2D structure twisted to a magic angle and overlaid on a hexagonal boron nitride (hBN) surface. The hBN surface is

extremely flat and non-interacting with graphene structures, minimizing the effects of defects, phonons, and other

scattering sites, allowing the properties of the graphene structures to be probed. (c) Diagrams showing the temperature

vs. carrier density on the conductance. It can be seen that simply by gating the devices, the 2D structures can be shifted

between the superconducting regions and the Mott insulating region. The two different plots are for two different twist

angles and moiré patterns. (d) DFT calculated band structure of a bilayer graphene structure where only the Dirac

point is shown. The following plot demonstrates band flattening, a consequence of not only sufficiently strong

interlayer coupling, but also the correct twist angle. (b)-(d) are reprinted with permission from 66. (e) Demonstration

for how an AFM can be used to controllably manipulate the twist angle in 2D structures. Two moiré patterns are shown,

demonstrating the changing angle-dependent supercell. (f)-(h) AFM images of an hBN/graphene structure where the

upper hBN material has been controllably rotated to numerous angles. This work was reprinted with permission from 160.

Page 42: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

42

Figure 1.15 shows two methods to produce 2D structures with user-defined twist angles. The first

method is to mechanically transfer one layer on top of another layer. This is the most popular

method with numerous publications expanding on the technique.161–164 It has shown consistently

to produce clean interfaces, and advanced methods even allow high accuracy alignment.164 A

weakness of the method is that it is very difficult to scale, suggesting it is not commercially viable.

Although recent work has used advanced robotics and machine learning to scale the mechanical

transfer process for the production of 2D structures.163 Figure 1.15e-h demonstrates a very new

technique with high rotational accuracy and user-defined angles with in-situ control. It uses an

AFM to controllably rotate the top material. Previous has also shown that 2D structures can be

fabricated by growing one material on top of the other.30,103,165,166 When one layer is grown on top

of the other, the resulting 2D structures often have a strong interlayer coupling, while maintaining

chemical purity, but twist angle control is difficult to achieve. This limitation is discussed further

Chapter 3 and Chapter 4 about Bi2Se3/TMD and graphene/MoS2 2D heterostructures, respectively.

In summary, 2D structures are more than a sum of the parent materials, but are also a product of

the interlayer interaction, which can induce new properties. Not only is the communication and

exchange of electrons between the parent materials important, but so is the twist angle, which

determines the moiré pattern and moiré supercell. This interlayer coupling can be controlled to

tailor properties and engineer application-specific materials.

Page 43: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

43

Chapter 2: Instrumentation and Synthesis

This chapter discusses the methods, techniques, equipment, and instrumentation required to

complete the experiments and measurements discussed throughout the dissertation. The below

descriptions are a summary of the technique or equipment, and not intended to replace the user

manual or be a thorough guide. Additionally, the descriptions are biased toward the applicability

and interest of 2D structures (all the techniques and equipment were primarily applied to 2D

materials), and caution should be taken before generalizing the expected results to other materials

or systems.

2.1 Growing 2D materials using Chemical Vapor Deposition (CVD)

Growth of TMD and TMD-alloy 2D crystals

All 2D TMDs, as well as 2D TMD-alloys, discussed in this dissertation were grown using vapor-

phase chalcogenization (VPC), a subset technique apart of the chemical vapor deposition (CVD)

family.31 The growth setup consisted of quartz tubes [1 inch (2.54 cm) in diameter] in a horizontal

tube furnace (Lindberg/Blue M). A quartz boat, containing a thin layer of either MoO2 or WO2

powder (3 mg or 15mg, respectively) with SiO2/Si (MTI Corporation) substrates suspended over

the powder with the growth side facing down, was placed in the hot center of the furnace. Either

Sulfur, selenium, or both powders (150 mg or 50mg, respectively) were used, depending on the

material being grown. To grow the alloy MoSe2-2xS2x, both powders were used at the same time.

Sulfur was placed near the insulating edge of the furnace upstream, whereas selenium was placed

~1cm past the insulating edge of the furnace upstream. The setup was pumped down and purged

with argon gas before it was filled with an Ar atmosphere. Downstream was then opened to

atmosphere, in addition to a constant 200 standard cubic centimeter per minute (SCCM) Ar flow.

The furnace was heated to different temperatures and at variable rates, depending on the material

being grown. The growth was conducted in two stages, 1st-stage and 2nd-stage, where 2nd stage

Page 44: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

44

would start once the 1st-stage temperature was reached. See the below table for material specific

growth information. After the elapsed time, the furnace was opened and allowed to cool rapidly.

Table 2.1 Monolayer TMD growth parameters using vapor-phase chalcogenization (VPC).

2D Crystal 1st Rate (°C/min.) 1st Temp. (°C) 2nd Rate (°C/min.) 2nd Temp. (°C) Hold time (min.)

MoS2 50 500 5 712 20

MoSe2 50 200 20 740 10

WS2 50 1080 5 1150 25

MoSe2(1-x)S2x 50 100 30 750 10

Bi2Se3 growth was performed in a nearly identical vapor-phase deposition setup, except a heating

wrap was coiled around the quartz tube at the down-stream end, leaving no gap between the

furnace and the heating wrap. The Bi2Se3 powder (50 mg) was placed in the hot center of the

furnace. The 2D TMD or 2D TMD-alloy substrate was placed downstream ~0.75cm from the

boundary between the furnace and the heating wrap. The system was pumped down to a base

pressure of ~10 mtorr before a 35-SCCM Ar flow was introduced, raising the growth pressure to

~490 mtorr. The heating wrap was set to a temperature of 245°C, and a temperature controller (J-

KEM Scientific Model Apollo) ensured it remained within ±2°C. The furnace was heated at a rate

of 50°C/min to 530°C and then held there for 20-25 min depending on the desired thickness. Once

growth was completed, the furnace was opened and the temperature controller was de-

energized, allowing the setup to cool rapidly.

2.2 Mechanically transferring 2D materials to create 2D structures

2D Structures can be constructed using numerous techniques, but one of the most popular

techniques is mechanically transferring 2D materials on top of each other. Although the method

often requires significant time and skill, and is unlikely to ever be commercially attractive, its

most attractive quality is that it can be universally applied to the construction of any 2D structure,

and is likely to produce the 2D structure with sufficient effort. This is in contrast to techniques

that fabricate 2D structures by growing 2D materials on top of each other, where the 2D structure

Page 45: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

45

only forms if it is the energetically preferred state. Such methods that rely on dynamics of the

natural world usually require a higher initial time and material investment, and sometimes are

only possible inside synthetic environments (e.g. unusual temperatures or pressures). Despite

these drawbacks, if 2D structures can be grown, vice mechanically constructed, they are

significantly more like to be economically viable.

Most mechanical transfer techniques use a specific material to lift a monolayer material off the

substrate it was grown on, and then transfer it onto a new substrate. In this dissertation, all

transfers used Polymethyl methacrylate (PMMA) to lift-off and transfer the 2D materials. After

the PMMA was placed onto the new substrate, is was dissolved using acetone, leaving the 2D

materials behind. Amplifying information with sufficient detail to enable reproduction of the

technique can be found in the methods section of the following reference.167 A weakness of this

method is that dissolving the PMMA often leaves significant residue. An example of this can be

seen in Figure 5.5, where the visually-non-linear spots are possibly PMMA or acetone remaining.

Folding and tearing can also be seen in various TEM images throughout the dissertation,

suggesting that this method can damage the 2D heterostructures. An improved method that has

even shown promise to be commercially viable is the dry transfer method, which uses

viscoeleastic compounds to both pick-up and place the 2D material down, removing the need to

solvents, thereby keeping the 2D materials clean and intact.161 Viscoelastic compounds are critical

for this method, as they behave like a solid over short time frames, but a liquid over long time

frames, thereby breaking the adhesion symmetry and allowing the same compound to both pick-

up and drop-off.

2.3 Optical setup for Raman and photoluminescence spectroscopy

Instrumentation

Raman and PL spectra were measured using a Renishaw Raman microscope equipped with a

488nm laser and a grating of 1800 lines/mm. A ×100 or ×150 objective focused the laser to

diffraction-limited spot size. TEM images and SAED patterns were collected from a JEOL 2010F

Page 46: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

46

operated at 200 kV. AFM images were taken from a NanoMagnetics Instruments Ambient AFM.

All Raman, PL, AFM, and UV-Vis experiments were performed under ambient condition.

2.4 Manipulating interlayer coupling in Bi2Se3/TMD 2D heterostructures using

laser exposure and thermal annealing

Laser Exposure Experiments

The 2D heterostructures were exposed to a 488nm continuous-wave laser at ambient temperature,

pressure, and atmosphere, and the atmosphere pressure was sea level. Previous work has shown

that the response of the 2D heterostructures to laser exposure is dependent on the partial pressure

of oxygen, suggesting environments at lower pressures might yield a different response.104 The

laser exposure recipe (i.e. power density and exposure time) determine the 2D heterostructure’s

response. The laser spot was approximately 1μm in diameter. Each type of 2D heterostructure

(i.e. Bi2Se3/MoS2, Bi2Se3/MoSe2, Bi2Se3/WS2, and Bi2Se3/MoSe2-2xS2x) responds differently to the same

recipe, where Bi2Se3/WS2 was the most sensitive and Bi2Se3/MoSe2 was the least sensitive.

Additionally, the Bi2Se3 layer count dictates the response, where a great number of layers requires

a more intense recipe to get the same response. The below table provides amplifying information

of the laser power required to induce approximately the same response across 2D

heterostructures, assuming the exposure is applied continuously without interruption.

Table 2.2: Lowest laser exposure powers for all Bi2Se3/TMD 2D heterostructures tested.

2D Heterostructure – 1 Layer Bi2Se3 grown on 1 layer of TMD Laser spot power

Bi2Se3/WS2 0.91μW

Bi2Se3/MoS2 60μW

Bi2Se3/MoSe2-2xS2x 70-90μW

Bi2Se3/MoSe2 80μW

The below table expands upon the parameters used for different experiments, both for exposure

and data acquisition, shown in the manuscript. All the experiments were conducted at ambient

Page 47: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

47

temperature (~22°C), pressure, and atmosphere. When “continuous” is listed it refers that the

data was collected while the exposure portion of the experiment was taking place.

Table 2.3: Detailed laser exposure parameters for experiments described in Chapter 4.

Experiment Experiment Power Experiment Time Data acquisition power Data acquisition time

Figure 5.6b 168μW 1 - 6 seconds N/A N/A

Figure 5.6e-f 0.91μW 30 seconds Continuous Continuous

Figure 5.13a-b 3.8µW - 1430µW 30 sec – 3 minutes 0.91μW 2 minutes

Figure 5.13c 50μW 1 minute Continuous Continuous

Figure 5.16 80μW 12 minutes 0.91μW 2 minutes

It should be noted that the as-grown samples, to the best of our knowledge, have clean interfaces.

Bi2Se3 grows highly crystalline across the whole TMD (i.e. micron sized domains), suggesting the

foreign atoms are not present to impede or modify the growth. Further, the laser exposure and

Raman results have been highly consistent across independently grown samples, suggesting the

interface is the same or very similar. Since it is unlikely that the samples would be equally “dirty”,

the likeliest explanation is that the interface is clean.

Thermal treatment experiments

All thermal treatment experiments were performed between 240-245°C for 3 hours. All

experiments used a flow rate of 3 SCCM, except the thermal treatment under air, where no flow

rate was used. The five environments were pure Ar, pure N2, N2+H2Ovapor, dry air (21% O2 and

79% N2), and air. All setups, except N2+H2Ovapor, were pumped down and filled with the

respective gas prior to thermal treatment. The N2+H2Ovapor environment was created by flowing

N2 at 3 SCCM, while several boats with deionized H2O were present in the tube. The heating of

the tube caused the H2O to evaporate. The downstream side “rained” significantly during the

entire thermal treatment process and H2O was still present in most of the boats, verifying that

sufficient H2Ovapor was present throughout the thermal treatment process.

Page 48: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

48

2.5 Atomic Force Microscope (AFM)

An atomic force microscope (AFM) was used in this dissertation to determine the height of the

2D materials being studied, and therefrom infer the number of layers present. This is critical

information, as the layer count determines the properties of the material. For example, as

discussed in Chapter 1, MoS2 in direct bandgap in monolayer morphology, but becomes indirect

bandgap as the layer count is increased. Previous work demonstrates that AFMs can be used for

not just a variety of other measurements relevant to 2D materials, but also applications, including

manipulating the twist angle of 2D structures,160 removing interlayer contaminants to strengthen

the coupling,145 and measuring the tensile break strength of a 2D material by pushing on a

suspended sample until it permanently deforms.4 A review of the AFM working principles, as

well as applications to 2D materials, can be found at the following reference.24

2.6 Scanning Electron Microscope (SEM)

The scanning electron microscope (SEM) is used in this dissertation to image 2D materials, write

patterns for electron-beam lithography, and manipulate the properties of Bi2Se3/TMD 2D

heterostructures. SEMs illuminate samples with a focused beam of electrons, and then measure

the electrons that emanate from the material. It is able to control the area illuminated by the

electron beam with exceptional precision using magnetic coils that steer the charged electrons.

Although the electrons interact with the sample in a variety mechanisms, only two types of

mechanisms are primarily used: backscattered electrons, and secondary electrons. Backscattered

electrons are reflected from the material. The size of the nucleus is the primary determinant for

the amount of backscattered electrons. Hence, this technique can be used to probe the atomic

composition of materials. Secondary electrons are not reflected from the material, but are actually

ejected from the material due to the energy of the incoming electron beam. More specifically,

when the electron beam interacts with the material, it imparts energy to the electrons in the

material, which will eject from the material if sufficient energy is imparted. The lower the ejection

Page 49: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

49

energy barrier is (i.e. the work function), the higher the probability an electron will be ejected,

suggesting that samples which appear brighter have a lower work function. Previous work has

shown that the SEM can be used to measure the work function of 2D materials.168 As shown in

Chapter 3, an SEM was used to probe the comparative work functions of Bi2Se3/MoS2 and

Bi2Se3/MoSe2 2D heterostructures before and after SEM-treatment, where the results suggest SEM-

treatment lowers the work function of the 2D structures. An explanation of the working principles

and physics of an SEM can be found at the following reference.169

2.7 Lithography: E-Beam, Photo, and Laser

Lithography was used extensively in this dissertation to build devices and put down markers for

wafers used to grow 2D materials. Additionally, it was found that both gold and titanium

promote the growth of 2D materials. Hence, lithography was also used to fabricate substrates

more favorable to the growth of 2D materials.

The exact instructions to apply lithography are extensive and beyond the scope of this

dissertation. Further, the instructions vary depending on the photoresist used, the operating

parameters of the equipment being used, and 2D material being work with. More specifically,

sensitivity of 2D materials to chemicals (e.g. photoresist, and Polymethyl methacrylate – PMMA)

varies, and specific processes need to be found for each material. For example, graphene is highly

sensitive to PMMA, and it is very difficult to remove it from the graphene surface. A consequence

is applying PMMA to graphene significantly dopes it and modifies its properties.

Laser lithography was developed during this dissertation with significant advantages over both

electron-beam lithography and photolithography. The 457nm laser in the laboratory coupled to

piezoelectric stages, allowing the laser spot location to be controlled with micron-scale precision.

Programs in LabView were written that allowed the writing of electrodes with micron-feature

size, and large pads to allow wires to be attached to the sample. The technology allows devices

on micron-scale 2D materials to be fabricated without the resource constraints demanded by

electron-beam lithography, or the pattern/mask constraints of photolithography. I demonstrated

Page 50: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

50

that titanium and gold could be deposited into the micron-scale channels, and it will remain on

the SiO2 wafer after stripping the photoresist away. This technique used chemicals readily

available in industry, suggesting it can be economically scaled. Unfortunately, the photoresist

showed indications it was damaging the Bi2Se3 layer of Bi2Se3/MoS2 2D heterostructures,

suggesting that further research is required to make it viable for device fabrication.

2.8 Transmission Electron Microscope (TEM)

A transmission electron microscope (TEM) was used in this dissertation to image 2D materials,

determine crystallinity using diffraction, and manipulate the twist angle and structure using

concentrated electron-beam exposure. The operation and physical concepts of a TEM are highly

complex and beyond the scope of this dissertation. The TEM used in this dissertation is the JEOL

2010F, where the manual can be found freely available online; however, the exact operation is

unique to each setup and institution. The general working principles and physics of a TEM can

be found at the following reference.170

Section 2.9 and Chapter 3 discusses the methods and techniques used to manipulate the twist

angle and crystal structure of 2D heterostructures using a focused electron-beam. Chapter 6

discusses the principles of selected area electron diffraction (SAED) in greater detail.

2.9 Electron Beam Treatment (EBT) Studies: Manipulating the twist angle and

interlayer interface

Prior to select area electron diffraction (SAED) EBT, the 2D heterostructures needed to be

transferred to a transmission electron microscope (TEM) grid. The TEM grids used were Pelco

QUANTFOIL® Holey Carbon film. PMMA C4 was first spin-coated at 4000 rpm for 60 s on the

SiO2 wafer with the 2D heterostructures and baked 180 C for 1:30 min. Then the chip was

immersed in 1 M KOH solution for 4 hours. Obtained PMMA and heterostructure film transferred

to new substrate. This was followed by acetone and IPA cleaning to remove PMMA residues. An

optical picture 2D heterostructures transferred onto the TEM grid are below.

Page 51: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

51

Figure 2.1: Optical Picture of Bi2Se3 2D heterostructures transferred onto a TEM grid

Once the grid was inside the TEM, a heterostructure suspended over a hole was found, zoomed-

in, and then focused. Then the 200nm-diameter aperture was inserted and TEM was switched to

SAED mode. Then the brightness was decreased to the lowest value, and an SAED image was

taken. To EBT the 2D heterostructure, the brightness knob was rapidly increased, while

maintaining the focus, until the desired value was obtained. The brightness was maintained for

the desired length of time as instructed by the recipe. After the recipe completion, the brightness

was rapidly diminished while maintaining focus. The operator had the phosphorous screen

lowered throughout the entire evolution and observed the SAED pattern carefully. The pattern

did not change as the brightness was rapidly increased or decreased, verifying that the imaged

location did not shift. Once the brightness was restored to its lowest value and the image focused,

the screen was raised and an image was taken. This process allowed snap-shots of the EBT

induced structural rearrangement to be observed.

The movement of the SAED spots would be demonstrated best by taking a movie of the SAED

pattern while EBT was being applied; however, this was not possible with the TEM used in this

work (JEOL 2010F). The TEM camera is exceptionally sensitive, limiting the exposure time to one

Page 52: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

52

second and the current density to the lowest value. EBTs are not allowed while the camera is

imaging. A phosphorous screen was lowered to block the electron beam while EBTs were applied,

allowing the crystal’s reaction to be observed and recorded by the TEM operator. At periodic

intervals the operator would decrease the beam current, restore identical imaging settings, and

record an SAED pattern, thereby documenting snap-shots of the EBT-induced changes.

The SEM EBT studies were done on the SiO2 wafer that the 2D heterostructures were grown on.

No transferring needed. Once inside the SEM, the SEM was zoomed-in and focused until near

the desired recipe, and pictures were taken. Once ready for the EBT recipe, the accelerating

voltage was modified and an area was scanned. The scanned area was calculated using the

equations in Chapter 3 to ensure that the electron beam current density was being applied. More

specifically, by modulating the area being scanned, the EBT current density was modulated. After

applying the EBT, the SEM scanning location was moved to a different part of the SiO2 wafer for

~20 minutes to allow for the 2D heterostructure to discharge. After ~20 minutes, the SEM imaged

the same area with the same settings.

2.10 Computational Methods

Lorentzian Fitting of the Photoluminescence Spectra for Exciton Analysis

The PL spectra were fit with two Lorentzian functions and a linear background, following

guidance in previously published work,33 where the Lorentzian’s represent the A exciton and A-

trion. The fitting was accomplished using the Matlab Fit() function. The quality of the fit was

accessed using the R-squared value and visual assessment.

Monte Carlo Simulations described in Chapter 5

Page 53: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

53

Monte Carlo simulations were done using Matlab software by MathWorks. 151 bins were made

representing the possible number of twist angles, which was 0-30° at increments of 0.2°. Then a

specific number of points were randomly assigned a twist angle and placed inside one of the bins.

The number of points assigned a twist angle was dependent on the number of experimental

points in the histogram (discussed in Chapter 3 and Chapter 9.2). 4-sigma criteria were applied

to determine whether a bin was statistically significant. The simulations were run 1,000,000

instances.

2.11 Device Fabrication

Bi2Se3/MoS2 2D heterostructures grown on 285 nm Si/SiO2 were transferred to an identical chip

that had titanium/gold markers, by the PMMA mechanical transfer method. First, PMMA C4 was

spin coated at 4000 rpm for 60 s and baked 180 C for 1:30 min. Then the chip was immersed in 1

M KOH solution for 4 hours. Obtained PMMA and heterostructure film transferred to new

substrate. This was followed by acetone and IPA cleaning to remove PMMA residues.

The field effect transistor (FET) devices were made on 285 nm Si/SiO2 substrate by Electron-beam

lithography using PMMA C4 or A4. The electrodes (5 nm Ti/50 nm Au) were deposited by e-

beam evaporator with rate deposition 1 and 3 Å/s, respectively. Lift off process was performed

with acetone followed by IPA cleaning.

Page 54: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

54

Chapter 3: Probing the interlayer interaction between

dissimilar 2D heterostructures by in-situ manipulation of

twist angle and structure

2D heterostructures are more than the sum of the two parent crystals - the interlayer interaction

facilitates the exchange of electrons and charge that can induce new properties.66,67,103,149,155,171 These

artificial materials hold enormous promise for new science and tailorable applications, due in

part to their diversity in properties.172–174 They often contain properties from both parent layers,

as well as new ones that result from the interlayer interaction.66,67,103,149,155 Recent works have

shown that the effect of the interlayer interaction can be substantial, and even dominate the

effective electronic and optical properties. Despite its importance, the interlayer interaction is not

fully understood and there are a limited number of tools to probe it. It has been shown to create

interlayer excitons,175,176 change the perceived color,150 shift e-h recombination paths,103 modify the

band structure,144 and even induce superconductivity.66 The importance of the interlayer

interaction has given rise to research that attempts to tune and manipulate it;149,177–179 however,

this is challenging to accomplish experimentally, and even theoretical descriptions are

incomplete. This work can be found published at the following reference.180

“Interlayer interaction”, sometimes called “interlayer coupling”, are broad terms that encompass

the known and unknown parameters affecting the exchange of electrons and charge. The atomic

registry (or interlayer nearest neighbors),159,181 interlayer spacing,144 interlayer charge

redistribution,103,182 intercalation of atoms,81,183 and interlayer electron energy barrier184 are some

parameters that affect the interlayer interaction and the 2D heterostructures properties. When

quantifying and studying this interaction, an important metric is the heterostructure’s stability

(i.e. the energy required to disrupt the current configuration, including twist angle and crystal

structure). Higher stability implies increased bonding (or charge redistribution) between the

layers, which facilitates increased electron and charge transfer, the primary “communication”

medium that leads to the manifestation of new properties. Hence, studying the stability will

reveal information about the interlayer interaction. Two notable parameters that affect the

Page 55: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

55

stability are the layer-to-layer atomic registry, and the chemical compositions. These parameters

determine how strongly one of the layers adhere to the other. Conversely, the ease with which

one could manipulate one layer with respect to the other, would be a measure of their interlayer

strength. However, there is currently no facile method for directly manipulating one layer of a

bilayer structure with respect to the other, for example in-situ relative twisting. There are several

impressive methods to manipulate the twist angle, including optically aligned mechanical

transfers,161,163 thermal annealing induced self-alignment,147,185 growing 2D heterostructures at

preferred twist angles,103,186–188 and AFM controlled twisting;160 however, each of these methods

has limitations. The relative orientation obtained by the first three methods cannot be altered once

the 2D heterostructure is fabricated, and the fourth has only been demonstrated for bulk-

monolayer heterostructures (i.e. one material is bulk). A method for in-situ modification of twist

angles would provide valuable insights into the nature of the interlayer interactions in 2D

heterostructures.

In this work, we investigate the interlayer interaction between two chemically and

crystallographically-dissimilar 2D layered materials, using statistical analysis of both as-grown

and in-situ “twist-angle” manipulated samples. We grew 1-2 layers of Bi2Se3 on top of several

different TMDs (MoSe2, MoS2, and WS2) and a TMD alloy (MoSe2-2xS2x), to test the how far the

chemical composition of the nearest and next-nearest interacting layer of atoms at the interface

impact the interaction strength. We found that the family of 2D heterostructures involving pure

TMDs all prefer (to varying degrees) to grow rotationally aligned at or near a twist angle of 0°,

with TMD-specific angular-distribution patterns. While the rotationally-aligned growth is a

possible consequence of van der Waals epitaxy, the angular distribution between 0-30° is

dependent on both interfacial surface atoms (i.e. chalcogens) and sub-surface atoms on the TMD

side (i.e. the transition metals), suggesting that the interlayer interactions penetrate beyond the

interfacial chalcogen layers. Additionally, the as-grown heterostructures corresponding to the

TMD alloys showed little or no affinity towards the 0° twist angle. The alloy vs. non-alloy TMD

results provide compelling evidence that long-range chemical order (i.e. of the same atomic

species) is critical for obtaining van der Waals epitaxy.

Page 56: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

56

Figure 3.1: Electron beam probes interlayer interaction in 2D heterostructures. (a) A vertically incident electron beam

treatment (EBT) imparts thermal energy into a 2D heterostructure that – at sufficiently high current densities – induces

it to overcome the interlayer interaction and rearrange. (b) Low current densities break only the weak interlayer bonds

associated with semi-stable grains, resulting in recrystallizing (i.e. grain migration to stable configurations). (c)

Intermediate current densities overcome most interlayer bonding, but cannot disrupt the stronger intra-layer bonding,

enabling significant grain rotation (twist). (d) Highest current densities break both the inter- and intra-layer bonding,

disrupting the crystallographic order and leading to the formation of nano-crystals. The interlayer stability was

quantified by varying the EBT recipe parameters (i.e. current density, time interval, and accelerating voltage) and

observing when the structure began to morph (or rearrange). By varying the beam current and analyzing the

parameters that induce each type of behavior, the interlayer interaction and configurational stability could be

compared.

More interestingly, we show for the first time that a tightly focused beam of electrons on

suspended Bi2Se3/TMD 2D heterostructures can selectively rotate the Bi2Se3 layer in-situ, and that

it is possible to do so on a large number of samples to obtain statistical relevance. This approach

enables a direct visualization and quantification of twist-angle-dependent interlayer bonding and

charge redistribution.182,189 We note that in an earlier study, we had seen the first evidence of Bi2Se3

layers rotating on MoS2 monolayers, triggered by a focused laser under ambient conditions.103

This present investigation was done using focused electron beams under high-vacuum,

circumventing any role of chemicals, and hence the observations are purely intrinsic measures of

the interlayer interactions. Figure 3.1a schematically summarizes our electron-beam treatment

(EBT) method, performed inside a JEOL 2010F high-resolution TEM. The intensity of the EBT was

controlled using the “brightness” feature of the TEM, and the crystallographic orientation of the

two layers in of the TMDs were determined using selected area electron diffraction (SAED)

Page 57: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

57

pattern spots. Under the application of an e-beam, the TMD lattice remains unchanged (except

for very high-density e-beams, discussed later), while the Bi2Se3 lattice showed signs of “rotation”

with respect to their original positions, the spots visibly “morphing” through continuous,

discrete, or diffused rotations. The interlayer stability could be quantified by varying the EBT

recipe parameters (i.e. current density, time interval, and accelerating voltage) and observing

when the structure began to morph (or rearrange). By investigating 12-16 samples, we were able

to categorize the behavior of these SAED spots into three broad categories: disruption, twisting,

or recrystallizing. More specifically, depending on the heterostructure being investigated, “low”,

“medium”, and “high” intensity EBT’s resulted in either breaking down of the Bi2Se3 layer into

smaller “grains”, twisting of these “grains” with respect to the underlying crystallographic axis

of the TMD, or recrystallizing of multiple grains into a lesser number. Figure 3.1b-d schematically

outlines these categories. Using this novel approach, we find that among the Bi2Se3/pure-TMD

junctions, the Bi2Se3/MoSe2 heterojunction appeared to be the most stable and Bi2Se3/WS2 the least;

while the Bi2Se3/TMD-alloy system shows the weakest interlayer interactions. Finally, we show

that sufficiently high doses of EBT appear to break the interlayer bonds and induce the Bi2Se3 to

form nanoparticles with a broad photoluminescence centered between 1.95-2.1eV.

The behavior of these 2D heterostructures is dependent on their interaction; however, what

comprises “weak” interactions between layers – the hallmark of layered materials – has no clear

definition or demarcation in the community. Electronically weak interactions would be an

outcome of a low overlap of electronic orbitals between the two participating layers, leading to

only incremental changes in the electronic properties, while mechanically it would be a measure

of how much energy is lowered when two free layers are brought into contact. These two

approaches are addressing the same question from two different perspectives. This work directly

addresses the mechanical interaction by allowing the interfaces go out of their lowest energy

configuration under EBT. We believe that by comparing the “ease” with which we are able to do

so should be a measure of their relative strengths, and that is presented here.

Page 58: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

58

Figure 3.2: As-grown 2D heterostructures prior to Electron Beam Treatment (EBT). (a) Optical image of a Bi2Se3/MoSe2

vertical heterostructure on SiO2, where 1-2 layers of Bi2Se3 were grown on a monolayer MoSe2 crystal (see Figure 3.3)

using vapor phase deposition. (b) Typical TEM SAED pattern of a Bi2Se3/MoSe2 heterostructure with the corresponding

diffraction spots labeled. The orange line shows the Bi2Se3 grew near-aligned on the MoSe2. The distinct dots indicate

well-formed crystals. (c) AFM scan of a Bi2Se3/MoSe2-2xS2x heterostructure, and (d) the corresponding line profile (left

to right). The scan is typical of a Bi2Se3/TMD 2D heterostructure: it has near-continuous monolayer Bi2Se3 with bilayer

islands. MoSe2-2xS2x grows crystalline with a lattice constant between MoSe2 and MoS2 (Figure 3.4), suggesting the

selenium and sulfur are randomly placed. Despite the loss of chemical periodicity, Bi2Se3 still grows crystalline (Figure

3.4).

Figure 3.2a shows an optical image of a Bi2Se3/MoSe2 2D heterstructure on SiO2, where 1-2 layers

of Bi2Se3 were uniformaly grown on top of a monolayer MoSe2 flake using vapor-phase

deposition.31 The original monolayer MoSe2 crystal (Figure 3.3) was grown using vapor-phase

chalcogenization (VPC), described in the methods section. Figure 3.2b is a typical TEM selected

area electron diffraction (SAED) pattern with the two parent crystals labeled. The distinct dots

indicate both materials are highly crystalline, and the orange line shows the Bi2Se3 grew aligned

along the 0° with respect to the underlying MoSe2 crystal. Figure 3.2c is an AFM scan of a

Bi2Se3/MoSe2-2xS2x heterostructure, and Figure 3.2d is the corresponding line profile. The AFM

scan is typical of any of the Bi2Se3/TMD 2D heterostructures: it contains near-continuous

monolayer Bi2Se3 with islands of bilayer regions.

Page 59: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

59

Figure 3.3: Representative Bi2Se3/TMD heterostructures. (a)-(b) Optical images of monolayer MoSe2 and a Bi2Se3/MoSe2

vertical heterostructure, respectively. Figure S1a shows an optical image of a monolayer MoSe2 flake grown on SiO2

using vapor-phase chalcogenization (VPC),31 its well-defined geometry suggesting a single-crystal structure. (b) shows

the same crystal after 1-2 layers of Bi2Se3 were uniformly grown on top using chemical vapor-phase deposition (see

methods section). The purple spots surrounding the 2D heterostructure are Bi2Se3 nanocrystals, which has been

previously demonstrated.52,190 (c) Side profile of a heterostructure with 1L Bi2Se3 and 1L TMD.

We are also able to infer information about the TMD-Bi2Se3 interlayer interaction by comparing it

to Bi2Se3 growth on SiO2, a material commonly used as a platform to grow 2D heterostructures.

First, when lower concentrations of vapor-phase bismuth and selenium atoms are generated,

Bi2Se3 grows only on the TMD (Figure 3.2c); however, at higher concentrations, the Bi2Se3 begins

to grow on the SiO2 as well (purple “spots” surounding heterostructure in Figure 3.2a). Second,

Bi2Se3 grows as multilayered nanocrystals on SiO2 (Figure 3.4e), but prefers to grow monolayer

and highly epitaxial/crystalline on TMDs (Figure 3.4a-d). Together the data suggests there is a

substantial interlayer interaction that stabilizes the Bi2Se3 in its few-layered morphology.

Page 60: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

60

Figure 3.4: TEM SAED Patterns of four 2D heterostructures studied, as well as Bi2Se3 grown on SiO2. (a) Bi2Se3/MoSe2,

(b) Bi2Se3/MoSe2-2xS2x, (c) Bi2Se3/MoS2, (d) Bi2Se3/WS2, and (e) Bi2Se3 grown on SiO2. The distinct dots indicate both

materials are highly crystalline. Interestingly, the MoSe2-2xS2x grows highly crystalline despite the fact that it is an alloy.

The both photoluminescence peak center, as well as the lattice constant, was between the MoS2 and MoSe2 values,

strongly suggesting that an alloy was grown, vice only a single material. We find the interesting fact that although

Bi2Se3 grows as multilayered nanocrystals when grown directly on the SiO2 (Figure 3.4e); however, it prefers to grow

as monolayers, and highly epitaxial/crystalline when grown on TMDs (Figure 3.4a-d), suggesting there is a substantial

interlayer interaction that stabilizes the Bi2Se3 in its few-layered morphology

Figure 3.5: TEM images of different 2D heterostructures. (a)-(c) TEM images of the same Bi2Se3/WS2 crystal from

different perspectives. The blemish in the photo (see top of (c)) is due to laser treatment, a method used to gently

Page 61: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

61

disrupt the crystal structure.103 Laser treatment was affected in order to disrupt the crystal structure and break the

Bi2Se3 into smaller grains, enabling the demonstration that EBT can anneal. (d) Bi2Se3/MoSe2 2D heterostructure. (e)

Bi2Se3/MoS2 2D heterostructure.

Figure 3.5 shows a variety of TEM images of different 2D heterostructures. The islands of Bi2Se3

can be seen, suggesting the electrons undergo a phase change or change in scattering between

different regions. Figure 3.5a-c show a 2D heterostructure that was exposed to a laser in an air in

a local region, where the change is evident. The laser appears to have induced tearing. It was

found that a 2D material suspended on a TEM grid was significantly more suscentible to tearing

than one grown on an SiO2/Si substrate. This is likely a result of the fact that these 2D materials

are suspended across micron-sized holes, but only approximately two nanometers tall,

suggesting that the bonding is exceptionally strong to hold the materials togetehr. When a laser

is applied, the materials heats locally and the strain from the suspended sample causes tearing.

The change in contrast of the region indicates changes in the material, possibly due to the

intercalation of oxygen or the formation of new materials due to oxygen absorption. Figure 3.5d-

e show that tearing and folding can occur during the transfer process.

Page 62: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

62

Figure 3.6 Tuning the interlayer interaction through chemistry. Histograms of the twist angles (Bi2Se3 vs. the TMD) in

as-grown 2D heterostructures (as labeled), with the following interlayer chemistry: (a) Se-Se interface with

molybdenum. (b) Se-S interface with molybdenum. (c) Se-S interface with tungsten. (d) Se-Se/S-alloy interface with

molybdenum. Despite their large crystallographic dissimilarities, Bi2Se3 prefers to grow near 0° when chemical

periodicity is present in the underlying material (i.e. non-alloy heterostructures), but prefers 12°-20° when it is absent

(i.e. MoSe2-2xS2x), suggesting chemical periodicity strengthens the interlayer interaction at 0°. The frequency at 0°

decreases from (a)-(c) suggesting Se-Se interfaces and molybdenum-based TMDs form stronger interlayer interactions

at this angle. Although Monte Carlo simulations with a “4-σ” criteria (see Section 3.1 later in this chapter for more

details) reveal that most of the data outside of 0° is statistically insignificant quantitatively, we find qualitatively that

the interlayers with TMDs “weaker” interactions were found to have larger angular spread in their “twist” angles. As

discussed later, experimental data indicates that the defect density is low.

The growth of 2D materials is governed by the interlayer interaction, and we use the twist angle

of the as-grown heterostructures as our first parameter to indicate the interlayer interaction

strength. In other words, if Bi2Se3 grows at preferred, statistically significant twist angles, they

represent energy-minimized configurations; conversely, if the twist angle distribution is more

random (i.e. “spread out”), we will infer weaker, non-direction specific interactions. The

measured twist angle distributions for each type of heterostructures grown were collected from

75-150 “measurements” taken across 15-40 flakes, for each heterostructure type. The resulting

angular distribution are shown as histograms in Figure 3.6a-d, corresponding to Bi2Se3/MoSe2,

Bi2Se3/MoS2, Bi2Se3/WS2, and Bi2Se3/MoSe2(1-x)S2x heterostructures, respectively. The insets in each

Page 63: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

63

figure show the relative arrangement of atoms in the nearest and next-nearest layers at the

interface. We see that as the TMD and corresponding interlayer interaction change, so does the

angular distribution. Bi2Se3 has a 79.7% probability of growing at a 0° twist angle on MoSe2 (Figure

3.6a), whereas it only has a 41.0% probability on MoS2 (Figure 3.6b), a 24.5% probability on WS2

(Figure 3.6c), and a near 0% probability on MoSe2(1-x)S2x. The twist angles outside of 0° were found

to be statistically insignificant based on Monte Carlo simulations (see Section 5.1 later in the

chapter for more information), suggesting the stability landscape outside of 0° is relatively flat

with only weak preference, although we can see some qualitative correlation between how weak

the response is, and how far “spread out” the distribution is. The stability and energy of the

interlayer interaction is in part determined by both the spatial density and energy of the interlayer

bonds which mediate the amount of charge redistribution/transfer between the layers. Also, the

bond energy is dependent on the interatomic distance, where a sufficiently long distance prevents

bond formation or significantly weakens it, thereby limiting bond formation to nearest

neighbors.191–193 The relative positioning, or atomic registry, of the atoms is instructed by the twist

angle. In summary, angles with greater stability are more likely to be observed, suggesting a 0°

twist is the most stable configuration for the non-alloy heterostructures.

Comparing the distribution in Figure 3.6a and Figure 3.6b, it is clear that the Se-Se nearest-layer

interface (Bi2Se3/MoSe2, Figure 3.6a) results in a much stronger inter-layer interaction compared

to the Se-S interface (Bi2Se3/MoS2, Figure 3.6b), as inferred from the near halving of the

rotationally-oriented (0°) percentage population of the latter. The more surprising result,

however, is how much the chemistry of the “next-nearest-layer” atoms affects the interlayer-

interactions, as seen by comparing the distributions in Figure 3.6b and Figure 3.6c, which

represent dissimilar “next-nearest-neighbors” only: i.e. -Mo-S-Se-, vs. -W-S-Se-. Once again, the

percentage population of rotationally-aligned heterostructures fall roughly by half, denoting that

W-based TMDs interact less strongly with Bi2Se3, compared to Mo-based ones. We note that when

molybdenum is replaced with tungsten, the in-plane lattice constant increases by less than 1%

(Figure 3.4), and hence it is unlikely that the relatively small change to the in-plane lattice constant

could alone induce such a large change to the stability at 0°, suggesting next-nearest-neighbor

Page 64: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

64

chemistry is strongly influential in determining the interlayer interaction in these

heterostructures.

Finally, we find that the Bi2Se3/MoSe2-2xS2x distribution (Figure 3.6d) is distinctly different from all

other heterostructures, despite the fact that the same atomic compounds were used (i.e. Mo, Se,

and S), and that the TMD-alloy has long-range crystallographic order using lattice constants

comparable to the other heterostructures (Figure 3.4). Comparing Figure 3.6d, we see that the

Bi2Se3 layer showed no tendency for forming rotationally aligned heterostructures (although the

underlying alloy-TMD had a well-defined, uniform crystallographic orientation), with the

orientations spread widely across the entire measured range. Since the only noteworthy

difference is that the chemical nature of the TMD alloy was random, i.e. the Se and S atoms are

likely randomly distributed with the TMD lattice, our results suggest that long-range chemical

periodicity, not just crystallographic periodicity, is critical for strong interlayer interaction, and

possible van der Waals epitaxy. Raman spectroscopy results (Figure 3.19), as well as recent

publications by our group,30,31,103,104 indicate the 2D heterostructures studied in this work had a

low number of defects, which would affect the interlayer interaction (see Section 5.2 of this

Chapter for more information). Taken together, the data from as-grown heterostructures suggests

the interlayer interaction (at 0°), arranged from strongest to weakest, is: Bi2Se3/MoSe2, Bi2Se3/MoS2,

Bi2Se3/WS2, and Bi2Se3/MoSe2-2xS2x, and we will use this result to test the efficacy of the inference

drawn from the EBT-treatment, and compare how far they are in agreement, as discussed next.

Our findings indicate that the similarity of the touching layers, i.e. Se-Se interactions leads to a

stronger interaction between the two parent crystals compared to Se-S layers. On the other hand,

for the Se-S case, the Mo-based TMD interacted more strongly than the W-based TMD. It was not

possible to pinpoint exactly which aspects (e.g. size, electronegativity, lattice parameters, density

of nearest-neighbor atoms within a moiré supercell) play the most dominant role in these two test

cases. We found little or no interaction in the case of the alloy, possibly due to its chemical

randomness. From these, we conclude that the so called “weak” interaction that is supposed to

be a hallmark of layered materials is at best a loose description, since similar categories of

materials can have vastly different interactions, as revealed by our work.

