probing excitonic mechanics in suspended and …

168
The Pennsylvania State University The Graduate School PROBING EXCITONIC MECHANICS IN SUSPENDED AND STRAINED TRANSITION METAL DICHALCOGENIDES MONOLAYERS A Dissertation in Physics by Hongchao Xie © 2019 Hongchao Xie Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy August 2019

Upload: others

Post on 20-May-2022

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

The Pennsylvania State University

The Graduate School

PROBING EXCITONIC MECHANICS IN SUSPENDED AND STRAINED TRANSITION

METAL DICHALCOGENIDES MONOLAYERS

A Dissertation in

Physics

by

Hongchao Xie

© 2019 Hongchao Xie

Submitted in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

August 2019

Page 2: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

The dissertation of Hongchao Xie was reviewed and approved* by the following:

Chaoxing Liu

Associate Professor of Physics

Dissertation Advisor, Chair of Committee

Kin Fai Mak

Associate Professor of Physics, Cornell University

Adjunct Associate Professor of Physics, Penn State University

Mikael C. Rechtsman

Downsbrough Early Career Development Professor of Physics

Venkatraman Gopalan

Professor of Materials Science & Engineering, Professor of Physics

Richard W. Robinett

Professor of Physics

Associate Head for Undergraduate and Graduate Students

*Signatures are on file in the Graduate School

Page 3: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

iii

ABSTRACT

Over the past decade, the interest in two-dimensional (2D) materials, especially for

atomically thin transition metal dichalcogenide (TMD) semiconductors, had dramatically thrived

for both fundamental science and practical applications. The reduced dielectric screening in 2D

mainly attributes to the strong excitonic effect in atomically thin TMD semiconductors. This

pronounced exciton feature can maintain at room temperature, which indicates strong light-matter

interaction and possible optoelectronic application using monolayer semiconductors. Meanwhile,

the absence of inversion symmetry and out-of-plane mirror symmetry jointly endows carriers in

monolayer TMDs with a new valley degree of freedom (DOF). Namely, in hexagonally-arranged

lattice of 2D materials, electrons that residing at band edges of K and K’ valleys can carry opposite

valley magnetic moments and Berry curvatures, which allows the further control of valley-indexed

carriers with polarized light, electrical and magnetic fields. Besides, the large strain sustainability

of monolayer TMDs gives rise to mechanically tunable band gap with 70 meV redshift per 1%

strain up to recorded 10% applied strain. Thus, the interaction of macroscopic mechanical means

with valley electrons makes monolayer TMD semiconductor a promising platform to implement

novel valley-controlled mechanical devices. This motivates the experimental studies demonstrated

in this dissertation.

In this dissertation, we investigate the valley contrasting coupling between optoelectronic

carriers (exciton & flowing electrons) and mechanics in a monolayer TMD semiconductor. In the

first parts (Chapter 1&2), I will present emerging properties of TMD monolayers and discuss

interesting physics that can study after suspending or straining these atomically thin materials. The

fabrication and measurement of typical TMD suspended devices will also be demonstrated in

details.

Page 4: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

iv

In the secondary part (Chapter 3), we demonstrate robust exciton bistability by continuous-

wave optical excitation in a suspended monolayer WSe2 at a much lower intensity level of 103

W/cm2. The observed bistability is originated from a photothermal mechanism, which can provide

both optical nonlinearity and internal passive feedback in a simple cavity-less structure. This is

supported by detailed excitation wavelength and power dependence studies of the sample

reflectance, as well as by numerical simulation including the temperature-dependent optical

response of monolayer WSe2. Furthermore, under a finite magnetic field, the bistability becomes

valley dependent and controllable not only by light intensity but also by light helicity due to the

exciton valley Zeeman effect, which open up an exciting opportunity in controlling light with light

using monolayer materials.

In the following part (Chapter 4), we report the observation of exciton-optomechanical

coupling in a suspended monolayer MoSe2 mechanical resonator. In particular, we have observed

light-induced damping and anti-damping of mechanical vibrations and modulation of the

mechanical spring constant by moderate optical pumping near the exciton resonance with variable

detuning. The observed exciton-optomechanical coupling strength is also highly gate-tunable. Our

observations can be fully explained by a model based on photothermal backaction and gate-induced

mirror symmetry breaking in the device structure. The observation of gate-tunable exciton-

optomechanical coupling in a monolayer semiconductor may find novel applications in

nanoelectromechanical systems (NEMS) and in exciton-optomechanics.