Page 65: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

65

The effect of e-beam on the heterostructures was investigated using selected area electron

diffraction (SAED) patterns of the samples in situ. SAED is conventionally used to probe a

material’s crystal structure in reciprocal space, where well defined patterns of bright spots

indicates a sample of high crystallinity.170,194 The material’s crystal structure determines the

pattern generated, while the lattice constant determines the pattern’s size - smaller lattice

constants produce larger patterns.170,194 When probed along the z-direction, Bi2Se3 and the TMDs

produce similar hexagonal SAED patterns; however, Bi2Se3 has a smaller pattern because its

lattice constant is larger (see Figure 3.2and Figure 3.4). The bright spots are spaced every 60° along

the same radii; however, due to rotational symmetry (i.e. –θ = θ), only twist angles between 0-30°

are unique.

Figure 3.7 Manipulating the twist angle in-situ using an electron beam. (a)-(d) TEM SAED images of a Bi2Se3/MoS2

heterostructure subject to consecutive TEM electron beam treatments (EBTs). (a) Initially two Bi2Se3 spots are observed,

where the spot at 27° is 2.58× times brighter than the spots at 21°. (b) Post 1st EBT: The spot at 21° splits into 3 smaller

spots, and a new spot appears at 3.7°. (c) Post 2nd EBT: The 3 spots near 21° combine into two, and the 27° spot splits

into two. (d) Post 3rd EBT: Both sets of Bi2Se3 spots settle at 24°, and are nearly equally bright - 8.4 vs. 8.0 photon counts

for orange and blue, respectively – suggesting the 24° twist angle is more stable than either 21° or 27. The 3.7° spot

disappeared, suggesting the grain was perturbed out of the imaging area. A vertically-incident electron beam can twist,

split, and merge Bi2Se3 grains grown on a TMD (more demonstrations in Figure 3.8 and Figure 3.9).

Page 66: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

66

When the TEM e-beam is incident on the heterostructures, it induces structural changes in the

SAED spots, and the nature of these changes depend on the e-beam current density (and hence

the rate of energy transfer to the crystals). Higher densities break bonds and disrupt the

crystallinity, whereas lower densities recrystallize by gently perturbing the material into a more

stable state with greater crystallinity. Densities in between exhibit more complex behavior

because the power is sufficient to loosen weak bonds (e.g. interlayer and van der Waals), but not

strong bonds (e.g. intra-material). Appropriately-tuned densities tend to loosen the interlayer

bonding while simultaneously perturbing Bi2Se3 grains to rotate relative to the TMD. The

movement can be observed in-situ using a TEM in SAED mode, revealing information about the

interlayer interaction landscape (e.g. the relative stability of the twist angles). It is reasonable to

assume that under EBT treatment, the cumulative time spent at each twist angle configuration

correlates to their relative stability, i.e. more time equates to higher stability.

We first present results for EBT treatment for a fixed beam current, using multiple exposures. The

SAED images in Figure 3.7 show the evolution of a Bi2Se3/MoS2 2D heterostructure subject to

multiple EBTs of the same recipe (e.g. the same electron beam current density was applied),

demonstrating in-situ manipulation of the twist angle. Figure 3.7a shows the initial configuration,

where two Bi2Se3 spots at 21° and 27° are identified by orange and blue arrows, respectively. Here,

the angles were measured with respect to the nearest stable MoS2 spot, shown by similarly-

colored dashed lines, resulting in the 21° spot measured clockwise and the 27° spot is counter-

clockwise from their nearest reference MoS2 spot. We associate the relative brightness of spots

with the relative areas of a material associated with the spot, where increased brightness

correlates to increased amounts.194 Initially the 27° spot is 2.58× brighter than the 21° spot,

suggesting the size of the Bi2Se3 crystallite under the TEM e-beam corresponding the 27°is bigger.

Under the EBT, the spots moved in real-time. Although these spot-movements could be seen real-

time on the phosphorescent screen of the TEM, unfortunately, they could not be recorded

videographically owing to an instrument limitation. Hence, we present snapshots of the system

between EBT treatments.

Page 67: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

67

The first EBT (Figure 3.7b) disrupted the Bi2Se3 spot at 21°, splitting it into three closely aligned

spots (18°, 21°, and 24°) that appeared to move within narrow angular ranges, indicating they

were only at semi-stable twist angles. The spot at 27° did not appear to move. Interestingly, a spot

at 3.7° appeared during the EBT; however, due to its rapid emergence, we are not sure of its

origins. This uncertainty led us to question whether the imaged location shifted, which is possible

if either the electron beam or the sample moved during the EBT. We tested for this, but found the

TEM to be probing the same area before and after EBT, leading us to conclude the changes are

due to movement of Bi2Se3 grains. We believe that a grain moved into the imaging field of view

and settled at a twist angle of 3.7°. This is a probable scenario since the electron beam

simultaneously treats areas inside and outside of the imaged location. While operating the TEM

in SAED mode, an aperture is inserted below the sample to diminish the spot intensities and

protect the camera. The aperture only reduces the surface area being imaged, and does not affect

the electron beam, its interaction with the sample, or the area being subject to EBT.

The second EBT (Figure 3.7c) merged two of the orange spots (the left spot at 18° merged with

the middle spot at 21°), and disrupted the spot at 27° to split into two spots (24° and 27°,

respectively). The third EBT (Figure 3.7d) merged the two orange spots (21° merged with 24°),

and merged the two blue spots (27° merged with 24°), resulting in only two spots that are both at

24°, but at opposing rotations. More specifically, the orange spot is twisted 24° to the right, and

the blue spot is twisted 24° to the left. Very interestingly, both spots at 24° are nearly equally

bright, where the orange spot is only 1.05× brighter (8.4 vs. 8.0 photon counts), respectively. This

is a significant change from the initial SAED image, where the blue spot was 2.58× brighter,

suggesting the system evolved into a statistical equilibrium where both the right- and left-circular

24° twist angles contain the same amount of Bi2Se3 crystal. These results are in strong agreement

with the fact that these 2D heterostructures have rotational symmetry (i.e. –θ = θ). Without any

external forces to break the symmetry, it appears that the EBT evenly split the Bi2Se3 crystals

between the two 24° twist angles. The spot at 3.7° also disappears, possibly because it moved out

of the imaging area. Our investigations hence reveal that although there were some quasi-stable

configurations (located at 18°, 21°, and 24°), the fact that both spots settled at 24° suggests that

24° is the most stable twist angle in the Bi2Se3/MoS2 heterostructure.

Page 68: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

68

Figure 3.8 In-situ crystal manipulation of Bi2Se3/MoSe2 using an electron beam. A Bi2Se3/MoSe2 2D heterostructure was

transferred to a TEM grid and laser treated, a previously demonstrated method to gently disrupt the crystal structure.103

The laser treatment disrupted the heterostructure enabling the demonstration that the heterostructure can be annealed

using an electron beam. A high current density was applied to the above heterostructure. Instead of becoming more

polycrystalline as seen with Bi2Se3/MoS2 (Figure 3.7) and Bi2Se3/WS2 (Figure 5), the heterostructure was annealed,

suggesting Bi2Se3/MoSe2 has a stronger interlayer interaction than the other heterostructures. Additionally, in (d)-(f)

the smaller MoSe2 grains cease to anneal into the large crystal, and oscillate near their current angle, suggesting those

grains are at a relatively stable twist angle.

Page 69: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

69

Figure 3.9 Bi2Se3/WS2 controllable crystal disruption, annealing, and rotation in response to EBT. (a) Initial

measurement. 2D heterostructure was disrupted using laser-treatment, a process where a focused laser beam is applied

Page 70: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

70

to the sample in an environment with O2. (b)-(f) Eight electron beam treatments (EBTs) were consecutively applied.

Precisely determining the current density through the sample is difficult; however, the brightness can be controlled,

where 0 is the lowest and 10 is the highest. All measurements were taken at the lowest brightness. The ratio of spot

brightness is used, vice the absolute brightness, to mitigate equipment and operator error. It can be seen that the

heterostructure behavior can be controlled and is determined by the current density. Higher densities disrupt the

crystallinity [(b), (d), (h), and (f)], whereas lower densities anneal [(c), (e), and (g)]. Very small values do not affect

noticeable changes (f). (g) The Bi2Se3 grain rotated into the 0° relative angle, a configuration believed to be more stable

than the other angles based on epitaxial growth studies (Figure 3.6). As mention in this Chapter, determining accurate

current density values is beyond the scope of this project due to the large number of variables present in focusing the

electron beam, and limitations in the equipment sensitivity. As purely a guide, we provide the values measured with

no aperture or sample inserted: the lowest value was 0.0pA/cm2 and highest was 33.4pA/cm2. Inserting the aperture

diminishes the beam intensity where it is no longer within range of the sensor.

Figure 3.10 Controllable crystal rearrangement in-situ using an electron beam. Electron beam treatments (EBTs) were

consecutively applied to a Bi2Se3/WS2 2D heterostructure, demonstrating the structural rearrangement possibilities. The

electron beam current density was adjusted across EBTs to demonstrate controllable structural rearrangement. (a)

Initial SAED image of a Bi2Se3/WS2 2D heterostructure, where the two Bi2Se3 spots and the WS2 crystal are identified

with green, red, and blue arrows, respectively. The relative intensities (i.e. photon counts) of the Bi2Se3 spots compared

to the WS2 spot are shown in the corresponding graph panel. (b) SAED image after two consecutive EBTs (60% followed

by 20%), where the red-arrow spot has vanished and the green-arrow spot is noticeably more intense (see

corresponding panel). (c) SAED image after applying a third EBT (60%), which disrupted the structure, inducing the

green-arrow spot to split into two. (d) SAED image after applying three additional EBTs (20%, 10%, and 20%), which

recrystallized the heterostructure and twisted the red-arrow spot into the 0° angle, a known stable twist angle (Figure

3.6). Further info can be found in Figure 3.8 and Figure 3.9, which also shows how very high current densities will

significantly disrupt the Bi2Se3/WS2 heterostructure. Approximate current density values are shown in Figure 3.9.

Page 71: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

71

We next present results for variable EBT beam currents on the same sample. Figure 3.10a-d shows

SAED images of the same location on a Bi2Se3/WS2 heterostructure, where multiple consecutive

EBTs of varying electron beam current density were applied, demonstrating controllable

structural rearrangement of the 2D heterostructure. The current density was controlled using the

TEM “brightness” knob, and is expressed as a percentage of the “brightness” knob’s possible

range. Higher brightness values correspond to increased current density, which disrupts the

heterostructure’s crystallinity. Lower brightness values recrystallize the heterostructure and

increase the crystallinity. Due to equipment limitations, the exact current densities are difficult to

measure; however, a guide of approximate values can is shown in Figure 3.9b. Figure 3.10a-d

present snapshots of SAED patterns following representative EBT treatments, while the graph in

Figure 3.10e compares the variation of the Bi2Se3 spot intensity (i.e. the photon counts) values

relative to the corresponding WS2 spot.170,195 Numerous EBTs of various current density values

were consecutively applied to the same location. The EBT current density used is stated below

each panel as a percentage of the possible range. The complete set of SAED images after each

treatment dose has been presented in Figure 3.8 and Figure 3.9.

Figure 3.10a is the SAED image of the pristine sample, where the two weaker Bi2Se3 spots and a

stronger spot corresponding to the WS2 crystal are identified by green, red and blue arrows,

respectively. The corresponding panel in Figure 3.10e shows the relative intensity values of both

Bi2Se3 spots to the WS2 spot, where the green spot is nearly twice as intense as the red spot. Figure

3.10b shows the result after two EBTs were applied (60% and 20%, respectively), which perturbed

the red-arrow spot to merge with the green-arrow spot.

A single high-density (60%) EBT was then applied (Figure 3.10c), which split the green spot into

two. The new spot (clockwise from the green spot) was labeled with the red arrow. This was done

to facilitate analysis, even though the Bi2Se3 crystals/atoms that correspond to this spot are likely

not the same as the original red-arrow spot. Next, as shown in Figure 3.10d, three low-density

EBTs were applied (20%, 10%, and 20%, respectively), perturbing the red spot to rotate into the

0° twist angle and become brighter, as seen in the corresponding Figure 3.10e panel. 0° is a known

Page 72: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

72

stable twist angle for Bi2Se3/WS2 2D heterostructures (Figure 3.6), suggesting that the low-density

EBTs recrystallized the 2D heterostructure into its most stable twist angle.

Figure 3.10e demonstrates several important results. When EBTs of 60% current density intensity

are applied, the combined brightness of the Bi2Se3 spots decreased, suggesting the

heterostructure’s crystallinity is being disrupted,170,195 since amorphous, or non-crystalline,

materials do not produce well defined SAED spots or rings. Conversely, when lower EBT current

density values are applied (20%), the combined brightness increases, suggesting the

heterostructure’s crystallinity is being recrystallized and increased. EBTs of 10% did not appear

to have a noticeable effect, suggesting the power was too low to affect change. As EBTs were

applied, the relative brightness between the red and green-arrow spots changed, demonstrating

structural reconfiguration and movement of crystals. The most interesting reconfiguration is

shown in Figure 3.10d where the red spot rotated to the 0° twist angle, a known stable twist angle,

a strong indication that the interlayer interaction significantly influences the behavior, and

suggests the efficacy of using this method to study the interlayer interaction. As shown in Figure

3.8 and Figure 3.9, EBTs of 100% current density values will rapidly disrupt the heterostructure’s

crystallinity, and even showed the ability to disrupt the WS2. In this manner, it was possible to

discern values of EBT-brightness that could controllably manipulate SAED spots, in-situ.

Figure 3.11 Summary of different 2D heterostructure’s response to an EBT intensity. Lower EBT intensities (i.e.

“brightness” knob settings) tend to recrystallize the 2D heterostructures and increase the crystallinity, while higher

EBT intensities tend to disrupt. The EBT intensity is correlated to the current density and the amount of power imparted

to the 2D heterostructure. It can be seen that Bi2Se3/MoSe2 is the most stable, while Bi2Se3/WS2 is the least.

Page 73: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

73

Figure 3.11 shows the responses of three different 2D heterostructures (Bi2Se3/MoSe2, Bi2Se3/MoS2,

and Bi2Se3/WS2) when different EBT intensities were applied, allowing their interlayer

interactions to be probed and compared (further info in Figure 3.7, Figure 3.8, Figure 3.9, and

Figure 3.10). These results are in agreement with the epitaxial growth studies (Figure 3.6), which

indicate that Bi2Se3/MoSe2 has the strongest interlayer interaction and Bi2Se3/WS2 the weakest.

While high current densities disrupted both Bi2Se3/MoS2 and Bi2Se3/WS2 (Figure 3.10), they

recrystallized Bi2Se3/MoSe2 heterostructures (Figure 3.7, Figure 3.8, and Figure 3.9), rotating a

significant portion of the Bi2Se3 spots into the 0° twist angle and increasing the crystallinity, in

agreement with earlier results that showed the 0° twist angle is very stable for Bi2Se3/MoSe2

heterostructures (Figure 3.6).

Previous work has demonstrated thermal heating is able to rotate and merge monolayer MoS2

nanoparticles,196 suggesting the EBT dose is being converted into thermal energy and inducing

the observed changes. Additionally, it has been previously shown that vertically incident TEM

electron irradiation is able to rotate and move crystallites and nano-particles.197 Together this

suggests that the mechanism for recrystallization, twisting, and disruption is due to heating by

electron beam irradiation.

To the best of our knowledge, this is the first demonstration of the in-situ manipulation of a 2D

heterostructure’s twist angle inside a TEM, a capability with significant potential when combined

with TEM measurement methods. More specifically, electron scattering inside a TEM has

demonstrated numerous impressive capabilities, including probing the interlayer van der Waals

electronic structure,182 measure chemical induced charge redistribution,198 probe interlayer energy

in 2D heterostructures,189 measure the interlayer distance,12 and probe crystallographic ordering

and charge redistribution of intercalated atoms.81 Current methods have not demonstrated the

ability to manipulate the twist angle on the TEM grid, limiting the amount of data that can be

reasonably collected. In contrast, this method allows the twist angle to be manipulated in-situ

facilitating numerous data points from the same sample. Further, this method can be done

Page 74: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

74

rapidly, highly locally, it is site selectable, and even nanoscale samples can be probed. Combining

these methods with the ability to control the twist angle in-situ can hence enable significant

progress in our understanding of 2D heterostructures whose properties can be engineered using

the twist angle.

As discussed in more detail in Section 5.2 of this chapter, we do not believe that EBT induces a

material transformation. Since SAED spots remain distinct before and after the EBT treatment, it

is difficult to assign this transformation to any “permanent” changes during the EBT.

Additionally, any intermediate “phase transition” is unlikely as the “re-oriented” spots represent

identical lattice parameters after the EBT, and we never see any other crystal structure no matter

what the extent of EBT was. Lastly, chemical modification such as oxidation is unlikely since all

the experiments were done at ultra-high vacuum and does not explain why the Bi2Se3 crystal

reappears after EBT.

Figure 3.12 Manipulation of the optical and photoluminescence properties using an electron beam. (a)-(b) Optical and

SEM images, respectively, of Bi2Se3/MoS2 heterostructures. Arrows mark locations where electron beam treatment

(EBT) was applied using a scanning electron microscope (SEM). Red arrows correspond to higher energy densities

(≥8.44x106 J·m-2) and green arrows to lower energy densities (<3.05 x106 J·m-2). While lower densities do not appear

to induce any changes, higher doses shift the heterostructure’s perceived color from purple to green and induce a new

PL peak between 1.95-2.1eV, shown in (c) and Figure 3.14). They also appear to alter the material’s texture and work

function based on SEM image analysis (Figure 3.15). (c) PL of as-grown Bi2Se3/MoS2, and the PL of a Bi2Se3/MoS2

heterostructure after high-dose EBT was applied. High-dose EBT induces a broad peak centered between 1.95-2.1eV,

which corresponds to Bi2Se3 nanoparticles. It was established in Figure 3.7, Figure 3.8, Figure 3.9, Figure 3.10, and

Figure 3.11 that high density EBT rearranges the structure and disrupts the Bi2Se3 to become nano-crystalline. All this

evidence suggests that high density SEM-EBT induces the Bi2Se3 to form nanoparticles. Here we demonstrate a

technique that allows for submicron pattern writing using focused electron beams.

Page 75: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

75

As discussed in Figure 3.12, EBT recipes contain three important parameters: current density,

accelerating voltage, and time interval. Using a TEM to apply EBT allows the crystal structure to

be observed in-situ; however, precisely measuring the electron beam current density or

manipulating the accelerating voltage are difficult, limiting our ability to quantify the beam

current. To overcome this, we also applied EBT using a scanning electron microscope (SEM),

whose electron beam parameters can be precisely measured and manipulated. Moreover, SEM

beams allow us to “write” complex patterns with nanoscale dimensions,199 and raster over a

greater spatial region, thereby allowing us to EBT-modify samples with regions large enough for

their properties to be investigated optically (i.e. under an optical microscope). The three

parameters (current density, accelerating voltage, and time interval) can be independently

controlled using an SEM. Increasing any of these parameters raises the energy dosage delivered

and, thereby, the likelihood of structural modification and the emergence of new properties. We

next show that sufficiently intense SEM-EBT recipes induce three notable changes: the perceived

color under broad-band illumination shifts from purple to bluish-green, emergence of a new,

broadband photoluminescence, and a possible modification of work function of the treated area.

Figure 3.12a-b are optical and SEM images, respectively, of Bi2Se3/MoS2 2D heterostructures (1-2

layers of Bi2Se3 CVD grown on monolayer MoS2) subject to different SEM-EBT recipes, where the

current density was manipulated, while the time interval (60 seconds) and accelerating voltage

(5keV) were held constant. The exposed areas are identified by red and green arrows,

representing recipes using higher vs. lower current densities, respectively, with values noted in

the figure caption. Optical and SEM images of the locations prior to EBT can be seen in Figure

3.13. Changes were only observed for the recipes using higher current (or energy) densities

(≥8.44x106 J·m-2, red arrows) and not for the lower densities ((<3.05 x106 J·m-2, green arrows),

suggesting a threshold current density, or energy dose, is required to induce change, which is in

agreement with the results in Figure 3.10, where threshold current densities (>10% “brightness”)

were required to induce structural rearrangement.

Page 76: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

76

The beam current increases slightly with increasing accelerating voltage. Example values are:

152.8pA, 157.3pA, and 160.2 for 5keV, 5.5keV, and 6keV, respectively. The values in Equation 1

and Equation 2 were used to calculate the dosing densities in Figure 3.10. We found treating

Bi2Se3/MoSe2 for 3 minutes at ~19.7A·m-2, changes were observed at accelerating voltages greater

than 5.5keV (2.01x107 J·m-2), and no changes were observed at 5keV (1.78x107 J·m-2). We found

treating Bi2Se3/MoS2 heterostructures for 60 seconds at 5keV, changes were observed to start

occurring at values greater than 8.44x106 J·m-2, and no changes were observed at densities less

than 3.05x106 J·m-2. As demonstrated in the movie in the SI, Bi2Se3/MoSe2 was affected with SEM-

EBT. It can be seen that the structural modification is not instantaneous, but happens slow and

steady over a set period of time. In the case of the movie, the change was observed over the course

of a minute. All the experiments came to the same conclusion: changes are only induced when

the electron beam current density crosses a certain threshold, and hence we can conclude that this

represented intrinsic energy barriers to changes.

𝐸𝑙𝑒𝑐𝑡𝑟𝑜𝑛 𝐵𝑒𝑎𝑚 𝐷𝑜𝑠𝑖𝑛𝑔 𝐷𝑒𝑛𝑠𝑖𝑡𝑦 = 𝐵𝑒𝑎𝑚 𝐶𝑢𝑟𝑟𝑒𝑛𝑡

𝑆𝑢𝑟𝑓𝑎𝑐𝑒 𝐴𝑟𝑒𝑎∗ 𝑇𝑖𝑚𝑒 ∗ 𝐴𝑐𝑐𝑒𝑙𝑒𝑟𝑎𝑡𝑖𝑛𝑔 𝑉𝑜𝑙𝑡𝑎𝑔𝑒 Equation 3.1

𝐸𝑙𝑒𝑐𝑡𝑟𝑜𝑛 𝐵𝑒𝑎𝑚 𝐷𝑜𝑠𝑖𝑛𝑔 𝐷𝑒𝑛𝑠𝑖𝑡𝑦 = 152.8𝑝𝐴∗60𝑠∗5𝑘

𝑆𝑢𝑟𝑓𝑎𝑐𝑒 𝐴𝑟𝑒𝑎=

4.58𝑥10−5𝐽

𝑆𝑢𝑟𝑓𝑎𝑐𝑒 𝐴𝑟𝑒𝑎 Equation 3.2

Figure 3.12c shows representative PL spectra from as-grown and EBT-affected Bi2Se3/MoS2

heterostructures, respectively (see Figure 3.14 for similar results in Bi2Se3/MoSe2) respectively.

Very interestingly, high-dose EBT induces the emergence of a new broad PL peak centered

between 1.95-2.1eV, and which corresponds to those seen previously reported Bi2Se3

nanoparticles.52,190 It was shown in Figure 3.7, Figure 3.8, Figure 3.9, Figure 3.10, and Figure 3.11

that high-dose EBT can disrupt the heterostructure’s crystalline order and induce the formation

of smaller Bi2Se3 crystallites. Both these observations suggest the emergence of the PL might be

due to the formation of Bi2Se3 nanoparticles. Low-dose EBT does not appear to have an effect on

the PL, which is in agreement with the optical and SEM images in Figure 3.12a-b, where no

changes were detected. Figure 3.15 shows an SEM image of as-grown vs. EBT-affected

Page 77: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

77

Bi2Se3/MoSe2, demonstrating high-dose EBT modifies the texture and contrast on the 2D

heterostructure’s SEM image, suggesting possible changes to the grain size, structure or work

function. Materials with a lower work function will emit more electrons and appear brighter.168,200

Figure 3.13 Manipulation of the optical and photoluminescence properties using an electron beam. (a)-(b) Optical and

SEM images of as-grown Bi2se3/MoS2 heterostructures, respectively. (c)-(d) Arrows mark locations that were electron

beam treated (EBT), where red corresponds to higher current densities (≥8.44x106 J·m-2) and green to lower (3.05 x106

J·m-2.

Page 78: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

78

Figure 3.14 Bi2Se3/MoSe2 affected with electron beam treatment (EBT). (a) Photoluminescence spectra for 1 layer MoSe2,

1-2 layer Bi2Se3 grown on MoSe2, post high dose EBT, and the same location after being laser treated. It can be seen that

EBT induces a new broad PL peak centered about 2.1eV. The new broad PL peak corresponds to previously reported

Bi2Se3 nanoparticles, suggesting that high-dose EBT induces the formation of Bi2Se3 nanoparticles.52,190 Laser treatment

diminishes the PL intensity, suggesting that the nanoparticles are metastable, and when perturbed by a laser, they will

morph into a structure that has a less intense PL. (b) SEM image of as-grown Bi2Se3/MoSe2. (c) Same heterostructure

after low-dose EBT (5keV), where no texture or work function changes were observed (green arrow identifies the

location). (d) Same heterostructure after a higher dose EBT (5.5keV), where texture and work function changes were

observed (red arrow), suggesting the EBT recipe determines whether the heterostructure is altered, as well as the rate

of change. More specifically, when treating Bi2Se3/MoSe2 for 3 minutes at ~19.7A·m-2, changes were observed at

accelerating voltages greater than 5.5keV (2.01x107 J·m-2), and no changes were observed at 5keV (1.78x107 J·m-2).

The time interval is an important parameter of EBT because it reveals the rate of structural

rearrangement and change. Nearly all the changes observed for the EBT experiments, both SAED

and SEM, were slow and steady enough to be observed by in real time, occurring over time

periods of several seconds to several minutes. Hence, in case of the SEM-EBT experiments, it was

possible to obtain a videographic representation of the changes; a movie of a Bi2Se3/MoSe2 2D

heterostructure undergoing SEM-EBT is provided at the following reference (Probing the

Page 79: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

79

Interlayer Interaction between Dissimilar 2D Heterostructures by In-Situ Rearrangement of their

Interface in 2D Materials), demonstrating the steady rate of change.

Figure 3.15 Texture and work function induced changes from EBT. Bi2Se3/MoSe2 affected with electron beam treatment

(EBT). Green lines show the area that was integrated over to produce the Brightness/Texture curve as a function of

distance (above the image). The area that was affected by EBT is the black-bordered box. It can be seen that the middle

looks “melted” with larger grains, compared to the undisturbed, as-grown areas outside the box. High-dose EBT

“smooths” the heterostructure’s texture, suggesting the Bi2Se3 grains increase in size by merging. This is reflected in

the curve where the EBT area is smoother and contains three broad peaks (red arrows), whereas the as-grown areas

(blue brackets) have more noise and contain only one broad peak (left side). These differences suggest EBT modifies

the structure. The EBT affected area has a diminished brightness compared to the as-grown area, suggesting EBT

increases the work function of the material. It has been shown that the work function can be measured by comparing

the relative brightness in an SEM image: Materials with a lower work function will emit more electrons and appear

brighter, while materials with higher work functions will eject less electrons, making them appear dimmer on the SEM

image.168,200 The effects of beam charging were mitigated by allowing the material to “rest” and discharge uninterrupted

for 20 minutes after the EBT. It should be noted that recipes with higher doses would occasionally create a black border

around the area affected, likely due to increased treatment time while the beam is “turning-around” and preparing to

move in the opposite direction. We believe that it is either the result of more intense material modification, carbon

build-up,201 or charging.202

Page 80: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

80

To understand the formation of the nanocrystals, we note that Bi2Se3 prefers to grow poly/nano-

crystalline on the SiO2 (Figure 3.4); however, it prefers to grow crystalline on the TMD, suggesting

the TMD provides a surface that stabilizes the Bi2Se3 through the interlayer interaction. We believe

that as the interlayer bonding it broken, the Bi2se3 becomes metastable and begins to breaks into

smaller grains, in agreement with the TEM-EBT experiments in Figure 3.7, Figure 3.8, Figure 3.9,

and Figure 3.10. Further, the quantity and relative brightness of the Bi2Se3 spots in those figures

suggests that high-dose EBT is able to break the Bi2Se3 into nano-scale particles. Previous work

demonstrated that Bi2Se3 particles tens of nanometers across have a PL,52,190 suggesting high-dose

EBT is able to disrupt the Bi2Se3 layer into forming nanoparticles that induce a different perceived

color (Figure 3.12a and Figure 3.13), work function (Figure 3.15), and PL spectra (Figure 3.12c and

Figure 3.14). Our results indicate that the mechanism for the grain rotation is that energy is

imparted to the grains via the electron beam, and once sufficient energy is imparted to overcome

bonding, the grains will rotate and move. Although this may appear to be thermal motion, we

are cautious to label it as such, since the interlayer interaction applies a different set of forces and

constraints than a typical “Brownian motion particle” would experience.

In conclusion for this section, mono- to few-layer Bi2Se3 can be grown epitaxial on a family of

TMDs (including an alloy), suggesting an influential interlayer interaction is present between the

2D materials. Comparing the angular distributions between the heterostructures provides insight

into their interlayer interaction, and the influence of certain parameters (e.g. twist angle, atomic

registry, chemistry). For example, when chemical periodicity is present (i.e. non-alloys), the most

stable twist angle is at 0°. We then demonstrated a new method to study and quantify the

interlayer interaction in-situ by varying the current density of focused electron beams in either a

TEM or SEM. While in SAED mode, the heterostructure could be disrupted using higher

densities, recrystallized using lower densities, or rotated in-situ into different twist angles using

values in between. EBT measurements indicate Bi2Se3/MoSe2 has the strongest interlayer

interaction and Bi2Se3/WS2 the least, results that are in agreement with the epitaxial growth study

(Figure 3). An SEM allows for greater control of the electron beam parameters (e.g. current density

Page 81: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

81

and accelerating voltage), enabling improved quantification of the EBT and subsequent changes.

Specific dosing thresholds were measured for when heterostructure disruption begins,

experimental results that can improve bonding strength estimates. Very interestingly, the recipe

can be tuned to change the optical properties, as well as induce the formation of Bi2Se3

nanoparticles, allowing the writing of submicron patterns that photoluminesce. This work

provides insight into the interlayer interaction in 2D heterostructures, an effect that has been

shown to dramatically transform the properties of 2D materials, but is not understood well due

to the limited number of experimental tools to probe it. This work demonstrates a facile technique

to manipulate the twist angle in-situ, which when combined with previously published TEM-

based techniques, is a platform to rapidly collect electronic structure information of 2D materials

facilitating statistically significant results, and supports the development of 2D heterostructures

whose properties are engineered using the twist angle. Lastly, the demonstration of submicron

pattern writing using SEM-EBT holds unique applications for information storage and photonics.

3.1 Monte Carlo Simulations Reveal Statistical Significance of Histograms

The following section explains in more detail the Monte Carlo simulations done to measure the

statistical significance of the histogram results in Figure 3.6. It can be seen that the instance

noise floor was found, which is the boundary established by the four-sigma criteria. When the

instance noise floor is one, the histogram is above the noise floor if one value is above it,

suggesting that the interlayer interaction is inducing a preference at that value. The two

instance noise floor means that when two bars are above that value, then the spectrum is above

the noise floor and the values are not random. The figure captions below describe the Monte

Carlo simulations, where additional information can be found in Chapter 2: Instrumentation

and Synthesis.

Page 82: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

82

Figure 3.16 Monte Carlo Simulation: Instance Noise Floor Values. Tuning the interlayer interaction through chemistry.

Histograms of the angular separation (Bi2Se3 vs. the TMD) in 2D heterostructures. (a) Se-Se interface with molybdenum.

(b) Se-S interface with molybdenum. (c) Se-S interface with tungsten. (d) Se-alloy interface with molybdenum. Bi2Se3

prefers to grow near 0° when chemical periodicity is present (i.e. non-alloy heterostructures), but prefers 12°-20° when

absent (i.e. MoSe2-2xS2x), suggesting chemical periodicity strengthens the interlayer interaction at 0°. The frequency at

0° decreases from (a)-(c) suggesting Se-Se interfaces and molybdenum-based TMDs form stronger interlayer

interactions. Monte Carlo simulations with a 4-sigma criteria reveal the data outside of 0° is statistically insignificant,

with the exception of 3.85% and 4.30% in (c) and (d), respectively. The relative positions of the interlayer atoms at 0°

might allow for stronger bonding.

Page 83: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

83

Figure 3.17 Monte Carlo Simulation data to identify the instance noise floors (INF) and statistical significance of the

histograms showing the as-grown 2D heterostructure twist angles. The table shows the values used for running Monte

Carlo simulations, as well as calculating the noise floor for each material. Simulations were run using the data points

outside of zero (right column); however, the results were divided by the total number (left column) to calculate the

percentage values shown in this chapter. To get the percentage we divided the number in each bin by the total number.

(a) Representative simulation and distribution from randomly placing 216 points into 151 bins. The bins represent the

range from 0°-30° at increments of 0.2°. This simulation represents a potential distribution if there was no interlayer

interaction to break the angular symmetry and induce angular preference. (b) 1 million simulations like that in (a) were

run, and this graph shows the simulation’s tallest bin vs. frequency. For example, the tallest bin was 9 for 2,287 of the

distributions.

Page 84: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

84

Figure 3.18 Data from the Monte Carlo simulations used to calculate the instance noise floor and statistical significance

of the histograms showing the as-grown 2D heterostructure twist angles. The instance noise floor (INF) sets a threshold

when data is statistically significant, assuming 4-sigma criteria. For example, 1 INF – 3.50% means individual peaks at

or below this percentage are expected to occur at least 63 times out of 1 million. Multiple INFs were calculated to verify

the validity of percentages that had numerous instances, such as 4.30% in (d) which is below the 1 INF, but has a

sufficiently high number of instances to make it statistically significant.

3.2 Material Transformation vs. Twist Angle Manipulation

In this section we address the question whether EBT is a material transformation, vice a twist

angle manipulation. If there is some kind of “Material Transformation” taking place during the

EBT, we ask ourselves, what is the nature of the transformation? Specifically, what physical

process could take a portion of the Bi2Se3 crystal (the part under the electron beam), transform it

into “something else”, and then transform it back into a Bi2Se3 crystal again, but with a different

orientation? Since SAED spots remain distinct before and after the EBT treatment, it is difficult to

assign this transformation to any “permanent” changes during the EBT, for example

amorphization. Also, any intermediate “phase transition” is unlikely as the “re-oriented” spots

represent identical lattice parameters after the EBT, and we never see any other crystal structure

no matter what the extent of EBT was. The other possibility to be considered is chemical

modification such as oxidation: This is unlikely since all the experiments were done at ultra-high

Page 85: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

85

vacuum and does not explain why the Bi2Se3 crystal reappears after EBT – since oxidation is a

thermodynamically-drive process that can go only in one direction under EBT.

The two possible candidates would be (a) if the EBT was somehow causing the crystal to non-

equilibrium non-crystalline phase, perhaps akin to 2D liquid like phase where the atoms re-orient

themselves randomly, but do not leave the zone, and then recrystallize when the EBT is turned

off, and (b) the sub-crystallite (under the electron beam) rotates as a whole, much like how a tiny

gold nanoparticle can rotate under EBT of a TEM.197 Both of these possibilities can allow the

recovery of the Bi2Se3 phase post EBT, and the material transformation as a whole could be a

combination of these two changes, different at different EBT treatment doses etc.

We do note, however, that the cumulative Bi2Se3 spot brightness was diminished, possibly

because it starts to become polycrystalline, reducing the efficiency with which constructive

interference can happen within each SAED spot. And finally, we also found that when sufficiently

high current densities are used, the 2D heterostructure appears to lose its highly crystalline

structure and break into nanocrystals, an indication that the sample is being structurally altered

and possibly degraded. With this in mind, we concur that at high current densities the 2D

heterostructures undergo crystallographic degradation.

3.3 Negligible Role of Potential Defects on the Interlayer Interaction

Chemical vapor deposition (CVD) has previously been shown to grow 2D crystals with more

defects than exfoliated samples, and that these defects can substantially influence the

properties.203,204 Further, defects at the surface would likely have a some impact on the interlayer

interaction because the chemistry would be different, for example grain boundaries could result

in the periodicity being interrupted. Hence, it is important to address what impact defects might

have on the system. Although we are not able to provide the defect density generated in either

the monolayer TMDs or the Bi2Se3, we are able to provide insight into the crystal quality and

consistency of the samples, suggesting that our observations are most likely dominated by the

chemistry of the materials and not that of the defects.

Page 86: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

86

Crystal Quality: Based on SAED, TEM images, Raman spectroscopy, and photoluminescence

measurements, the quality of our crystals is comparable to published results, suggesting a low

density of defects, and that a majority of the interlayer interaction is mediated by the non-defect

ordering. Previous work on 2D materials has shown there is a relationship between the defect

density and the Raman response. The Raman response of our samples (see representative spectra

below) is in agreement with that of numerous exfoliated samples, and published work studying

this question, suggesting we have a defect density less than LD=24nm (i.e. one defect every

24nm).205,206

Additionally, previous publications from our group have completed a thorough characterization

of the different monolayer TMDs grown in our laboratory, where the results suggest there are

minimal defects in our samples.30,31,103,104,187 We are proud to report we were able to observe very

subtle effects that are known to be sensitive to the sample quality. For example, the Raman spectra

observed for monolayer MoSe2 matched strongly with the density functional theory (DFT)

predictions.30 In another paper, high-resolution TEM imaging is able to observe the moiré pattern

and identify atomic number in MoS2/MoSe2 2D heterostructures.187

Figure 3.19: Representative Raman spectra of a Bi2Se3/MoS2 2D heterostructure , where all the peaks have a strong

intensity and can be clearly identified, suggesting that the defect density is low. This Raman spectra can also be seen

in an upcoming work in ACS Applied Materials & Interfaces.

Page 87: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

87

Growths Yield Reproducible Samples: The 2D crystals grown in this lab are relatively

reproducible and consistent, suggesting a majority of the samples formed equivalent interlayer

interactions. If this is the case, the primary variable would be the electron beam, suggesting the

conclusions for the effect of an electron beam are reproducible.

Hence, although it was not possible to measure or identify the nature of defects at the interface,

Raman spectroscopy, TEM imaging, SAED, and PL suggest the defect density is low. Further, the

samples are reproducible, suggesting the interlayer interaction was equivalent across the 2D

heterostructures probed. This is important because it means, although the defect density cannot

be probed, its effect on the 2D heterostructures is at least consistent.

And finally, if we assume that the defect density remains consistent for all the samples, then by

comparing the heterostructures with different nearest and/or next-nearest neighbors

automatically eliminates the possibility that defects are playing a major role in the observations.

This is perhaps most strongly demonstrated when we observe noticeable differences even when

the next-nearest neighbor layer is changed (i.e. the interface along with defects remain the same).

From this, we conclude that the role of defects is not dominant in our observation.

3.4 Discussion Grain Rotation Mechanisms

Our results indicate that the mechanism for the grain rotation is that energy is imparted to the

grains via the electron beam, and once sufficient energy is imparted to overcome bonding, the

grains will rotate and move. Although this may appear to be thermal motion, we are cautious to

label it as such, since the interlayer interaction applies a different set of forces and constraints

than a typical “Brownian motion particle” would experience.

As shown in Figure 6, the current density determines the response of the 2D heterostructures. For

example, Bi2Se3/WS2 appears to have the weakest interlayer interaction, while Bi2Se3/MoSe2 the

strongest. Applying a certain current density will have differing effects. Further, it can be seen

that each material has a range where the 2D heterostructure is neither annealed nor disrupted,

Page 88: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

88

but is twisted. This is because the powers are sufficient to overcome the interlayer bonding, but

not intense enough to overcome the intra-layer bonding.

We note that as the crystal is made smaller its properties might change. For example,

photoluminescence dependent on the size of Bi2Se3 nanoparticles has been reported. 52,190 This is

a possible explanation for the change in color and properties, suggesting that the interlayer

coupling may play only a partial role, or no role at all.

Page 89: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

89

Chapter 4: Oxygen-Induced In-Situ Manipulation of the

Interlayer Coupling and Exciton Recombination in

Bi2Se3/MoS2 2D Heterostructures

Research in 2D heterostructures continues to gain enormous interest for their potential to advance

both fundamental and application-oriented research. They have contributed to fields as diverse

as transistors,207–209 optoelectronics,210–212 information storage,103,213,214 plasmonics,215–217

photocatalysis,218–220 capacitors,221,222 biosensors,109 spintronics,113,223 high-density lithium

storage,81,82 and superconductivity.66,67,144 The far-reaching success of 2D heterostructures is in part

due to the large spectrum of properties they have demonstrated. A 2D heterostructure’s

properties are more than a sum of the parent 2D materials, but are also a product of the interlayer

interaction, which can be manipulated to engineer new capabilities. It has been shown that the

twist angle,224 interlayer spacing,144 thermal annealing,148,150,225 and intercalation of

molecules127,129,135 all influence the interlayer coupling. There are no tools to directly probe the

interlayer coupling strength; however, its impact can be inferred by the extent it influences the

properties. In this paper we demonstrate that the interlayer coupling in Bi2Se3/MoS2 2D

heterostructures can be tuned by regulating the oxygen presence in the atmosphere, while

applying controlled laser or thermal energy doses. The coupling strength is inferred using

changes in the photoluminescence (PL) intensity, where lower PL corresponds to higher coupling.

Our data suggests the interlayer coupling is modulated by diffusively intercalating and de-

intercalating oxygen molecules. The work shown in this chapter can be found at the following

reference.226

Page 90: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

90

Figure 4.1: Figure briefly summarizing the capabilities demonstrated in chapter 4. Note that oxygen is likely

intercalating between the layers to induce the change in photoluminescence, a finding not shown pictorially above.