In the last part of this dissertation (Chapter 5), we present the study of magnetization purely

originated from the valley DOF in strained MoS2 monolayers. By breaking the three-fold rotational

symmetry in single-layer MoS2 via a uniaxial stress, we have demonstrated the pure electrical

generation of valley magnetization in this material, and its direct imaging by Kerr rotation

microscopy. The observed out-of-plane magnetization is independent of in-plane magnetic field,

Page 5: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

v

linearly proportional to the in-plane current density, and optimized when the current is orthogonal

to the strain-induced piezoelectric field. These results are fully consistent with a theoretical model

of valley magnetoelectricity driven by Berry curvature effects. Furthermore, the effect persists at

room temperature, opening possibilities for practical valleytronic devices.

Page 6: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

vi

TABLE OF CONTENTS

LIST OF FIGURES ................................................................................................................. ix

LIST OF TABLES ................................................................................................................. xvii

ACKNOWLEDGEMENTS ................................................................................................... xviii

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

1.1 Monolayer transition metal dichalcogenides (TMDs) ............................................... 3 1.1.1 Crossover from indirect to direct band gap ..................................................... 5 1.1.2 Electronic band structure ................................................................................. 7

1.1.3 Valley-dependent optical selection rules ....................................................... 10 1.1.4 Valley magnetic response .............................................................................. 11 1.1.5 Exciton .......................................................................................................... 12 1.1.6 Mechanical properties ................................................................................... 15

1.2 Nanoelectromechanical Systems (NEMS) .............................................................. 16 1.2.1 2D NEMS ...................................................................................................... 16 1.2.2 Mechanical model of membrane ................................................................... 18

1.3 Optomechanics ........................................................................................................ 23 1.3.1 Modeling optomechanical coupling .............................................................. 24 1.3.2 Optomechanical backaction .......................................................................... 27

Chapter 2 Experimental reviews ........................................................................................... 31

2.1 Device fabrication .................................................................................................... 31 2.1.1 Sample preparation ........................................................................................ 31

2.1.2 Van der Waals assembly ............................................................................... 33

2.1.3 Suspend atomically thin membrane .............................................................. 36 2.2 Optical spectroscopy of monolayer semiconductors ................................................ 37

2.2.1 Photoluminescence spectroscopy .................................................................. 38

2.2.2 Optical reflection spectroscopy ..................................................................... 39

2.2.3 Magnetic circular dichroism spectroscopy .................................................... 41 2.2.4 Characterization of mechanics in 2D NEMS ................................................ 42

Chapter 3 Valley selective exciton bistability in suspended monolayer WSe2 ..................... 45

3.1 Optical bistability ..................................................................................................... 46 3.2 Experimental setup and basic characterization of suspended WSe2 monolayer ...... 47

3.3 Observation of optical bistability in suspended monolayer WSe2 ........................... 50 3.3.1 Power dependence of monochromatic reflectance spectra across exciton

resonance ........................................................................................................ 50

3.3.2 Wavelength dependence of broadband reflectance spectra across exciton

resonance ........................................................................................................ 52

3.4 Modelling mechanism of the observed optical bistability ....................................... 54 3.4.1 Simulation of photothermal model ................................................................ 55

Page 7: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

vii

3.4.2 Comparison to the observed bistable exciton resonance ............................... 58

3.5 Valley contrasting bi-exciton resonance and control of light by helicity ................. 59 3.6 Conclusion ............................................................................................................... 62

Chapter 4 Tunable exciton-optomechanical coupling in suspended monolayer MoSe2 ....... 64

4.1 Exciton-optomechanical backaction through strain ................................................. 65 4.2 Experimental setup ................................................................................................... 66

4.3 Observation of exciton-optomechanical coupling in monolayer MoSe2 resonator .. 68 4.3.1 Optical response of strained MoSe2 resonator ............................................... 69

4.3.2 Spectral dependence of mechanical response................................................ 70

4.3.3 Photothermal backaction of exciton-optomechanical coupling .................... 72

4.4 Modelling exciton-optomechanical coupling in MoSe2 resonator ........................... 77 4.4.1 Gate dependence of exciton spectral shift ..................................................... 78

4.4.2 Gate dependence of mechanical resonance ................................................... 79

4.4.3 Modelling photothermal backaction in strained MoSe2 resonator ................ 80

4.5 Tunable exciton-optomechanical effects .................................................................. 83 4.5.1 Gate dependence of exciton-optomechanical coupling ................................. 83

4.5.2 Power dependence of exciton-optomechanical coupling .............................. 84

4.6 Detection of thermal vibrations of monolayer MoSe2 at cryogenic temperatures .... 86

4.7 Conclusion ............................................................................................................... 87

Chapter 5 Valley magnetoelectricity in single-layer MoS2 ................................................... 89

5.1 Magnetization from valley degree of freedom ......................................................... 90 5.2 Experimental setup ................................................................................................... 93

5.2.1 Device fabrication ......................................................................................... 94

5.2.2 Second-harmonic generation (SHG) ............................................................. 95

5.2.3 Photoluminescence (PL) spectroscopy .......................................................... 96

5.2.4 Magneto-optic Kerr rotation (KR) spectroscopy ........................................... 96

5.3 Observation of valley magnetoelectric effect in single-layer MoS2 ......................... 97 5.3.1 Valley magnetization in single-layer MoS2 ................................................... 97

5.3.2 Dependences of bias current and in-plane magnetic field ............................. 99

5.3.3 Hamiltonian model for valley magnetoelectric effect in strained

monolayer MoS2 ............................................................................................. 100

5.3.4 Dependences of the current and strain directions with respect to the

crystalline axis ................................................................................................ 101

5.3.5 Temperature dependence of valley magnetoelectric effect ........................... 104

5.4 Extended support experiments for the observed valley magnetoelectricity ............. 106 5.4.1 Dependence of Kerr rotation on probe polarization ...................................... 107

5.4.2 Dependence of Kerr rotation angle on probe photon energy ........................ 108

5.4.3 Kerr rotation under pure gate modulation ..................................................... 109

5.4.4 Simulation of valley magnetoelectric effect in Corbino disk geometry ........ 109

5.4.5 Effects of conductivity anisotropy ................................................................ 111

5.4.6 Comparison to the effect of trigonal warping of the band structure .............. 113

5.4.7 Valley Hall effect in single-layer MoS2 ........................................................ 114

5.5 Microscopic model of valley magnetization in strained and biased MoS2 .............. 116

Page 8: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

viii

5.5.1 Derivation of the piezoelectric field in single-layer MoS2 under a uniaxial

stress ............................................................................................................... 116

5.5.2 Electronic band structure of strained single-layer MoS2 ............................... 118

5.5.3 Valley magnetization ..................................................................................... 120

5.5.4 Circular dichroism ......................................................................................... 124

5.5.5 Comparison of microscopic theory with experiment .................................... 125

5.6 Conclusion ............................................................................................................... 127

Chapter 6 Summary and future perspectives ........................................................................ 128

Bibliography .......................................................................................................................... 132

Page 9: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

ix

LIST OF FIGURES

Figure 1-1: Trigonal prismatic structure of monolayer TMDs in 2H phase. The honeycomb

lattice structure with broken inversion symmetry and the first Brillouin zone with the

high symmetry points are shown. Figure is obtained from Ref. [23]. .............................. 4

Figure 1-2: Calculated electronic band structure of atomically thin MoS2. (a) Atomic

orbital projection of band structures for monolayer MoS2 from first-principle

predictions, without spin-orbit coupling. Fermi energy is set to zero. Symbol size is

proportional to its contribution to corresponding state. (1) Contributions from Mo-d

orbital. (2) Total contributions from p orbitals, dominated by S atoms. (3) Total s

orbitals. (b) The simulated band structure of bulk, quadri-layer, bilayer and single-

layer MoS2 (from left to right). A transition from indirect band gap to direct band gap

semiconductor occurs at monolayer limit. Figures are obtained from Ref. [30], [31]. .... 6