The interlayer coupling in 2D materials is considered a promising parameter for designing

materials with tailored properties; however, despite the significant interest, the interlayer

coupling is not well understood, which is in part due to the lack of experimental techniques with

the ability to precisely manipulate it. Some previous experimental work has focused on using

global thermal annealing to manipulate the coupling; however, this method has not

demonstrated precision, and often requires several hours to complete.148,150,225 Ion irradiation,227

in-plane strain,151 and lateral pressure144,146 have all been shown to increase the interlayer coupling

by decreasing the interlayer separation; however, ion irradiation has also been shown to damage

2D materials and induce defects228–230, and applying strain or pressure alters the 2D material’s

lattice parameters and properties,231–233 thereby introducing uncertainty. Work to manipulate the

interlayer coupling via electric fields has only been theoretical, and follow-up experimental work

is needed to confirm it.234–237 In this paper we demonstrate a facile method to tune the interlayer

coupling in-situ with high-spatial resolution, all using ambient conditions and tools commonly

found in a 2D research laboratory.

Monolayer MoS2 is known, among other things, for having a bright PL due to the formation of

tightly-bound excitons.33 However, when only one layer of Bi2Se3 is grown on the monolayer MoS2

using vapor-phase deposition, the bright PL is >99% quenched because the interlayer coupling

induces a non-radiative exciton recombination pathway.103 In this paper we demonstrate that the

Page 91: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

91

PL can be controllably increased and decreased at small increments of only a few percent,

suggesting that the coupling is being diminished or strengthened, respectively, at the same small

increments. This is done by applying energy to the material in either an oxygen-present or

oxygen-absent environment, which we believe facilitates the diffusive intercalation or de-

intercalation of oxygen molecules. Our results suggest that the diffusive intercalation of oxygen

molecules induces the monolayer MoS2 to behave more electronically independent, thereby

restoring the radiative recombination pathway and facilitating the emergence of the signature PL

peak.

The intercalation of molecules between layers has been shown to disrupt their coupling, leading

to change in the properties.81,127,129,133–135,142,238–241 For example, it has been shown that when several

layer MoS2 is intercalated with lithium or quaternary ammonium molecules, the signature

monolayer photoluminescence emerges, suggesting the intercalated molecules induce the

MoS2 layers to behave electronically independent, as if they were monolayer.107,129,238 It has also

been shown that oxygen can easily intercalate between 2D crystals and their substrates,

decoupling the two materials and inducing them to behave more “freestanding” (i.e.

electronically independent).127,133–135,142,239,240

A natural question arises as to why oxygen, vice other molecules in the atmosphere, is

manipulating the interlayer coupling in Bi2Se3/MoS2 2D heterostructure. Bi2Se3 is well-known for

its oxygen-affinity, where even trace amounts of oxygen dramatically affect its exotic topological

properties.60 Not only can oxygen react with Bi2Se3 to form a native oxide and new bonding,55,59

but it has even been shown that oxygen is able to passively diffuse through multiple layers of

Bi2Se3.242 The interaction of Bi2Se3 and oxygen is complex and has been well-studied. A thorough

literature review and how previous work relates to our findings is provided later.

In this work, we obtain highly controllable, reversible, and site-selectable switching between

direct (i.e. radiative and photoluminescent) and indirect (i.e. non-radiative) exciton-

recombination pathways in Bi2Se3/MoS2 2D heterostructures (1-3 layers of Bi2Se3 grown on

monolayer MoS2). Our approach allows permanent, in-situ, electrode-less, and use-specific

Page 92: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

92

programming of the interlayer coupling and PL intensity. Exciton recombination dynamics could

be switched (between radiative and non-radiative) by thermally treating the heterostructures in

oxygen-present (e.g. air) vs. inert atmospheres (i.e. Ar or N2). Alternately, the switching could be

localized in a highly controllable manner at ambient temperatures using a continuous wave

488nm laser (as before, in air or under N2), which allowed site-selective reversible manipulation

of different regions of the same 2D heterostructures. Additionally, our results suggest that these

2D heterostructures might have applications as standard temperature-pressure high-density

oxygen storage devices, potentially storing 69 kg/m3 (a factor of 52 times the density of O2 gas at

1 atm).

Figure 4.2: As-grown Bi2Se3/MoS2 vertical 2D heterostructures. (a) Optical image of a Bi2Se3/MoS2 vertical 2D

heterostructure, where 1 layer of Bi2Se3 was uniformly grown on a monolayer MoS2 crystal using vapor-phase

deposition. (b) Side-view diagram of the heterostructure. (c) Representative atomic force microscope (AFM) scan of a

Bi2Se3/MoS2 2D heterostructure, demonstrating monolayer Bi2Se3 (with trilayer islands) grew uniformly across the

entire MoS2 crystal. (e) Typical drain current (IDS) vs. gate voltage (VG¬¬) in back-gated bare MoS2 and Bi2Se3/MoS2

devices, respectively. The bare MoS2 device is n-doped into the conduction band, while the right-shifted data-curve of

the heterostructure suggests reduction of n-type carriers and lowering of the Fermi level into the band gap. (f)-(g)

Bi2Se3/MoS2 2D heterostructure with the lattice parameters labeled of both parent crystals. The distinct dots indicate

both materials are highly crystalline. The Bi2Se3 tends to grow crystallographically aligned (i.e. twist angle is 0°). (b) A

laser was applied while in an oxygen-present environment (i.e. air), and it was found to induce the Bi2Se3 to become

nano- and poly-crystalline. As discussed later, at higher energy doses the system undergoes more robust changes that

Page 93: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

93

are not easily reversed. This is possibly due to the Bi2Se3 becoming nano-crystalline, which creates more grain

boundaries. Oxygen is known to more easily diffuse through grain boundaries (discussed thoroughly later), which

would explain the variation in the PL’s rate-of-change.

Figure 4.2a shows an optical image of Bi2Se3/MoS2 2D heterostructure (1 layer MoS2 + 1 layer

Bi2Se3) on SiO2, where the monolayer MoS2 was grown using vapor-phase chalcogenization

(VPC)31 and the Bi2Se3 was grown on top using vapor-phase deposition. Figure 4.2b is a side-view

diagram of a typical Bi2Se3/MoS2 2D heterostructure. Despite the huge lattice mismatch (e.g. 2.74Å

to 3.57Å, see Figure 4.2f), uniform layers of Bi2Se3 grow with high regularity on top of the MoS2

crystal, suggesting strong van der Waals epitaxy-mediated growth between the two component

layers.243 Figure 4.2e shows the transfer characteristics of back-gated monolayer MoS2 and

Bi2Se3/MoS2 FET devices. The right-shift of the threshold gate voltage in the 2D heterostructure

indicates a relative downshift of the effective Fermi level due to reduction of excess n-type carriers

(by about ~9.50x1012 cm-2). The device response looks neither like monolayer MoS2, nor few layer

Bi2Se3 (a semi-metal),244 suggesting there is sufficient interlayer coupling to modify the electronic

structure of both materials.

Figure 4.3: Manipulating the interlayer coupling using oxygen and energy. (a) Optical image of a Bi2Se3/MoS2 vertical

2D heterostructure where energy was applied locally using a focused laser, while the sample was in an oxygen-present

Page 94: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

94

environment (in air). The optical properties of the affected area were altered going from purple to white, allowing

affected locations to be easily identified. (b) PL spectra from the same spot before and after energy was applied,

demonstrating how the interlayer coupling can be manipulated. Initially, the PL spectra was flat; however, after

applying a focused laser for 8 minutes at 168μW, a PL spectra corresponding to monolayer MoS2 appeared, suggesting

the interlayer coupling was disrupted, allowing the MoS2 layer became more electronically independent. Electronically

independent monolayer MoS2 has a signature PL peak due to its tightly bound excitons. (c) The laser was now applied

to a different location (orange arrow) on the same sample at the same power (168μW for 12 minutes), but the

environment was changed from oxygen to nitrogen. Interestingly, the color change seen in (a) was not observed. The

inset shows the setup where N2 gas was flown across the sample to displace the oxygen. (d) PL spectra from before

and after energy was applied show no perceivable appearance of PL, suggesting the interlayer coupling was not

affected. Thermal treatment experiments discussed later demonstrate applying heat in an oxygen environment disrupts

the interlayer interaction, and that the other components in air (e.g. N2, H20, CO2) do not appear to affect the interlayer

coupling. (e) PL spectra of a Bi2Se3/MoS2 2D heterostructure before energy was applied, as well as after several

consecutive doses, demonstrating how the PL intensity is dependent on the total dose.

Figure 4.3a is an optical image of a Bi2Se3/MoS2 vertical 2D heterostructure, where energy was

applied locally using a focused laser (168μW for 8 minutes), while the sample was in an oxygen-

present environment (i.e. air). The red arrow identifies the affected location, which underwent a

change in color from purple to white. Figure 4.3b shows its PL spectra before and after the energy

dose was applied. Characteristic B2Se3/MoS2 2D heterostructures do not have a PL, or only a very

weak PL, because the interlayer coupling creates a non-radiative recombination path for the

tightly-bound excitons that exist in the monolayer MoS2.103 However, if energy is applied to the

heterostructure in the form of either thermal treatment or a laser, while the heterostructure is in

an O2-present environment, a PL spectrum emerges that corresponds to that of monolayer MoS2,

suggesting the interlayer coupling was disrupted and the signature MoS2 excitons are

recombining along radiative pathways. Further, Raman spectroscopy measurements show the

intensity of the monolayer MoS2 modes increasing after laser exposure in air, suggesting the

interlayer coupling has been diminished (see Figure 5.10, Figure 5.11, and Figure 4.16). Previous

work has demonstrated a correlation between the intensity of the Raman modes and the

interlayer coupling strength.103,123,126,140

Page 95: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

95

Figure 4.4: Atomic force microscope (AFM) images of 2D heterostructure thermally treated in air. The images show

that thermally treating in air does not induce obvious changes to the Bi2Se3 topography or step height despite the fact

that there are large optical changes. A Bi2Se3/MoSe2-2xS2x 2D heterostructure was used above.

Figure 4.5: AFM analysis of 2D heterostructure: as-grown vs. laser exposed in air. Laser exposure in air does not remove

the Bi2Se3, a key finding demonstrating the exciton recombination pathway switching is more subtle and not simply

Page 96: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

96

because the Bi2Se3 is being removed. Interestingly, the effect of the laser exposure on the topography is difficult to

discern, where the pre- and post-air-laser-treatment AFM images look nearly identical. No clear changes were detected,

despite the fact that the optical and photoluminescence properties are drastically altered. A Bi2Se3/MoSe2-2xS2x 2D

heterostructure was used above.

Atomic force microscope (AFM) measurements shown in Figure 4.4 and Figure 4.5 demonstrate

that the Bi2Se3 remains on the MoS2 after a color change has been induced and the PL remerges,

verifying that the change is not due simply to the removal of Bi2Se3. A different spot on the same

sample (Figure 4.3c, yellow arrow) was exposed to the same power (168μW for 12 minutes), while

being continuously purged with N2 gas (inset Figure 4.3c), thereby removing oxygen from the

vicinity of the exposed spot. We find that even with an increased laser dose (12 min. vs. 8 min.

exposure), there was neither a perceivable color change, nor emergence of the PL peak (Figure

4.3d), suggesting the interlayer coupling was not affected. Figure 4.3e demonstrates how the PL

intensity is dependent on the energy dose applied, suggesting that the interlayer coupling can be

manipulated in small increments.

Figure 4.6: Oxygen-induced manipulation of exciton dynamics using laser setup. (a) Optical image of a Bi2Se3/MoS2

heterostructure. The arrow indicates a spot whose color changed after being air-laser-exposed. (b) PL spectra from the

Page 97: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

97

same spot before and after the air-laser-exposure, demonstrating the predicted behavior (i.e. PL recovery). (c) A

different location (orange arrow) on the same sample was N2-laser-exposed by flooding the environment with N2 gas

flow and displacing the air. N2-laser-exposure inhibits the color change observed in (a). Inset shows the setup used. (d)

PL spectra from before and after N2-laser-exposure, showing no perceivable appearance of PL after N2-laser-exposure.

Laser exposure and thermal treatment experiments were separately performed in synthetic dry O2+N2 mixture and

moisture-saturate N2 (shown later) from which any role of the other components of air, e.g. N2, H20, CO2 could be

eliminated. (e) The same spot in (c) was air-laser-exposed. The color-change is perceivable, along with the appearance

of the PL spectrum, shown in (f), demonstrating that extended N2-laser-exposure had no obvious deleterious effect on

the switching of the exciton recombination pathways.

After demonstrating that removing the partial pressure of oxygen disrupts change, the

atmosphere was switched back to air without changing the laser power or moving the sample -

thereby re-introducing oxygen without changing the specific location being probed - resulting in

a color change and a strong PL peak recovery (shown in Figure 4.6), confirming that the

atmosphere plays a critical role in the observed changes. Below are numerous experiments that

unequivocally demonstrate that oxygen is required to induce the changes, and not a different gas

in the atmosphere.

Page 98: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

98

Figure 4.7: 2D heterostructures thermally treated in Ar (i.e. an O2-free environment) at 240°C for 3 hours. 1-3 layers of

Bi2Se3 were grown on monolayer MoS2 using chemical vapor deposition (CVD). The photoluminescence was not only

quenched after each thermal treatment session, but it was also controllably increased using laser exposure in air,

demsonstrating write-read-erase-reuse capabilities.

Page 99: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

99

Figure 4.8: 2D heterostructure thermally treated in dry air (i.e. 79% N2 + 21% O2). The behavior is similar to that of the

air-annealing experiments in that the PL intensity increases when dry-air-annealed and then decreases again after N2-

annealing. However, it is different in two subtle points: (1) after dry-air-annealing, air-laser-treatment does not affect

the PL; and (2) a possible new PL signal emerges in the upper energies (~2.0eV), where the PL plateaus to a higher

intensity than that seen at 1.7eV (i.e. the PL spectra is not symmetric, showing higher intensities at higher energies).

Page 100: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

100

Figure 4.9: 2D heterostructure thermally treated in an N2+H2Ovapor environment. This experiment established that water

vapor does not react with the heterostructures to induce the optical and photoluminescence changes observed during

laser exposure in air (i.e. water vapor is not necessary to switch exciton recombination pathway). In fact, the

heterostructure will reverse both the air-laser-treatment induced optical and photoluminescence changes, when

annealed in an N2 environment saturated with water vapor. As seen above, five locations were air-laser-treated prior

to annealing, and an optical change was induced in each on. Annealing reversed the optical change in all five spots.

The two larger spots seen in the right two optical images were induced post-anneal, demonstrating it was still possible

to switch the recombination pathway. A Bi2Se3/MoSe2-2xS2x 2D heterostructure was used above.

Page 101: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

101

Figure 4.10: 2D heterostructures first thermally treated in air, and then in N2. Photoluminescence spectra at different

locations. The heterostructures are monolayer MoSe2-2xS2x with 1-2 layers of Bi2Se3 CVD grown on top. (Top Optical

Images) Note how the air-annealed heterostructures appear a lighter shade of purple and have less contrast, compared

to the as-grown heterostructures. The heterostructures were then N2-annealed, resulting in the air-annealed optical

changes being partially reversed (i.e. there is a darker shade of purple and more contrast). (Bottom PL spectra) Left to

right are the as-grown, air-thermally-treated, and N2-thermally-treated (following air-thermal-treatment), respectively.

Each graph has the pre- and post-air-laser-exposed PL spectra. The data between annealing sessions was taken at

different locations (specified by the red arrows). Air-annealed samples will have a brighter PL than the as-grown

samples, but it will be weaker than a laser-treated as-grown heterostructure. The PL intensity of air-annealed

heterostructures will decrease when air-laser-treated. A Bi2Se3/MoSe2-2xS2x 2D heterostructure was used above.

Page 102: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

102

Figure 4.11: 2D heterostructures thermally treated in air and then in N2. PL spectra at the same location. The

heterostructures are monolayer MoSe2-2xS2x with 1-2 layers of Bi2Se3 CVD grown on top. (Bottom PL spectra) The data

between annealing sessions was taken at the same location. The air-laser-exposed location maintained approximately

the same intensity; however, the peak blue-shifted. The PL intensity of air-thermally-treated heterostructures will

decrease when air-laser-treated, which is the same behavior observed when recipes with very high powers and energies

are used. After N2-thermal-treatment, the PL will be quenched, and air-laser-exposure induces the PL upward, which

is the same behavior observed in as-grown heterostructures. A Bi2Se3/MoSe2-2xS2x 2D heterostructure was used above.

Detailed investigations in different gas environments shown in Figure 4.7, Figure 4.8, Figure 4.9,

Figure 4.10 and Figure 4.11 establish that oxygen (O2), and not nitrogen, H2O vapor, or carbon

dioxide, is required to induce the changes and manipulate the interlayer coupling. Thermal

treatment at 240°C was able to modulate the interlayer coupling, suggesting that the mechanism

actuates when sufficient heat is present, and illumination with a particular wavelength is not

Page 103: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

103

required. Later in the paper we demonstrate that the interlayer coupling can be tuned with high

precision by controlling both the energy applied and the partial pressure of oxygen in the

surrounding atmosphere.

Photo-excited e-h pairs in mono-layered transition metal dichalcogenides (TMDs) form tightly-

bound neutral and charged excitons.33 In direct-gap monolayer TMDs, they recombine

radiatively, producing well-known PL spectra. In few-layered and thicker TMDs the quasiparticle

band gaps are indirect, hence the KΓ indirect (non-radiative) recombination pathway becomes

more favorable, resulting in progressively suppressed PL.245 Indirect recombination in certain

heterostructures can similarly be non-radiative, if the excitons formed near a k

-vector in the

reciprocal lattice of one layer finds the most favorable recombination pathway via a lower-

energy-state that is located at a different k

-point in the reciprocal lattice of the second layer. In

all the as-grown Bi2Se3/MoS2 2D heterostructures, the PL spectra were strongly quenched,

suggesting the interlayer coupling was inducing non-radiative recombination pathways for the

excitons. Upon application of energy (either heat or laser) in an oxygen-present environment, all

heterostructures recovered their radiative recombination pathways, suggesting the interlayer

coupling was disrupted, thereby inducing the MoS2 layer to behave more electronically

independent. Next we show that the PL can be quenched again by applying energy (via either

thermal treatment or laser exposure) in an O2-free environment, demonstrating that the effects

could be reversed and the interlayer coupling restored.

Page 104: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

104

Figure 4.12: Tuning the interlayer coupling and photoluminescence intensity. (a) Variation of PL intensity under

alternating air and nitrogen environments while a focused laser (i.e. energy) is applied, demonstrating that a nitrogen

environment is able to restore the interlayer coupling. The solid black arrow shows how the baseline PL reading

monotonically grows after repeated environment cycling, and the PL’s rate-of-change varies, suggesting the 2D

heterostructure may not return to its initial (as-grown) state. The dashed blue arrow shows the overall PL intensity

growth factor, up to 215×, achievable by this cycling approach. (b) Correlation between laser (i.e. energy) exposure (in

air) and the resulting PL intensity of a Bi2Se3/MoS2 2D heterostructure using a calibrated recipe (50 µW, 6 s doses

followed by 1 µW, 60 s for collecting data), demonstrating a controlled disruption of interlayer coupling. Under this

dose recipe, the radiative recombination was found to grow approximately logarithmically, and a vast majority of

exposures result in a <5% change in intensity (see inset). See Figure 4.13 for other possible behaviors when the recipe

is changed. Inset: histogram of the change steps under each exposure. Both (a) and (b) taken together demonstrate an

unprecedented degree of controlled manipulation of the interlayer coupling and PL achievable. Later we demonstrate

this is possible in a site-selectable manner as well.

Remarkably, not only are we able to disrupt the interlayer coupling and increase the PL intensity,

but we can also restore the interlayer coupling and decrease the PL intensity, which we believe is

due to the intercalation and de-intercalation of oxygen. Figure 4.12a shows the change in PL

intensity of a Bi2Se3/MoS2 2D heterostructure that is oxygenated and deoxygenated several cycles

by switching the ambient atmosphere between air and nitrogen. This controllable switching of

radiative and non-radiative exciton recombination pathways, demonstrates the ability to both

disrupt and restore the interlayer coupling. This process can be cycled several times,

underscoring the fact that at the initial stage, under identical laser power, the oxygenation is

reversed by the mere removal of the O2 partial pressure in the ambient. This indicates that at least

initially, the oxygenation process is diffusive and does not form chemical bonds. After a few

cycles, the maximum PL intensity grows by as much as 215× and stops quenching fully,

Page 105: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

105

suggesting that other more permanent changes occur at higher energy dosage (discussed later in

the paper).

Figure 4.12b shows the variation of PL intensity measured after repeated doses (t=6s) of combined

exposure to air (at ambient pressure) and laser power (at 50 µW). At this dose-value, the PL was

found to grow approximately logarithmically with combined exposure (along with incident

energy), suggesting the interlayer coupling can be tuned with high precision. The logarithmic

shape is in agreement with Fick’s law of diffusion, which states the diffusion flux will decrease

as the system reaches equilibrium, and whose general solution has an exponential form.

Additionally, Fick’s law has been used to describe the intercalation process, including the

intercalation of lithium into vertically stacked bilayer graphene.82 The inset shows that the

intensity-change (IPL) can be as low as~5%, suggesting possible applications as low-cost

atomically-thin laser calorimeters or photon-counters. Figure 4.13 demonstrates other possible

behaviors when the recipe is changed.

Figure 4.13: High Tunability and control of interlayer coupling and PL intensity. Rate of change of PL is dependent on

the recipe of energy applied. A Bi2Se3/MoS2 crystal was air-laser-treated using five different recipes, where the power

was held constant power, but time was changed. Different locations on the same Bi2Se3/MoS2 crystal were studied,

Page 106: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

106

allowing the results to be better compared. The results show that 6 sec. at 48.6µW produced the most consistent results.

The difference in curve shapes is due to competing factors: too much air-laser-treatment results in degradation of the

heterostructure; however, too little power does not induce sufficient change to be permanent. Additionally, using very

low powers (~0.91µW) causes the PL intensity to decline. I believe this is because the powers perturb the

heterostructure into a better coupling, which squeezes the intercalated oxygen out. It has been shown previously that

changing the interlayer coupling will change the rate of diffusion of oxygen intercalating between graphene and the

surface (discussed more in detail later).

Figure 4.13 demonstrates that changing the recipe changes the response of the Bi2Se3/TMD 2D

heterostructure. The power was held constant while the time was exposure time changed. The

same 2D heterostructure crystal was used, where only different locations on the same crystal were

probed, thereby mitigating different samples as a variable. The red arrows show the different

locations that were probed, where the laser and oxygen induced color changes can be seen. The

lack of clear logarithmic behavior when time intervals outside of six seconds were used, suggests

that the one minute data collection period began to partially reverse the effects and disrupt the

predictable logarithmic behavior.

Figure 4.14: DFT calculations predict intercalated O2 diminishes interlayer coupling (a) DFT calculations of a

rotationally aligned (i.e. twist angle is 0°) Bi2Se3/MoS2 superlattice predict significant charge redistribution into the

interlayer region, and an influential interlayer coupling. This is discussed further in Chapter 6. (b) However, when O2

molecules are placed in the interface between the layers, DFT calculations predict the average interlayer separation

increases from 3.57Å to 4.18Å (17% increase), diminishing the interlayer coupling. These results are in agreement with

previous studies where intercalated O2 between a 2D material and the substrate induces the 2D material to behave

electronically independent (i.e. “freestanding”). Further, intercalation and de-intercalation are diffusion-dependent

processes, offering an explanation for why the PL intensity rises and falls so quickly as the environment changes (Figure

4.12). DFT calculations predict an interlayer separation ~4.7 times larger than the diameter of an oxygen atom,

suggesting sufficient space exists for O2 molecules to reside. The above structure was calculated by students in Prof.

Arun Bansil’s group and reproduced with their permission.

Page 107: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

107

The rapid changes observed in Figure 4.12 by simply switching the environment between air (i.e.

O2-present) and nitrogen (i.e. O2-absent) suggest that oxygen is diffusing in and out of the 2D

heterostructure. As seen in Figure 4.14, the O2 molecule is relatively small compared to the

interlayer spacing. Further, it has been shown that O2 molecules are able to permeate into Bi2Se3,

demonstrating they are able to fit between the atoms.55,59,242 Density functional theory (DFT)

calculations predict that the interlayer coupling in a Bi2Se3/MoS2 2D heterostructure induces

significant charge redistribution in the interlayer region, hybridizing nearest neighbor atoms to

form bonds, and that it is influential in modifying the bandstructure.103 Interestingly, when O2

molecules are placed in the interlayer region, DFT calculations predict that the materials will

begin to separate, diminishing the interlayer coupling, and inducing the layers to behave more

electronically independent. It can be seen in Figure 4.14 that the interlayer spacing increases on

average by 17% when five O2 molecules are intercalated. Together these results hint at the

interesting possibility that these atomically-thin layers may be excellent candidates for oxygen

storage devices, potentially storing 69 kg/m3 (a factor of 52 times the density of O2 gas at 1 atm).

Figure 4.15: Site-selective manipulation of interlayer coupling and photoluminescence. (a) Schematic of a method to

write patterns with site-selected interlayer coupling strength on Bi2Se3/TMD 2D heterostructures using a focused laser.

(b) Optical image of a laser-written pattern on a Bi2Se3/MoS2 2D heterostructure. The letters “NEU” were “drawn”

using different exposure times. (c) Fluorescence microscope image of the same sample (excitation λ=488 nm). We note

Page 108: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

108

that using a focused laser beam, excitons of selected regions could be programmed to recombine radiatively (bright

regions) or non-radiatively (dark regions), where the size of the affected area is dependent on the laser spot geometery

and recipe used (i.e. the power and exposure duration). The smallest “radiative” regions were below a micron in

diameter. This method allows for the rapid manipulation and measurement of the interlayer coupling with high spatial

resolution, facilitating experiments that produce statistically significant results.

Figure 4.15 demonstrates a low-cost method to rapidly and precisely manipulate the interlayer

coupling of Bi2Se3/MoS2 2D heterostructures with submicron spatial resolution. Figure 4.15a is a

schematic showing how a focused laser beam can be used to apply calibrated energy doses to

precisely manipulate the interlayer coupling with high (submicron) spatial resolution. Figure

4.15b shows the optical image where the method in Figure 4.15a was used to trace the letters

“NEU” using different exposure times, demonstrating a facile method to manipulate the

interlayer coupling. The laser-written sample was then imaged using a fluorescence microscope

(Figure 5c, λex=488 nm), demonstrating the ability to rapidly measure the degree of interlayer

coupling disruption. The ability to rapidly write and read patterns with varying interlayer

coupling strength, facilitates experiments that produce statistically significant results. The

fluorescence image also demonstrates site-selective light emission with a spatial size close to that

of the incident laser spot, enabling photoluminescing pixels (PLPs) tailored down to sub-micron

diameters. These site-programmable, color-selectable, atomically-thin, micron-scale PLPs (with

effective volumes ~10-21 m3) are attractive for optical and optoelectronic applications that require

ultra-small form-factors.

We now discuss the various possible mechanisms that could lead to the observed phenomena.

As mentioned earlier, we tested the effect of all the primary gaseous species present in air and

found it to be limited to oxygen (Figure 4.7, Figure 4.8, Figure 4.9, Figure 4.10, and Figure 4.11),

suggesting only oxygen is able to intercalate. While attempting to understand why oxygen would

have such a dramatic effect, vice the other gases, we found an large amount literature that

discusses the impact of oxygen on Bi2Se3 and other materials in its class (e.g. Bi2Te3), suggesting,

under different circumstances, both diffusive behavior as well as bond-forming mechanisms.55,58–

Page 109: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

109

60,242,246–249 In sharp contrast, we were unable to find any literature that demonstrates evidence

of absorption or chemical reaction with nitrogen, argon, or carbon dioxide, suggesting these

molecules are significantly less reactive/interactive with Bi2Se3. The collective body of work

appears to indicate the energetics facilitate these interactions with oxygen, vice the other gases.

From this, we conclude that interaction of oxygen, mediated by the bismuth selenide layer is the

most likely manner the changes are initiated. Further, oxygen has been shown to easily intercalate

between 2D crystals and their substrates, decoupling the two materials, and inducing them to

behave more “freestanding” (i.e. electronically independent).127,133–135,140–142,239,240 Indeed, our DFT

calculations of a Bi2Se3/MoS2 2D heterostructure predict that intercalated O2 molecules force the

materials apart, enlarging the interlayer separation by 17%, thereby diminishing their interlayer

hybridization (see DFT calculations later in the chapter - Figure 4.18). This is further supported

by our Raman measurements, which show that laser exposure in air increases the intensity of the

monolayer MoS2 modes (see Figure 5.10, Figure 5.11, and Figure 4.16).

Figure 4.16: Raman response of Bi2Se3/MoS2 pre and post laser exposure. The Bi2Se3 peaks decrease, suggesting the

modes are being suppressed. This is possibly related to the observation that laser exposure in air breaks the Bi2Se3

into nanocrystals, thereby diminishing its crystallinity. The MoS2 peaks, very interestingly, increase, suggesting that

the coupling has been diminished, in agreement with previous observations.103 Amplifying information and Raman

studies of other 2D heterostructures can be found in Chapter 3, Figure 5.10 and Figure 5.11.

Page 110: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

110

Previous work has shown that the intercalation of oxygen can diminish the interlayer coupling

and modify the Raman modes.140 Furthermore, intercalation is a diffusive process that has been

shown to be reversible,134,135,142,240 which could explain why laser exposure in an O2-free

environment is able to rapidly reverse the changes. Most notably, though, despite the large

change in the observed return of the MoS2 PL, the Raman peaks for both bismuth selenide and

molybdenum disulfide are near-identically retained before and after the laser treatment,

suggesting very little structural, electronic or chemical changes to each of these layers,

individually. For this reason, physical separation of the two layers due to diffusion of oxygen into

the interlayer region appears to be a very reasonable underlying mechanism.

If in fact oxygen is diffusing into the interlayer, what path(s) does it take? A number of past works

have explored the intercalation of atoms in micron-scale graphene and other 2D materials, where

it has shown to be highly dependent on the number of grain boundaries.81,82,127,134–139,142,250,251

More specifically, a single grain boundary of these systems was sufficient to enable large-scale

intercalation in micron-scale samples, strongly suggesting that, a high concentration of

edges/grain boundaries under the laser spot might not be needed for the observed rapid

intercalation/deintercalation. Bi2Se3 grows highly crystalline on the underlying MoS2, and when

the 2D heterostructure is then exposed to a laser in air, the Bi2Se3 becomes polycrystalline (Figure

4.2), suggesting the formation of grain boundaries. Together these results suggest (1) Intercalation

via grain boundaries is well-known, and do not necessarily require a high density of grain

boundaries to produce observable effects, and (2), these intercalations result in a de-coupling of

the parent layers electronically. While that is not a direct evidence that intercalation of oxygen is

taking place, in our case, these are strong indicators of the most possible mechanism (more

detailed explanation is below).

Other possibilities include the formation of chemical bonds, trapped oxygen at edge sites, or new

materials, however, these are less likely, at least at the initial stages of our experiments. It is well-

established that Bi2Se3 is sensitive to oxygen, including the partial pressure of oxygen in air.55,58–

Page 111: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

111

60,246–248 Studies have also shown that oxygen interacting with Bi2Se3 can form a native oxide

layer. It is unlikely that native oxide growth is the primary mechanism, at least initially, because

of the rapid rate of change and the easy reversibility of PL. We see quite easily that the oxygen

insertion and release could be achieved (keeping all other parameters constant) by simply

changing the partial pressure of oxygen in the surrounding environment. Additionally, previous

studies measured topographical changes due to native-oxide growth in Bi2Se3, 55 whereas we

found no such height changes (Figure 4.4 and Figure 4.5). Oxidation of Bi2Se3 has also been shown

to reveal new Raman spectral peaks.252–254 In contrast, our Raman spectroscopic measurements

show no new peaks (Figure 5.10, Figure 5.11, and Figure 4.16). These, taken together with the

oxygenation/release time-scales point more strongly in favor of a diffusive mechanism. Further,

the logarithmic behavior when low powers are applied (Figure 4.12b) is in agreement with Fick’s

law of diffusion, which also has a logarithmic shape.

In conclusion for this chapter, the interlayer coupling in vertically stacked 2D materials has

demonstrated the ability to dramatically alter their properties, and is considered a promising

parameter for designing materials with tailored capabilities. The Bi2Se3/MoS2 2D heterostructure

an ideal platform to study the interlayer interaction of 2D materials because it offers the ability to

tune the interlayer coupling in-situ and with high-spatial resolution. From a fundamental

perspective, in addition to rich excitonic physics, this system interplays strong spin-orbit coupling

in non-centro-symmetric crystal structures, and hence could potentially demonstrate novel

correlated, spin and valley physics.255–257 In addition, as site-programmable, color-selectable,

atomically-thin, micron-scale and intensity-tunable photoluminescing pixels (PLPs), this system

could be attractive for ultrathin and flexible optical information storage devices, color converters,

micro-cavity-lasers, and other photonic, plasmonic and optoelectronic applications.39,258–260 The

strong oxygen-selectivity of these heterostructures could also be potentially used as low-cost

oxygen-sensors and photon/power meters. We also showed data that suggested these materials

could be used as O2 storage devices, potentially storing 69 kg/m3 (a factor of 52 times the density

of O2 gas at 1 atm).

Page 112: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

112

4.1 Discussion of Possible exciton recombination pathways

This family of 2D heterostructures has produced intriguing data that speaks both to the promising

applications, as well as the complexity of the underlying science. There are a number of well-

studied mechanisms in literature that may apply because they are in agreement with the data.

Out findings indicate the Bi2Se3/TMD 2D heterostructure family potentially has multiple

competing mechanisms. Below we describe the mechanisms that we think could be present, and

then justify our conclusions with our observations and references. To best convey our thought

process, we list all the key observations (KOs) first.

Key Observations

1. The PL of the TMD is over 99% quenched when only one layer of Bi2Se3 is CVD grown on

the TMD. One layer of Bi2Se3 is not thick enough to reflect or absorb the incoming and

outgoing photons, meaning the Bi2se3 introduces a non-radiative electron-hole

recombination path.

2. Laser exposure in an air induces several changes :

a. PL intensity increases.

b. PL intensity can be precisely tuned (i.e. recovered with high control) over several

orders of magnitude.

c. Affected regions undergo a perceived color change, appearing brighter and with

less contrast, compared to as-grown heterostructures.

d. PL peak position is in the same location as the monolayer TMD’s PL peak position

from pre-Bi2Se3 growth, strongly indicating that the recovered PL observed is from

the excitons and trions in the TMD. This would indicate that air-laser-treatment

removes the non-radiative electron-hole recombination pathway, allowing the

excitons and trions to recombine at the K-point in the TMD, where it is direct

bandgap.

Page 113: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

113

e. The laser exposure recipe (i.e. laser power and time interval) will affect the rate of

change of the PL intensity and peak position shift.

3. The laser induced changes can be spatial determined with submicron precision (i.e. the

laser spot size).

4. Oxygen is required to induce changes when a laser or thermal treatment are applied.

5. Thermal treatment or laser-exposure in an oxygen-free environment reverses the oxygen-

induced changes.

6. Bi2Se3/MoS2 2D heterostructures are p-type compared to pristine monolayer MoS2.

7. Laser exposure does not remove Bi2Se3 from the surface, as seen by AFM scans.

8. Higher energy doses appear to induce more permanent changes.

9. Bi2Se3 grows crystalline and with long-range order on the TMD, suggesting strong van

der Waals epitaxy-mediated growth between the two component layers.

10. Density functional theory (DFT) calculations of the Bi2Se3/MoS2 2D heterostructure predict

that intercalated O2 molecules will increase the interlayer separation, disrupt the

interlayer bonding, and diminish the interlayer coupling, thereby inducing the two

materials to behave more “free-standing”.

Exciton recombination pathways that may be present

#1 –A straddled or staggered bandgap induces the photoluminescence quenching

Well-coupled 2D heterostructures have been shown to have overlapping bandgaps,149,261 which

we believe is happening when Bi2Se3 is grown on the TMDs, for reasons explained in key

observation 9 (KO-9). There are three different types of overlapping bandgaps: broken, straddled,

and staggered. Device data showed that Bi2Se3/MoS2 is p-type, meaning the bandgap is not broken

(broken bandgaps behave metallic), but is forming either a straddled or a staggered bandgap. All

four TMD’s studied in this work are wide bandgap compared to Bi2Se3 (>1.6eV to ~0.3eV),60

meaning both a straddled and a staggered bandgap would introduce a non-radiative electron-

hole recombination pathway.

Page 114: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

114

Mechanism #1 is in agreement with the PL quenching (KO-1) and why Bi2Se3/MoS2 is p-type

(KO-6).

Figure 4.17: Possible non-radiative electron-hole recombination pathways leading to PL quenching in as-grown 2D

heterostructures. Both straddled and staggered bandgaps are in agreement with the data and explain the

photoluminescence quenching observed because they introduce a non-radiative recombination pathway. Although a

broken bandgap also has a non-radiative recombination pathway, it disagrees with KO-6; broken band gaps are

metallic, whereas our device was p-type.

#2 – O2 intercalates between the monolayer MoS2 and the Bi2Se3, and then diminishes the

interlayer coupling

Laser exposure in air induces numerous dramatic optical and PL changes; however, the Bi2Se3 is

not removed (KO-7), as shown with AFM. We know that whatever the mechanism is, it is highly

local (KO-3) and that O2 is required for the process (KO-4).We believe that O2 intercalates into the

interlayer spacing of the heterostructure, where it disrupts the interlayer coupling, leading to the

changes in properties observed. It has been shown that O2 intercalating between 2D materials and

their surface decouples the 2D material, making quasi free-standing.127,134,135,137,142,239,240 In this case,

it decouples the materials, allowing the TMD to regain the radiative recombination pathway.

Additionally, each O2 molecule is only able to disrupt the interlayer coupling locally, meaning

the amount of disruption in a certain area is dependent on the number of O2 molecules; the more

O2 there is, the greater the interlayer coupling disruption. This mechanism (or radiative

recombination pathway) is in agreement with numerous key observations, stated below:

Page 115: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

115

KO-2a & KO-2d: As the interlayer coupling is disrupted, the non-radiative e-h recombination

path will no longer be allowed, permitting e-h excitonic pairs to radiatively recombine in the

TMD.

KO-2b: The amount of recovery is dependent on the amount of O2 that intercalates, explaining

the high control.

KO-2e: Changing the laser-exposure recipe will affect the diffusion of O2, thereby affecting the

rate of change.

KO-5: The changes can be reversed by thermal treatment or laser exposure in an O2 free

environment, potentially due to deintercalation.

KO-8: Laser exposure breaks the Bi2Se3 up into small grains. It has been previously shown that

oxygen intercalation is highly edge-limited and dependent on the number of grain

boundaries,127,134,135,137,239,240 suggesting increasing the number of grain boundaries will affect the

rate of diffusion.

KO-10: DFT calculations predict intercalated O2 will disrupt the interlayer interaction.

A number of past works have explored the intercalation of oxygen and other atoms in micron-

scale graphene and other 2D materials, where it has shown to be highly dependent on the number

of grain boundaries, and consistently shown to have an interlayer decoupling effect.81,82,127,134–

139,142,250,251 More specifically, a single grain boundary of these systems was sufficient to enable

large-scale intercalation in micron-scale samples, strongly suggesting that, a high concentration

of edges/grain boundaries under the laser spot might not be needed for the observed rapid

intercalation/deintercalation. As shown in Figure 4.2, Bi2Se3 grows highly crystalline on the

underlying TMD, and when the 2D heterostructure is then exposed to a laser in air, the Bi2Se3

becomes polycrystalline, suggesting the formation of grain boundaries. Together these results

suggest (1) Intercalation via grain boundaries is well-known, and do not necessarily require a

high density of grain boundaries to produce observable effects, and (2), these intercalations result

Page 116: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

116

in a de-coupling of the parent layers electronically. While that is not a direct evidence that

intercalation of oxygen is taking place, in our case, these are strong indicators of the most possible

mechanism.

Figure 4.18: DFT calculations comparing a Bi2Se/MoS2 superlattice with and without O2 intercalated. The calculations

predict that O2 intercalation will increase the interlayer separation, disrupt the interlayer bonding, and diminish the

interlayer interaction. Note how the interlayer separation increases after adding O2 molecules, going from an average

separation of 3.57Å to an average separation of 4.18Å, a 17% increase. Interestingly, the O2 molecules create an uneven

landscape of selenium atoms at the interface, pushing different atoms to different separation values. The above 2D

heterostructure is rotationally aligned (i.e. twist angle is 0°), where 3 Bi2Se3 unit cells are the same length as 4 MoS2 unit

cells

Figure 4.19: DFT calculations showing the location of the O2. These atomically-thin layers may be excellent candidates

for oxygen storage devices, potentially storing 69 kg/m3 ¬(a factor of 52 times the density of O2 gas at 1 atm). The

Page 117: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

117

Bi2Se3/MoS2 superlattice has a predicted volume of 1.936nm3 (1.247x1.086x1.43 nm), and the combined mass of five O2

molecules is 1.328x10-25 kg, yielding a density of 68.6 kg/m3.

Figure 4.20: Diagrams demonstrating the creation of grain boundaries, facilitating O2 intercalation into the interlayer

region. The upper diagram shows the heterostructure in an O2 environment prior to laser-exposure, and the lower

diagram is during laser-exposure.