Figure 1-3: Optical selection rules and valley magnetic response. (a)Electronic bands (here

for W compounds) around K and K’ points, which are spin-split by the spin-orbit

interactions. The spin (up and down arrows) and valley (K and K’) degree of freedom

are locked together. Azimuthal quantum number (m) for each band is shown, and the

valley-dependent optical selection rules are depicted. (b) Zeeman shift of the K and

K’ valleys under an out-of-plane magnetic field. The dashed and solid lines represent

bands of opposite spin (split by spin-orbit coupling). 𝜇𝑠, 𝜇𝑉 and 𝜇𝑂 are the spin, valley

and intra-atomic orbital magnetic moments, respectively. Figures are obtained from

Ref. [23], [33]. .................................................................................................................. 10

Figure 1-4: Strong excitonic effect in 2D TMDs. (a) Schematics of electrons and holes

bound into excitons for the three-dimensional bulk and a quasi-two-dimensional

monolayer. The changes in the dielectric environment are indicated schematically by

different dielectric constants 휀3𝐷 and 휀2𝐷 and by the vacuum permittivity 휀0 .(b)

Absorption spectrum of monolayer MoS2 at 10K (solid green line). A and B are

exciton resonances corresponding to transitions from the two spin-split valence bands

to the conduction bands. The blue dashed line shows the absorbance (arbitrary units)

if excitonic effects were absent. The inset illustrates the Coulomb bounding between

an optically generated electron-hole pair, forming a bound exciton. Figures are

obtained from Ref. [23], [44]. .......................................................................................... 14

Figure 1-5: NEMS devices made of 2D materials. By fabricating suspended samples over

a trench structure, intrinsic properties include (a) electrical transport (e.g. carrier

mobility), (b) optical response (PL& linear absorbance) (c) mechanical properties (e.g.

Young’s modulus) and (d) thermal transport (thermal conductivity) of 2D materials

were systematically studied, which would be challenged for samples on substrate.

Also, 2D NEMS is a good platform to study interplay between nanomechanics and

other interesting DOF (e.g. charge, spin, valley etc.) in real 2D limit. Figures are

obtained from Ref. [10], [12], [29], [58]. ......................................................................... 18

Page 10: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

x

Figure 1-6: Schematic of the cavity optomechanical system. A mechanical oscillator

comprises one end of a Fabry-Perot cavity, coupling mechanical motion to the optical

cavity length. Figure is obtained from Ref. [108]. ........................................................... 24

Figure 1-7: Schematics of back-action effects in cavity optomechanics. (a) Schematic

diagram of light-induced force experienced by the end mirror as a function of

displacement in Fabry-Perot system. The finite decay rate of confined photons induces

a delayed response to mirror motion. (b) “P-V” style schematic depicting the work

done by the radiation force during one cycle of oscillation. The work is given by the

enclosed area swept in the force-displacement diagram, which is due to the retardation

of the force (finite cavity decay rate). The work is negative or positive, depending on

whether one is on the red-detuned or blue-detuned side of resonance. This thus gives

rise to damping or amplification, respectively. Figure is obtained and modified from

Ref. [109]. ........................................................................................................................ 28

Figure 2-1: Optical images of monolayer TMDs exfoliated on (a) silicon substrate with

285nm oxide layer atop and (b) PDMS substrate doubly-capsulated with thin hBN

flakes. ............................................................................................................................... 33

Figure 2-2: Process flow for heterostructure fabrication using PC as adhesive polymer.

This process flow can be used to fabricate heterostructure FET or NEMS device. The

final heterostructure maintains the stacking order. Step 1: Attach a PC/PPC/PDMS

stamp on the glass slide and place it on micro-manipulator of transfer stage; Step 2:

First layer pick-up; Step 3: Unload the bottom substrate and load the other exfoliated

substrate; Step 4: Second layer pick-up, Step 5: Place a new prepatterned substrate

with electrodes for final transfer; Step 6: The new prepatterned substrate makes a

contact with heterostructure and PC stamp at 150 C; Step 7: PC and heterostructure

detached from handle substrate and transferred to the new prepatterned substrate.

Figure is obtained from Ref. [137]. .................................................................................. 35

Figure 2-3: PDMS deterministic dry transfer method for 2D NEMS ..................................... 37

Figure 2-4: Schematics of microscopic confocal system in reflection geometry................... 39

Figure 2-5: Interferometry technique to characterize mechanics of 2D NEMS. (a)

Schematic of reflected beams from both suspended membrane and silicon surface. (b)

The normalized reflectance as a function of displacement of membrane. The blue

arrows indicate the transduction between displacement of membrane and reflectance.

The oscillation of membrane can periodically modulate the sample reflectance (brown

dash lines). Inset: gate dependence of normalized reflectance of suspended sample

with 450 nm probe wavelength. The red solid line is prediction with Eqns. (2-3) to (2-

6).. .................................................................................................................................... 43

Figure 3-1: Schematic of optical bistability. 𝐼𝑓 and 𝐼𝑏 are the critical intensities for

transmittance bistability under forward and backward intensity sweeps, respectively.

Figure is obtained from Ref. [150]. .................................................................................. 46

Page 11: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

xi

Figure 3-2: Experimental setup and suspended WSe2 device. (a) Sketch of experimental

setup. The suspended samples are placed in an optical cryostat. Optical excitation,

either broadband radiation or monochromatic beam, is illuminated onto sample. For

single-colored source, both the incident and reflected laser light are monitored by a

pair of photodiodes. The reflected broadband spectrum from the sample is recorded

by a dispersive spectrometer. (b) Schematic side view of a suspended monolayer WSe2

sample over a drumhead trench of 8 μm in diameter and 600 nm in depth on a Si

substrate. The substrate has a 600-nm oxide layer and pre-patterned Ti/Au electrodes.

The inset is an optical image of a sample illuminated by white light. Gold bars are

electrodes and the dark ring is the edge of the trench. ..................................................... 48

Figure 3-3: Temperature dependence of exciton resonance in suspended WSe2 monolayer.

(a) Reflectance spectrum measured at 4 - 330 K. A sharp exciton resonance is

observed at low temperatures. The resonance redshifts and broadens with increasing

temperature. (b) Simulated reflectance spectrum at 4 - 330 K. (c) Exciton peak

wavelength and linewidth (symbols) as a function of temperature used in the

simulation of (b). Solid lines are fits to Eqns. (3-2) and (3-3). ........................................ 49

Figure 3-4: Optical bistability in suspended monolayer WSe2. (a) Reflectance spectrum

of sample #1 at 4 K. (b) Absorbance spectrum described by a single Lorentzian that

best fits the reflectance spectrum of (a). (c, d) Reflectance at representative

wavelengths (708, 712, 720, 725 and 728 nm) under forward and backward power

sweeps from experiment (c) and simulation (d). The wavelengths are marked by

vertical dashed lines in (a, b)............................................................................................ 51

Figure 3-5: Bistable excitonic reflection spectra under monochromatic pump. (a, b)

Contour plots of reflectance of sample #2 at 4 K as a function of probe wavelength

(bottom axis) and pump wavelength (left axis). The pump wavelength is swept

forward (i.e. increases) in (a) and backward (i.e. decreases) in (b). The pump power is

fixed at 100 μW. The diagonal dashed lines are the residual pumps. (c, d) The

extracted exciton peak wavelength (c) and linewidth (d) from the experimental

reflectance spectra. (e, f) Simulation of the parameters shown in (c, d) based on the

photo-thermal effect as described in the main text. ......................................................... 53

Figure 3-6: (a) Schematic of the photo-thermal effect. The exciton resonance redshifts

under optical excitation for both blue- and red-detuned optical excitation from the

exciton resonance (shown by blue and red vertical arrows, respectively). For the red-

detuned optical excitation, the exciton resonance runs towards the pump and the effect

is cumulative. This internal passive feedback together with nonlinearity can lead to

optical bistability. (b) Simulated average sample temperature as a function of

excitation power at 708 nm (blue-detuned from the exciton resonance) and 728 nm

(red-detuned from the exciton resonance). 𝑃𝑓 and 𝑃𝑏 are the critical powers for

temperature instability under forward and backward power sweeps, respectively. (c)

Simulated average sample temperature as a function of wavelength at 1, 10 and 100

μW. 𝜆𝑓 and 𝜆𝑏 are the critical wavelengths for temperature instability under forward

and backward wavelength sweeps, respectively. The region between the critical

powers (wavelengths) is bistable. .................................................................................... 57

Page 12: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

xii

Figure 3-7: Wavelength dependence of the sample reflectance for forward (blue) and

backward (red) wavelength sweeps. (a, b, c) are experiment and (d, e, f) are simulation.