Heterostructure prior to air-laser-treatment

Heterostructure after and during air-laser-treatment

Page 118: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

118

#3 – Potential of edge chemistry to affect the interlayer coupling

We do not believe that oxygen forms chemical bonds, at least initially, since this would contradict

the easy release by simple reversal of the oxygen partial pressure, reversal timescale (Figure 4.12)

and the absence of any oxide-related Raman peaks (Figure 4.16). On the other hand, edges or

edge chemistry might provide shallow trap sites that arise out of complex interplay between the

edge energy and surface energy of the underlying layers. Such shallow trap sites might provide

the reversible behavior seen in our system, without the need for formation of bonds. However,

the likelihood that the mere physical presence of trapped oxygen at the grain boundaries can

(without forming any chemical bonds) completely change the optical properties of the Bi2Se3

layers is possibly low.

4.2 Discussion on interaction of Bi2Se3 with oxygen

We consider why we believe that the intercalation should be limited to oxygen and not occur for

the other gaseous species. To address this, we note that the possible effect of all these gases,

including water vapor, were tested for their effect on the 2D heterostructures, by investigating

our samples dry air (79% N2 +21% O2), humid nitrogen (H2Ovapor+N2), nitrogen, argon, and air. By

process of elimination it was found that only O2 had the observed effect.

While attempting to understand why oxygen would have such a dramatic effect, vice the other

gases, we found an overwhelming amount literature that discussed the impact of oxygen on

Bi2Se3 and other materials in its class (e.g. Bi2Te3), suggesting, under different circumstances, both

diffusive behavior as well as bond-forming mechanisms.55,58–60,242,246–249 In sharp contrast, we were

unable to find any literature that demonstrates evidence of absorption or chemical reaction with

nitrogen, argon, or carbon dioxide, suggesting these molecules are significantly less

reactive/interactive with Bi2Se3. The collective body of work appears to indicate the energetics

facilitate these interactions with oxygen, vice the other gases.

Page 119: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

119

On one hand, some of the above-mentioned findings suggest that oxygen is able to readily absorb

into Bi2Se3. In particular, Bansal et al.242 showed that oxygen can diffuse through even 12 QLs of

Bi2Se3 epitaxially grown on pure Si under atmospheric exposure, and will completely oxidize the

underlying Si substrate into its amorphous oxide. Remarkably, no structural deformation of Bi2Se3

was observed (i.e. no change in lattice parameters or appearance of new bonds etc.). The

experiment was done using high resolution cross-sectional TEM. At the same time, no such

diffusive behavior is known to occur for oxygen in TMDs. Taken together, this suggests that the

possibility that oxygen, diffusing through the Bi2Se3 layers, can diffuse all the way till it meets the

Bi2Se3 interface. Our DFT calculations were utilized to interrogate the stability of such

intercalation and their effect, if any, on inter-layer separation distance.

On the other hand, the other possibility, i.e. that of the formation of chemical bonds was less

likely, at least at the initial stages of our experiments. It is well-established that Bi2Se3 is sensitive

to oxygen, including the partial pressure of oxygen in air.55,58–60,246–248 Studies have also shown that

oxygen interacting with Bi2Se3 can form a native oxide layer. It is unlikely that native oxide

growth is the primary mechanism, at least initially, because of the rapid rate of change and the

easy reversibility of PL. We see quite easily that the oxygen insertion and release could be

achieved (keeping all other parameters constant) by simply changing the partial pressure of

oxygen in the surrounding environment. Further, oxidation of Bi2Se3 has been shown to reveal

new Raman spectral peaks.252–254 In contrast, our Raman spectroscopic measurements which show

no new peaks. These, taken together with the oxygenation/release time-scales point more strongly

in favor of a diffusive mechanism. Further, the logarithmic behavior when low powers are

applied (Figure 4.12b) is in agreement with Fick’s law of diffusion, which also has a logarithmic

shape.

If in fact oxygen is diffusing into the interlayer, what is the mechanism(s), and what path(s) does

it take? To answer this question as best we can, we have conducted considerable literature

research and compared the conclusions to our data. Below we will lay out the possible

mechanisms we found for how oxygen might intercalate into the interlayer region, and then

explain why we think the most likely mechanism is through grain boundaries.

Page 120: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

120

Previous research has shown that atoms are able to diffuse into material through line-defects (in

the body of the crystals, not grain boundaries) using a process called defect-assisted

diffusion.262,263 Although the defects facilitate diffusion, the process of atoms diffusing into

crystals is still relatively slow occurring mostly over timescales of hours, especially if the diffusion

is required to travers several nanometers into the crystal and only assisted with little heating.

In contrast, we observed relatively fast rates of change in PL, suggesting a relatively fast rate of

diffusion. Further, we are able to affect these changes at relatively low laser powers (83.1μW),

and we can even affect them when more than five layers of Bi2Se3 are grown on top. Together this

data would suggest line-defect-assisted diffusion is unlikely to be the primary mechanism, and

grain-boundary assisted intercalation is the most likely mechanism.

Lastly, ARPES measurements have shown that under very low exposures (<0.1 Langmuirs),

oxygen inclusion hole-dopes ultraclean Bi2Se3 samples.60 However, the exposure level in our

system is ~109 orders of magnitude larger, and so we don’t expect doping to be a dominant cause

behind the observed switching of recombination pathways.

Page 121: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

121

Chapter 5: Tunable Photoluminescence in Bi2Se3/TMD

2D Heterostructures for Potential Write-Read-Erase-Reuse

Applications

In recent times, atomically-thin, layered 2D transition metal dichalcogenides (TMDs) have

attracted enormous attention both due to their remarkable electronic,258 optical,26 spin and valley

properties,264 and their potential for diverse flexible nanoelectronic,265 optoelectronic,266 sensing,40

and energy applications.100 Carriers in these systems form tightly-bound neutral and charged

excitons,33,267 and monolayer TMDs show strong photoluminescence (PL) peaks which result from

a direct-recombination of these excitons. Controlling the population and recombination energy

of these excitons is extremely important for both fundamental and applied research, and hence,

methods for direct manipulation of excitons is a topic of rapidly-growing interest. This work can

be found at the following reference.268

Figure 5.1: Figure briefly summarizing the capabilities demonstrated in this chapter.

In this chapter, we demonstrate highly controllable, reversible, and site-selectable switching

between radiative (i.e. photoluminescent) and non-radiative exciton-recombination pathways in

Page 122: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

122

Bi2Se3/TMD 2D heterostructures, possibly due to the regulated intercalation and deintercalation

of oxygen (Figure 5.2 and Chapter 4).104 Further, our approach allows permanent, in-situ,

electrode-less, and use-specific programming of both the radiative exciton populations and their

recombination energy values. To demonstrate this, we have directly synthesized a family of 2D

vertical heterostructures consisting of mono/bi/tri-layers of Bi2Se3 grown on top of several

monolayer TMDs (WS2, MoSe2, and MoS2) and a TMD alloy (MoSe2-2xS2x). The monolayer TMD’s

signature PL is quenched in all as-grown 2D heterostructures, indicating the interlayer coupling

is inducing non-radiative recombination pathways for the TMD’s excitons. However, the PL

intensity can be recovered at small increments by exposing them to calibrated laser doses in air,

suggesting the interlayer coupling is being controllably disrupted. Interestingly, not only does

the intensity increase, but the PL peak position shifts with increasing laser exposure as well,

suggesting multiple excitonic parameters are being manipulated. Our work indicates that both

the neutral and charged exciton binding energies shift, as well as the ratio of their populations,

suggesting these parameters can be tuned within a range. Finally, all the laser-induced changes

can be reversed by thermally annealing the 2D heterostructures in Ar or N2, suggesting the

interlayer coupling can be restored. Taken together, this family of 2D heterostructures offers a

wide range of emission energy values in the visible (1.5eV<Eph<2eV), where complex patterns

with submicron features and user defined the exciton parameters (e.g. PL intensity, peak position,

binding energies, relative populations) can be “written” into the material, suggesting interesting

optoelectronic applications.

Page 123: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

123

Figure 5.2: Reversible oxygen absorption and its role in manipulating the photoluminescence intensity (a) Optical

image of a Bi2Se3/MoS2 vertical 2D heterostructure exposed to a laser in an oxygen-present environment (i.e. air), where

the optical properties of the affected area were altered from purple to white, identified by a red arrow. (b) PL spectra

from the same spot before and after laser exposure. (c) The laser was now applied to a different location (orange arrow)

on the same sample at the same power (168μW), but the environment was changed from oxygen to nitrogen.

Interestingly, the color change seen in (a) was not observed. (d) Spectra before and after the laser was applied showing

no perceivable PL or change, suggesting the sample was not affected. (e) Optical image of the setup where N2 gas was

flown across the sample to displace the air and oxygen. (f) Variation of PL intensity under alternating air and N2

environments while a focused laser is applied, demonstrating that the PL intensity is dependent on the environment,

and can be reversed by simply switching the environment.

Our previous work demonstrated that when Bi2Se3 is grown on monolayer MoS2, the resulting

2D heterostructure forms a strong interlayer coupling that facilitates electron exchange and leads

to the formation of a new band structure.103 More specifically, Bi2Se3 grows at preferred twist

angles and moiré patterns that form supercells with their own band structure distinct from those

of Bi2Se3 and MoS2. Density functional theory (DFT) calculations predict the formation of

interlayer hybridization and an indirect bandgap, providing an explanation for the PL

suppression observed in the Bi2Se3/TMD 2D heterostructures studied in this work. The laser

exposure induced PL recovery in Bi2Se3/MoS2 2D heterostructures has appears to result from a

reversible absorption/desorption of oxygen, which electronically decouples the Bi2Se3 and MoS2

Page 124: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

124

layers, and allows the MoS2 to regain PL (Figure 5.2).104 In this work, we show that this

phenomenon can be extended to a family of Bi2Se3/TMD structures.

Figure 5.3 As-grown Bi2Se3/TMD 2D heterostructures (a) and (b) Optical images of monolayer MoSe2 and a

Bi2Se3/MoSe2 vertical 2D heterostructure, respectively. (c) AFM image of the red box in (b), demonstrating monolayer

Bi2Se3 (with bilayer islands) grew uniformly across the entire MoSe2 crystal, completely covering it. Inset: Line profile

of location shown by black line. (d) Side-view diagram of a heterostructure. Vapor-phase chalcogenization (VPC) is

used to grow monolayer TMDs on SiO2, and then vapor-phase deposition (VPD) is used to grow Bi2Se3 on the

monolayer TMDs.

Figure 5.3a shows an optical image of monolayer MoSe2 grown on SiO2 using vapor-phase

chalcogenization (VPC).31 Figure 5.3b shows a Bi2Se3/MoSe2 vertical 2D heterostructure, where 1-

2 layers of Bi2Se3 were uniformly grown on top of the TMD (see atomic force microscope (AFM)

image in Figure 5.3c), using vapor-phase deposition. Figure 5.3d is a side-view schematic of a

typical Bi2Se3/TMD 2D heterostructure. Despite the huge lattice mismatch between Bi2Se3 and

each of the TMDs studied (e.g. 3.62Å and 2.89Å for Bi2Se3 and MoSe2, respectively), uniform 1-2

layers of Bi2Se3 grow with high regularity and crystallinity on top of all the TMDs (see diffraction

images in Figure 5.4), suggesting strong van der Waals epitaxy-mediated growth between the

two component layers.243

Page 125: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

125

Figure 5.4: Representative SAED images of as-grown and laser-exposed 2D heterostructures : Bi2Se3/MoSe2 and

Bi2Se3/WS2. The laser-exposure makes the Bi2Se3 nano-crystalline and polycrystalline. The TMD will also become nano-

and polycrystalline if a sufficiently high laser power is used.

As shown in Figure 5.4, the highly crystalline growth of the Bi2Se3, as well as the consistent

response to laser exposure across independent samples, suggests that the as-grown 2D

heterostructure interface is clean and absent of foreign atoms. Bi2Se3 grows highly crystalline

across the whole TMD (i.e. micron sized domains), suggesting the foreign atoms are not present

to impede or modify the growth. Further, the laser exposure and Raman results have been highly

consistent across independently grown samples, suggesting the interface is the same or very

similar. Since it is unlikely that the samples would be equally “dirty”, the likeliest explanation is

that the interface is clean.

Page 126: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

126

Figure 5.5: Bi2Se3/MoSe2(1-x)S2xalloy SAED and TEM images. The SAED image shows how both the monolayer MoSe2-2xS2x

and the Bi2Se3 grow as crystals, indicating that there is strong epitaxial mediated growth, and an interaction between

the materials. Bi2Se3 seems to prefer to grow at 15°; although, other angles have been recorded. The TEM image shows

how MoSe2-2xS2x grows as a triangle, further evidence that it grows as a crystal. The square in the middle of the image

is an MoO2 crystal that became the seed for the MoSe2-2xS2x crystal. We are uncertain what the darker spots on the 2D

heterostructure are, but believe they are either the Bi2Se3 (it possible grew as an incomplete growth) or they are

undissolved PMMA (PMMA absorbs electrons, which is why they are so dark).

Figure 5.5 shows a TEM diffraction image of a Bi2Se3/MoSe2-2xS2x 2D heterostructure, where the

clear spots suggest that well-formed crystals are present. It should be noted that the Bi2Se3 did

not grow crystallographically aligned, but grew at a twist angle. As discussed in Chapter 5 in

more detail, the lack of chemical periodicity (i.e. the Se and S atoms are randomly distributed)

appears to remove the aligned state (i.e. 0° twist angle) as the lowest energy twist angle, where

most growths appear to be centered around 15°. The TEM image is to show that the atoms

maintain their triangular shapes after being mechanically transferred onto the TEM grids,

suggesting that they maintain their crystallographic order and structure. The dark spots on the

TEM image are likely reminiscent PMMA and not 2D heterostructure induced phase changes to

the electrons.

Page 127: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

127

Figure 5.6: Manipulating the optical and excitonic properties in Bi2Se3/TMD 2D heterostructures (a) Optical image of

the same Bi2Se3/MoSe2 2D heterostructure from Figure 5.3b after exposure to a laser showing a perceptible change in

color. (b) Optical image of a different Bi2Se3/MoSe2 2D heterostructure with a laser-written grid of “bytes”,

demonstrating the ability to write complex patterns with submicron features using a low-power laser. The signature

PL peaks of monolayer TMDs, which are suppressed in all as-grown heterostructures, appear strongly at these laser-

exposed spots, and re-quench when annealed in N2 or Ar (discussed later). (c)-(d) Normalized PL spectra of the entire

range of 2D heterostructures studied. (c) The monolayer TMDs on SiO2 prior to the Bi2Se3 growth. (d) The

corresponding 2D heterostructures with both as-grown and laser-treatment spectra, demonstrating the wide range of

photon energies over which the PL can be strongly tuned. Since the PL is a direct measure of the radiative excitons and

trions - whose recombination energies are indicated with arrows in (c) - these samples reflect a family of radiative

exciton population-tunable 2D materials, with wide-band selectable values in the visible photon energies. Diamond-

and circle-capped arrows correspond to as-grown and laser exposed 2D heterostructures, respectively. (e) PL spectra

from a Bi2Se3/WS2 2D heterostructure exposed to calibrated energy doses, as well as the PL spectrum from the

corresponding monolayer WS2 crystal prior to Bi2Se3 growth, demonstrating the PL intensity and peak position are

tunable. (f) Bi2Se3/WS2 2D heterostructure PL peaks from (e), where the dose-dependent peak position evolution is

better displayed.

Figure 5.6a is the same Bi2Se3/MoSe2 2D heterostructure as in Figure 5.3b after being exposed to a

laser at multiple locations (identified by the dark purple to light purple color change). Similar

optical changes were observed in all 2D heterostructures tested (Figure 5.7 and Figure 5.16).

Interestingly, the color change remains even after the 2D heterostructure has been kept

unprotected in air for several months, suggesting the laser-induced modifications are stable. AFM

scans from before and after laser exposure demonstrate that the changes are not due to Bi2Se3

simply being removed (Figure 5.9). Complex patterns with submicron features, including a grid

Page 128: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

128

of “bytes” (Figure 5.6b) and alphabetical letters (Figure 5.8), can be written using a low-power

laser.

Figure 5.7: Demonstration that laser exposure in air effects are reversed through annealing in either Ar or N2 (i.e. an

O2-free environment) for Bi2Se3/MoSe2 and Bi2Se3/WS2 2D heterostructures. The optical and PL effects from laser-

treatment in air can be reversed by annealing samples at 240° for 3 hours in an O2-free environment. After annealing

samples exposed to a low laser dosage, they often appear and behave indistinguishable to pristine samples; if higher

doses are applied, the annealing may not fully reverse the effects, possibly due to changes in the crystal structure. The

samples above were a monolayer TMD with 1-2 layers of Bi2Se3 deposited using vapor-phase deposition (VPD).

Page 129: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

129

Figure 5.8: Complex patterns and grids can be controllable written into Bi2Se3/TMD 2D heterostructures with

submicron precision. (a) The letters “NEU” are written into Bi2Se3/MoS2 using a more elegant font. (b) The letters

“NEU” were written into Bi2Se3/MoSe2, as well as an elongated “N” to show how submicron spots are used to write

the patterns.

Four different Bi2Se3/TMD 2D heterostructures were synthesized, where the normalized

representative PL spectra of each are shown in Figure 5.6c. Interestingly, when only one layer of

Bi2Se3 is grown on any of the monolayer TMDs using vapor-phase deposition, the bright PL

quenches >99% (Figure 5.6d and Figure 5.6e), suggesting the interlayer coupling between the

Bi2Se3 and the TMD induces a non-radiative recombination path for the excitons. Previous work

demonstrated this in Bi2Se3/MoS2 2D heterostructures, and used DFT calculations to predict the

formation of interlayer hybridization and an indirect bandgap, providing an explanation for the

PL suppression.103

Page 130: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

130

Figure 5.9: Atomic Force Microscopy (AFM) analysis of Bi2Se3/MoSe2 and Bi2Se3/WS2 : as-grown vs. laser exposed in air

vs. O2-free annealed. (a) AFM image of the as-grown Bi2Se3/MoSe2 sample shown in (e), Figure 1b, and Figure 1c. The

2nm step height reveals that there is one tri-layer of MoSe2 with 1-2 continuous quintuple-layers of Bi2Se3 on top. The

nanostructure was then exposed to a laser in air (see (f)) and then it was O2-free annealed in Ar (see (b) and (g)). (b)

AFM image of the Bi2Se3/MoSe2 sample shown in (g), which was exposed to a laser in air and then annealed in Ar. It

can be seen that the step height is still 2nm, indicating that the Bi2Se3 remains on the TMD after laser exposure and

annealing. (c) AFM image of the laser exposed Bi2Se3/MoSe2 2D nanostructure shown in (h). This AFM image reveals

two important findings: (1) the Bi2Se3 is not removed through laser exposure, and (2) the Bi2Se3 independently grows

at multiple nucleation sites. The large optical change induced by laser exposure (blue arrow) is clearly visible in (h);

however, the AFM image shows no step downward, indicating that the Bi2Se3 is still present. The Bi2Se3 is not uniformly

and continuously spread on the MoSe2, which is what is seen in (a) and (b), but is instead distinct islands between 1-3

layers thick. The Bi2Se3 deposits at multiple nucleation sites independently, and then the grains will begin to growth

together to form a continuous layer. (d) AFM image of the as-grown Bi2Se3/WS2 sample shown in (l). The sample has

>98% Bi2Se3 coverage, with islands of bilayer. The line profiles in (d), (j) and (k) verify this. (e) As-grown Bi2Se3/MoSe2,

which was then then exposed to a laser in air (see (f)). Laser exposure induces a perceived color change in the sample,

which is very remarkably reversed when the 2D nanostructure is O2-free annealed (see (g)). (h) Optical image of the

laser exposed Bi2Se3/MoSe2 nanostructure in (c), where the laser-treated area is recognized by a distinct change in

color (blue arrow). The lines on the side are markers used to identify the sample. (j), (k), and (l) are line profiles and an

optical image, respectively, associated with (d).

Very interestingly, when any of the Bi2Se3/TMD 2D heterostructures are exposed to a laser in air,

the PL intensity increases and the peak position shifts, due to a disruption of the interlayer

coupling, which induces the monolayer TMD to behave more electronically independent,

facilitating radiative exciton recombination. Figure 5.6d shows the large change in PL intensity

and recombination energy between as-grown and laser-exposed Bi2Se3/TMD 2D heterostructures.

All the spectra have been normalized, and the markers show the change in peak position. Figure

5.6are representative PL spectra from a Bi2Se3/WS2 2D heterostructure exposed to increasing laser

Page 131: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

131

energy doses, as well as from that of a monolayer WS2 crystal prior to Bi2Se3 growth. The 2D

heterostructure’s PL is over 99% quenched compared to the initial monolayer WS2 crystal. As

low-energy doses are applied, the PL intensity increases at a controlled rate and the peak position

shifts, due to changes in the exciton dynamics (discussed later).

Figure 5.10: Raman spectroscopy study of Bi2Se3/MoSe2 : as-grown vs. laser exposure in air vs. O2-free annealing.

MoSe2 and Silicon peaks. The Bi2Se3, MoSe2 and silicon Raman peaks are labeled above. An analysis of how the Bi2Se3

Raman peaks evolve as a function of laser exposure in air and O2-free annealing is shown in next. Lase exposure in air

causes both the MoSe2 and silicon peaks to increase. This is in agreement with our claim that laser exposure in air

diminishes the interlayer coupling, and induces each material to behave more “free-standing”. When the TMD is more

“free-standing”, its atoms will vibrate more freely – they will be less confined by the Bi2Se3 – thereby increasing the

intensity of the Raman peaks. The silicon peaks likely increase because the laser exposure causes the nanostructure to

become more transparent, previously shown to the case with Bi2Se3/MoS2 2D heterostructures. Annealing will either

have a negligible effect on the Raman response (see first annealing), or it can reverse the effects of the laser-treatment

(see second annealing). The acronym “LT” refers to laser-treatment (LT), the process where a laser is exposed to the 2D

nanostructure.

Page 132: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

132

Figure 5.11: Raman spectroscopy study of Bi2Se3/MoSe2 and Bi2Se3/WS2: as-grown vs. air-laser-treatment vs. O2-free

annealing. Bi2Se3 peaks. (Left Graphs) The Bi2Se3 Raman response contains two outer peaks, and a middle peak

between the two outer peaks. When the sample is exposed to a laser in air, the two outer peaks will decrease in intensity

and the inner peak will increase. Similar to the MoSe2 and silicon peaks, the first O2-free annealing had a negligible

effect on the Bi2Se3 peaks; however, the second O2-free annealing partially reversed the effects of the laser exposure by

increasing the intensity of the outer peak. These changes in the Bi2Se3 Raman modes might be related to the structural

changes shown next. (Right Graph) Raman spectroscopy study of Bi2Se3/WS2: as-grown vs. air-laser-treatment. Laser

exposure in air on Bi2Se3/WS2 has a similar impact as on Bi2Se3/MoSe2: the Bi2Se3 peaks decrease in intensity, and the

WS2 and silicon peaks increase in intensity.

It has been previously shown that the Raman modes are influenced by the interlayer

coupling.123,126 We found that laser exposure enhances the monolayer TMD’s Raman peaks,

suggesting the interlayer coupling is being diminished and the TMD is more independent (Figure

5.10 and Figure 5.11), in agreement with our PL findings. The changes in the Raman modes could

possibly be related to the introduction of oxygen in the interlayer region (Figure 5.2), or possibly

related to the changes in structure induced in the Bi2Se3 (Figure 5.12).

Page 133: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

133

Figure 5.12: Bi2Se3/WS2 SAED images after laser exposure in air, and then annealed in Ar. Laser exposure (or laser-

treatment) in air induces the Bi2Se3 to become more nano- and polycrystalline. We show that annealing in Ar or N2 (i.e.

an O2-free environment) tends to reverse the laser induced structural changes to Bi2Se3 (i.e. the Bi2Se3 becomes more

crystalline); however, a near full reversal was only observed in nanostructures treated with a low dosage, indicating

that large structural changes are difficult to fully reverse. The TEM image shows the location that was laser-exposed

and annealed.

Figure 5.13: Tuning the PL intensity and peak position (a) PL intensity and (b) PL peak position of a Bi2Se3/MoS2 2D

heterostructure exposed to 68 energy doses (via laser) of varying recipes (i.e. varying time increments and laser

powers), demonstrating the wide range of PL intensities and peak positions obtainable. The PL intensity increased by

a factor of 64x, while the peak position varied by 37meV. The experiment was conducted such that after applying a

pre-determined recipe, the laser power was reduced to the lowest setting to collect the 2D heterostructure’s entire PL

Page 134: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

134

spectrum for exciton analysis (discussed later). (c) When a laser is continuously applied at low-power (e.g. 50 µW) to

a Bi2Se3/MoSe2-2xS2x 2D heterostructure, the PL intensity increases consistently along a logarithmic curve shape before

plateauing, demonstrating that the exciton dynamics can be tuned with high precision.

By exposing the 2D heterostructures to discrete laser doses of varying recipes (i.e. varying time

increments and laser powers), the PL intensity (Figure 5.13a) and peak position (Figure 5.13b) can

be modified over a wide range. Conversely, if the laser in continuously applied without

interruption at a low power (50 µW), the range of PL intensity was found to be smaller, but also

increase approximately logarithmically (Figure 5.13c), suggesting the response is dependent not

only on the total energy dose, but also how it is applied. As shown next, the shift in PL peak

position is likely due to changes in both the exciton binding energies and the ratio of the excitonic

populations, suggesting the exciton dynamics can be manipulated.

Figure 5.14: Manipulation of both exciton populations and their recombination energy (a) Deconvolution of a typical

PL peak into its two Lorentzian contributions - which represent the exciton and trion quasiparticles - from a Bi2Se3/MoS2

2D heterostructure exposed to a laser in air. The Lorentzian’s peak position is a measure of the recombination energy

(i.e. binding energy), while the peak intensity represents the population of quasiparticles that radiate. (b)-(e) are

extracted by analyzing the PL spectra from the experiment in Figure 5.13a-b. (b) Relative variation of the radiative A-

exciton population versus laser dose. During the experiment, the laser power had to be increased from 3.8µW to

1430µW to cover the large range (3 decades) of dosing within reasonable time. Applying these changes resulted in

sharper increases, as well as decreases, in the trends. (c) Variation of the A exciton recombination energy under the

same range of doses. A large variation of 35 meV could be obtained in this sample that included both red shifts and

blue shifts. (d) Relative variation of the radiative A- -trion population versus laser dose. (e) Variation of the A- -trion

recombination energy under the same range of doses. Like the exciton, a large variation of about 35 meV could be

obtained in this sample that included both red shifts and blue shifts. The altering trends in both exciton and trion

population suggests multiple mechanisms may cause the changes seen, with the possibility of significant sample

degradation at the highest doses.

Page 135: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

135

Analysis of the PL spectra provides a broader context for the laser-induced changes within the

encapsulated TMD films under Bi2Se3. Owing to the possibility that both the population of

radiative excitons, as well as the value of their recombination energy (i.e. Eph, the energy of the

emitted photons) can be manipulated by external probes, 2D materials have become extremely

attractive for atomically-thin and tunable functional components in LEDs and lasers,39 optical

modulators,260 photonic crystals,260 and optical resonators.259 In the presence of applied electrical,33

optical,269 mechanical231 and magnetic fields,270 the excitonic populations and recombination

energy values can be manipulated; however, the changes vanish when the fields do. In contrast,

our approach allows permanent, in-situ, electrode-less, and use-specific programming of both

radiative exciton population and recombination energy values. Figure 5.14a shows a typical PL

peak arising from two known excitonic contributions (A exciton and A- trion) in the Bi2Se3/MoS2

2D heterostructure.33

Page 136: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

136

Figure 5.15: Representative A and A- fits for photoluminescence spectra shown in Figure 5.14, showing the good quality

of the fits. All R-squared values are >0.98, and 56/60 of the fits have values >0.995, a quantitative demonstration they

are a good fit. Two randomly chosen intervals (every 6 and every 10) show the progression. Fitting was done using

Matlab. Guidance for this method can be seen in 29.

Figure 5.15 shows additional examples of the deconvolution of the PL spectra. Figure 5.14b-e

show the variation peak amplitudes and recombination energies of the A exciton and A- trion for

the experiments described in Figure 5.13a-b. In addition to the 16× - 18× change in the population

of radiative excitons, there is a large (up to ~35 meV) shift of the excitonic recombination energies.

This reflects a significantly larger degree of manipulation compared to some of the previously-

reported approaches,33,259,267,271,272 suggesting sizable changes in the electronic behavior of the

heterostructure. Specifically, the recombination energy for excitons (as measured using PL

spectroscopy) is )1(

bgph EEE , where gE is the quasi-particle band gap of the heterostructure,

Page 137: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

137

and the ground-state binding energy1

bE is given by22

41 2

eEb ,273 where is the dielectric constant

of the surrounding medium, and )/(1 11 he mm is the reduced mass of the e-h system. From

this formalism, we conclude that laser-exposure imparts large changes to either: (i) the band gap

and/or (ii) the binding energy of the excitons. In our heterostructures, the overall changes

reflected both red and blue shifts in the excitonic energy values, suggesting that multiple

competing mechanisms could be present. The question arises, what are these mechanisms?

Indeed, previous work has shown that the surrounding dielectric environment influences the

exciton binding energy.272 Additionally, our concurrent work reveals that that oxygen is critical

to induce the expected changes when a laser is applied to a Bi2Se3/MoS2 2D heterostructure, and

suggests oxygen molecules might be intercalating between the layers.104 Together this suggests

diffusive oxygen might be modifying the surrounding dielectric, thereby shifting the exciton

binding energies.

Figure 5.16: Annealing reverses the changes to optical and excitonic properties. (a) Sequentially-acquired PL spectra

representative of the same location, and (b) the corresponding optical images for: the initial monolayer MoSe2-2xS2x

Page 138: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

138

crystal, the same monolayer crystal with 1-2 layers of Bi2Se3 grown on top, and the same 2D heterostructure subject to

alternating laser exposure and thermal annealing. Laser exposure in air induced the expected PL reemergence and

color change in each instance, while thermal annealing in Ar or N2 reverses those changes. The process can be repeated

numerous iterations with little to no noticeable degradation to the 2D heterostructure. The red arrows indicate the

locations that were exposed to a laser.

Figure 5.16 demonstrates that thermal annealing of the samples in either Ar or N2 environments

can reverse the laser induced changes to the optical and excitonic properties, where little to no

degradation was observed when sufficiently low-energy doses were applied, suggesting the 2D

heterostructures have repeat-use capability. Figure 5.16a shows sequentially-acquired PL spectra

representative of the same location, and Figure 5.16b shows the corresponding optical images.

Initially a monolayer MoSe2-2xS2x crystal was grown with a bright PL spectra centered at 1.74eV,

which is in between monolayer MoS2 (~1.83eV) and monolayer MoSe2 (~1.53eV), suggesting it is

an alloy of the two. 1-2 layers of Bi2Se3 were then grown on top, inducing a color change and

quenching the monolayer TMD’s PL. SAED images suggest both the monolayer MoSe2-2xS2x and

the Bi2Se3 grown on top are highly crystalline (Figure 5.5). The 2D heterostructure as then exposed

to a laser while in air, inducing the expected color change and facilitating reemergence of the

monolayer TMD-alloy’s signature PL. The 2D heterostructure was then annealed in an Ar

environment at 240° for 3 hours, reversing the changes to the optical and excitonic properties.

This sequence of laser exposure followed by annealing was repeated twice more, where each

iteration produced the anticipated results, and there appeared to be little to no noticeable

degradation, suggesting the 2D heterostructures have repeat-use capability.

In conclusion for this chapter, the Bi2Se3/TMD 2D heterostructure family is a complex system that

has demonstrated promise as both a research tool and a material with a direct connection to real-

world applications. To our knowledge, this is the first demonstration of a family of semiconductor

2D heterostructures whose photoluminescence can be reversibly tuned in such a controllable

manner, in-situ, and over such a wide range of photon energies (1.5 eV<Eph<2eV). The in-situ

tuning and wide range of photon energies allows researchers to fabricate customized excitons.

Page 139: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

139

We demonstrated that a focused laser can be used to “write” complex patterns with submicron

features, and that annealing “erases” the laser-induced changes. There is little to no noticeably

degradation when low-powers are controllably applied, suggesting these materials have repeat-

use capability. These 2D heterostructures might be useful as ultrathin and flexible optical

information storage devices, color converters, micro-cavity-lasers, and other photonic, plasmonic

and optoelectronic applications.39,258–260 Additionally, this system interplays strong spin-orbit

coupling in non-centro-symmetric crystal structures, and could potentially be used to study

correlated, spin and valley physics.255–257 Hence, we believe our work paves the way for rich new

science and technology research.

5.1 Laser-Annealing in Air, and Strong Stability of As-Grown Configuration

The below figures demonstrate that a laser applied at very low powers (0.91μW) continuously

for several minutes is able to anneal the 2D heterostructures and reverse the PL intensity changes

induced by higher laser powers. What is very surprising is that this can even be accomplished in

air, which contains oxygen. These results suggest that the very low powers are able to slowly

perturb the oxygen to leave the interlayer region (or wherever it is residing) and increase the

interlayer coupling strength. This might be because the laser perturbs the Bi2Se3 into a more stable

twist angle and configuration, which then is more likely to squeeze the oxygen out. It has been

previously shown with graphene on metal substrates that the twist angle affects the rate of

oxygen absorption, and certain twist angles that are particularly stable will even encourage the

oxygen to de-intercalate.

The other figure below shows how the PL changes over a very long experiment where very low

powers are used to decrease the PL and higher powers are used to slowly raise it. Interestingly,

the PL will plateau at unexpected intervals, suggesting that a stable configuration was reached.

Additionally, and very interestingly, the as-grown configuration appears to be more stable than

the other post-laser-exposure configurations, in agreement with the assumption that these

Page 140: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

140

materials would grow at their most energetically stable state. We found that a higher power was

required to break-out of the initial as-grown state.

Figure 5.17 Very low laser power can anneal Bi2Se3/TMD 2D heterostructures. The H-MoS2 heterostructure was laser-

treated upward sufficiently, and then the power was turned down, such that a very low power was being applied to

the sample. It was observed that the PL intensity began to consistently decrease, until it plateaued. Then an impulse of

high power was applied and the PL increased, which confirmed that the PL intensity did not simply decrease because

the TMD was damaged. Next, the same low power laser-treatment recipe was applied, and the PL intensity once again

began to decrease until it plateaued; however, this time it plateaued to a higher PL intensity. Next, a lower power laser-

treatment recipe was applied, and the PL intensity increased controllably, revealing three observations: (1) the

underlying TMD was not noticeably damaged anywhere in this experiment, (2) the low-power laser-treatments are

likely annealing the sample, and (3) a laser can be used to tune the PL intensity up or down.

(Bottom right inset) The histogram counts the change in PL intensity between measurements (IN-IN-1=ΔI) that were

obtained during the first annealing session (i.e. between the blue and purple arrows). The distribution shows that the

behavior of the system is statistical, and that very low laser power does not always cause the heterostructure’s PL

intensity to decrease; however, the overall trend is downward. This can be compared to the distribution shown in

figure 3, which is shifted to the right because the laser-treatment recipe used a higher power. Comparing these two

distributions reveals that the moiré heterostructures have a statistical response to laser-treatments, and that the peak

shifts left or right depending on the laser-treatment recipe. Higher powers and longer time intervals shift the peak to

the right, and lower powers and short time intervals do the opposite.

Page 141: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

141

Figure 5.18 Probing the PL stability of moiré heterostructures. The PL of a moiré heterostructure can be tuned

through laser-treatment and annealing. When the interface between the TMD and Bi2Se3 is disrupted, the Bi2Se3 will

break into individual grains, and these grains will begin to move (i.e. rotate, strain, shift position, and shift z-distance

to the TMD). As they move, they will form different moiré patterns with the underlying TMD, which will change the

interaction between the two compounds and affect the luminescent properties (i.e. the PL will be turned ‘on’ or ‘off’

depending on how the TMD and Bi2Se3 interact). The DFT results revealed that the different moiré patterns

(commensurate unit cells) will have different formation energies, meaning some will be more energetically stable

than others, thereby making those patterns more favorable. This favoritism was observed in the histogram (figure S9)

where certain relative-rotations are statistically preferred.

(Left side) The PL intensity was quantitatively determined to be stable eight separate instances (i.e. it plateaued),

even though the laser-treatment recipe (the time interval and power) had not changed, leading to the question: why

would the PL intensity suddenly stabilize? The PL was changing in response to the laser-treatments, which means

the moiré heterostructure itself was changing. The heterostructure will change and evolve until it falls into a pseudo-

stable state, which forces the PL intensity to remain stable for several laser-treatments. The stable points were

identified using the following rigorous quantitative criteria: (1) four consecutive measurements all need to be within

1.25% of their mean value, and (2) the first and last values need to be within 1.25% of each other.

(Right side) Pristine and annealed moiré heterostructures are in the energetically most favorable configuration,

which is why they are more stable than their laser-treated counterparts. Annealing allows a crystal to find its most

energetically favorable configuration. Here the H-MoS2 was laser-treated with a certain recipe, and the PL only

increased on average by 14 counts. Then a short impulse of a higher laser power was applied to the sample, which

caused some bonds in the sample to break. After the high energy burst, the same, original laser-treatment recipe was

applied, but this time the heterostructure’s PL increased by 96 counts on average, a 586% increase from before, which

indicates that the heterostructure was more sensitive after a larger energy impulse was applied. The heterostructure

was more sensitive because the heterostructure was taken out of its most stable configuration.

Page 142: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

142

Chapter 6: Evidence of a purely electronic two-

dimensional lattice at room temperature

Interlayer coupling can trigger a variety of phenomena in vertical 2D heterostructures, including

creating new excitons (both interlayer and moiré),212,274 modifying e-h recombination pathways,103

altering Raman vibrational modes,123 shifting the absorption profile,184 enabling super-dense

storage of lithium,81 and inducing superconductivity.66 In many instances, these new properties

are strongly modulated by the formation of moiré superlattices, where the periodic variations of

atomic registry between the two layers generate a new “lattice” with significantly modified band

structure, electronic, optical and magnetic properties.66,86 Since it is possible to artificially stack 2D

layers with arbitrary twist angles, understanding and controlling the enhanced interactions at

sites of high atomic registry in twisted 2D bilayers has become the central focus of research by

many theoretical 115 and experimental efforts 156 in recent times. These superlattices can be

“visualized” in real space in high-resolution TEM or STM measurements, both of which measure

the sum of intensities (or currents) of electrons, either transmitted through or tunneled-out from

various lattice sites, but neither of which capture the electron amplitude and phase information.

In contrast, high energy selected area electron diffraction (SAED) experiments, which are

sensitive to the amplitude and phase of electrons as they scatter from various lattice sites, have

so far never been reported to show diffraction patterns that corresponding to superlattice

periodicities in 2D materials. The wavelength of electrons are orders of magnitude smaller than

the effective separation between the layers (therefore reducing the possibility of interference

between the two atomic lattices) and electronic charge redistribution between layers (if any), are

too weak to scatter high-energy electron beams. As a result, nearly all past work on 2D moiré

lattices use 2D FFT of real-space images as a means to simulate moiré “diffraction” patterns,

rather than showing their real appearance in, e.g. a SAED measurement. In sharp contrast, we

show, for the first time, direct observation of SAED patterns whose positions appear to

correspond to the moiré superlattices in a 2D heterostructure, at room temperature. We argue

Page 143: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

143

that these are not arriving from an interference of electrons from the atomic superlattice, but

rather, from a completely new periodic lattice of scattering sites that are purely electronic in

nature. This work can be found at the following reference.275,276

Figure 6.1 Direct observation of a new lattice in selected area electron diffraction of vertically stacked Bi2Se3/TMD 2D

heterostructures. (a) and (b) are optical images of monolayer WS2 and a Bi2Se3/WS2 2D heterostructure, respectively. A

contrast is observed on the 2D heterostructure because Bi2Se3 grew monolayer with bilayer islands. (c) Atomic Force

Microscope (AFM) scan showing three layers of Bi2Se3 grew evenly on monolayer MoSe2-2xS2x, an alloy of MoS2 and

MoSe2. (d) Diagram representing the four different 2D heterostructures studied in this paper: Bi2Se3 grown on MoSe2,

MoS2, WS2, and MoSe2-2xS2x. The TMD is grown on SiO2, and then Bi2Se3 is grown on top of the TMD. (e) Select area

electron diffraction (SAED) image of a Bi2Se3/WS2 2D heterostructure with the in-plane lattice parameters labeled.

Lattice parameters were calculated by measuring the spot separation. The distinct dots (labelled in blue and green)

indicate the TMD and Bi2Se3, respectively, both grow highly crystalline. A third, new set of diffraction spots (red label)

is present that follows the geometric moiré superlattice pattern of the 2D heterostructure and is ascribed to a lattice of

purely electronic charge pools residing at the interface (see text).

High energy (200keV) SAED investigations were performed on 2D heterostructures containing 1-

3 layers of Bi2Se3 grown on a variety of monolayer transition metal dichalcogenides (TMDs) -

MoSe2, MoS2, WS2, and MoSe2-2xS2x – with various crystallographic twist-angles, and some of the

results were confirmed using first-principles density functional theory (DFT). Our past and

ongoing work show quenching of photoluminescence in these heterostructures, and their

reversible behavior when laser-treated in oxygen environment.103,104 Here, we focus on the

appearance of new diffraction spots in some of these structures.

Page 144: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

144

Figure 6.1a shows an optical image of monolayer WS2 grown on SiO2 using vapor-phase

chalcogenization.31 Bi2Se3 samples were then grown on top of such WS2 flakes using vapor-phase

deposition. Figure 6.1b shows a Bi2Se3/WS2 2D heterostructure with 1-2 layers of Bi2Se3 (i.e.

monolayer with bilayer islands). Figure 6.1c is an atomic force microscope (AFM) scan of a

Bi2Se3/MoSe2-2xS2x 2D heterostructure. Interestingly, three layers of Bi2Se3 grew on the monolayer

TMD, while none grew on the SiO2, suggesting the TMD substrate stabilizes the Bi2Se3 in its few

layer morphology and is a preferred surface over the SiO2. Figure 6.1d schematically shows how

the two different materials stack up on the substrate. Figure 6.1e is a high energy SAED image of

a Bi2Se3/WS2 2D heterostructure with the diffraction patterns (with a six-fold symmetry)

corresponding to the two parent crystals Bi2Se3 and WS2 as labeled along with their corresponding

lattice parameters. In addition, a third set of spots (also with a six-fold symmetry) labeled “New

lattice” can be clearly seen with smaller reciprocal lattice dimensions (and hence much larger real-

space dimensions), which has the same lattice structure as the moiré superlattice of the parent

crystals, and which we attribute to a separately-formed purely electronic lattice. We next discuss

the justifications for such an assignment.

To begin with, the detected extra set of diffraction pattern could not be associated with any known

atomic basis, but always appear to have a real-space equivalent to the moiré superlattice structure

of the 2D heterostructures being probed. This observation is demonstrated for a typical

heterostructure in Figure 6.2 (also section 6.2 in this chapter). First, to simulate the real-space

superlattice, hexagonal lattices representing the TMD and Bi2Se3 layers scaled to experimentally

measured lattice parameters and twist angles, were digitally generated, twisted, and then

superimposed into a composite image. An example of such a simulated superlattice created

(using the lattice parameters and twist angle value from Figure 6.1e), is presented in Figure 6.2a,

where the blue lines and the red rhombus have been used to visualize and measure the lattice

spacing and unit cell of the clearly visible moiré superlattice. Next, a 2D fast Fourier transform

(FFT) was performed on this superposed image, to simulate a set of SAED spots (Figure 6.2b). We

note the digital superposition of the real-space lattice flattens them into a single plane, and

contains no layer-separation information in the perpendicular direction, which allows the 2D FFT

to retain the superlattice-induced spots. In other words, the top-down view (along the z-axis) is

Page 145: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

145

how the SAED electrons would perceive the 2D heterostructure if the two parent lattices were merged

into the same plane. Conceptually, this 2D FFT of the superimposed “heterostructure”, mimics

what the SAED patterns would look like, had there been enough overlap between the real-space

lattices such that the diffraction pattern from the first lattice could interfere with the second one.

Figure 6.2c shows an overlap of Figure 6.2b on to the actual SAED pattern shown earlier in Figure

6.1e. In all cases investigated, the 2D FFT-generated and experimentally-obtained bright spots are

in near perfect alignment – including the new lattice - suggesting that the experimentally

observed additional spots have a direct correlation with the structure of the moiré superlattice of

the parent crystals. The Diophantine equation shown in section 6.2 (Equation 6.1) was used to

calculate the corresponding commensurate unit cell (Figure 6.2d), verifying that the SAED spots

are in fact correlated to the moiré superlattice pattern, and not some other artifact.277 And finally,

the moiré lattice was absent when the TMD and Bi2Se3 lattices were individually Fourier

transformed and overlaid (Figure 6.3), demonstrating it is not an artifact from either lattice

individually.

Figure 6.2 Electronic moiré lattice scattering sites formed by the interlayer coupling. (a) Simplified real-space model of

the 2D heterostructure in Figure 6.1e. Two lattices with in-plane lattice parameters representing the WS2 and Bi2Se3

crystals were twisted and then overlaid, thereby collapsing both lattices into the same 2D plane, over-writing pixels,

and inducing the emergence of a moiré pattern (see methods section). Moiré patterns only manifest when there is an

interaction between two lattices. The moiré pattern lattice spacing and unit cell are identified by the blue and red lines,

respectively. (b) 2D Fourier transform of the same model with the lattices labeled, where the blue, green, and red

hexagons correspond to the WS2, Bi2Se3, and moiré superlattice, respectively. (c) 2D Fourier transform (orange shading)

overlaid on the SAED image seen in Figure 6.1e (white spots), where the near-perfect alignment of the spots suggests

Page 146: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

146

the model captures the system sufficiently. Interestingly, the moiré superlattice spots match the experimental new

lattice spots, suggesting the new lattice forms in congruently with sites of high atomic registry in the moiré superlattices

of Bi2Se3 and WS2. The inset with red arrows indicate patterns from the new lattice that are hidden under the red

hexagon in the main image due to overlap of the modeling and experimental spots. (d) Table with the parameters from

the mathematically calculated commensurate unit cell (see section 6.2 and section 6.3 for more information), the model,

and the experimentally measured values. All the values are nearly identical and well within the error, demonstrating

agreement between theory, modeling, and experiment.

Figure 6.3 The moiré SAED spots are not a product of one of the individual lattices. (a) The Fourier transform of a

model of a Bi2Se3 lattice, and (b) and the Fourier transform of a representative monolayer TMD lattice. (c) TMD

transform was colored in red and overlaid on the Bi2Se3 transform. None of the images contain the moiré superlattice

spots, demonstrating that an effective multiplication and interaction of the lattices is needed to produce them (see

section 6.2 for more information).

It is unlikely that the additional SAED spots are a result that electrons diffracted by the upper

crystal is further diffracted by the lower crystal. Mathematically, the formation of moiré patterns

– or moiré effects – requires an interaction (or non-linear operation) between the two parent

lattices (see section 6.2 for details).277,278 For example, when images of periodic structures are

overlaid, pixels are over-written as both images are collapsed onto the same surface. This over-

writing of pixels, vice simply summing their intensities, reflects a non-linear “superposition” that

produces a visual moiré pattern. In contrast, when 2D crystals are stacked, the different layers do

not reside in the same plane, but are separated by ~ angstroms. The possibility of interferences

between the two layers is hence remote, since the layer-center separations (~10Å) is massive

(~400×) compared to the wavelength of the electrons (0.0251Å at 200 keV), suggesting the

electrons perceive the two materials independently. Further, the very low scattering angles of

Page 147: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

147

high energy electrons in SAED (~0.6°) usually means very little chance of any "double diffraction"

for our atomically-thin Bi2Se3/TMD heterostructures since, separated at a distance of ~3.5Å, the

lateral spread of an electron scattered from the first crystal would be ~ 0.0367 Å at the second

layer, which is far smaller than the lattice parameter of the second crystal. Hence, we conclude

that observed new diffraction spots are not resulting from composite interferences of diffraction

patterns from the atomic lattice of the parent crystals, but rather, are due to diffraction from a

completely new lattice that has the same “structure” as the moiré superlattice of the parent

crystals. Since electrons scatter elastically from concentrated charge pools, this implies that a

separate lattice of charge pools must have been created in these heterostructures that induces

these new spots.

These results lead to three important questions: (1) What do the electronic moiré lattice scattering

sites look like, (2) where do they reside, and (3) why does the lattice follow the moiré pattern, vice

a different structure? We addressed these questions by performing first principles density

functional theory (DFT) calculations, which are known to provide structural information and

electronic charge distribution in solids with relatively high accuracy. Indeed, our DFT

calculations not only confirm the formation of a purely electronic 2-D lattice in the interlayer

region, but also whose unit cell size and relative rotation are in agreement with the

experimentally observed moiré pattern in our system.

Page 148: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

148

Figure 6.4 Charge pools form between nearest interlayer neighbors. (a) DFT calculated interlayer charge redistribution

due to the interlayer coupling of a twisted 1L Bi2Se3 + 1L WS2 2D heterostructure, using the lattice parameters described

in Figure 6.2d. The image is a cut of the plane lying equidistant from both materials. The nearest neighbor atoms are

included in the image. It can be seen that the charge redistribution is concentrated between nearest interlayer neighbors.

(b) Cross-section view for the same DFT calculated 2D heterostructure, showing charge pools form between nearest

neighbors. Since the atomic registry (or nearest neighbors) is determined by the moiré pattern, the charge pools form

a purely electronic lattice that follows the moiré pattern. (c) Table with the parameters from the experimentally

measured third-crystal and the DFT calculated interlayer electronic lattice, demonstrating their complete agreement.

Together the experimental and theoretical work suggest that the high energy electrons could be scattering from the

DFT-predicted interlayer charge pools.

Figure 6.4 shows DFT calculations for the predicted charge redistribution into the interlayer

region, due to the interlayer coupling in a 2D heterostructure (1-layer Bi2Se3 and 1-layer WS2 twist

to 10.9° - to match closely with the structure shown in Figure 6.2c). DFT calculations for other 2D

heterostructures, along with the experimental data, are detailed in section 6.3. Figure 6.4a is a

charge density plot of a plane in the interlayer region equidistant from the Bi2Se3 and TMD layers,

where red specifies higher charge density. The charge congregates between nearest neighbor

atoms (shown in the image), revealing possible bonding that forms between the layers. Figure

6.4b is a cross-section image of the same DFT calculation, showing both the Bi2Se3 and WS2 layers,

as well as the shape and relative displacement of the predicted interlayer charge pooling. The

charge pools form a lattice with long-range order, where the unit cell size and twist angle (see

Figure 6.4c) are in complete agreement with the experimentally observed new diffraction pattern

Page 149: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

149

in Figure 6.1e. Together these results provide strong evidence that the new set of spots seen in

our heterostructures are very likely a .planar lattice of purely electronic charge pools that reside

midway between the two parent crystals in the Bi2Se3/WS2 2D heterostructure twisted to 10.9°.

DFT calculations were also performed on two additional 2D heterostructures - Bi2Se3/MoSe2 (0°

twist) and Bi2Se3/MoS2 (0° twist). For all 2D heterostructures DFT calculations confirmed the

formation of an electronic lattice in the interlayer region, whose parameters (i.e. unit cell size and

twist angle) are in agreement with the moiré patterns of the corresponding parent crystals (see

section 6.3). Section 6.3 shows experimental and modeling data for Bi2Se3/MoSe2-2xS2x 2D

heterostructures at various twist angles, and a Bi2Se3/WS2 2D heterostructure at 4.7°, where the

experimental and modeling data are in near perfect agreement. The experimental observation of

new SAED patterns that match perfectly with various DFT-predicted electronic lattices across

multiple different 2D heterostructures and numerous twist angles, further verifies our

hypothesis, and suggests that the formation of a purely electronic 2-D lattice (at least at certain

twist angles) even at room temperature might be characteristic trait of the investigated family of

2D heterostructures.

To summarize, strong interactions at sites of high atomic registry in Bi2Se3/TMD heterostructures,

enable charges to overcome the interlayer separation and form a purely electronic lattice situated

between the two parent crystals, and concentrated at these sites, as revealed by additional SAED

patterns, and confirmed by DFT calculations. This is surprising because it is traditionally

assumed that such high energy (200 keV) SAED should only be able to reveal atomic structural

information, owing to the large scattering cross sections of the atoms,194 but are not sensitive to

bonding or weak charge distributions. Hence, the presence of the distinct diffraction spots

suggests significant charge redistribution into concentrated pools, as seen in our DFT

calculations. To the best of our knowledge, this is the first demonstration of high energy SAED

electrons being used to reveal non-atomic electronic scattering sites, and evidence of a purely

electronic lattice at the interface of 2D materials. Very likely, the formation of interlayer charge

pools between nearest interlayer neighbors play an important role in the interlayer hybridization

Page 150: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

150

and coupling in these systems. Our findings provide significant new insights into how 2D layers

interact, and access to the 2D electronic lattice could open up opportunities for researching the

physical properties of novel purely electronic periodic systems devoid of any atoms or defects.

Control of such systems will very likely lead to new developments in nanoelectronics and

optoelectronics, as well as other applications that depend on interfaces between 2D layers.

113,129,160,279

6.1 Diophantine equation and guidance on mathematically calculating the

electronic moiré lattice commensurate unit cell.

The below equations were used to mathematically calculate the moiré superlattice commensurate

unit cell parameters, where m, n, r, and s are the unit cell parameters.277,280 For the equations to be

satisfied, all the lattice parameters must be integers. Equation 1 is Diophantine and used to

calculate the moiré superlattice commensurate unit cell length. Equation 2 was used in Figure 6.5.

Equation 3 is used to calculate the relative rotation of the moiré superlattice relative to the Bi2Se3

crystal. Equation 1 and Equation 3 were used to calculate the parameters shown in Figure 2d and

all the figures in section 6.2.

𝐿𝑒𝑛𝑔𝑡ℎmoiré = 𝑎𝐵𝑖2𝑆𝑒3√𝑚2 + 𝑛2 − 𝑚𝑛 = 𝑎𝑇𝑀𝐷√𝑟2 + 𝑠2 − 𝑟𝑠 Equation

6.1

𝑙𝑎𝑡𝑡𝑖𝑐𝑒 𝑟𝑎𝑡𝑖𝑜 =𝑎𝐵𝑖2𝑆𝑒3

𝑎𝑇𝑀𝐷 Equation 6.2

𝑐𝑜𝑠(𝜑moiré,𝐵𝑖2𝑆𝑒3) =𝑚−

𝑛

2

√𝑚2+𝑛2−𝑚𝑛 Equation 6.3

Next we will discuss the mathematics of how a moiré pattern is formed because it provides

insight into the interlayer coupling. Moiré patterns only develop when two lattices interact

through an effective multiplication - simply summing two lattices does not produce a moiré

pattern.152,277,278,280 The Fourier transform is a linear operation: ℱ[𝑎𝑓(𝑥) + 𝑏𝑔(𝑥)] = 𝑎ℱ[𝑓(𝑥)] +

𝑏ℱ[𝑔(𝑥)], where 𝑓(𝑥) and 𝑔(𝑥) are functions. Summing two functions does not produce any new

frequencies. However, when two non-identical functions are multiplied, they will produce a new

Page 151: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

151

frequency: ℱ[𝑎𝑓(𝑥) · 𝑏𝑔(𝑥)] = 𝑎𝑏𝐹[𝑓(𝑥)] ∗ 𝐹[𝑔(𝑥)] ≠ 𝑎𝐹[𝑓(𝑥)] + 𝑏𝐹[𝑔(𝑥)], where ∗ is the

convolution operation. This new frequency is the moiré pattern. The presence of moiré

superlattice spots when using SAED indicates that the two materials are being “multiplied”, a

process facilitated through the interlayer coupling.

Figure 6.5 Guidance on calculating the moiré superlattice commensurate unit cell. (a) Matlab generated model of the

2D heterostructure in Figure 1e. (b) Fourier transform of the model in (a). (c) The Fourier transform was shaded orange

and then overlaid on the experimental SAED data shown in Figure 1e (white spots). The near perfect alignment

indicates that the model sufficiently captures the system. (d) Plot showing 2D heterostructure configurations that

produce smaller (m, n, r, and s are all less than 10) moiré superlattice commensurate unit cells, using the relative

rotation (or twist angle) and ratio of lattice constants (Eq.2) are input parameters.277,280 The intersection of dark or light

lines indicate regions where a commensurate unit cell is more like to found. See the corresponding references for more

thorough guidance on interpreting the graph.277,280 The red arrow highlights the location that corresponds to the 2D

heterostructure in (a), (b), and (c), indicating that the moiré superlattice parameters have been identified.

6.2 Data demonstrating that the electronic moiré lattice was observed in other

2D heterostructures

SAED is a powerful tool used to observe the reciprocal space of samples. If a periodic

arrangement of scattering sites with a sufficient high cross exists (e.g. a lattice), spots will

appear.194 Spots closer to the center correspond to crystals with a larger periodicity, which is why

Page 152: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

152

the monolayer TMD has the largest radius and the electronic moiré lattice has the smallest. Other

electron diffraction methods (e.g. convergent beam electron diffraction - CBED) have been used

to reveal bonding,281 and low energy electrons (~236eV) have revealed the moiré pattern in bilayer

graphene structures; however, both these situations involve different physics and scattering

mechanisms. CBED functions with different assumptions, and is able to reveal significantly more

information.

2D Structure moiré patterns have been previously shown using TEM imaging; however, the

electron-material interaction mechanisms between imaging and SAED are different, allowing

each method to provide unique information. TEM imaging is a different process involving both

inelastic and elastic scattering, as well as charge induced phase shifts. SAED, on the other hand,

only detects electrons elastically scattered at low angles (~0.6°), where bright spots are the product

of periodically spaced scattering sites.194 Observation of periodic bright spots in SAED is strong

evidence that a “lattice” of scattering sites exists. Further, the scattering cross section is dependent

on not only the total charge, but also the shape,194 suggesting the scattering sites have well-formed

shapes (i.e. the charge redistribution was not diffusive). These conclusions are in agreement with

previously published synchrotron X-ray diffraction experiments, which showed charge

accumulation between the layers of bulk TiS2, suggesting the “weak” van de Waals bonding

induces sufficient charge redistribution that leads to the formation of a polar bond.282

The observation of SAED spots using high energy electrons, suggests the electronic moiré lattice

has a relatively large scattering cross section. It has been previously shown that, when used in

conjunction with DFT, scattering data is able to measure the magnitude and shape of charge

redistribution.281,282 Although performing such a robust analysis is beyond the scope of this paper,

we are able to infer a relative range of scattering cross section values by comparing the relative

brightness (i.e. photon counts) of the SAED spots. The electronic moiré lattice SAED spots are 1-

6% the intensity of the monolayer TMD spots, suggesting an approximate range of values that

would be very large for non-atomic scattering sites. The scattering cross section is determined not

only by the magnitude of charge, but also by the distribution,170,194 where well-formed sites are

more likely to scatter than very diffuse clouds.

Page 153: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

153

Figure 6.6 Experimental, moiré modeling, and DFT data for a Bi2Se3/MoS2 2D heterostructure. Taken together, the data

suggests high energy electrons are diffracting off interlayer coupling induced scattering sites, which DFT predicts to

be located between nearest interlayer neighbors. (a) Real-space model of the 2D heterostructure in (b) using the

techniques described in section 6.2. (b) High-voltage (200keV) SAED experimental data with the Fourier transformed

model from (a) overlaid on top. The strong overlap of the spots suggests the model and experiment are in agreement,

and that the model captures the system sufficiently. (c) Table with the data summarized, demonstrating that theory

and experiment are in very good agreement. (d) DFT calculated interlayer charge distribution in the interlayer region

between both layers. (e) Cross-section view of a DFT calculated 2D heterostructure with the interlayer coupling

induced charge pools shown. The DFT calculated 2D heterostructure is different from the parameters in (d) due to

computational limitations; however, the calculated structure is very similar where the twist angle was made 0°, vice

1.7°, making the lattice parameters (3, 0, 4, 0 for m, n, r, and s, respectively). The DFT models do not contain the

experimentally measured twist angles of 1.4° and 1.7° for the Bi2Se3/MoSe2 and Bi2Se3/MoS2 2D heterostructures,

respectively, because the superlattice required at such twist angles was computationally excessive.

Page 154: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

154

Figure 6.7 Experimental, moiré modeling, and DFT data for a Bi2Se3/MoSe2 2D heterostructure. Taken together, the

data suggests high-voltage electrons are diffracting off interlayer coupling induced scattering sites, which DFT

predicts to be located between nearest interlayer neighbors. (a) Real-space model of the 2D heterostructure in (b)

using the techniques described in Section 6.2. (b) High-voltage (200keV) SAED experimental data with the Fourier

transformed model from (a) overlaid on top. The strong overlap of the spots suggests the model and experiment are

in agreement, and that the model captures the system sufficiently. (c) Table with the data summarized,

demonstrating that theory and experiment are in very good agreement. (d) DFT calculated interlayer charge

distribution in the interlayer region between both layers. (e) Cross-section view of a DFT calculated 2D

heterostructure with the interlayer coupling induced charge pools shown. The DFT calculated 2D heterostructure is

different from the parameters in (d) due to computational limitations; however, the calculated structure is very

similar where the twist angle was made 0°, vice 1.7°, making the lattice parameters (4, 0, 5, 0 for m, n, r, and s,

respectively).

Page 155: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

155

Figure 6.8 Experimental and moiré modeling images for different Bi2Se3/MoSe2-2xS2x 2D heterostructures at various

twist angles. The data demonstrates that the method can not only be applied to arbitrary twist angles, but can also be

used on alloy-based 2D structures. Despite the fact that the sulfur and selenium atoms appear to be randomly

distributed, the monolayer TMD still forms a crystal, and this crystal is able to form a significant interlayer coupling

with the Bi2Se3. Interestingly, the Bi2Se3 appeared to grow with slightly different lattice constants across different

monolayer TMDs, but kept the same lattice constant when growing on MoSe2-2xS2x, despite the fact that the MoSe2-

2xS2x lattice constant fluctuated. The fluctuation in MoSe2-2xS2x is believed to be due to the changing sulfur and

selenium concentrations. Photoluminescence values were always between signature monolayer MoS2 and MoSe2

values, but varied across samples, suggesting different crystals contained different ratios of sulfur to selenium, which

offers an explanation why the monolayer MoSe2-2xS2x lattice constant fluctuated. DFT calculations on the alloy 2D

heterostructures were not completed due to computational limitations. To the best of our knowledge, the sulfur and

selenium atoms are randomly distributed, therefore requiring very large supercells that can capture the random

placement.

Page 156: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

156

Figure 6.9 Experimental and moiré modeling images for a Bi2Se3/WS2 2D heterostructure (with a different twist angle

from that in the main text). The image is presented on its own to better show the moiré superlattice spots, and the

strong agreement between the model and the experimental data.

Figure 6.10 Experimental and moiré modeling images for Bi2Se3/WS2 2D heterostructures from Figure 2. A complete

explanation can be found in Figure 2. The information lattice information from solving the Diophantine equation is

included in this figure (i.e. the m, n, r, and s data).

Page 157: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

157

6.3 Brief explanation of double diffraction with respect to electron diffraction

As shown in this chapter, a third crystal was observed using SAED, and this crystal cannot be

correlated to any known atomic basis. Using supporting DFT results, it is argued that this

unknown crystal is due to periodic charge redistribution inside the interlayer region due to the

interlayer coupling. The SAED evidence is not a direct probe of the charge crystal, but an

indication that it exists. With this in mind, it could be that a different effect is responsible for the

crystal’s diffraction spots. For completeness of the work, a significant literature review was done

to search for other possible explanations, which could provide guidance as to the source of the

diffraction spots. The only report of diffraction spots that correspond to the moiré pattern found

were due to an effect called “double diffraction”. We applied this concept to our 2D

heterostructure system and determined that it does not apply. Despite the fact that it does not

apply, it is briefly discussed in this dissertation because it is the only other documented source of

moiré pattern diffraction spots, suggesting that the spots we observe are simply due to this effect.

Double diffraction is explained briefly in Figure 6.11 below. In summary, double diffraction is

when the top material of a heterostructure projects a diffraction pattern onto the bottom material.

As it projects the diffraction pattern, the bright spots could happen to overlap with certain atoms

of the lower material in a periodic fashion. An atom that is illuminated with a bright spot will

carry a weighting with a higher value than the neighboring atoms, suggesting it appears brighter

or more intense than the neighboring atoms. The periodically spaced atoms with additional

intensity will then produce their own diffraction pattern spots that correspond to their spacing.

The observation of double diffraction spots requires not only a heterostructure that is oriented at

the correct twist angle, but the electron wavelength must also correspond, so that the diffraction

onto the lower material matches (the diffraction pattern spacing is dependent on the initial

electron wavelength).

Page 158: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

158

Figure 6.11: Simplified description of double diffraction due to TEM electrons. Electrons are incident on a bulk 3D

heterostructure as a plane wave, where they begin to scatter and diffract from the upper material. If the upper material

is sufficiently large, the diffraction response will project onto the lower material (shown as the diffraction peaks in the

black box). Certain atoms are periodic intervals will overlap with the diffraction response, whereupon they will be

illuminated with an electron intensity. This illumination provides specific atoms at periodically spaced intervals with

a higher weighting, thereby inducing a disproportionate response from them and producing a new set of diffraction

spots.

Double diffraction is not possible with 2D heterostructures due to the small size of the sample

and the low scattering angle of the elastic (i.e. kinetic energy conserving) interaction, which

together prohibit neighboring atoms from interacting to produce a diffraction pattern on the

lower material. More specifically, double diffraction presents when the diffraction pattern of the

upper material is projected onto the lower material and the bright spots periodically align with

the atoms of the lower material. However, to project a diffraction pattern, the electrons scattering

from neighboring atoms must interact, which is not possible considering the atomically thin

nature of the samples. SAED only allows low scattering angles (0.6°) due to the elastic scattering

requirement, which requires approximately 100nm of distance before the scattered electrons of

neighboring atoms interact. With this in mind, it is highly unlikely that a diffraction pattern of

the upper material can be projected onto the lower material, suggesting that double diffraction is

not a probably explanation for the third crystal.

Page 159: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

159

Chapter 7: Conclusion and Discussion

This dissertation discusses methods to probe and manipulate the interlayer coupling in 2D

structures, using primarily bismuth selenide (Bi2Se3) and transition metal dichalcogenides (TMD)

2D heterostructures as a platform. Both Bi2Se3 and TMDs are layered materials with the same

hexagonal structure. Bi2Se3 is also a topological insulator whose properties are strongly

dependent on the number of layers. Monolayer TMDs are best known for being direct bandgap

with tightly bound excitons. As demonstrated in this dissertation, when Bi2Se3 and monolayer

TMDs are combined to form 2D heterostructures, a strong interlayer coupling is formed that

modifies the properties of both materials. The interlayer coupling is then manipulated using

focused electron beams in a vacuum, and, possibly, the controlled intercalation and

deintercalation of oxygen.

The properties of monolayer materials – or 2D materials - are often dramatically different than

their bulk counter parts, in part due to quantum confinement along the z-direction. For example,

monolayer MoS2 has a direct bandgap, while a bilayer MoS2 2D structure has an indirect bandgap,

indicating that simply stacking the second layer induces changes to the band structure. These

changes are a product of the interlayer coupling, which facilitates communication and interaction

between the layers. On the other hand, if the interlayer coupling could be diminished, thereby

disrupting the communication and interaction, the layers would become electronically

independent and gain the properties of their monolayer morphology. Hence, if the interlayer

coupling could be controllably modulated, the properties of materials could in principle be

manipulated and tailored for specific applications. All this inspires the following questions: What

are the physical changes induced by the interlayer coupling that induce changes to the properties?

Can the interlayer coupling be manipulated across a range, vice simply “on/off”? What are

primary factors that determine the interlayer coupling strength?

Page 160: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

160

The interlayer coupling in 2D structures induces charge redistribution, which can facilitate the

formation of bonds between the layers and affect the bonding within the material. The interlayer

coupling induced changes to bonding can be probed using Raman spectroscopy, which measures

the vibrational modes of the atoms. It was found that the intensity of the Raman modes of the

parent materials in a well-coupled 2D heterostructure are relatively, suggesting the charge

redistribution is affecting the bonding, possibly also due to the formation of new bonds. When

the interlayer coupling is diminished, the intensity of those same Raman modes increases, in

agreement with theory, and demonstrating a measurement technique to quantify the interlayer

coupling strength. The charge redistribution can also be studied using transmission electron

microscope (TEM) imaging and selected area electron diffraction (SAED), where the scattering of

electrons incident onto the sample is determined by the charge redistribution, allowing changes

induced by the interlayer coupling to be studied. These experimental results can then be

compared to density functional theory (DFT) calculations, which theoretically predict the induced

charge redistribution.

This dissertation describes two methods to controllably manipulate the interlayer coupling with

Bi2Se3/TMD 2D heterostructures. The first method uses a focused electron beam in a vacuum to

impart energy with spatial resolution. If sufficient energy is imparted, the interlayer bonds break

allowing the parent materials to become more electronically independent. The second method

uses both oxygen in the surrounding atmosphere and thermal energy to possibly controllably

intercalate and deintercalate oxygen between the layers. The intercalation of oxygen between

layers, or between 2D materials and their substrates, has been shown to diminish the interlayer

coupling strength, facilitating the 2D materials to behave “free standing” or electronically

independent. Further, this dissertation demonstrates that the interlayer coupling strength is not

simply “on/off”, but can be a range of values, suggesting it can be controllably manipulated to

engineer materials with application tailored properties.

The dissertation found that when Bi2Se3 is grown on monolayer TMDs, certain twist angles are

preferred, suggesting they are energetically more favorable. DFT calculations performed by Prof.

Page 161: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

161

Bansil’s group, collaborators specializing in theoretical condensed matter physics, predicted that

the twist angle and resulting moiré superlattice are critical in determining the energetic stability

and the formation of an interlayer charge crystal, in agreement with the experimental findings.

Further, it was found that the preferred twist angles are dependent on the underlying monolayer

TMD, suggesting that atomic chemistry is an influential factor in determining the interlayer

coupling strength.

Chapter 1 is a background and overview of monolayer materials and 2D structures. It discusses

the physics of 2D materials and provides a significant number of references for readers seeking

more detail. Chapter 1 was written for readers with no knowledge of 2D materials, and only a

general technical background, facilitating accessibility for the general population. This chapter

discuss the individual properties of monolayer TMDs, as well as monolayer to few layer Bi2Se3.

Previous work on the interlayer coupling are summarized, and noteworthy, high-impact papers

are discussed in more detail, providing a significant background into the primary topic of this

dissertation.

Chapter 2 is the instrumentation and synthesis of all the important techniques and equipment

used during this dissertation. The chapter is split into eleven sections, where each section is a

different technique or equipment operation. Each section provides a brief overview of the

method, why it is important to this dissertation, important details on how it was applied, and, in

some cases, a reference containing amplifying information.

Chapter 3 describes work using a focused electron beam to manipulate the twist angle, structure,

and interlayer interaction of Bi2Se3/TMD 2D heterostructures. A focused electron beam imparts

energy into the material, which when large enough will cause interlayer bonds to break. The

changes to the twist angle and structure were studied using SAED, while changes to the interlayer

coupling were studied using an SEM and laser setup with Raman and photoluminescence

spectroscopy capability. The interlayer coupling strength was also studied by analyzing the

statistical likelihood of different as-grown 2D heterostructures.

Page 162: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

162

Chapter 4 demonstrates that the interlayer coupling in Bi2Se3/MoS2 2D heterostructures can be

controllably manipulated using the partial pressure of oxygen in the surrounding atmosphere

and the application of energy. The results suggest that oxygen is being controllably intercalated

and deintercalated between the layers, thereby affecting the strength of the interlayer coupling.

Further, the interlayer coupling can be manipulated with submicron precision using a focused

laser.

Chapter 5 demonstrates color-tunable, photoluminescing pixels (PLPs) can be fabricated using a

family of Bi2Se3/TMD 2D heterostructures. The exciton dynamics in the monolayer TMDs can be

tuned by manipulating the interlayer coupling, where the photoluminescence peak position can

be controllably shifted. Analysis suggests that both the binding energy, as well as the ratio of the

trion/exciton, are inducing the shift in the peak position. This chapter shows that the same

behavior is observed in a family of Bi2Se3/TMD 2D heterostructures and is not confined to

Bi2Se3/MoS2, suggesting PLPs with a range of colors can be fabricated. The 2D heterostructures

demonstrated in this chapter are: Bi2Se3/MoS2, Bi2Se3/MoSe2, Bi2Se3/WS2, and Bi2Se3/MoSe2-2xS2x, an

alloy.

Chapter 6 discusses the possible formation of a charge crystal in the interlayer region, providing

insight into the interlayer coupling induced charge redistribution. Using SAED, the crystal

structure of the 2D heterostructure was studied, where three distinct crystals were observed:

Bi2Se3, the TMD, and a new larger lattice distinct from both parent materials. The new crystal

corresponded to the moiré superlattice formed by the parent crystals. The moiré pattern is

dependent on the lattice constants and the twist angle, and does not correspond to any known

atomic basis. SAED captures the elastic scattering of electrons at very low scattering angles, where

only periodic scattering sites with large cross sections are detected. Previous work indicates that

only atoms have a sufficiently large cross section to generate a signal; however, no atomic basis

corresponds to the new crystal. DFT calculations completed by Prof. Bansil’s group predict the

formation of a charge crystal in the interlayer region, in agreement with the experimental results.

Together these results suggest that the interlayer coupling is redistributing charge into

Page 163: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

163

concentrated pools in the interlayer region, which have a sufficiently large cross section to be

detected.

The appendix discusses non-Bi2Se3 2D structures fabricated, where data suggests the interlayer

coupling is weaker than those formed in Bi2Se3/TMD 2D heterostructures. The following 2D

structures are studied: graphene/MoS2, bilayer graphene, Sb2Se3/MoS2, and Bi2Te3/MoS2. The

appendix also contains programming code written in the Matlab language, and used to analyze

data in the dissertation.

Page 164: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

164

Chapter 8: Dissemination

8.1 Publications

Hennighausen Z.B., Lane C., Benabbas A., Mendez K., Eggenberger M., Champion P.,

Robinson J., Kar S. Oxygen-induced in-situ manipulation of the interlayer coupling and exciton

recombination in Bi2Se3/MoS2 2D heterostructures. ACS Applied Materials and Interfaces (11)

2019, DOI: 10.1021/acsami.9b02929

Hennighausen Z.B., Kar S. Probing the interlayer interaction between dissimilar 2D

heterostructures by in-situ rearrangement of their interface. 2D Materials (6) 2019, DOI:

10.1088/2053-1583/ab136e

Keyvani-Someh E., Hennighausen Z.B., Lee W., Igwe R.C., Krami M.E., Kar S., Fenniri H.

Organic Photovoltaics with Stacked Graphene Anodes. ACS Applied Energy Materials 2018,

DOI: 10.1021/acsaem.7b00020

Vargas A., Liu F., Lane C., Rubin D., Bilgin I., Hennighausen Z.B., DeCapua M., Bansil A., Kar

S. Tunable and laser- reconfigurable 2D heterocrystals obtained by epitaxial stacking of

crystallographically incommensurate Bi2Se3 and MoS2 atomic layers. Science Advances 2017,

DOI: 10.1126/sciadv.1601741

Poudel Y., Lim G., Moazzezi M., Hennighausen Z.B., Rostovstev Y., D’Souza F., Kar S., Neogi

A. Active control of coherent dynamics in hybrid plasmonics MoS2 monolayers with dressed

phonons. ACS Photonics 2019, DOI: 10.1021/acsphotonics.8b01436

Hennighausen, Z.B.; Lane, C.; Buda, I. G.; Mathur, V.; Bansil, A.; Kar, S. Evidence of a Purely

Electronic Two-Dimensional Lattice at the Interface of TMD/Bi2Se3 Heterostructures. Nanoscale

2019. https://doi.org/10.1039/C9NR04412D.

Hennighausen, Z.B.; Bilgin, I.; Casey, C.; Kar, S. Widely Tunable Bi2Se3/Transition Metal

Dichalcogenide 2D Heterostructures for Write-Read-Erase-Reuse Applications. 2D Materials.

2019, 6 (4), 041003. https://doi.org/10.1088/2053-1583/ab33b5.

Mathur V., M., Jaeger E., Talapatra S., Kar S., and Hennighausen Z.B. Computationally

Identified In Vivo Partial Restoration of Neural Activity Across Severed Spinal Cords Bridged

with Carbon Nanotubes. (Under Revision).

Poudel Y., Seetharaman S., Hennighausen Z.B., Kar S., D’Souza F., Neogi A. Interlayer exciton-

phonon coupling modulated absorption and emission from bulk GaN-monolayer MoS2

Heterostructure. (Under Review).

Page 165: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

165

Isherwood L., Hennighausen Z.B., Kar S., Baidak A. Heavy Ion-Induced Defect Production in

MoS2 Produced by Mechanical Exfoliation and Chemical Vapour Deposition. (Submitted).

8.2 Patents

Kar S., Vargas A., Liu F., Lane C.A., Rubin D., Bansil A., Buda G., Hennighausen Z.B.,

inventors.

Northeastern University, assignee. Tunable and Reconfigurable Atomically Thin

Heterostructures. United States of America PCT/US2017/048191. 2017 August 23

Kar S., Hennighausen Z.B., inventors. Northeastern University, assignee. Two Dimensional

Materials for Use in Ultra High Density Information Storage and Sensor Devices. United States

of America Provisional Patent Application. ADN/19815-504-INV-19032. 2018 December 18

8.3 Presentations - Conferences

Hennighausen, Z.B., et al., Direct observation and in-situ control of interfaces between

heterostructures of chemically and structurally distinct 2D materials. APS March Meeting 2019,

Session K15

Hennighausen, Z.B., et al., O2-induced in-situ manipulation of exciton recombination pathways

in 2D heterostructures: Submicron, intensity-programmable pixels with rapid write-read-erase

capability, as well as 2D O2-sensor applications. APS March Meeting 2018, Session F36

Hennighausen, Z.B., et al., Reconfigurable Electronics and Optics at the Interface between

Dissimilar 2D Materials. 2017 MRS Fall Boston, Session NM04.02

Hennighausen, Z.B., et al., 2D Heterocrystals: Vertical stacking of rotationally aligned

electronically, structurally and chemically dissimilar 2D materials. APS March Meeting 2017,

Session C32

8.4 Presentations – Invited Talks

Hennighausen, Z.B., Probing and Manipulating the Interlayer Coupling in Bi2Se3/Transition

Metal Dichalcogenide 2D Heterostructures: Site-Selectable, in-situ Tuning of

Photoluminescence. Ludwig-Maximilian-Universitaet (LMU). June 2019

Hennighausen, Z.B., Engineering the properties of 2D materials via in-situ manipulation of the

interlayer coupling in vertically stacked 2D structures. Naval Research Laboratory (NRL).

December 2018

Page 166: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

166

Hennighausen, Z.B., In-situ Interface Engineering between 2D Materials: Tunable and

reversible properties with submicron spatial control (laser pattern-writing) of vertically stacked

TMD-Bi2Se3 heterostructures. National Institute of Science and Technology (NIST). May 2018

Hennighausen, Z.B., Graduate School and Opportunities in Science for CG Officers: Cadet to

2D Materials Researcher. U.S. Coast Guard Academy Symposia. April 2017

Page 167: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

167

Chapter 9: Appendix

9.1 Non-Bi2Se3 2D Heterostructures

9.1.1 Graphene/MoS2 Heterostructures

Graphene was the first monolayer material reported,283 and monolayer MoS2 is believed to be the

first monolayer transition metal dichalcogenides reported. 44 Both these materials have incredible

properties independently and have experienced tremendous growth and research since they

were first reported.25 A natural next step has been to combine the two materials to utilize the

comparative advantages of each material.109,284–286 For example, graphene has an exceptionally

high conductance, and it is relatively transparent. Monolayer MoS2 is photosensitive due to its

direct bandgap.

During this Ph.D. (or Doktorarbeit) monolayer and bulk MoS2 was grown on top of graphene to

create MoS2/graphene 2D heterostructures. The MoS2 was found to grow epitaxial on the

graphene, suggesting that the two materials are interacting and that there is a non-zero interlayer

coupling. Previous work has shown that there is charge transfer between MoS2 and graphene that

is large enough to quench the monolayer MoS2 signature PL.287 Our results are in agreement,

where no PL was detected when MoS2 was grown on graphene. Raman spectroscopy studies

showed a clear MoS2 signature, suggesting it was present. Below are SEM images, showing clear

indications of MoS2. The differences in work function between different layers of MoS2 and

graphene, including their interaction with electrons, allows contrast to be observed and

differences to be identified. It can be seen that islands of different graphene layers grew, and that

MoS2 grew on top, where triangles are characteristic of MoS2, while graphene prefers to grow as

a hexagon.

Page 168: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

168

Figure 9.1 MoS2 grown on top of graphene to create MoS2/graphene 2D heterostructures. The darker locations are

graphene (see blue arrow), suggesting graphene interacts less with the SEM electron beam or has a higher work

function since less electrons are generated. The lighter areas are MoS2, suggesting more electrons are being scattered

or ejected. The red arrows highlight locations that are clearly MoS2 due to both their lighter shading and the triangular

nature. MoS2 generally grows as a triangular crystal. The MoS2 grew in the graphene cracks, suggesting the cracks are

nucleation sites. The green arrow shows a crack where MoS2 grew. Raman and photoluminescence spectroscopy show

clear monolayer MoS2 signatures, suggesting it is present. Backscattered imaging was used to take the following image.

Figure 9.2 MoS2 grown on graphene using the secondary electron SEM imaging mode. Blue arrow shows three layers

of graphene. The green arrow shows two layers of graphene. These are islands that grew in the chemical vapor

Page 169: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

169

deposition (CVD) process. The red arrow shows the MoS2 crystals. The remaining area is graphene. Secondary electron

imaging was used for this image.

As shown in both Figure 9.1 and Figure 9.2, the MoS2 grew in the cracks and grain boundaries

between graphene grains, suggesting that the grain boundaries a nucleation sites. The changes in

contrast and color labels between the two images is due to collecting backscattered electrons vs.

secondary electrons. Backscattered electrons are more dependent on the size of the nucleus,

which is why MoS2 is brighter in Figure 9.1. Figure 9.2 is collected secondary electrons which are

ejected depending on the work function of the material. Since graphene has a lower work function

- it is bandgap-less material – it will appear brighter, while MoS2, which has a large bandgap, will

be darker.

9.1.2 Graphene/Graphene Heterostructures

In the beginning of the Ph.D. (or Doktorarbeit) a significant amount of graphene was grown. A

project was completed in collaboration with the first author, Dr. Ehsan Keyvani-Someh, where

graphene was mechanically stacked and used to fabricate low-cost solar cells. The graphene

served as a conductive and transparent electrode. The work found that two layers of graphene

are the most ideal 2D structure when using them as a transparent electrode in the structure

fabricated. Two layers were the ideal balance between transparency and conductivity. As the

number of layers increases, the transparency decreases, while the conductivity increases. The

change in transparency and conductivity is non-linear as the layer count changes, suggesting the

interlayer coupling is influencing the properties. The work showed that low-cost solar cells can

be fabricated whose performance is lower than commercially available products, but are cost

competitive due to the low manufacturing costs. Manufacturing high-quality graphene is

difficult, whereas low-quality graphene is much easier, suggesting that research into low-quality

graphene technology is more commercially realistic. Below are images and figures taken with

permission from the following reference.17

Page 170: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

170

Figure 9.3 Artistic diagrams demonstrating mechanically transferring economical graphene produces a conducting

path. Growing economical graphene is often separated spatially and not connected between electrodes. However,

mechanically transferring two of these growths creates a connecting path between spatially separated electrodes, as it

is statistically very likely that the islands will overlap in a configuration that allows for electrons to travel between

well-spaced electrodes. The following images were taken with permission from the following reference.17

While predictably growing a single and continuous layer of graphene is difficult and expensive,

growing separated islands is more economical and can be done predictably. However, this

method produces graphene with a high-resistance as there is often not a direct path for the

electrons. In this work we showed that mechanically transferring two of these economical

growths on top of each other is statistically very likely to produce a well-connected path that has

a sufficiently low resistance. Vice the electrons traveling continuously in-plane from one electrode

to the other, this method conducts electrons through different graphene grains vertically stacked.

This method is only possible since the interlayer coupling between graphene layers is known to

be sufficiently low to facilitate electron transfer between the layers.

Page 171: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

171

Figure 9.4 Images of as-grown graphene mechanically transferred onto glass. (a) shows graphene grown as islands. (b)

Shows monolayer graphene with islands of bilayer grown on top. (c) Monolayer graphene where only Poly-methyl

methacrylate (PMMA) residue can be seen. Removing PMMA from graphene is known to be difficult as they form a

strong coupling. The scale bars are all 10μm. The following images were taken with permission from the following

reference.17

9.1.3 MoS2/Sb2Se3 Heterostructures

Antimony selenide, Sb2Se3, similar to Bi2Se3 in that antimony (Sb) is in the same column as

bismuth, suggesting that Sb2Se3 would have overlapping properties to Bi2Se3. Sb2Se3 was grown

on top of monolayer MoS2 with the objective of studying whether similar effects are observed as

the Bi2Se3/MoS2 2D heterostructures. It was found that despite the fact several nanometers of

Sb2Se3 were grown on top, suggesting that the monolayer MoS2 is fully covered, the signature PL

was not fully quenched, suggesting that the interlayer coupling is leaving the excitons

significantly intact. This indicates that either the interlayer coupling is weak, or it does not affect

the monolayer MoS2. The height of the Sb2Se3 was verified with AFM measurements, which are

not provided in this Chapter.

Page 172: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

172

Figure 9.5 Optical images of monolayer MoS2 prior to growth of Sb2Se3. (a) and (b) show MoS2 triangles, where the blue

spots in the middle are the nucleation sites. (c) and (d) show thin films, where rods or tubes of sulfurized MoO2 crystal

grew around (note the blue lines).

Figure 9.5 shows optical images of monolayer MoS2 prior to growth of Sb2Se3. The PL and Raman

measurements are signature of monolayer MoS2, suggesting it is of sufficient quality (Figure 9.7).

MoO2 rods and/or tubes that have been sulfurized can be seen, a consequence of the chemical

vapor deposition growth. Figure 9.6 are optical images after Sb2Se3 was grown on top, where the

perceived color changed from purple to teal. AFM scans showed that several nanometers of Sb2Se3

grew on top indicating that there was full coverage. It can be seen that the Sb2Se3 appear to prefer

to grow on the MoS2 substrate, vice the SiO2, suggesting that an interlayer coupling makes the

growth more energetically favorable. Figure 9.6d shows an area where Sb2Se3 grew on SiO2, and

then was exposed to a laser in air. The white lines are due to the laser exposure, suggesting that

it is affecting the Sb2Se3, possibly also due to oxygen in the air. This same effect was observed in

Bi2Se3 that was grown on SiO2.

Page 173: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

173

Figure 9.6 Optical images of heterostructures where Sb2Se3 was grown on top of monolayer MoS2. The growth modifies

the perceived color from purple (see Figure 9.5) to a teal color. Additionally, it appears as though the Sb2Se3 prefers to

grow on the MoS2, suggesting there is an interlayer interaction that makes this substrate more favorable for growth. (a)

and (b) are optical images of the heterostructures. (c) is a close-up of the same location in (a) and (b). (d) shows an area

where Sb2Se3 grew on bare SiO2. The patterns with the color white were “drawn” using a 488nm laser, suggesting that

the Sb2Se3 is sensitive to laser exposure in air. This is in agreement with the sensitivity of Bi2Se3 to oxygen and laser

exposure. Further, these same lines and patterns can be drawn into Bi2Se3 that is grown on SiO2.

Figure 9.7 Photoluminescence and Raman spectroscopy from Sb2Se3/MoS2 heterostructures. (a) PL of monolayer MoS2

and after Sb2Se3 was grown on that area. Interestingly, the PL shifts to a higher power, suggesting that the ratio of

excitons, or the binding energy, have been modified by the Sb2Se3. Despite the fact that several nanometers of Sb2Se3

were grown on top, the PL has only been reduced to ~25% of its pre-growth value, suggesting that the interlayer

interaction is not significantly modifying the exciton dynamics. (b) PL spectra have consecutive laser exposure in air,

where shifts in the peak energy and PL intensity are observed. Although the changes are not as pronounced as in the

Page 174: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

174

Bi2Se3/MoS2 2D heterostructures, the changes are relatively large, suggesting that alterations in the interlayer coupling

are happening. Such wild changes are not expected for laser exposure of a monolayer MoS2 crystal. (c) Raman spectra

of a Sb2Se3/MoS2 heterostructure before and after laser exposure. The only change detected is that the MoS2 peaks got

stronger, suggesting the interlayer coupling was weakened, in agreement with the Bi2Se3/MoS2 results.

Figure 9.7 show photoluminescence (PL) and Raman spectra of monolayer MoS2 and the

Sb2Se3/MoS2 heterostructure. Figure 9.7a shows the PL spectra of monolayer MoS2 and then the

Sb2Se3/MoS2 heterostructure, where a clear decrease in PL intensity is observed, but not nearly

the same quenching observed for Bi2Se3/MoS2 2D heterostructures. Interestingly, the PL peak

position is shifted higher, suggesting that either lower energy excitons (such as the trion) are

mitigated, or the binding energy of the excitons has increased. The shift in the PL peak suggests

that an influential interlayer coupling is present. Figure 9.7b the same Sb2Se3/MoS2

heterostructure after exposure to a laser in air. Although the same PL recovery is not observed,

significant changes in the PL peak position are, along with moderate changes in the PL intensity.

These changes are not observed in MoS2 when exposed to a laser in air, suggesting that changes

to the heterostructure are being affected. Figure 9.7c shows the PL spectra of as-grown and laser

exposed Sb2Se3/MoS2 heterostructure, where the Sb2Se3 peaks cannot be seen. Additionally, the

Raman spectra of MoS2 appears to increase in intensity with laser exposure, suggesting that the

interlayer coupling has been diminished, in agreement with the Bi2Se3/MoS2 2D heterostructure.

Page 175: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

175

Figure 9.8 TEM SAED images and TEM images of Sb2se3/MoS2 heterostructures. (a) and (b) are SAED images, where

the crystal structure of MoS2 can be clearly identified, but the Sb2Se3 crystal structure cannot be identified, suggesting

that it did not grow as a crystal. Further, the bright diffuse halo surrounding the beam blocker is indicative of

amorphous materials, which normally have periodicity that induces such patterns. (b) shows numerous MoS2 crystal

grains, where the spots are varying brightness. This is either due to folding, as shown in (d), or due to a polycrystalline

section, as shown in Figure 9.5. (c) and (d) are TEM images, where breaking and folding due to the transfer process can

be seen. The granulated nature of the heterostructure, indicates that Sb2Se3 is on top.

Figure 9.8 shows transmission electron microscope (TEM) select area electron diffraction (SAED)

images and TEM images. Figure 9.8a and Figure 9.8b show SAED images, where the MoS2 spots

can be clearly identified, but the Sb2Se3 spots are noticeably absent, despite the fact that several

nanometers of Sb2Se3 grew on top, suggesting it grew amorphous. Additionally, the bright diffuse

ring around the center is indicative of amorphous materials, which have semi periodicity that

produces such rings. Sb2Se3 is known to have a different crystal structure from either MoS2 or

Page 176: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

176

Bi2Se3, which is possibly why it cannot be identified and/or grew amorphous. Figure 9.8c and

Figure 9.8d show TEM images that have folding and tearing of the heterostructure from the

transfer process.

9.1.4 MoS2/Bi2Te3 Heterostructures

Bismuth telluride (Bi2Te3) was grown on top of monolayer MoS2 and then characterized using

optical microscopy, photoluminescence spectroscopy, and Raman spectroscopy. The results

indicate that Bi2Te3 and monolayer MoS2 likely form a very weak or near-zero interlayer coupling,

in contrast to the strong interlayer coupling observed between Bi2Se3 and MoS2.

Figure 9.9 Optical images of monolayer MoS2 and Bi2Te3 grown on monolayer MoS2. (a) and (b) are monolayer MoS2.

The triangular shapes suggest they are well-formed crystals. (c) Bi2Te3 was grown on top on the MoS2, and it grew as

gold bars/rods/tubes, in contrast to the Bi2Se3 or Sb2Se3 results, which grew continuous and smooth.

As shown in Figure 9.9, Bi2Te3 grew in a rod/tubular structure on the MoS2, in contrast to both

Bi2Se3 and Sb2Se3 which grew continuous and smooth on MoS2, suggesting that epitaxial growth

is not present. Further, the growth of Bi2Te3 appears to be the same whether on SiO2 or MoS2,

suggesting the interlayer coupling is not strong enough to modify or alter the growth dynamics.

Figure 9.9a and Figure 9.9b show optical images of as-grown monolayer and bulk MoS2. The

monolayer MoS2 is a light purple color, while the blue and bright spots are multilayer or bulk

MoS2. The triangular structure suggests that the MoS2 is crystalline. Figure 9.9c shows the Bi2Te3

grown on top, where the tubular/rod structure can be seen. The tubes appear to have grown

several microns in length.

Page 177: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

177

Figure 9.10 Photoluminescence and Raman spectra of Bi2Te3/MoS2 heterostructures. (a) PL spectra from monolayer

MoS2 and Bi2Te3 heterostructures, where a clear decrease in PL intensity is observed after Bi2Te3 was grown on top, but

not nearly to the same extent (or quenching) as with Bi2Se3/MoS2 2D heterostructures. The presence of a bright PL

spectra suggests that the interlayer coupling is weaker than that between Bi2Se3 and MoS2. There does not appear to be

a significant shift in the PL peak, suggesting the exciton ratio is approximately the same as in monolayer MoS2. (b)

Raman spectra of pre and post Bi2Te3 growth on monolayer MoS2, where no significant changes were detected. Bi2Te3

modes have demonstrated a weak intensity and appear at low Raman shift values, suggesting that the equipment may

not measure to sufficiently low values.

Figure 9.10 shows the PL and Raman spectra of monolayer MoS2 and a Bi2Te3/MoS2

heterostructure. Figure 9.10a shows the PL spectra, where a clear decrease in PL intensity is

observed to approximately a 15% of the original value. Although this is a significant decrease, it

is not close to the quenching observed in Bi2Se3/MoS2 2D heterostructures. The PL decrease is

possibly not due to an influential interlayer coupling, but by how much light is transmitted

through the upper Bi2Te3 material. The decrease in PL could also be due to a weak interlayer

coupling that facilitates partial charge transfer. Interestingly, no significant changes to the Raman

spectrum were recorded between pre and post Bi2Te3 growth, suggesting that the MoS2 modes

are unaffected. Further, there does not appear to be any new modes that might correspond to

Bi2Te3. This is not surprising as Bi2Te3 Raman modes have been documented to have a low

intensity and are at a low Raman shift of approximately 100 1/cm, suggesting that our equipment

may not be sensitive to detect them.288

Page 178: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

178

9.2 Programming Code

Below are codes for two programs that were especially important for the work presented in this

dissertation. Both codes are written in the Matlab language with comments kept to provide

amplifying information.

9.2.1 Fitting Lorentzian Functions to Exciton Photoluminescence Spectra

% Versuche zwei L Formeln mit meine nummer anzupassen

% % Define the Lorentzian % dualLorenz = @(x, a1, b1, a2, b2)lorenz(x, a1, b1)+lorenz(x, a2, b2); % % Parameterise the % objFn = @(p, x, y) sum( (y - dualLorenz(x, p(1), p(2), p(3), p(4))) .^ 2 ); % % Use fsolve to % p0 = [100 0.05 0.05 100]; % p = fsolve( @(p) objFn(p, X, Y), p0 );

%Import all the file names files=dir('*.csv');

%Sort by name. Make sure all the files are in the correct order before %running program [~,ind]=sort({files.name}); files=files(ind);

% Matrix for the final values % File name/ a/ b/ Constant/ gamma1/ gamma2/ Slope/ x01/x02/ Max1/ Max2/

MaxTot/ MaxeV/

values=zeros(length(files),11); valuesC = cell(length(files)+1,11); vgof2=zeros(length(files),2);

% % Make Column headers % valuesC(1,1) = {'File name'}; valuesC(1,2) = {'a'}; valuesC(1,3) = {'b'}; % valuesC(1,5) = {'gamma1'}; valuesC(1,6) = {'gamma2'}; valuesC(1,8) =

{'x01'}; % valuesC(1,9) = {'x02'}; valuesC(1,10) = {'Max1'}; valuesC(1,11) = {'Max2'}; % valuesC(1,12) = {'MaxTot'}; valuesC(1,13) = {'MaxeV'}; % valuesC(1,7) = {'linear Slope'}; valuesC(1,4) = {'linear constant'};

Page 179: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

179

% Make Column headers valuesC(1,1) = {'File name'}; valuesC(1,2) = {'a'}; valuesC(1,3) = {'b'}; valuesC(1,4) = {'gamma1'}; valuesC(1,5) = {'gamma2'}; valuesC(1,6) = {'x01'}; valuesC(1,7) = {'x02'}; valuesC(1,8) = {'Max1'}; valuesC(1,9) = {'Max2'}; valuesC(1,10) = {'MaxTot'}; valuesC(1,11) = {'MaxeV'};

% Array to store SSE values SSE = zeros(length(files),2);

% Script to loop through the files, find a fit, and then output the values for cnt=1:length(files) tmpFileVals = dlmread(files(cnt).name); % Store vals to a temp array valuesC(cnt+1,1)={files(cnt).name};

% Start Fitting algo x = tmpFileVals(:,1); % input these files into algorthm y = tmpFileVals(:,2);

% Initial parameters a = 100; b = 200; gamma1 = 0.05; gamma2 = 0.04; x01 = 1.82; x02 = 1.79; za = 10; zb = 90; start = [a, b, gamma1, gamma2, x01, x02, za, zb];

% Upper and lower bounds low = [0 0 0 0 1.7 1.7 -1000 -2000]; upp = [3000 3000 .15 .15 2.0 2.0 1000 2000];

% 'Robust', 'LAR', opcjedopasowania = fitoptions('Method', 'NonlinearLeastSquares',

'Lower',low, ... 'Startpoint', start, 'Upper', upp, 'MaxIter', 10000); typdopasowania = fittype('a*0.5*gamma1/(pi()*((x-

x01)^2+0.25*gamma1*gamma1))+b*0.5*gamma2/(pi()*((x-

x02)^2+0.25*gamma2*gamma2))+za*x+zb', 'options', opcjedopasowania); [fitobject,gof] =fit(x,y,typdopasowania); tmpCoefVal=coeffvalues(fitobject);

vgof2(cnt,1)=getfield(gof,'sse'); vgof2(cnt,2)=getfield(gof,'rsquare');

% Loop through all coeff values; store to larger value array for cnt2 = 1:6 values(cnt,cnt2)=tmpCoefVal(cnt2); valuesC(cnt+1,cnt2+1)={tmpCoefVal(cnt2)}; end

a=tmpCoefVal(1); b=tmpCoefVal(2); gamma1=tmpCoefVal(3); gamma2=tmpCoefVal(4); x01=tmpCoefVal(5); x02=tmpCoefVal(6); za = tmpCoefVal(7); zb = tmpCoefVal(8);

% Compute the Lorentzian Max values values(cnt,7)=a*2/(pi()*gamma1); values(cnt,8)=b*2/(pi()*gamma2);

Page 180: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

180

valuesC(cnt+1,8)={a*2/(pi()*gamma1)};

valuesC(cnt+1,9)={b*2/(pi()*gamma2)};

% Compute the overall Max value; anpassung is the fit anpassung1 = a.*0.5.*gamma1./(pi().*((x-x01).^2+0.25.*gamma1.*gamma1)); anpassung2 = b.*0.5.*gamma2./(pi().*((x-x02).^2+0.25.*gamma2.*gamma2)); anpassung = anpassung1+anpassung2+za*x+zb; [M,I] = max(anpassung); values(cnt,9)=M; values(cnt,10)=x(I); valuesC(cnt+1,10)={M}; valuesC(cnt+1,11)={x(I)};

% Put all trions and excitons in the correct columns if values(cnt,6) < values(cnt,5) %if values(cnt,7) < values(cnt,6) % switch first three values for cnt3 = 1:2:7 tmpVal=values(cnt,cnt3); values(cnt,cnt3)=values(cnt,cnt3+1); values(cnt,cnt3+1)=tmpVal;

tmpValC=valuesC(cnt+1,cnt3+1); valuesC(cnt+1,cnt3+1)=valuesC(cnt+1,cnt3+2); valuesC(cnt+1,cnt3+2)=tmpValC; end

% tmpVal=values(cnt,1); % values(cnt,1)=values(cnt,1+1); % values(cnt,1+1)=tmpVal; % tmpValC=valuesC(cnt+1,1+1); % valuesC(cnt+1,1+1)=valuesC(cnt+1,1+2); % valuesC(cnt+1,1+2)=tmpValC; % % tmpVal=values(cnt,4); % values(cnt,4)=values(cnt,4+1); % values(cnt,4+1)=tmpVal; % tmpValC=valuesC(cnt+1,4+1); % valuesC(cnt+1,4+1)=valuesC(cnt+1,4+2); % valuesC(cnt+1,4+2)=tmpValC; % % tmpVal=values(cnt,7); % values(cnt,7)=values(cnt,7+1); % values(cnt,7+1)=tmpVal; % tmpValC=valuesC(cnt+1,7+1); % valuesC(cnt+1,7+1)=valuesC(cnt+1,7+2); % valuesC(cnt+1,7+2)=tmpValC; % % tmpVal=values(cnt,9); % values(cnt,9)=values(cnt,9+1); % values(cnt,9+1)=tmpVal; % tmpValC=valuesC(cnt+1,9+1); % valuesC(cnt+1,9+1)=valuesC(cnt+1,9+2); % valuesC(cnt+1,9+2)=tmpValC;

Page 181: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

181

% tmpVal=values(cnt,1); % values(cnt,1)=values(cnt,1+1); % values(cnt,1+1)=tmpVal; % tmpValC=valuesC(cnt+1,1+1); % valuesC(cnt+1,1+1)=valuesC(cnt+1,1+2); % valuesC(cnt+1,1+2)=tmpValC; % for cnt3=4:2:8 % tmpVal=values(cnt,cnt3); % values(cnt,cnt3)=values(cnt,cnt3+1); % values(cnt,cnt3+1)=tmpVal; % tmpValC=valuesC(cnt+1,cnt3+1); % valuesC(cnt+1,cnt3+1)=valuesC(cnt+1,cnt3+2); % valuesC(cnt+1,cnt3+2)=tmpValC; % end

% % switch first three values % for cnt4 = 7:7 % tmpVal=values(cnt,cnt4); % values(cnt,cnt4)=values(cnt,cnt4+1); % values(cnt,cnt4+1)=tmpVal; % % tmpValC=valuesC(cnt+1,cnt4+1); % valuesC(cnt+1,cnt4+1)=valuesC(cnt,cnt4+2); % valuesC(cnt+1,cnt4+2)=tmpValC; % end

% Re-Compute the anpassung for making the graphs anpassung2 = a.*0.5.*gamma1./(pi().*((x-x01).^2+0.25.*gamma1.*gamma1)); anpassung1 = b.*0.5.*gamma2./(pi().*((x-x02).^2+0.25.*gamma2.*gamma2)); end

anpassung=anpassung1+anpassung2+za*x+zb;

% % Create jpeg files of all the figures % cntStr = num2str(cnt); % %fig=figure();set(gfc,'Visible', 'off');% suppresses the figure from

popping up % fig=figure();set(fig,'DefaultFigureVisible', 'off');% suppresses the

figure from popping up % plot(x,y,x,anpassung1,x,anpassung2,x,anpassung); % saveas(fig,cntStr,'jpeg'); % %figure();set(gcf,'Visible', 'on');% Allows the figure to appear again % %fig=figure();set(fig,'DefaultFigureVisible', 'on');% suppresses the

figure from popping up

% % Test the goodness of fit % SSE_m = sqrt((anpassung-y).^2); % SSE(cnt,1) = sum(SSE_m)/length(SSE_m);

% % Initial parameters % ap = Anpassp(cnt,1); bp = Anpassp(cnt,4); % gamma1p = Anpassp(cnt,2); gamma2p = Anpassp(cnt,5);

Page 182: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

182

% x01p = Anpassp(cnt,3); x02p = Anpassp(cnt,6); % % % Compute the overall Max value; anpassung is the fit % anpassung1p = ap.*0.5.*gamma1p./(pi().*((x-

x01p).^2+0.25.*gamma1p.*gamma1p)); % anpassung2p = bp.*0.5.*gamma2p./(pi().*((x-

x02p).^2+0.25.*gamma2p.*gamma2p)); % anpassungp = anpassung1p+anpassung2p; % % SSE_p = sqrt((anpassungp-y).^2); % SSE(cnt,2) = sum(SSE_p)/length(SSE_p); % % Create jpeg files of all the figures cntStr = num2str(cnt); %fig=figure();set(gfc,'Visible', 'off');% suppresses the figure from

popping up fig=figure();set(fig,'DefaultFigureVisible', 'off');% suppresses the

figure from popping up plot(x,y,x,anpassung1,x,anpassung2,x,za*x+zb,x,anpassung); saveas(fig,cntStr,'jpeg'); %figure();set(gcf,'Visible', 'on');% Allows the figure to appear again %fig=figure();set(fig,'DefaultFigureVisible', 'on');% suppresses the

figure from popping up

end

9.2.2 Raw Search for Commensurate Unit Cells and Moiré Patterns

Daten = zeros(7,11);

n_c=1; mx = 10; Xme_fin = 1.28; Xme_int = 1.23;

%1.27 and 1.34

for m = 0:1:mx for n = 0:1:mx for r = 0:1:mx for s = 0:1:mx diph_eq = (r*r+s*s-r*s)/(m*m+n*n-m*n);

% put the Xme and Xme_fin equations here if diph_eq > Xme_int*Xme_int && diph_eq< Xme_fin*Xme_fin &&

r>=s && m>=n Daten(n_c,1) = m; Daten(n_c,2) = n; Daten(n_c,3) = r; Daten(n_c,4) = s; Daten(n_c,5) = diph_eq; Daten(n_c,6) = 180*acos((m-n/2)/sqrt(m*m+n*n-m*n))/pi(); Daten(n_c,7) = abs(180*(acos((m-n/2)/sqrt(m*m+n*n-m*n))-

acos((r-s/2)/sqrt(r*r+s*s-r*s)))/pi());

Page 183: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

183

if Daten(n_c,7)>30 && Daten(n_c,7)<=60 Daten(n_c,7) = 60-Daten(n_c,7); elseif Daten(n_c,7)>60 Daten(n_c,7) = Daten(n_c,7)-60; end

Daten(n_c,8) = sqrt(m*m+n*n-m*n)*.418; %*.276 *.344

Daten(n_c,10) = m+n+m*n; Daten(n_c,11) = r+s+r*s;

n_c=n_c+1; end end end end end

% histogram(Daten(:,7),300)

% Plot % Xme_fin = 1.364; Xme_int = 1.252394; Xstep = 0.0001; theta_fin = 0.5310; theta_int = -0.0100; theta_step = 0.0001; theta_axis = theta_int:theta_step:theta_fin; theta_axis = theta_axis*180/pi(); x_me_axis = Xme_int:Xstep:Xme_fin;

% Create empty arrays x_tot_cnt = round((Xme_fin-Xme_int)/Xstep); theta_tot_cnt = round((theta_fin-theta_int)/theta_step); %moire_calc = zeros(x_tot_cnt, theta_tot_cnt);

%imagesc(theta_axis, x_me_axis, moire_calc) %colormap(gray)

hold on; Daten(:,5)=sqrt(Daten(:,5)); Daten(:,7)=Daten(:,7); y = transpose(Daten(:,5)); x = transpose(Daten(:,7));

sz = 700; % Assume that there is 1% error on either side c = ['k'];

scatter(x,y,sz,c)

title('') xlabel('') ylabel('') set(gca,'xticklabel',[],'yticklabel',[],'ytick',[1.26 1.28 1.30 1.32 1.34

1.36]) % % % title('Hws Moire Pattern - 3rd order - 2.1% Strain for Both') % % xlabel('Relative Crystal Rotation (Deg)')

Page 184: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

184

% % ylabel('x = Bi2Se3/TMD')

Page 185: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

185

References

(1) History of Carbon https://www.tf.uni-

kiel.de/matwis/amat/iss/kap_4/advanced/t4_2_1.html (accessed Feb 20, 2019).

(2) Mahan, G. D. Many-Particle Physics, 3rd ed.; Physics of Solids and Liquids; Springer US,

2000.

(3) Andre Geim - Nobel Lecture: Random Walk to Graphene

https://www.nobelprize.org/nobel_prizes/physics/laureates/2010/geim-lecture.html

(accessed Jul 26, 2018).

(4) Lee, C.; Wei, X.; Kysar, J. W.; Hone, J. Measurement of the Elastic Properties and

Intrinsic Strength of Monolayer Graphene. Science 2008, 321 (5887), 385–388.

https://doi.org/10.1126/science.1157996.

(5) Geim, A. K.; Novoselov, K. S. The Rise of Graphene. Nat. Mater. 2007, 6 (3), 183–191.

https://doi.org/10.1038/nmat1849.

(6) Balandin, A. A.; Ghosh, S.; Bao, W.; Calizo, I.; Teweldebrhan, D.; Miao, F.; Lau, C. N.

Superior Thermal Conductivity of Single-Layer Graphene. Nano Lett. 2008, 8 (3), 902–

907. https://doi.org/10.1021/nl0731872.

(7) Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.; Katsnelson, M. I.; Grigorieva, I.

V.; Dubonos, S. V.; Firsov, A. A. Two-Dimensional Gas of Massless Dirac Fermions in

Graphene. Nature 2005, 438 (7065), 197–200. https://doi.org/10.1038/nature04233.

(8) Rafiee, J.; Mi, X.; Gullapalli, H.; Thomas, A. V.; Yavari, F.; Shi, Y.; Ajayan, P. M.;

Koratkar, N. A. Wetting Transparency of Graphene. Nat. Mater. 2012, 11 (3), 217–222.

https://doi.org/10.1038/nmat3228.

(9) An, X.; Liu, F.; Jung, Y. J.; Kar, S. Tunable Graphene–Silicon Heterojunctions for

Ultrasensitive Photodetection. Nano Lett. 2013, 13 (3), 909–916.

https://doi.org/10.1021/nl303682j.

(10) Rodrigo, D.; Limaj, O.; Janner, D.; Etezadi, D.; Abajo, F. J. G. de; Pruneri, V.; Altug, H.

Mid-Infrared Plasmonic Biosensing with Graphene. Science 2015, 349 (6244), 165–168.

https://doi.org/10.1126/science.aab2051.

(11) Qian, Z.; Liu, F.; Hui, Y.; Kar, S.; Rinaldi, M. Graphene as a Massless Electrode for

Ultrahigh-Frequency Piezoelectric Nanoelectromechanical Systems. Nano Lett. 2015, 15

(7), 4599–4604. https://doi.org/10.1021/acs.nanolett.5b01208.

(12) Latychevskaia, T.; Escher, C.; Fink, H.-W. Moiré Structures in Twisted Bilayer Graphene

Studied by Transmission Electron Microscopy. Ultramicroscopy 2019, 197, 46–52.

https://doi.org/10.1016/j.ultramic.2018.11.009.

(13) Robinson, J. T.; Perkins, F. K.; Snow, E. S.; Wei, Z.; Sheehan, P. E. Reduced Graphene

Oxide Molecular Sensors. Nano Lett. 2008, 8 (10), 3137–3140.

https://doi.org/10.1021/nl8013007.

(14) Bae, S.; Kim, H.; Lee, Y.; Xu, X.; Park, J.-S.; Zheng, Y.; Balakrishnan, J.; Lei, T.; Ri Kim, H.;

Song, Y. I.; et al. Roll-to-Roll Production of 30-Inch Graphene Films for Transparent

Electrodes. Nat. Nanotechnol. 2010, 5 (8), 574–578. https://doi.org/10.1038/nnano.2010.132.

Page 186: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

186

(15) Das, T.; Sharma, B. K.; Katiyar, A. K.; Ahn, J.-H. Graphene-Based Flexible and Wearable

Electronics. J. Semicond. 2018, 39 (1), 011007. https://doi.org/10.1088/1674-

4926/39/1/011007.

(16) Zang, X.; Chen, Q.; Li, P.; He, Y.; Li, X.; Zhu, M.; Li, X.; Wang, K.; Zhong, M.; Wu, D.; et

al. Highly Flexible and Adaptable, All-Solid-State Supercapacitors Based on Graphene

Woven-Fabric Film Electrodes. Small 2014, 10 (13), 2583–2588.

https://doi.org/10.1002/smll.201303738.

(17) Keyvani-Someh, E.; Hennighausen, Z.; Lee, W.; Igwe, R. C. K.; Kramdi, M. E.; Kar, S.;

Fenniri, H. Organic Photovoltaics with Stacked Graphene Anodes. ACS Appl. Energy

Mater. 2018, 1 (1), 17–21. https://doi.org/10.1021/acsaem.7b00020.

(18) Hao, J.; Li, B.; Jung, H. Y.; Liu, F.; Hong, S.; Jung, Y. J.; Kar, S. Vapor-Phase-Gating-

Induced Ultrasensitive Ion Detection in Graphene and Single-Walled Carbon Nanotube

Networks. Adv. Mater. 29 (23), 1606883. https://doi.org/10.1002/adma.201606883.

(19) Merchant, C. A.; Healy, K.; Wanunu, M.; Ray, V.; Peterman, N.; Bartel, J.; Fischbein, M.

D.; Venta, K.; Luo, Z.; Johnson, A. T. C.; et al. DNA Translocation through Graphene

Nanopores. Nano Lett. 2010, 10 (8), 2915–2921. https://doi.org/10.1021/nl101046t.

(20) Ozturk, B.; de-Luna-Bugallo, A.; Panaitescu, E.; Chiaramonti, A. N.; Liu, F.; Vargas, A.;

Jiang, X.; Kharche, N.; Yavuzcetin, O.; Alnaji, M.; et al. Atomically Thin Layers of B–N–

C–O with Tunable Composition. Sci. Adv. 2015, 1 (6), e1500094.

https://doi.org/10.1126/sciadv.1500094.

(21) Hu, M.; Yao, Z.; Wang, X. Graphene-Based Nanomaterials for Catalysis. Ind. Eng. Chem.

Res. 2017, 56 (13), 3477–3502. https://doi.org/10.1021/acs.iecr.6b05048.

(22) Graphene Flagship http://graphene-flagship.eu:80/ (accessed Feb 20, 2019).

(23) Ren, S.; Li, M.; Smith, J. M.; DeTolla, L. J.; Furth, P. A. Low-Volume Jet Injection for

Intradermal Immunization in Rabbits. BMC Biotechnol. 2002, 2 (1), 10.

https://doi.org/10.1186/1472-6750-2-10.

(24) Zhang, H.; Huang, J.; Wang, Y.; Liu, R.; Huai, X.; Jiang, J.; Anfuso, C. Atomic Force

Microscopy for Two-Dimensional Materials: A Tutorial Review. Opt. Commun. 2018, 406,

3–17. https://doi.org/10.1016/j.optcom.2017.05.015.

(25) Zeng, M.; Xiao, Y.; Liu, J.; Yang, K.; Fu, L. Exploring Two-Dimensional Materials toward

the Next-Generation Circuits: From Monomer Design to Assembly Control. Chem. Rev.

2018. https://doi.org/10.1021/acs.chemrev.7b00633.

(26) 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), 116–130.

https://doi.org/10.1016/j.mattod.2016.10.002.

(27) Amani, M.; Lien, D.-H.; Kiriya, D.; Xiao, J.; Azcatl, A.; Noh, J.; Madhvapathy, S. R.;

Addou, R.; Kc, S.; Dubey, M.; et al. Near-Unity Photoluminescence Quantum Yield in

MoS2. Science 2015, 350 (6264), 1065–1068. https://doi.org/10.1126/science.aad2114.

(28) Mak, K. F.; Shan, J. Photonics and Optoelectronics of 2D Semiconductor Transition

Metal Dichalcogenides. Nat. Photonics 2016, 10 (4), 216–226.

https://doi.org/10.1038/nphoton.2015.282.

Page 187: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

187

(29) Mouri, S.; Miyauchi, Y.; Matsuda, K. Tunable Photoluminescence of Monolayer MoS2

via Chemical Doping. Nano Lett. 2013, 13 (12), 5944–5948.

https://doi.org/10.1021/nl403036h.

(30) Bilgin, I.; Raeliarijaona, A. S.; Lucking, M. C.; Hodge, S. C.; Mohite, A. D.; de Luna

Bugallo, A.; Terrones, H.; Kar, S. Resonant Raman and Exciton Coupling in High-

Quality Single Crystals of Atomically Thin Molybdenum Diselenide Grown by Vapor-

Phase Chalcogenization. ACS Nano 2018, 12 (1), 740–750.

https://doi.org/10.1021/acsnano.7b07933.

(31) Bilgin, I.; Liu, F.; Vargas, A.; Winchester, A.; Man, M. K. L.; Upmanyu, M.; Dani, K. M.;

Gupta, G.; Talapatra, S.; Mohite, A. D.; et al. Chemical Vapor Deposition Synthesized

Atomically Thin Molybdenum Disulfide with Optoelectronic-Grade Crystalline Quality.

ACS Nano 2015, 9 (9), 8822–8832. https://doi.org/10.1021/acsnano.5b02019.

(32) Chernikov, A.; van der Zande, A. M.; Hill, H. M.; Rigosi, A. F.; Velauthapillai, A.; Hone,

J.; Heinz, T. F. Electrical Tuning of Exciton Binding Energies in Monolayer WS2. Phys.

Rev. Lett. 2015, 115 (12), 126802. https://doi.org/10.1103/PhysRevLett.115.126802.

(33) Mak, K. F.; He, K.; Lee, C.; Lee, G. H.; Hone, J.; Heinz, T. F.; Shan, J. Tightly Bound

Trions in Monolayer MoS2. Nat. Mater. 2013, 12 (3), 207–211.

https://doi.org/10.1038/nmat3505.

(34) Xi, X.; Zhao, L.; Wang, Z.; Berger, H.; Forró, L.; Shan, J.; Mak, K. F. Strongly Enhanced

Charge-Density-Wave Order in Monolayer NbSe2. Nat. Nanotechnol. 2015, 10 (9), 765–

769. https://doi.org/10.1038/nnano.2015.143.

(35) Wang, H.; Huang, X.; Lin, J.; Cui, J.; Chen, Y.; Zhu, C.; Liu, F.; Zeng, Q.; Zhou, J.; Yu, P.;

et al. High-Quality Monolayer Superconductor NbSe 2 Grown by Chemical Vapour

Deposition. Nat. Commun. 2017, 8 (1), 394. https://doi.org/10.1038/s41467-017-00427-5.

(36) Ugeda, M. M.; Bradley, A. J.; Zhang, Y.; Onishi, S.; Chen, Y.; Ruan, W.; Ojeda-

Aristizabal, C.; Ryu, H.; Edmonds, M. T.; Tsai, H.-Z.; et al. Characterization of Collective

Ground States in Single-Layer NbSe2. Nat. Phys. 2016, 12 (1), 92–97.

https://doi.org/10.1038/nphys3527.

(37) Soumyanarayanan, A.; Yee, M. M.; He, Y.; Wezel, J. van; Rahn, D. J.; Rossnagel, K.;

Hudson, E. W.; Norman, M. R.; Hoffman, J. E. Quantum Phase Transition from

Triangular to Stripe Charge Order in NbSe2. Proc. Natl. Acad. Sci. 2013, 110 (5), 1623–

1627. https://doi.org/10.1073/pnas.1211387110.

(38) Geim, A. K.; Grigorieva, I. V. Van Der Waals Heterostructures. Nature 2013, 499 (7459),

419–425. https://doi.org/10.1038/nature12385.

(39) Ye, Y.; Wong, Z. J.; Lu, X.; Ni, X.; Zhu, H.; Chen, X.; Wang, Y.; Zhang, X. Monolayer

Excitonic Laser. Nat. Photonics 2015, 9 (11), 733–737.

https://doi.org/10.1038/nphoton.2015.197.

(40) Perkins, F. K.; Friedman, A. L.; Cobas, E.; Campbell, P. M.; Jernigan, G. G.; Jonker, B. T.

Chemical Vapor Sensing with Monolayer MoS2. Nano Lett. 2013, 13 (2), 668–673.

https://doi.org/10.1021/nl3043079.

(41) Yang, S.; Jiang, C.; Wei, S. Gas Sensing in 2D Materials. Appl. Phys. Rev. 2017, 4 (2),

021304. https://doi.org/10.1063/1.4983310.

Page 188: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

188

(42) Park, H.; Han, G.; Lee, S. W.; Lee, H.; Jeong, S. H.; Naqi, M.; AlMutairi, A.; Kim, Y. J.;

Lee, J.; Kim, W.; et al. Label-Free and Recalibrated Multilayer MoS2 Biosensor for Point-

of-Care Diagnostics. ACS Appl. Mater. Interfaces 2017, 9 (50), 43490–43497.

https://doi.org/10.1021/acsami.7b14479.

(43) Nayeri, M.; Moradinasab, M.; Fathipour, M. The Transport and Optical Sensing

Properties of MoS2, MoSe2, WS2 and WSe2 Semiconducting Transition Metal

Dichalcogenides. Semicond. Sci. Technol. 2018, 33 (2), 025002. https://doi.org/10.1088/1361-

6641/aaa168.

(44) Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis, A. Single-Layer MoS2

Transistors. Nat. Nanotechnol. 2011, 6 (3), 147–150.

https://doi.org/10.1038/nnano.2010.279.

(45) Costanzo, D.; Jo, S.; Berger, H.; Morpurgo, A. F. Gate-Induced Superconductivity in

Atomically Thin MoS2 Crystals. Nat. Nanotechnol. 2016, 11 (4), nnano.2015.314.

https://doi.org/10.1038/nnano.2015.314.

(46) Wu, W.; Wang, L.; Li, Y.; Zhang, F.; Lin, L.; Niu, S.; Chenet, D.; Zhang, X.; Hao, Y.;

Heinz, T. F.; et al. Piezoelectricity of Single-Atomic-Layer MoS2 for Energy Conversion

and Piezotronics. Nature 2014, 514 (7523), 470–474. https://doi.org/10.1038/nature13792.

(47) Poudel, Y.; Lim, G. N.; Moazzezi, M.; Hennighausen, Z.; Rostovtsev, Y.; D’Souza, F.;

Kar, S.; Neogi, A. Active Control of Coherent Dynamics in Hybrid Plasmonic MoS2

Monolayers with Dressed Phonons. ACS Photonics 2019.

https://doi.org/10.1021/acsphotonics.8b01436.

(48) Lu, J. M.; Zheliuk, O.; Leermakers, I.; Yuan, N. F. Q.; Zeitler, U.; Law, K. T.; Ye, J. T.

Evidence for Two-Dimensional Ising Superconductivity in Gated MoS2. Science 2015,

350 (6266), 1353–1357. https://doi.org/10.1126/science.aab2277.

(49) Waduge, P.; Bilgin, I.; Larkin, J.; Henley, R. Y.; Goodfellow, K.; Graham, A. C.; Bell, D.

C.; Vamivakas, N.; Kar, S.; Wanunu, M. Direct and Scalable Deposition of Atomically

Thin Low-Noise MoS2 Membranes on Apertures. ACS Nano 2015, 9 (7), 7352–7359.

https://doi.org/10.1021/acsnano.5b02369.

(50) Edwards, G. M.; Wilford, F. H.; Liu, X.; Hennighausen, L.; Djiane, J.; Streuli, C. H.

Regulation of Mammary Differentiation by Extracellular Matrix Involves Protein-

Tyrosine Phosphatases. J. Biol. Chem. 1998, 273 (16), 9495–9500.

https://doi.org/10.1074/jbc.273.16.9495.

(51) Xia, Y.; Qian, D.; Hsieh, D.; Wray, L.; Pal, A.; Lin, H.; Bansil, A.; Grauer, D.; Hor, Y. S.;

Cava, R. J.; et al. Observation of a Large-Gap Topological-Insulator Class with a Single

Dirac Cone on the Surface. Nat. Phys. 2009, 5 (6), 398–402.

https://doi.org/10.1038/nphys1274.

(52) Vargas, A.; Liu, F.; Kar, S. Giant Enhancement of Light Emission from Nanoscale Bi2Se3.

Appl. Phys. Lett. 2015, 106 (24), 243107. https://doi.org/10.1063/1.4922729.

(53) Sun, Y.; Cheng, H.; Gao, S.; Liu, Q.; Sun, Z.; Xiao, C.; Wu, C.; Wei, S.; Xie, Y. Atomically

Thick Bismuth Selenide Freestanding Single Layers Achieving Enhanced Thermoelectric

Energy Harvesting. J. Am. Chem. Soc. 2012, 134 (50), 20294–20297.

https://doi.org/10.1021/ja3102049.

Page 189: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

189

(54) Mamontov, E. Fast Oxygen Diffusion in Bismuth Oxide Probed by Quasielastic Neutron

Scattering. Solid State Ion. 2016, 296, 158–162. https://doi.org/10.1016/j.ssi.2016.09.022.

(55) Green, A. J.; Dey, S.; An, Y. Q.; O’Brien, B.; O’Mullane, S.; Thiel, B.; Diebold, A. C.

Surface Oxidation of the Topological Insulator Bi2Se3. J. Vac. Sci. Technol. A 2016, 34 (6),

061403. https://doi.org/10.1116/1.4964637.

(56) Edmonds, M. T.; Hellerstedt, J. T.; Tadich, A.; Schenk, A.; O’Donnell, K. M.; Tosado, J.;

Butch, N. P.; Syers, P.; Paglione, J.; Fuhrer, M. S. Stability and Surface Reconstruction of

Topological Insulator Bi2Se3 on Exposure to Atmosphere. J. Phys. Chem. C 2014, 118 (35),

20413–20419. https://doi.org/10.1021/jp506089b.

(57) Ambrosi, A.; Sofer, Z.; Luxa, J.; Pumera, M. Exfoliation of Layered Topological

Insulators Bi2Se3 and Bi2Te3 via Electrochemistry. ACS Nano 2016, 10 (12), 11442–11448.

https://doi.org/10.1021/acsnano.6b07096.

(58) Wang, X.; Bian, G.; Miller, T.; Chiang, T.-C. Fragility of Surface States and Robustness of

Topological Order in Bi2Se3 against Oxidation. Phys. Rev. Lett. 2012, 108 (9), 096404.

https://doi.org/10.1103/PhysRevLett.108.096404.

(59) Kong, D.; Cha, J. J.; Lai, K.; Peng, H.; Analytis, J. G.; Meister, S.; Chen, Y.; Zhang, H.-J.;

Fisher, I. R.; Shen, Z.-X.; et al. Rapid Surface Oxidation as a Source of Surface

Degradation Factor for Bi2Se3. ACS Nano 2011, 5 (6), 4698–4703.

https://doi.org/10.1021/nn200556h.

(60) Chen, Y. L.; Chu, J.-H.; Analytis, J. G.; Liu, Z. K.; Igarashi, K.; Kuo, H.-H.; Qi, X. L.; Mo,

S. K.; Moore, R. G.; Lu, D. H.; et al. Massive Dirac Fermion on the Surface of a

Magnetically Doped Topological Insulator. Science 2010, 329 (5992), 659–662.

https://doi.org/10.1126/science.1189924.

(61) Tian, W.; Yu, W.; Shi, J.; Wang, Y. The Property, Preparation and Application of

Topological Insulators: A Review. Materials 2017, 10 (7).

https://doi.org/10.3390/ma10070814.

(62) Music, D.; Chang, K.; Schmidt, P.; Braun, F. N.; Heller, M.; Hermsen, S.; Pöllmann, P. J.;

Schulzendorff, T.; Wagner, C. On Atomic Mechanisms Governing the Oxidation of

Bi2Te3. J. Phys. Condens. Matter 2017, 29 (48), 485705. https://doi.org/10.1088/1361-

648X/aa945f.

(63) Chen, Y. L.; Analytis, J. G.; Chu, J.-H.; Liu, Z. K.; Mo, S.-K.; Qi, X. L.; Zhang, H. J.; Lu, D.

H.; Dai, X.; Fang, Z.; et al. Experimental Realization of a Three-Dimensional Topological

Insulator, Bi2Te3. Science 2009, 325 (5937), 178–181.

https://doi.org/10.1126/science.1173034.

(64) Li, Y.-J.; Sun, Q.-Q.; Chen, L.; Zhou, P.; Wang, P.-F.; Ding, S.-J.; Zhang, D. W. Hexagonal

Boron Nitride Intercalated Multi-Layer Graphene: A Possible Ultimate Solution to Ultra-

Scaled Interconnect Technology. AIP Adv. 2012, 2 (1), 012191.

https://doi.org/10.1063/1.3701267.

(65) Yakovkin, I. N. Dirac Cones in Graphene, Interlayer Interaction in Layered Materials,

and the Band Gap in MoS2. Crystals 2016, 6 (11), 143.

https://doi.org/10.3390/cryst6110143.

Page 190: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

190

(66) 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), 43–50. https://doi.org/10.1038/nature26160.

(67) Cao, Y.; Fatemi, V.; Demir, A.; Fang, S.; Tomarken, S. L.; Luo, J. Y.; Sanchez-Yamagishi,

J. D.; Watanabe, K.; Taniguchi, T.; Kaxiras, E.; et al. Correlated Insulator Behaviour at

Half-Filling in Magic-Angle Graphene Superlattices. Nature 2018, 556 (7699), 80–84.

https://doi.org/10.1038/nature26154.

(68) Morozov, S. V.; Novoselov, K. S.; Katsnelson, M. I.; Schedin, F.; Elias, D. C.; Jaszczak, J.

A.; Geim, A. K. Giant Intrinsic Carrier Mobilities in Graphene and Its Bilayer. Phys. Rev.

Lett. 2008, 100 (1), 016602. https://doi.org/10.1103/PhysRevLett.100.016602.

(69) Frank, I. W.; Tanenbaum, D. M.; van der Zande, A. M.; McEuen, P. L. Mechanical

Properties of Suspended Graphene Sheets. J. Vac. Sci. Technol. B Microelectron. Nanometer

Struct. Process. Meas. Phenom. 2007, 25 (6), 2558–2561. https://doi.org/10.1116/1.2789446.

(70) Kuzmenko, A. B.; Benfatto, L.; Cappelluti, E.; Crassee, I.; van der Marel, D.; Blake, P.;

Novoselov, K. S.; Geim, A. K. Gate Tunable Infrared Phonon Anomalies in Bilayer

Graphene. Phys. Rev. Lett. 2009, 103 (11), 116804.

https://doi.org/10.1103/PhysRevLett.103.116804.

(71) Ghosh, S.; Bao, W.; Nika, D. L.; Subrina, S.; Pokatilov, E. P.; Lau, C. N.; Balandin, A. A.

Dimensional Crossover of Thermal Transport in Few-Layer Graphene. Nat. Mater. 2010,

9 (7), 555–558. https://doi.org/10.1038/nmat2753.

(72) Nair, R. R.; Blake, P.; Grigorenko, A. N.; Novoselov, K. S.; Booth, T. J.; Stauber, T.; Peres,

N. M. R.; Geim, A. K. Fine Structure Constant Defines Visual Transparency of Graphene.

Science 2008, 320 (5881), 1308–1308. https://doi.org/10.1126/science.1156965.

(73) Bunch, J. S.; Verbridge, S. S.; Alden, J. S.; van der Zande, A. M.; Parpia, J. M.; Craighead,

H. G.; McEuen, P. L. Impermeable Atomic Membranes from Graphene Sheets. Nano Lett.

2008, 8 (8), 2458–2462. https://doi.org/10.1021/nl801457b.

(74) Elias, D. C.; Nair, R. R.; Mohiuddin, T. M. G.; Morozov, S. V.; Blake, P.; Halsall, M. P.;

Ferrari, A. C.; Boukhvalov, D. W.; Katsnelson, M. I.; Geim, A. K.; et al. Control of

Graphene’s Properties by Reversible Hydrogenation: Evidence for Graphane. Science

2009, 323 (5914), 610–613. https://doi.org/10.1126/science.1167130.

(75) Xia, F.; Farmer, D. B.; Lin, Y.; Avouris, P. Graphene Field-Effect Transistors with High

On/Off Current Ratio and Large Transport Band Gap at Room Temperature. Nano Lett.

2010, 10 (2), 715–718. https://doi.org/10.1021/nl9039636.

(76) Kar, S.; Nguyen, V. L.; Mohapatra, D. R.; Lee, Y. H.; Sood, A. K. Ultrafast Spectral

Photoresponse of Bilayer Graphene: Optical Pump–Terahertz Probe Spectroscopy. ACS

Nano 2018, 12 (2), 1785–1792. https://doi.org/10.1021/acsnano.7b08555.

(77) Wang, C.-R.; Lu, W.-S.; Hao, L.; Lee, W.-L.; Lee, T.-K.; Lin, F.; Cheng, I.-C.; Chen, J.-Z.

Enhanced Thermoelectric Power in Dual-Gated Bilayer Graphene. Phys. Rev. Lett. 2011,

107 (18), 186602. https://doi.org/10.1103/PhysRevLett.107.186602.

(78) Yan, H.; Li, X.; Chandra, B.; Tulevski, G.; Wu, Y.; Freitag, M.; Zhu, W.; Avouris, P.; Xia,

F. Tunable Infrared Plasmonic Devices Using Graphene/Insulator Stacks. Nat.

Nanotechnol. 2012, 7 (5), 330–334. https://doi.org/10.1038/nnano.2012.59.

Page 191: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

191

(79) Yan, J.; Kim, M.-H.; Elle, J. A.; Sushkov, A. B.; Jenkins, G. S.; Milchberg, H. M.; Fuhrer,

M. S.; Drew, H. D. Dual-Gated Bilayer Graphene Hot-Electron Bolometer. Nat.

Nanotechnol. 2012, 7 (7), 472–478. https://doi.org/10.1038/nnano.2012.88.

(80) Sugawara, K.; Kanetani, K.; Sato, T.; Takahashi, T. Fabrication of Li-Intercalated Bilayer

Graphene. AIP Adv. 2011, 1 (2), 022103. https://doi.org/10.1063/1.3582814.

(81) Kühne, M.; Börrnert, F.; Fecher, S.; Ghorbani-Asl, M.; Biskupek, J.; Samuelis, D.;

Krasheninnikov, A. V.; Kaiser, U.; Smet, J. H. Reversible Superdense Ordering of

Lithium between Two Graphene Sheets. Nature 2018, 564 (7735), 234.

https://doi.org/10.1038/s41586-018-0754-2.

(82) Kühne, M.; Paolucci, F.; Popovic, J.; Ostrovsky, P. M.; Maier, J.; Smet, J. H. Ultrafast

Lithium Diffusion in Bilayer Graphene. Nat. Nanotechnol. 2017, 12 (9), 895–900.

https://doi.org/10.1038/nnano.2017.108.

(83) Gong, L.; Young, R. J.; Kinloch, I. A.; Riaz, I.; Jalil, R.; Novoselov, K. S. Optimizing the

Reinforcement of Polymer-Based Nanocomposites by Graphene. ACS Nano 2012, 6 (3),

2086–2095. https://doi.org/10.1021/nn203917d.

(84) Chen, M.; Hsu, C.; Hsueh, T. Fabrication of Humidity Sensor Based on Bilayer

Graphene. IEEE Electron Device Lett. 2014, 35 (5), 590–592.

https://doi.org/10.1109/LED.2014.2310741.

(85) Li, G.; Luican, A.; Santos, J. M. B. L. dos; Neto, A. H. C.; Reina, A.; Kong, J.; Andrei, E. Y.

Observation of Van Hove Singularities in Twisted Graphene Layers. Nat. Phys. 2009, 6

(2), nphys1463. https://doi.org/10.1038/nphys1463.

(86) Dean, C. R.; Wang, L.; Maher, P.; Forsythe, C.; Ghahari, F.; Gao, Y.; Katoch, J.; Ishigami,

M.; Moon, P.; Koshino, M.; et al. Hofstadter’s Butterfly and the Fractal Quantum Hall

Effect in Moire Superlattices. Nature 2013, 497 (7451), 598–602.

https://doi.org/10.1038/nature12186.

(87) 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), 1427–1430. https://doi.org/10.1126/science.1237240.

(88) Ju, L.; Shi, Z.; Nair, N.; Lv, Y.; Jin, C.; Velasco Jr, J.; Ojeda-Aristizabal, C.; Bechtel, H. A.;

Martin, M. C.; Zettl, A.; et al. Topological Valley Transport at Bilayer Graphene Domain

Walls. Nature 2015, 520 (7549), 650–655. https://doi.org/10.1038/nature14364.

(89) Koren, E.; Leven, I.; Lörtscher, E.; Knoll, A.; Hod, O.; Duerig, U. Coherent

Commensurate Electronic States at the Interface between Misoriented Graphene Layers.

Nat. Nanotechnol. 2016, 11 (9), 752–757. https://doi.org/10.1038/nnano.2016.85.

(90) Basu, J.; RoyChaudhuri, C. Attomolar Sensitivity of FET Biosensor Based on Smooth and

Reliable Graphene Nanogrids. IEEE Electron Device Lett. 2016, 37, 492–495.

(91) Rozhkov, A. V.; Sboychakov, A. O.; Rakhmanov, A. L.; Nori, F. Electronic Properties of

Graphene-Based Bilayer Systems. Phys. Rep. 2016, 648, 1–104.

https://doi.org/10.1016/j.physrep.2016.07.003.

(92) McCann, E.; Koshino, M. The Electronic Properties of Bilayer Graphene. Rep. Prog. Phys.

2013, 76 (5), 056503. https://doi.org/10.1088/0034-4885/76/5/056503.

Page 192: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

192

(93) Padilha, J. E.; Peelaers, H.; Janotti, A.; Van de Walle, C. G. Nature and Evolution of the

Band-Edge States in ${\mathrm{MoS}}_{2}$: From Monolayer to Bulk. Phys. Rev. B 2014,

90 (20), 205420. https://doi.org/10.1103/PhysRevB.90.205420.

(94) Scheuschner, N.; Ochedowski, O.; Kaulitz, A.-M.; Gillen, R.; Schleberger, M.; Maultzsch,

J. Photoluminescence of Freestanding Single- and Few-Layer ${\text{MoS}}_{2}$. Phys.

Rev. B 2014, 89 (12), 125406. https://doi.org/10.1103/PhysRevB.89.125406.

(95) Kim, J. G.; Yun, W. S.; Jo, S.; Lee, J.; Cho, C.-H. Effect of Interlayer Interactions on

Exciton Luminescence in Atomic-Layered MoS2 Crystals. Sci. Rep. 2016, 6, 29813.

https://doi.org/10.1038/srep29813.

(96) Moody, G.; Schaibley, J.; Xu, X. Exciton Dynamics in Monolayer Transition Metal

Dichalcogenides [Invited]. JOSA B 2016, 33 (7), C39–C49.

https://doi.org/10.1364/JOSAB.33.000C39.

(97) Shi, H.; Yan, R.; Bertolazzi, S.; Brivio, J.; Gao, B.; Kis, A.; Jena, D.; Xing, H. G.; Huang, L.

Exciton Dynamics in Suspended Monolayer and Few-Layer MoS2 2D Crystals. ACS

Nano 2013, 7 (2), 1072–1080. https://doi.org/10.1021/nn303973r.

(98) Yu, J. H.; Lee, H. R.; Hong, S. S.; Kong, D.; Lee, H.-W.; Wang, H.; Xiong, F.; Wang, S.;

Cui, Y. Vertical Heterostructure of Two-Dimensional MoS2 and WSe2 with Vertically

Aligned Layers. Nano Lett. 2015, 15 (2), 1031–1035. https://doi.org/10.1021/nl503897h.

(99) Yu, X.; Rahmanudin, A.; Jeanbourquin, X. A.; Tsokkou, D.; Guijarro, N.; Banerji, N.;

Sivula, K. Hybrid Heterojunctions of Solution-Processed Semiconducting 2D Transition

Metal Dichalcogenides. ACS Energy Lett. 2017, 2 (2), 524–531.

https://doi.org/10.1021/acsenergylett.6b00707.

(100) Yu, X.; Sivula, K. Layered 2D Semiconducting Transition Metal Dichalcogenides for

Solar Energy Conversion. Curr. Opin. Electrochem. 2017, 2 (1), 97–103.

https://doi.org/10.1016/j.coelec.2017.03.007.

(101) Avsar, A.; Tan, J. Y.; Taychatanapat, T.; Balakrishnan, J.; Koon, G. K. W.; Yeo, Y.; Lahiri,

J.; Carvalho, A.; Rodin, A. S.; O’Farrell, E. C. T.; et al. Spin–Orbit Proximity Effect in

Graphene. Nat. Commun. 2014, 5, 4875. https://doi.org/10.1038/ncomms5875.

(102) Velusamy, D. B.; Kim, R. H.; Cha, S.; Huh, J.; Khazaeinezhad, R.; Kassani, S. H.; Song,

G.; Cho, S. M.; Cho, S. H.; Hwang, I.; et al. Flexible Transition Metal Dichalcogenide

Nanosheets for Band-Selective Photodetection. Nat. Commun. 2015, 6, 8063.

https://doi.org/10.1038/ncomms9063.

(103) Vargas, A.; Liu, F.; Lane, C.; Rubin, D.; Bilgin, I.; Hennighausen, Z.; DeCapua, M.;

Bansil, A.; Kar, S. Tunable and Laser-Reconfigurable 2D Heterocrystals Obtained by

Epitaxial Stacking of Crystallographically Incommensurate Bi2Se3 and MoS2 Atomic

Layers. Sci. Adv. 2017, 3 (7), e1601741. https://doi.org/10.1126/sciadv.1601741.

(104) Hennighausen, Z.; Lane, C.; Benabbas, A.; Mendez, K.; Eggenberger, M.; Champion, P.

M.; Robinson, J. T.; Bansil, A.; Kar, S. Oxygen-Induced in-Situ Manipulation of the

Interlayer Coupling and Exciton Recombination in Bi2Se3/MoS2 2D Heterostructures.

ArXiv190101310 Cond-Mat 2019.

(105) Baugher, B. W. H.; Churchill, H. O. H.; Yang, Y.; Jarillo-Herrero, P. Optoelectronic

Devices Based on Electrically Tunable p–n Diodes in a Monolayer Dichalcogenide. Nat.

Nanotechnol. 2014, 9 (4), 262–267. https://doi.org/10.1038/nnano.2014.25.

Page 193: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

193

(106) Poudel, Y.; Lim, G. N.; Moazzezi, M.; Hennighausen, Z.; Rostovtsev, Y.; DSouza, F.; Kar,

S.; Neogi, A. Active Control of Coherent Dynamics in Hybrid Plasmonic MoS2

Monolayers with Dressed Phonons. ArXiv181002056 Cond-Mat 2018.

(107) Xiong, F.; Wang, H.; Liu, X.; Sun, J.; Brongersma, M.; Pop, E.; Cui, Y. Li Intercalation in

MoS2: In Situ Observation of Its Dynamics and Tuning Optical and Electrical Properties.

Nano Lett. 2015, 15 (10), 6777–6784. https://doi.org/10.1021/acs.nanolett.5b02619.

(108) Joshi, V. K. Spintronics: A Contemporary Review of Emerging Electronics Devices. Eng.

Sci. Technol. Int. J. 2016, 19 (3), 1503–1513. https://doi.org/10.1016/j.jestch.2016.05.002.

(109) Loan, P. T. K.; Zhang, W.; Lin, C.-T.; Wei, K.-H.; Li, L.-J.; Chen, C.-H. Graphene/MoS2

Heterostructures for Ultrasensitive Detection of DNA Hybridisation. Adv. Mater. 2014,

26 (28), 4838–4844. https://doi.org/10.1002/adma.201401084.

(110) Lv, X.; Wei, W.; Zhao, P.; Li, J.; Huang, B.; Dai, Y. Tunable Schottky Contacts in

MSe2/NbSe2 (M = Mo and W) Heterostructures and Promising Application Potential in

Field-Effect Transistors. Phys. Chem. Chem. Phys. 2018, 20 (3), 1897–1903.

https://doi.org/10.1039/C7CP07546D.

(111) Wang, Z.; Zhang, T.; Ding, M.; Dong, B.; Li, Y.; Chen, M.; Li, X.; Huang, J.; Wang, H.;

Zhao, X.; et al. Electric-Field Control of Magnetism in a Few-Layered van Der Waals

Ferromagnetic Semiconductor. Nat. Nanotechnol. 2018, 13 (7), 554.

https://doi.org/10.1038/s41565-018-0186-z.

(112) Yang, H.; Ouyang, W.; Yan, X.; Li, Z.; Yu, R.; Yuan, W.; Luo, J.; Zhu, J. Bilayer MoS2

Quantum Dots with Tunable Magnetism and Spin. AIP Adv. 2018, 8 (11), 115103.

https://doi.org/10.1063/1.5045531.

(113) Song, T.; Cai, X.; Tu, M. W.-Y.; Zhang, X.; Huang, B.; Wilson, N. P.; Seyler, K. L.; Zhu, L.;

Taniguchi, T.; Watanabe, K.; et al. Giant Tunneling Magnetoresistance in Spin-Filter van

Der Waals Heterostructures. Science 2018, eaar4851.

https://doi.org/10.1126/science.aar4851.

(114) Kim, H. H.; Yang, B.; Patel, T.; Sfigakis, F.; Li, C.; Tian, S.; Lei, H.; Tsen, A. W. One

Million Percent Tunnel Magnetoresistance in a Magnetic van Der Waals

Heterostructure. Nano Lett. 2018, 18 (8), 4885–4890.

https://doi.org/10.1021/acs.nanolett.8b01552.

(115) Tong, Q.; Liu, F.; Xiao, J.; Yao, W. Skyrmions in the Moiré of van Der Waals 2D Magnets.

Nano Lett. 2018, 18 (11), 7194–7199. https://doi.org/10.1021/acs.nanolett.8b03315.

(116) Huh, W.; Jang, S.; Lee, J. Y.; Lee, D.; Lee, D.; Lee, J. M.; Park, H.-G.; Kim, J. C.; Jeong, H.

Y.; Wang, G.; et al. Synaptic Barristor Based on Phase-Engineered 2D Heterostructures.

Adv. Mater. 2018, 30 (35), 1801447. https://doi.org/10.1002/adma.201801447.

(117) Zhang, Y.; He, K.; Chang, C.-Z.; Song, C.-L.; Wang, L.-L.; Chen, X.; Jia, J.-F.; Fang, Z.;

Dai, X.; Shan, W.-Y.; et al. Crossover of the Three-Dimensional Topological Insulator

Bi2Se3 to the Two-Dimensional Limit. Nat. Phys. 2010, 6 (8), 584–588.

https://doi.org/10.1038/nphys1689.

(118) Hsieh, D.; Xia, Y.; Qian, D.; Wray, L.; Dil, J. H.; Meier, F.; Osterwalder, J.; Patthey, L.;

Checkelsky, J. G.; Ong, N. P.; et al. A Tunable Topological Insulator in the Spin Helical

Dirac Transport Regime. Nature 2009, 460 (7259), 1101–1105.

https://doi.org/10.1038/nature08234.

Page 194: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

194

(119) Kravchenko, S. V.; Kravchenko, G. V.; Furneaux, J. E.; Pudalov, V. M.; D’Iorio, M.

Possible Metal-Insulator Transition at B=0 in Two Dimensions. Phys. Rev. B 1994, 50 (11),

8039–8042. https://doi.org/10.1103/PhysRevB.50.8039.

(120) Buda, I. G.; Lane, C.; Barbiellini, B.; Ruzsinszky, A.; Sun, J.; Bansil, A. Characterization of

Thin Film Materials Using SCAN Meta-GGA, an Accurate Nonempirical Density

Functional. Sci. Rep. 2017, 7, 44766. https://doi.org/10.1038/srep44766.

(121) Furness, J. W.; Zhang, Y.; Lane, C.; Buda, I. G.; Barbiellini, B.; Markiewicz, R. S.; Bansil,

A.; Sun, J. An Accurate First-Principles Treatment of Doping-Dependent Electronic

Structure of High-Temperature Cuprate Superconductors. Commun. Phys. 2018, 1 (1), 11.

https://doi.org/10.1038/s42005-018-0009-4.

(122) Lane, C.; Furness, J. W.; Buda, I. G.; Zhang, Y.; Markiewicz, R. S.; Barbiellini, B.; Sun, J.;

Bansil, A. Antiferromagnetic Ground State of La2CuO4: A Parameter-Free Ab Initio

Description. Phys. Rev. B 2018, 98 (12), 125140.

https://doi.org/10.1103/PhysRevB.98.125140.

(123) Ding, L.; Ukhtary, M. S.; Chubarov, M.; Choudhury, T. H.; Zhang, F.; Yang, R.; Zhang,

A.; Fan, J. A.; Terrones, M.; Redwing, J. M.; et al. Understanding Interlayer Coupling in

TMD-HBN Heterostructure by Raman Spectroscopy. IEEE Trans. Electron Devices 2018,

65 (10), 4059–4067. https://doi.org/10.1109/TED.2018.2847230.

(124) Chen, M.; Zhou, B.; Wang, F.; Xu, L.; Jiang, K.; Shang, L.; Hu, Z.; Chu, J. Interlayer

Coupling and the Phase Transition Mechanism of Stacked MoS2/TaS2 Heterostructures

Discovered Using Temperature Dependent Raman and Photoluminescence

Spectroscopy. RSC Adv. 2018, 8 (39), 21968–21974. https://doi.org/10.1039/C8RA03436B.

(125) Eckmann, A.; Park, J.; Yang, H.; Elias, D.; Mayorov, A. S.; Yu, G.; Jalil, R.; Novoselov, K.

S.; Gorbachev, R. V.; Lazzeri, M.; et al. Raman Fingerprint of Aligned Graphene/h-BN

Superlattices. Nano Lett. 2013, 13 (11), 5242–5246. https://doi.org/10.1021/nl402679b.

(126) Yan, Y.; Li, F.; Gong, Y.; Yao, M.; Huang, X.; Fu, X.; Han, B.; Zhou, Q.; Cui, T. Interlayer

Coupling Affected Structural Stability in Ultrathin MoS2: An Investigation by High

Pressure Raman Spectroscopy. J. Phys. Chem. C 2016, 120 (43), 24992–24998.

https://doi.org/10.1021/acs.jpcc.6b06562.

(127) Du, Y.; Zhuang, J.; Wang, J.; Li, Z.; Liu, H.; Zhao, J.; Xu, X.; Feng, H.; Chen, L.; Wu, K.; et

al. Quasi-Freestanding Epitaxial Silicene on Ag(111) by Oxygen Intercalation. Sci. Adv.

2016, 2 (7), e1600067. https://doi.org/10.1126/sciadv.1600067.

(128) Trainer, D. J.; Putilov, A. V.; Wang, B.; Lane, C.; Saari, T.; Chang, T.-R.; Jeng, H.-T.; Lin,

H.; Xi, X.; Nieminen, J.; et al. Moiré Superlattices and 2D Electronic Properties of

Graphite/MoS 2 Heterostructures. J. Phys. Chem. Solids 2017.

https://doi.org/10.1016/j.jpcs.2017.10.034.

(129) Wang, C.; He, Q.; Halim, U.; Liu, Y.; Zhu, E.; Lin, Z.; Xiao, H.; Duan, X.; Feng, Z.; Cheng,

R.; et al. Monolayer Atomic Crystal Molecular Superlattices. Nature 2018, 555 (7695),

231–236. https://doi.org/10.1038/nature25774.

(130) Lincoln Vogel, F. Some Potential Applications for Intercalation Compounds of Graphite

with High Electrical Conductivity. Synth. Met. 1980, 1 (3), 279–286.

https://doi.org/10.1016/0379-6779(80)90017-X.

Page 195: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

195

(131) Whittingham, M. S.; Ebert, L. B. Applications of Intercalation Compounds. In Intercalated

Layered Materials; Lévy, F., Ed.; Physics and Chemistry of Materials with Layered

Structures; Springer Netherlands: Dordrecht, 1979; pp 533–562.

https://doi.org/10.1007/978-94-009-9415-7_10.

(132) Müller-Warmuth, W.; Schöllhorn, R. Progress in Intercalation Research; Springer Science &

Business Media, 2012.

(133) Álvarez-Fraga, L.; Rubio-Zuazo, J.; Jiménez-Villacorta, F.; Climent-Pascual, E.; Ramírez-

Jiménez, R.; Prieto, C.; de Andrés, A. Oxidation Mechanisms of Copper under

Graphene: The Role of Oxygen Encapsulation. Chem. Mater. 2017, 29 (7), 3257–3264.

https://doi.org/10.1021/acs.chemmater.7b00554.

(134) Bignardi, L.; Lacovig, P.; Dalmiglio, M. M.; Orlando, F.; Ghafari, A.; Luca Petaccia;

Baraldi, A.; Larciprete, R.; Lizzit, S. Key Role of Rotated Domains in Oxygen

Intercalation at Graphene on Ni(1 1 1). 2D Mater. 2017, 4 (2), 025106.

https://doi.org/10.1088/2053-1583/aa702c.

(135) Grånäs, E.; Knudsen, J.; Schröder, U. A.; Gerber, T.; Busse, C.; Arman, M. A.; Schulte, K.;

Andersen, J. N.; Michely, T. Oxygen Intercalation under Graphene on Ir(111):

Energetics, Kinetics, and the Role of Graphene Edges. ACS Nano 2012, 6 (11), 9951–9963.

https://doi.org/10.1021/nn303548z.

(136) Kowalski, G.; Tokarczyk, M.; Dąbrowski, P.; Ciepielewski, P.; Możdżonek, M.;

Strupiński, W.; Baranowski, J. M. New X-Ray Insight into Oxygen Intercalation in

Epitaxial Graphene Grown on 4H-SiC(0001). J. Appl. Phys. 2015, 117 (10), 105301.

https://doi.org/10.1063/1.4914161.

(137) Li, T.; Yarmoff, J. A. Intercalation and Desorption of Oxygen between Graphene and

Ru(0001) Studied with Helium Ion Scattering. Phys. Rev. B 2017, 96 (15).

https://doi.org/10.1103/PhysRevB.96.155441.

(138) Ligato, N.; Caputi, L. S.; Cupolillo, A. Oxygen Intercalation at the Graphene/Ni(111)

Interface: Evidences of Non-Metal Islands underneath Graphene Layer. Carbon 2016,

100, 258–264. https://doi.org/10.1016/j.carbon.2016.01.011.

(139) Ma, L.; Zeng, X. C.; Wang, J. Oxygen Intercalation of Graphene on Transition Metal

Substrate: An Edge-Limited Mechanism. J. Phys. Chem. Lett. 2015, 6 (20), 4099–4105.

https://doi.org/10.1021/acs.jpclett.5b01841.

(140) Oliveira, M. H.; Schumann, T.; Fromm, F.; Koch, R.; Ostler, M.; Ramsteiner, M.; Seyller,

T.; Lopes, J. M. J.; Riechert, H. Formation of High-Quality Quasi-Free-Standing Bilayer

Graphene on SiC(0001) by Oxygen Intercalation upon Annealing in Air. Carbon 2013, 52,

83–89. https://doi.org/10.1016/j.carbon.2012.09.008.

(141) Romero-Muñiz, C.; Martín-Recio, A.; Pou, P.; Gómez-Rodríguez, J. M.; Pérez, R.

Unveiling the Atomistic Mechanisms for Oxygen Intercalation in a Strongly Interacting

Graphene–Metal Interface. Phys. Chem. Chem. Phys. 2018, 20 (19), 13370–13378.

https://doi.org/10.1039/C8CP01032C.

(142) Ulstrup, S.; Lacovig, P.; Orlando, F.; Lizzit, D.; Bignardi, L.; Dalmiglio, M.; Bianchi, M.;

Mazzola, F.; Baraldi, A.; Larciprete, R.; et al. Photoemission Investigation of Oxygen

Intercalated Epitaxial Graphene on Ru(0001). Surf. Sci. 2018.

https://doi.org/10.1016/j.susc.2018.03.017.

Page 196: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

196

(143) Bao, W.; Wan, J.; Han, X.; Cai, X.; Zhu, H.; Kim, D.; Ma, D.; Xu, Y.; Munday, J. N.; Drew,

H. D.; et al. Approaching the Limits of Transparency and Conductivity in Graphitic

Materials through Lithium Intercalation. Nat. Commun. 2014, 5, 4224.

https://doi.org/10.1038/ncomms5224.

(144) Yankowitz, M.; Jung, J.; Laksono, E.; Leconte, N.; Chittari, B. L.; Watanabe, K.;

Taniguchi, T.; Adam, S.; Graf, D.; Dean, C. R. Dynamic Band-Structure Tuning of

Graphene Moiré Superlattices with Pressure. Nature 2018, 557 (7705), 404.

https://doi.org/10.1038/s41586-018-0107-1.

(145) Rosenberger, M. R.; Chuang, H.-J.; McCreary, K. M.; Hanbicki, A. T.; Sivaram, S. V.;

Jonker, B. T. Nano-“Squeegee” for the Creation of Clean 2D Material Interfaces. ACS

Appl. Mater. Interfaces 2018, 10 (12), 10379–10387. https://doi.org/10.1021/acsami.8b01224.

(146) Fu, X.; Li, F.; Lin, J.-F.; Gong, Y.; Huang, X.; Huang, Y.; Gao, H.; Zhou, Q.; Cui, T.

Coupling-Assisted Renormalization of Excitons and Vibrations in Compressed MoSe2–

WSe2 Heterostructure. J. Phys. Chem. C 2018, 122 (10), 5820–5828.

https://doi.org/10.1021/acs.jpcc.8b01453.

(147) Woods, C. R.; Withers, F.; Zhu, M. J.; Cao, Y.; Yu, G.; Kozikov, A.; Ben Shalom, M.;

Morozov, S. V.; van Wijk, M. M.; Fasolino, A.; et al. Macroscopic Self-Reorientation of

Interacting Two-Dimensional Crystals. Nat. Commun. 2016, 7, 10800.

https://doi.org/10.1038/ncomms10800.

(148) Wang, F.; Wang, J.; Guo, S.; Zhang, J.; Hu, Z.; Chu, J. Tuning Coupling Behavior of

Stacked Heterostructures Based on MoS2, WS2, and WSe2. Sci. Rep. 2017, 7, 44712.

https://doi.org/10.1038/srep44712.

(149) Chiu, M.-H.; Li, M.-Y.; Zhang, W.; Hsu, W.-T.; Chang, W.-H.; Terrones, M.; Terrones, H.;

Li, L.-J. Spectroscopic Signatures for Interlayer Coupling in MoS2–WSe2 van Der Waals

Stacking. ACS Nano 2014, 8 (9), 9649–9656. https://doi.org/10.1021/nn504229z.

(150) Alexeev, E. M.; Catanzaro, A.; Skrypka, O. V.; Nayak, P. K.; Ahn, S.; Pak, S.; Lee, J.;

Sohn, J. I.; Novoselov, K. S.; Shin, H. S.; et al. Imaging of Interlayer Coupling in van Der

Waals Heterostructures Using a Bright-Field Optical Microscope. Nano Lett. 2017, 17 (9),

5342–5349. https://doi.org/10.1021/acs.nanolett.7b01763.

(151) Pak, S.; Lee, J.; Lee, Y.-W.; Jang, A.-R.; Ahn, S.; Ma, K. Y.; Cho, Y.; Hong, J.; Lee, S.; Jeong,

H. Y.; et al. Strain-Mediated Interlayer Coupling Effects on the Excitonic Behaviors in an

Epitaxially Grown MoS2/WS2 van Der Waals Heterobilayer. Nano Lett. 2017, 17 (9),

5634–5640. https://doi.org/10.1021/acs.nanolett.7b02513.

(152) Lebanon, G.; Bruckstein, A. M. On Designing Moiré Patterns. In Visual Attention

Mechanisms; Cantoni, V., Marinaro, M., Petrosino, A., Eds.; Springer US: Boston, MA,

2002; pp 205–217. https://doi.org/10.1007/978-1-4615-0111-4_19.

(153) Xu, L.; Huang, W.-Q.; Hu, W.; Yang, K.; Zhou, B.-X.; Pan, A.; Huang, G.-F. Two-

Dimensional MoS2-Graphene-Based Multilayer van Der Waals Heterostructures:

Enhanced Charge Transfer and Optical Absorption, and Electric-Field Tunable Dirac

Point and Band Gap. Chem. Mater. 2017, 29 (13), 5504–5512.

https://doi.org/10.1021/acs.chemmater.7b00281.

Page 197: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

197

(154) Velický, M.; Toth, P. S. From Two-Dimensional Materials to Their Heterostructures: An

Electrochemist’s Perspective. Appl. Mater. Today 2017, 8, 68–103.

https://doi.org/10.1016/j.apmt.2017.05.003.

(155) Zhang, C.; Chuu, C.-P.; Ren, X.; Li, M.-Y.; Li, L.-J.; Jin, C.; Chou, M.-Y.; Shih, C.-K.

Interlayer Couplings, Moiré Patterns, and 2D Electronic Superlattices in MoS2/WSe2

Hetero-Bilayers. Sci. Adv. 2017, 3 (1), e1601459. https://doi.org/10.1126/sciadv.1601459.

(156) Yu, H.; Liu, G.-B.; Tang, J.; Xu, X.; Yao, W. Moiré Excitons: From Programmable

Quantum Emitter Arrays to Spin-Orbit–Coupled Artificial Lattices. Sci. Adv. 2017, 3 (11),

e1701696. https://doi.org/10.1126/sciadv.1701696.

(157) Ohta, T.; Robinson, J. T.; Feibelman, P. J.; Bostwick, A.; Rotenberg, E.; Beechem, T. E.

Evidence for Interlayer Coupling and Moir\’e Periodic Potentials in Twisted Bilayer

Graphene. Phys. Rev. Lett. 2012, 109 (18), 186807.

https://doi.org/10.1103/PhysRevLett.109.186807.

(158) Naik, M. H.; Jain, M. Ultraflat Bands and Shear Solitons in Moir\’e Patterns of Twisted

Bilayer Transition Metal Dichalcogenides. ArXiv180309240 Cond-Mat 2018.

(159) Bistritzer, R.; MacDonald, A. H. Moiré Bands in Twisted Double-Layer Graphene. Proc.

Natl. Acad. Sci. 2011, 108 (30), 12233–12237. https://doi.org/10.1073/pnas.1108174108.

(160) Ribeiro-Palau, R.; Zhang, C.; Watanabe, K.; Taniguchi, T.; Hone, J.; Dean, C. R. Twistable

Electronics with Dynamically Rotatable Heterostructures. Science 2018, 361 (6403), 690–

693. https://doi.org/10.1126/science.aat6981.

(161) Castellanos-Gomez, A.; Buscema, M.; Molenaar, R.; Singh, V.; Janssen, L.; Zant, H. S. J.

van der; Steele, G. A. Deterministic Transfer of Two-Dimensional Materials by All-Dry

Viscoelastic Stamping. 2D Mater. 2014, 1 (1), 011002. https://doi.org/10.1088/2053-

1583/1/1/011002.

(162) Wang, E.; Chen, G.; Wan, G.; Lu, X.; Chen, C.; Avila, J.; Fedorov, A. V.; Guangyu Zhang;

Asensio, M. C.; Zhang, Y.; et al. Electronic Structure of Transferred Graphene/h-BN van

Der Waals Heterostructures with Nonzero Stacking Angles by Nano-ARPES. J. Phys.

Condens. Matter 2016, 28 (44), 444002. https://doi.org/10.1088/0953-8984/28/44/444002.

(163) Masubuchi, S.; Morimoto, M.; Morikawa, S.; Onodera, M.; Asakawa, Y.; Watanabe, K.;

Taniguchi, T.; Machida, T. Autonomous Robotic Searching and Assembly of Two-

Dimensional Crystals to Build van Der Waals Superlattices. Nat. Commun. 2018, 9 (1),

1413. https://doi.org/10.1038/s41467-018-03723-w.

(164) Kim, K.; Yankowitz, M.; Fallahazad, B.; Kang, S.; Movva, H. C. P.; Huang, S.; Larentis,

S.; Corbet, C. M.; Taniguchi, T.; Watanabe, K.; et al. Van Der Waals Heterostructures

with High Accuracy Rotational Alignment. Nano Lett. 2016, 16 (3), 1989–1995.

https://doi.org/10.1021/acs.nanolett.5b05263.

(165) 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), 8368–8375. https://doi.org/10.1021/acsnano.5b03043.

(166) Gong, Y.; Lei, S.; Ye, G.; Li, B.; He, Y.; Keyshar, K.; Zhang, X.; Wang, Q.; Lou, J.; Liu, Z.;

et al. Two-Step Growth of Two-Dimensional WSe2/MoSe2 Heterostructures. Nano Lett.

2015, 15 (9), 6135–6141. https://doi.org/10.1021/acs.nanolett.5b02423.

Page 198: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

198

(167) Hennighausen, Z.; Kar, S. Probing the Interlayer Interaction between Dissimilar 2D

Heterostructures by In-Situ Rearrangement of Their Interface. 2D Mater. 2019.

https://doi.org/10.1088/2053-1583/ab136e.

(168) Zhou, Y.; Fox, D. S.; Maguire, P.; O’Connell, R.; Masters, R.; Rodenburg, C.; Wu, H.;

Dapor, M.; Chen, Y.; Zhang, H. Quantitative Secondary Electron Imaging for Work

Function Extraction at Atomic Level and Layer Identification of Graphene. Sci. Rep.

2016, 6, 21045. https://doi.org/10.1038/srep21045.

(169) Vernon-Parry, K. D. Scanning Electron Microscopy: An Introduction. III-Vs Rev. 2000, 13

(4), 40–44. https://doi.org/10.1016/S0961-1290(00)80006-X.

(170) Thomas, J.; Gemming, T. Analytical Transmission Electron Microscopy: An Introduction for

Operators; Springer Netherlands, 2014.

(171) Shi, Z.; Wang, X.; Sun, Y.; Li, Y.; Zhang, L. Interlayer Coupling in Two-Dimensional

Semiconductor Materials. Semicond. Sci. Technol. 2018, 33 (9), 093001.

https://doi.org/10.1088/1361-6641/aad6c3.

(172) Novoselov, K. S.; Mishchenko, A.; Carvalho, A.; Neto, A. H. C. 2D Materials and van

Der Waals Heterostructures. Science 2016, 353 (6298), aac9439.

https://doi.org/10.1126/science.aac9439.

(173) Androulidakis, C.; Zhang, K.; Robertson, M.; Tawfick, S. Tailoring the Mechanical

Properties of 2D Materials and Heterostructures. 2D Mater. 2018, 5 (3), 032005.

https://doi.org/10.1088/2053-1583/aac764.

(174) Liu, Y.; Weiss, N. O.; Duan, X.; Cheng, H.-C.; Huang, Y.; Duan, X. Van Der Waals

Heterostructures and Devices. Nat. Rev. Mater. 2016, 1 (9), 16042.

https://doi.org/10.1038/natrevmats.2016.42.

(175) Hanbicki, A. T.; Chuang, H.-J.; Rosenberger, M. R.; Hellberg, C. S.; Sivaram, S. V.;

McCreary, K. M.; Mazin, I. I.; Jonker, B. T. Double Indirect Interlayer Exciton in a

MoSe2/WSe2 van Der Waals Heterostructure. ACS Nano 2018, 12 (5), 4719–4726.

https://doi.org/10.1021/acsnano.8b01369.

(176) Rivera, P.; Schaibley, J. R.; Jones, A. M.; Ross, J. S.; Wu, S.; Aivazian, G.; Klement, P.;

Seyler, K.; Clark, G.; Ghimire, N. J.; et al. Observation of Long-Lived Interlayer Excitons

in Monolayer MoSe2–WSe2 Heterostructures. Nat. Commun. 2015, 6, 6242.

https://doi.org/10.1038/ncomms7242.

(177) Fang, H.; Battaglia, C.; Carraro, C.; Nemsak, S.; Ozdol, B.; Kang, J. S.; Bechtel, H. A.;

Desai, S. B.; Kronast, F.; Unal, A. A.; et al. Strong Interlayer Coupling in van Der Waals

Heterostructures Built from Single-Layer Chalcogenides. Proc. Natl. Acad. Sci. 2014, 111

(17), 6198–6202. https://doi.org/10.1073/pnas.1405435111.

(178) Heo, H.; Sung, J. H.; Cha, S.; Jang, B.-G.; Kim, J.-Y.; Jin, G.; Lee, D.; Ahn, J.-H.; Lee, M.-J.;

Shim, J. H.; et al. Interlayer Orientation-Dependent Light Absorption and Emission in

Monolayer Semiconductor Stacks. Nat. Commun. 2015, 6, 7372.

https://doi.org/10.1038/ncomms8372.

(179) Shinde, S. M.; Dhakal, K. P.; Chen, X.; Yun, W. S.; Lee, J.; Kim, H.; Ahn, J.-H. Stacking-

Controllable Interlayer Coupling and Symmetric Configuration of Multilayered MoS2.

NPG Asia Mater. 2018, 10 (2), e468. https://doi.org/10.1038/am.2017.226.

Page 199: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

199

(180) Hennighausen, Z.; Kar, S. Probing the Interlayer Interaction between Dissimilar 2D

Heterostructures by in Situ Rearrangement of Their Interface. 2D Mater. 2019, 6 (3),

035022. https://doi.org/10.1088/2053-1583/ab136e.

(181) Jung, J.; Raoux, A.; Qiao, Z.; MacDonald, A. H. Ab Initio Theory of Moir\’e Superlattice

Bands in Layered Two-Dimensional Materials. Phys. Rev. B 2014, 89 (20), 205414.

https://doi.org/10.1103/PhysRevB.89.205414.

(182) Yoo, H.; Engelke, R.; Carr, S.; Fang, S.; Zhang, K.; Cazeaux, P.; Sung, H.; Hovden, R.;

Tsen, A. W.; Taniguchi, T.; et al. Atomic and Electronic Reconstruction at van Der Waals

Interface in Twisted Bilayer Graphene. 37.

(183) O’Farrell, E. C. T.; Tan, J. Y.; Yeo, Y.; Koon, G. K. W.; Özyilmaz, B.; Watanabe, K.;

Taniguchi, T. Rashba Interaction and Local Magnetic Moments in a Graphene-BN

Heterostructure Intercalated with Au. Phys. Rev. Lett. 2016, 117 (7), 076603.

https://doi.org/10.1103/PhysRevLett.117.076603.

(184) Yuan, L.; Chung, T.-F.; Kuc, A.; Wan, Y.; Xu, Y.; Chen, Y. P.; Heine, T.; Huang, L.

Photocarrier Generation from Interlayer Charge-Transfer Transitions in WS2-Graphene

Heterostructures. Sci. Adv. 2018, 4 (2), e1700324. https://doi.org/10.1126/sciadv.1700324.

(185) Wang, D.; Chen, G.; Li, C.; Cheng, M.; Yang, W.; Wu, S.; Xie, G.; Zhang, J.; Zhao, J.; Lu,

X.; et al. Thermally Induced Graphene Rotation on Hexagonal Boron Nitride. Phys. Rev.

Lett. 2016, 116 (12), 126101. https://doi.org/10.1103/PhysRevLett.116.126101.

(186) Lin, Y.-C.; Chang, C.-Y. S.; Ghosh, R. K.; Li, J.; Zhu, H.; Addou, R.; Diaconescu, B.; Ohta,

T.; Peng, X.; Lu, N.; et al. Atomically Thin Heterostructures Based on Single-Layer

Tungsten Diselenide and Graphene. Nano Lett. 2014, 14 (12), 6936–6941.

https://doi.org/10.1021/nl503144a.

(187) Carozo, V.; Fujisawa, K.; Rao, R.; Kahn, E.; Cunha, J. R.; Zhang, T.; Daniel Rubin;

Salazar, M. F.; Bugallo, A. de L.; Kar, S.; et al. Excitonic Processes in Atomically-Thin

MoSe 2 /MoS 2 Vertical Heterostructures. 2D Mater. 2018, 5 (3), 031016.

https://doi.org/10.1088/2053-1583/aacbe8.

(188) Yang, N.; Choi, K.; Robertson, J.; Park, H. G. Layer-Selective Synthesis of Bilayer

Graphene via Chemical Vapor Deposition. 2D Mater. 2017, 4 (3), 035023.

https://doi.org/10.1088/2053-1583/aa805d.

(189) Zhang, K.; Tadmor, E. B. Structural and Electron Diffraction Scaling of Twisted

Graphene Bilayers. J. Mech. Phys. Solids 2018, 112, 225–238.

https://doi.org/10.1016/j.jmps.2017.12.005.

(190) Vargas, A.; Basak, S.; Liu, F.; Wang, B.; Panaitescu, E.; Lin, H.; Markiewicz, R.; Bansil, A.;

Kar, S. The Changing Colors of a Quantum-Confined Topological Insulator. ACS Nano

2014, 8 (2), 1222–1230. https://doi.org/10.1021/nn404013d.

(191) Qiu, X.; Ji, W. Illuminating Interlayer Interactions. Nat. Mater. 2018, 17 (3), 211.

https://doi.org/10.1038/s41563-018-0027-3.

(192) Le, N. B.; Huan, T. D.; Woods, L. M. Interlayer Interactions in van Der Waals

Heterostructures: Electron and Phonon Properties. ACS Appl. Mater. Interfaces 2016, 8 (9),

6286–6292. https://doi.org/10.1021/acsami.6b00285.

Page 200: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

200

(193) Leven, I.; Maaravi, T.; Azuri, I.; Kronik, L.; Hod, O. Interlayer Potential for Graphene/h-

BN Heterostructures. J. Chem. Theory Comput. 2016, 12 (6), 2896–2905.

https://doi.org/10.1021/acs.jctc.6b00147.

(194) Bendersky, L. A.; Gayle, F. W. Electron Diffraction Using Transmission Electron

Microscopy. J. Res. Natl. Inst. Stand. Technol. 2001, 106 (6), 997–1012.

https://doi.org/10.6028/jres.106.051.

(195) Schwarzer, R. A. The Determination of Local Texture by Electron Diffraction–A Tutorial

Review https://www.hindawi.com/journals/tsm/1993/696572/abs/ (accessed Nov 27,

2018). https://doi.org/10.1155/TSM.20.7.

(196) Fei, L.; Lei, S.; Zhang, W.-B.; Lu, W.; Lin, Z.; Lam, C. H.; Chai, Y.; Wang, Y. Direct TEM

Observations of Growth Mechanisms of Two-Dimensional MoS2 Flakes. Nat. Commun.

2016, 7, 12206. https://doi.org/10.1038/ncomms12206.

(197) Yuk, J. M.; Jeong, M.; Kim, S. Y.; Seo, H. K.; Kim, J.; Lee, J. Y. In Situ Atomic Imaging of

Coalescence of Au Nanoparticles on Graphene: Rotation and Grain Boundary

Migration. Chem. Commun. 2013, 49 (98), 11479–11481.

https://doi.org/10.1039/C3CC46545D.

(198) Meyer, J. C.; Kurasch, S.; Park, H. J.; Skakalova, V.; Künzel, D.; Groß, A.; Chuvilin, A.;

Algara-Siller, G.; Roth, S.; Iwasaki, T.; et al. Experimental Analysis of Charge

Redistribution Due to Chemical Bonding by High-Resolution Transmission Electron

Microscopy. Nat. Mater. 2011, 10 (3), 209–215. https://doi.org/10.1038/nmat2941.

(199) Chen, Y. Nanofabrication by Electron Beam Lithography and Its Applications: A

Review. Microelectron. Eng. 2015, 135, 57–72. https://doi.org/10.1016/j.mee.2015.02.042.

(200) Cazaux, J. Material Contrast in SEM: Fermi Energy and Work Function Effects.

Ultramicroscopy 2010, 110 (3), 242–253. https://doi.org/10.1016/j.ultramic.2009.12.002.

(201) Griffiths, A. J. V.; Walther, T. Quantification of Carbon Contamination under Electron

Beam Irradiation in a Scanning Transmission Electron Microscope and Its Suppression

by Plasma Cleaning. J. Phys. Conf. Ser. 2010, 241 (1), 012017. https://doi.org/10.1088/1742-

6596/241/1/012017.

(202) Shaffner, T. J.; Veld, R. D. V. “Charging” Effects in the Scanning Electron Microscope. J.

Phys. [E] 1971, 4 (9), 633. https://doi.org/10.1088/0022-3735/4/9/002.

(203) Zhou, W.; Zou, X.; Najmaei, S.; Liu, Z.; Shi, Y.; Kong, J.; Lou, J.; Ajayan, P. M.; Yakobson,

B. I.; Idrobo, J.-C. Intrinsic Structural Defects in Monolayer Molybdenum Disulfide.

Nano Lett. 2013, 13 (6), 2615–2622. https://doi.org/10.1021/nl4007479.

(204) Lin, Z.; Carvalho, B. R.; Kahn, E.; Lv, R.; Rao, R.; Terrones, H.; Pimenta, M. A.; Terrones,

M. Defect Engineering of Two-Dimensional Transition Metal Dichalcogenides. 2D

Mater. 2016, 3 (2), 022002. https://doi.org/10.1088/2053-1583/3/2/022002.

(205) Cançado, L. G.; Jorio, A.; Ferreira, E. H. M.; Stavale, F.; Achete, C. A.; Capaz, R. B.;

Moutinho, M. V. O.; Lombardo, A.; Kulmala, T. S.; Ferrari, A. C. Quantifying Defects in

Graphene via Raman Spectroscopy at Different Excitation Energies. Nano Lett. 2011, 11

(8), 3190–3196. https://doi.org/10.1021/nl201432g.

(206) Mignuzzi, S.; Pollard, A. J.; Bonini, N.; Brennan, B.; Gilmore, I. S.; Pimenta, M. A.;

Richards, D.; Roy, D. Effect of Disorder on Raman Scattering of Single-Layer MoS2.

Phys. Rev. B 2015, 91 (19), 195411. https://doi.org/10.1103/PhysRevB.91.195411.

Page 201: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

201

(207) Britnell, L.; Gorbachev, R. V.; Jalil, R.; Belle, B. D.; Schedin, F.; Mishchenko, A.;

Georgiou, T.; Katsnelson, M. I.; Eaves, L.; Morozov, S. V.; et al. Field-Effect Tunneling

Transistor Based on Vertical Graphene Heterostructures. Science 2012, 335 (6071), 947–

950. https://doi.org/10.1126/science.1218461.

(208) Georgiou, T.; Jalil, R.; Belle, B. D.; Britnell, L.; Gorbachev, R. V.; Morozov, S. V.; Kim, Y.-

J.; Gholinia, A.; Haigh, S. J.; Makarovsky, O.; et al. Vertical Field-Effect Transistor Based

on Graphene–WS2 Heterostructures for Flexible and Transparent Electronics. Nat.

Nanotechnol. 2012, 8 (2), nnano.2012.224. https://doi.org/10.1038/nnano.2012.224.

(209) Yu, W. J.; Li, Z.; Zhou, H.; Chen, Y.; Wang, Y.; Huang, Y.; Duan, X. Vertically Stacked

Multi-Heterostructures of Layered Materials for Logic Transistors and Complementary

Inverters. Nat. Mater. 2013, 12 (3), 246–252. https://doi.org/10.1038/nmat3518.

(210) Britnell, L.; Ribeiro, R. M.; Eckmann, A.; Jalil, R.; Belle, B. D.; Mishchenko, A.; Kim, Y.-J.;

Gorbachev, R. V.; Georgiou, T.; Morozov, S. V.; et al. Strong Light-Matter Interactions in

Heterostructures of Atomically Thin Films. Science 2013, 340 (6138), 1311–1314.

https://doi.org/10.1126/science.1235547.

(211) Yu, W. J.; Liu, Y.; Zhou, H.; Yin, A.; Li, Z.; Huang, Y.; Duan, X. Highly Efficient Gate-

Tunable Photocurrent Generation in Vertical Heterostructures of Layered Materials. Nat.

Nanotechnol. 2013, 8 (12), 952–958. https://doi.org/10.1038/nnano.2013.219.

(212) Ross, J. S.; Rivera, P.; Schaibley, J.; Lee-Wong, E.; Yu, H.; Taniguchi, T.; Watanabe, K.;

Yan, J.; Mandrus, D.; Cobden, D.; et al. Interlayer Exciton Optoelectronics in a 2D

Heterostructure p–n Junction. Nano Lett. 2017, 17 (2), 638–643.

https://doi.org/10.1021/acs.nanolett.6b03398.

(213) Hong, A. J.; Song, E. B.; Yu, H. S.; Allen, M. J.; Kim, J.; Fowler, J. D.; Wassei, J. K.; Park,

Y.; Wang, Y.; Zou, J.; et al. Graphene Flash Memory. ACS Nano 2011, 5 (10), 7812–7817.

https://doi.org/10.1021/nn201809k.

(214) Bertolazzi, S.; Krasnozhon, D.; Kis, A. Nonvolatile Memory Cells Based on

MoS2/Graphene Heterostructures. ACS Nano 2013, 7 (4), 3246–3252.

https://doi.org/10.1021/nn3059136.

(215) Agarwal, A.; Vitiello, M. S.; Viti, L.; Cupolillo, A.; Politano, A. Plasmonics with Two-

Dimensional Semiconductors: From Basic Research to Technological Applications.

Nanoscale 2018, 10 (19), 8938–8946. https://doi.org/10.1039/C8NR01395K.

(216) Iranzo, D. A.; Nanot, S.; Dias, E. J. C.; Epstein, I.; Peng, C.; Efetov, D. K.; Lundeberg, M.

B.; Parret, R.; Osmond, J.; Hong, J.-Y.; et al. Probing the Ultimate Plasmon Confinement

Limits with a van Der Waals Heterostructure. Science 2018, 360 (6386), 291–295.

https://doi.org/10.1126/science.aar8438.

(217) Wang, G.; Li, L.; Fan, W.; Wang, R.; Zhou, S.; Lü, J.-T.; Gan, L.; Zhai, T. Interlayer

Coupling Induced Infrared Response in WS2/MoS2 Heterostructures Enhanced by

Surface Plasmon Resonance. Adv. Funct. Mater. 2018, 28 (22), 1800339.

https://doi.org/10.1002/adfm.201800339.

(218) Zhu, M.; Sun, Z.; Fujitsuka, M.; Majima, T. Z-Scheme Photocatalytic Water Splitting on a

2D Heterostructure of Black Phosphorus/Bismuth Vanadate Using Visible Light. Angew.

Chem. Int. Ed. 2018, 57 (8), 2160–2164. https://doi.org/10.1002/anie.201711357.

Page 202: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

202

(219) Chen, Z.; Cao, J.; Yang, L.; Yin, W.; Wei, X. The Unique Photocatalysis Properties of a 2D

Vertical MoO 2 /WO 2 Heterostructure: A First-Principles Study. J. Phys. Appl. Phys.

2018, 51 (26), 265106. https://doi.org/10.1088/1361-6463/aac7d5.

(220) Zhang, J.; Zhu, Z.; Tang, Y.; Müllen, K.; Feng, X. Titania Nanosheet-Mediated

Construction of a Two-Dimensional Titania/Cadmium Sulfide Heterostructure for High

Hydrogen Evolution Activity. Adv. Mater. 2014, 26 (5), 734–738.

https://doi.org/10.1002/adma.201303571.

(221) Peng, L.; Peng, X.; Liu, B.; Wu, C.; Xie, Y.; Yu, G. Ultrathin Two-Dimensional

MnO2/Graphene Hybrid Nanostructures for High-Performance, Flexible Planar

Supercapacitors. Nano Lett. 2013, 13 (5), 2151–2157. https://doi.org/10.1021/nl400600x.

(222) Ramakrishnan, K.; Nithya, C.; Karvembu, R. Heterostructure of Two Different 2D

Materials Based on MoS2 Nanoflowers@rGO: An Electrode Material for Sodium-Ion

Capacitors. Nanoscale Adv. 2018. https://doi.org/10.1039/C8NA00104A.

(223) Dankert, A.; Dash, S. P. Electrical Gate Control of Spin Current in van Der Waals

Heterostructures at Room Temperature. Nat. Commun. 2017, 8, 16093.

https://doi.org/10.1038/ncomms16093.

(224) Yoo, H.; Engelke, R.; Carr, S.; Fang, S.; Zhang, K.; Cazeaux, P.; Sung, S. H.; Hovden, R.;

Tsen, A. W.; Taniguchi, T.; et al. Atomic and Electronic Reconstruction at van Der Waals

Interface in Twisted Bilayer Graphene. ArXiv180403806 Cond-Mat 2018.

(225) Tongay, S.; Fan, W.; Kang, J.; Park, J.; Koldemir, U.; Suh, J.; Narang, D. S.; Liu, K.; Ji, J.;

Li, J.; et al. Tuning Interlayer Coupling in Large-Area Heterostructures with CVD-

Grown MoS2 and WS2 Monolayers. Nano Lett. 2014, 14 (6), 3185–3190.

https://doi.org/10.1021/nl500515q.

(226) Hennighausen, Z.; Lane, C.; Benabbas, A.; Mendez, K.; Eggenberger, M.; Champion, P.

M.; Robinson, J. T.; Bansil, A.; Kar, S. Oxygen-Induced In Situ Manipulation of the

Interlayer Coupling and Exciton Recombination in Bi2Se3/MoS2 2D Heterostructures.

ACS Appl. Mater. Interfaces 2019, 11 (17), 15913–15921.

https://doi.org/10.1021/acsami.9b02929.

(227) Tan, Y.; Liu, X.; He, Z.; Liu, Y.; Zhao, M.; Zhang, H.; Chen, F. Tuning of Interlayer

Coupling in Large-Area Graphene/WSe2 van Der Waals Heterostructure via Ion

Irradiation: Optical Evidences and Photonic Applications. ACS Photonics 2017, 4 (6),

1531–1538. https://doi.org/10.1021/acsphotonics.7b00296.

(228) Ghorbani-Asl, M.; Kretschmer, S.; Spearot, D. E.; Krasheninnikov, A. V. Two-

Dimensional MoS 2 under Ion Irradiation: From Controlled Defect Production to

Electronic Structure Engineering. 2D Mater. 2017, 4 (2), 025078.

https://doi.org/10.1088/2053-1583/aa6b17.

(229) Kretschmer, S.; Maslov, M.; Ghaderzadeh, S.; Ghorbani-Asl, M.; Hlawacek, G.;

Krasheninnikov, A. V. Supported Two-Dimensional Materials under Ion Irradiation:

The Substrate Governs Defect Production. ACS Appl. Mater. Interfaces 2018, 10 (36),

30827–30836. https://doi.org/10.1021/acsami.8b08471.

(230) Maguire, P.; Fox, D. S.; Zhou, Y.; Wang, Q.; O’Brien, M.; Jadwiszczak, J.; Cullen, C. P.;

McManus, J.; Bateman, S.; McEvoy, N.; et al. Defect Sizing, Separation, and Substrate

Page 203: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

203

Effects in Ion-Irradiated Monolayer Two-Dimensional Materials. Phys. Rev. B 2018, 98

(13), 134109. https://doi.org/10.1103/PhysRevB.98.134109.

(231) Conley, H. J.; Wang, B.; Ziegler, J. I.; Haglund, R. F.; Pantelides, S. T.; Bolotin, K. I.

Bandgap Engineering of Strained Monolayer and Bilayer MoS2. Nano Lett. 2013, 13 (8),

3626–3630. https://doi.org/10.1021/nl4014748.

(232) Roldán, R.; Castellanos-Gomez, A.; Cappelluti, E.; Guinea, F. Strain Engineering in

Semiconducting Two-Dimensional Crystals. J. Phys. Condens. Matter 2015, 27 (31),

313201. https://doi.org/10.1088/0953-8984/27/31/313201.

(233) Ahn, G. H.; Amani, M.; Rasool, H.; Lien, D.-H.; Mastandrea, J. P.; Iii, J. W. A.; Dubey, M.;

Chrzan, D. C.; Minor, A. M.; Javey, A. Strain-Engineered Growth of Two-Dimensional

Materials. Nat. Commun. 2017, 8 (1), 608. https://doi.org/10.1038/s41467-017-00516-5.

(234) Hu, W.; Wang, T.; Zhang, R.; Yang, J. Effects of Interlayer Coupling and Electric Fields

on the Electronic Structures of Graphene and MoS2 Heterobilayers. J. Mater. Chem. C

2016, 4 (9), 1776–1781. https://doi.org/10.1039/C6TC00207B.

(235) Phuc, H. V.; Hieu, N. N.; Hoi, B. D.; Nguyen, C. V. Interlayer Coupling and Electric

Field Tunable Electronic Properties and Schottky Barrier in a Graphene/Bilayer-GaSe

van Der Waals Heterostructure. Phys. Chem. Chem. Phys. 2018, 20 (26), 17899–17908.

https://doi.org/10.1039/C8CP02190B.

(236) Zhang, W.; Chang, D.; Gao, Q.; Niu, C.; Li, C.; Wang, F.; Huang, X.; Xia, C.; Jia, Y.

Interlayer Coupling and External Electric Field Tunable Electronic Properties of a 2D

Type-I α-Tellurene/MoS2 Heterostructure. J. Mater. Chem. C 2018, 6 (38), 10256–10262.

https://doi.org/10.1039/C8TC03286F.

(237) Gao, S.; Yang, L.; Spataru, C. D. Interlayer Coupling and Gate-Tunable Excitons in

Transition Metal Dichalcogenide Heterostructures. Nano Lett. 2017, 17 (12), 7809–7813.

https://doi.org/10.1021/acs.nanolett.7b04021.

(238) Wang, Y.; Ou, J. Z.; Balendhran, S.; Chrimes, A. F.; Mortazavi, M.; Yao, D. D.; Field, M.

R.; Latham, K.; Bansal, V.; Friend, J. R.; et al. Electrochemical Control of

Photoluminescence in Two-Dimensional MoS2 Nanoflakes. ACS Nano 2013, 7 (11),

10083–10093. https://doi.org/10.1021/nn4041987.

(239) Kowalski, G.; Tokarczyk, M.; Dąbrowski, P.; Ciepielewski, P.; Możdżonek, M.;

Strupiński, W.; Baranowski, J. M. New X-Ray Insight into Oxygen Intercalation in

Epitaxial Graphene Grown on 4 H -SiC(0001). J. Appl. Phys. 2015, 117 (10), 105301.

https://doi.org/10.1063/1.4914161.

(240) Sutter, P.; Sadowski, J. T.; Sutter, E. A. Chemistry under Cover: Tuning Metal−Graphene

Interaction by Reactive Intercalation. J. Am. Chem. Soc. 2010, 132 (23), 8175–8179.

https://doi.org/10.1021/ja102398n.

(241) Wan, J.; Lacey, S. D.; Dai, J.; Bao, W.; Fuhrer, M. S.; Hu, L. Tuning Two-Dimensional

Nanomaterials by Intercalation: Materials, Properties and Applications. Chem. Soc. Rev.

2016, 45 (24), 6742–6765. https://doi.org/10.1039/C5CS00758E.

(242) Bansal, N.; Kim, Y. S.; Edrey, E.; Brahlek, M.; Horibe, Y.; Iida, K.; Tanimura, M.; Li, G.-

H.; Feng, T.; Lee, H.-D.; et al. Epitaxial Growth of Topological Insulator Bi2Se3 Film on

Si(111) with Atomically Sharp Interface. Thin Solid Films 2011, 520 (1), 224–229.

https://doi.org/10.1016/j.tsf.2011.07.033.

Page 204: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

204

(243) Walsh, L. A.; Hinkle, C. L. Van Der Waals Epitaxy: 2D Materials and Topological

Insulators. Appl. Mater. Today 2017, 9, 504–515.

https://doi.org/10.1016/j.apmt.2017.09.010.

(244) Kim, D.; Syers, P.; Butch, N. P.; Paglione, J.; Fuhrer, M. S. Coherent Topological

Transport on the Surface of Bi2Se3. Nat. Commun. 2013, 4, 2040.

https://doi.org/10.1038/ncomms3040.

(245) Splendiani, A.; Sun, L.; Zhang, Y.; Li, T.; Kim, J.; Chim, C.-Y.; Galli, G.; Wang, F.

Emerging Photoluminescence in Monolayer MoS2. Nano Lett. 2010, 10 (4), 1271–1275.

https://doi.org/10.1021/nl903868w.

(246) Yashina, L. V.; Sánchez-Barriga, J.; Scholz, M. R.; Volykhov, A. A.; Sirotina, A. P.;

Neudachina, V., S.; Tamm, M. E.; Varykhalov, A.; Marchenko, D.; Springholz, G.; et al.

Negligible Surface Reactivity of Topological Insulators Bi2Se3 and Bi2Te3 towards

Oxygen and Water. ACS Nano 2013, 7 (6), 5181–5191. https://doi.org/10.1021/nn400908b.

(247) Gao, L.; Li, H.; Ren, W.; Wang, G.; Li, H.; Zhou, Z.; Ji, H.; Niu, X.; Wang, Z. Patterning

Bi2Se3 Single-Crystalline Thin Films on Si(111) Substrates Using Strong Oxidizing

Acids. RSC Adv. 2017, 7 (51), 32294–32299. https://doi.org/10.1039/C7RA05317G.

(248) Yeh, Y.-C.; Ho, P.-H.; Wen, C.-Y.; Shu, G.-J.; Sankar, R.; Chou, F.-C.; Chen, C.-W. Growth

of the Bi2Se3 Surface Oxide for Metal–Semiconductor–Metal Device Applications. J.

Phys. Chem. C 2016, 120 (6), 3314–3318. https://doi.org/10.1021/acs.jpcc.5b10425.

(249) Music, D.; Chang, K.; Schmidt, P.; Braun, F. N.; Heller, M.; Hermsen, S.; Pöllmann, P. J.;

Schulzendorff, T.; Wagner, C. On Atomic Mechanisms Governing the Oxidation of Bi 2

Te 3. J. Phys. Condens. Matter 2017, 29 (48), 485705. https://doi.org/10.1088/1361-

648X/aa945f.

(250) Píš, I.; Nappini, S.; Bondino, F.; Menteş, T. O.; Sala, A.; Locatelli, A.; Magnano, E. Fe

Intercalation under Graphene and Hexagonal Boron Nitride In-Plane Heterostructure

on Pt(111). Carbon 2018, 134, 274–282. https://doi.org/10.1016/j.carbon.2018.03.086.

(251) Zhang, H.; Fu, Q.; Cui, Y.; Tan, D.; Bao, X. Growth Mechanism of Graphene on Ru(0001)

and O2 Adsorption on the Graphene/Ru(0001) Surface. J. Phys. Chem. C 2009, 113 (19),

8296–8301. https://doi.org/10.1021/jp810514u.

(252) Trentelman, K. A Note on the Characterization of Bismuth Black by Raman

Microspectroscopy. J. Raman Spectrosc. 2009, 40 (5), 585–589.

https://doi.org/10.1002/jrs.2184.

(253) Hardcastle, F. D.; Wachs, I. E. The Molecular Structure of Bismuth Oxide by Raman

Spectroscopy. J. Solid State Chem. 1992, 97 (2), 319–331. https://doi.org/10.1016/0022-

4596(92)90040-3.

(254) Steele, J. A.; Lewis, R. A. In Situ Micro-Raman Studies of Laser-Induced Bismuth

Oxidation Reveals Metastability of β-Bi2O3 Microislands. Opt. Mater. Express 2014, 4

(10), 2133–2142. https://doi.org/10.1364/OME.4.002133.

(255) Hamamoto, K.; Ezawa, M.; Kim, K. W.; Morimoto, T.; Nagaosa, N. Nonlinear Spin

Current Generation in Noncentrosymmetric Spin-Orbit Coupled Systems. Phys. Rev. B

2017, 95 (22), 224430. https://doi.org/10.1103/PhysRevB.95.224430.

Page 205: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

205

(256) Smidman, M.; Salamon, M. B.; Yuan, H. Q.; Agterberg, D. F. Superconductivity and

Spin–Orbit Coupling in Non-Centrosymmetric Materials: A Review. Rep. Prog. Phys.

2017, 80 (3), 036501. https://doi.org/10.1088/1361-6633/80/3/036501.

(257) Mera Acosta, C.; Babilonia, O.; Abdalla, L.; Fazzio, A. Unconventional Spin Texture in a

Noncentrosymmetric Quantum Spin Hall Insulator. Phys. Rev. B 2016, 94 (4), 041302.

https://doi.org/10.1103/PhysRevB.94.041302.

(258) Wang, Q. H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J. N.; Strano, M. S. Electronics and

Optoelectronics of Two-Dimensional Transition Metal Dichalcogenides. Nat.

Nanotechnol. 2012, 7 (11), 699–712. https://doi.org/10.1038/nnano.2012.193.

(259) Xiao, J.; Zhao, M.; Wang, Y.; Zhang, X. Excitons in Atomically Thin 2D Semiconductors

and Their Applications. Nanophotonics 2017, 6 (6). https://doi.org/10.1515/nanoph-2016-

0160.

(260) Zhang, H.; Lu, S. B.; Zheng, J.; Du, J.; Wen, S. C.; Tang, D. Y.; Loh, K. P. Molybdenum

Disulfide (MoS2) as a Broadband Saturable Absorber for Ultra-Fast Photonics. Opt.

Express 2014, 22 (6), 7249–7260. https://doi.org/10.1364/OE.22.007249.

(261) Xenogiannopoulou, E.; Tsipas, P.; E. Aretouli, K.; Tsoutsou, D.; A. Giamini, S.; Bazioti,

C.; P. Dimitrakopulos, G.; Komninou, P.; Brems, S.; Huyghebaert, C.; et al. High-Quality,

Large-Area MoSe 2 and MoSe 2 /Bi 2 Se 3 Heterostructures on AlN(0001)/Si(111)

Substrates by Molecular Beam Epitaxy. Nanoscale 2015, 7 (17), 7896–7905.

https://doi.org/10.1039/C4NR06874B.

(262) Schirra, P.; Lopez, G. M.; Fiorentini, V. Complexes, Clustering, and Native-Defect-

Assisted Diffusion of Aluminum in Silicon. Phys. Rev. B 2004, 70 (24), 245201.

https://doi.org/10.1103/PhysRevB.70.245201.

(263) Melis, C.; Lopez, G. M.; Fiorentini, V. Anomalous Energetics and Defect-Assisted

Diffusion of Ga in Silicon. Appl. Phys. Lett. 2004, 85 (21), 4902–4904.

https://doi.org/10.1063/1.1826230.

(264) Xie, L.; Cui, X. Manipulating Spin-Polarized Photocurrents in 2D Transition Metal

Dichalcogenides. Proc. Natl. Acad. Sci. 2016, 113 (14), 3746–3750.

https://doi.org/10.1073/pnas.1523012113.

(265) Akinwande, D.; Petrone, N.; Hone, J. Two-Dimensional Flexible Nanoelectronics. Nat.

Commun. 2014, 5, 5678. https://doi.org/10.1038/ncomms6678.

(266) Manzeli, S.; Ovchinnikov, D.; Pasquier, D.; Yazyev, O. V.; Kis, A. 2D Transition Metal

Dichalcogenides. Nat. Rev. Mater. 2017, 2 (8), 17033.

https://doi.org/10.1038/natrevmats.2017.33.

(267) Ugeda, M. M.; Bradley, A. J.; Shi, S.-F.; Jornada, F. H. da; Zhang, Y.; Qiu, D. Y.; Ruan,

W.; Mo, S.-K.; Hussain, Z.; Shen, Z.-X.; et al. Giant Bandgap Renormalization and

Excitonic Effects in a Monolayer Transition Metal Dichalcogenide Semiconductor. Nat.

Mater. 2014, 13 (12), 1091–1095. https://doi.org/10.1038/nmat4061.

(268) Hennighausen, Z.; Bilgin, I.; Casey, C.; Kar, S. Widely Tunable Bi2Se3/Transition Metal

Dichalcogenide 2D Heterostructures for Write-Read-Erase-Reuse Applications. 2D

Mater. 2019, 6 (4), 041003. https://doi.org/10.1088/2053-1583/ab33b5.

Page 206: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

206

(269) Kim, J.; Hong, X.; Jin, C.; Shi, S.-F.; Chang, C.-Y. S.; Chiu, M.-H.; Li, L.-J.; Wang, F.

Ultrafast Generation of Pseudo-Magnetic Field for Valley Excitons in WSe2 Monolayers.

Science 2014, 346 (6214), 1205–1208. https://doi.org/10.1126/science.1258122.

(270) Stier, A. V.; Wilson, N. P.; Clark, G.; Xu, X.; Crooker, S. A. Probing the Influence of

Dielectric Environment on Excitons in Monolayer WSe2: Insight from High Magnetic

Fields. Nano Lett. 2016, 16 (11), 7054–7060. https://doi.org/10.1021/acs.nanolett.6b03276.

(271) Hu, P.; Ye, J.; He, X.; Du, K.; Zhang, K. K.; Wang, X.; Xiong, Q.; Liu, Z.; Jiang, H.; Kloc, C.

Control of Radiative Exciton Recombination by Charge Transfer Induced Surface

Dipoles in MoS2 and WS2 Monolayers. Sci. Rep. 2016, 6, 24105.

https://doi.org/10.1038/srep24105.

(272) Lin, Y.; Ling, X.; Yu, L.; Huang, S.; Hsu, A. L.; Lee, Y.-H.; Kong, J.; Dresselhaus, M. S.;

Palacios, T. Dielectric Screening of Excitons and Trions in Single-Layer MoS2. Nano Lett.

2014, 14 (10), 5569–5576. https://doi.org/10.1021/nl501988y.

(273) Chernikov, A.; Berkelbach, T. C.; Hill, H. M.; Rigosi, A.; Li, Y.; Aslan, O. B.; Reichman,

D. R.; Hybertsen, M. S.; Heinz, T. F. Exciton Binding Energy and Nonhydrogenic

Rydberg Series in Monolayer WS2. Phys. Rev. Lett. 2014, 113 (7), 076802.

https://doi.org/10.1103/PhysRevLett.113.076802.

(274) Zhang, N.; Surrente, A.; Baranowski, M.; Maude, D. K.; Gant, P.; Castellanos-Gomez, A.;

Plochocka, P. Moiré Intralayer Excitons in a MoSe2/MoS2 Heterostructure. Nano Lett.

2018, 18 (12), 7651–7657. https://doi.org/10.1021/acs.nanolett.8b03266.

(275) Hennighausen, Z.; Lane, C.; Buda, I. G.; Mathur, V.; Bansil, A.; Kar, S. Evidence of a

Purely Electronic Two-Dimensional Lattice at the Interface of TMD/Bi2Se3

Heterostructures. Nanoscale 2019. https://doi.org/10.1039/C9NR04412D.

(276) Hennighausen, Z.; Lane, C.; Buda, I. G.; Mathur, V.; Bansil, A.; Kar, S. Strong Interlayer

Coupling Induced Charge Redistribution Forms Robust Non-Atomic Electron Scattering

Sites in Bi2Se3/Transition Metal Dichalcogenide 2D Heterostructures. Nanoscale 2019.

(277) Zeller, P.; Ma, X.; Günther, S. Indexing Moiré Patterns of Metal-Supported Graphene

and Related Systems: Strategies and Pitfalls. New J. Phys. 2017, 19 (1), 013015.

https://doi.org/10.1088/1367-2630/aa53c8.

(278) Lebanon, G.; Bruckstein, A. M. Variational Approach to Moiré Pattern Synthesis. JOSA

A 2001, 18 (6), 1371–1382. https://doi.org/10.1364/JOSAA.18.001371.

(279) Peng, L.; Zhu, Y.; Peng, X.; Fang, Z.; Chu, W.; Wang, Y.; Xie, Y.; Li, Y.; Cha, J. J.; Yu, G.

Effective Interlayer Engineering of Two-Dimensional VOPO4 Nanosheets via Controlled

Organic Intercalation for Improving Alkali Ion Storage. Nano Lett. 2017, 17 (10), 6273–

6279. https://doi.org/10.1021/acs.nanolett.7b02958.

(280) Zeller, P.; Günther, S. What Are the Possible Moiré Patterns of Graphene on

Hexagonally Packed Surfaces? Universal Solution for Hexagonal Coincidence Lattices,

Derived by a Geometric Construction. New J. Phys. 2014, 16 (8), 083028.

https://doi.org/10.1088/1367-2630/16/8/083028.

(281) Nakashima, P. N. H.; Smith, A. E.; Etheridge, J.; Muddle, B. C. The Bonding Electron

Density in Aluminum. Science 2011, 331 (6024), 1583–1586.

https://doi.org/10.1126/science.1198543.

Page 207: Probing and Manipulating the Interlayer Coupling in 2D ...m044kn658/fulltext.pdfstructures that include, graphene/MoS 2, graphene/graphene, Sb 2 Se 3 /MoS 2, and Bi 2 Te 3 /MoS 2.

207

(282) Kasai, H.; Tolborg, K.; Sist, M.; Zhang, J.; Hathwar, V. R.; Filsø, M. Ø.; Cenedese, S.;

Sugimoto, K.; Overgaard, J.; Nishibori, E.; et al. X-Ray Electron Density Investigation of

Chemical Bonding in van Der Waals Materials. Nat. Mater. 2018, 17 (3), 249.

https://doi.org/10.1038/s41563-017-0012-2.

(283) Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.; Zhang, Y.; Dubonos, S. V.;

Grigorieva, I. V.; Firsov, A. A. Electric Field Effect in Atomically Thin Carbon Films.

Science 2004, 306 (5696), 666–669. https://doi.org/10.1126/science.1102896.

(284) Wang, H.; Li, C.; Fang, P.; Zhang, Z.; Zhang, J. Z. Synthesis, Properties, and

Optoelectronic Applications of Two-Dimensional MoS2 and MoS2-Based

Heterostructures. Chem. Soc. Rev. 2018, 47 (16), 6101–6127.

https://doi.org/10.1039/C8CS00314A.

(285) Qiu, B.; Zhao, X.; Hu, G.; Yue, W.; Ren, J.; Yuan, X. Optical Properties of Graphene/MoS2

Heterostructure: First Principles Calculations. Nanomaterials 2018, 8 (11), 962.

https://doi.org/10.3390/nano8110962.

(286) Liao, M.; Wu, Z.-W.; Du, L.; Zhang, T.; Wei, Z.; Zhu, J.; Yu, H.; Tang, J.; Gu, L.; Xing, Y.;

et al. Twist Angle-Dependent Conductivities across MoS 2 /Graphene Heterojunctions.

Nat. Commun. 2018, 9 (1), 4068. https://doi.org/10.1038/s41467-018-06555-w.

(287) Lin, W.; Zhuang, P.; Chou, H.; Gu, Y.; Roberts, R.; Li, W.; Banerjee, S. K.; Cai, W.;

Akinwande, D. Electron Redistribution and Energy Transfer in Graphene/MoS2

Heterostructure. Appl. Phys. Lett. 2019, 114 (11), 113103.

https://doi.org/10.1063/1.5088512.

(288) Liang, Y.; Wang, W.; Zeng, B.; Zhang, G.; Huang, J.; Li, J.; Li, T.; Song, Y.; Zhang, X.

Raman Scattering Investigation of Bi2Te3 Hexagonal Nanoplates Prepared by a

Solvothermal Process in the Absence of NaOH. J. Alloys Compd. 2011, 509 (16), 5147–

5151. https://doi.org/10.1016/j.jallcom.2011.02.015.


Recommended