The excitation power is fixed at 1 μW (a, d), 10 μW (b, e) and 100 μW (c, f). ............... 59

Figure 3-8: Valley-dependent optical bistability and control of light by its helicity. (a)

Helicity-resolved reflectance of sample #3 at 4 K under an out-of-plane magnetic field

of 8 T. The K and K’ valley excitons are split by ~ 2 meV. The vertical dashed line

corresponds to 710 nm. (b) Helicity-resolved reflectance under power sweeps at 710

nm. There exists a finite power range between two bistability regions (between the

vertical dotted lines). (c) Magnetic circular dichroism (MCD) under forward and

backward power sweeps at 710 nm. (d) Helicity-controlled optical ‘switching’ in real

time at 710 nm and 39 μW. The helicity of light is modulated at 1 Hz by a liquid

crystal modulator (upper panel) and the sample reflectance follows the helicity

modulation with a change of reflectance by ~ 30% (lower panel). .................................. 61

Figure 4-1: Schematic illustration of the device geometry and experimental setup. The

mechanics of suspended membrane is characterized with optical interferometry

technique. And the incident and reflected beams are recorded by a normal and fast

photodetector, respectively. The static displacement 𝑧 and dynamical oscillation 𝛿𝑧 of membrane can transduce to reflection signals, which are measured with a

multimeter and a network analyzer correspondingly. Meanwhile, the vibration of

monolayer MoSe2 membrane generates a dynamical strain 𝛿𝜖 that give rise to an

oscillating spectral shift of exciton resonance δ𝜔𝑋. In addition, the illumination of the

sample with light near the exciton resonance can generate a photothermal force 𝐹𝑝ℎ.

As a result, the oscillation of monolayer semiconductor (oscillating exciton spectral

shift) can periodically modulate the photothermal force as δ𝐹𝑝ℎ. ................................... 68

Figure 4-2: Gate tuning of excitonic and mechanical resonances in suspended MoSe2

resonator. (a) Optical reflectance spectra of sample #1 are shown at selective gate

voltages (upper panel) and in contour plot (lower panel). The orange dashed line in

lower panel of (a) is prediction of Eqn. (4-6). (b) Mechanical resonance shift as a

function of gate voltages. The grey dashed line is the prediction of Eqn. (4-8). Inset:

A typical mechanical resonance of our resonator at 3.3K, and the corresponding

Lorentzian fit with Q ~ 20,000. ........................................................................................ 70

Figure 4-3: Interaction between excitonic and mechanical resonances in monolayer MoSe2

resonator. (a) Normalized reflectivity change pumped with 1 μW as a function of

driven frequency at selective wavelengths throughout the exciton resonance. The

spectra are displaced with equal spacing for clarity. Data was taken with 𝑉𝑔 = 40𝑉 at

3.3K. The illumination power is 1 μW. (b) The peak response of mechanical

resonance across exciton resonance (symbols with line) as a function of incident

wavelength (bottom axis) and exciton detuning Δ𝜔 (top axis). Data was extracted

from (a). The dashed red line is the illustration of spectral shift due to dynamical strain.

Inset: A sketch of a dynamical shift in the exciton resonance (δ𝜔𝑋). (c) The extracted

dynamic shift of exciton resonance δ𝜔𝑋 at selective gate voltages (symbols). The blue

solid line is obtained by multiplying quantity 𝑑𝜔𝑋

𝑑𝑉𝑔 extracted from the lower panel of

Fig.4-2 (a) with a dynamic voltage δ𝑉𝑔~ 70 mV. ............................................................ 72

Page 13: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

xiii

Figure 4-4: Light induced damping and anti-damping of mechanical modes across exciton

resonance in monolayer MoSe2 resonator. (a) Normalized reflectivity change pumped

with 10 μW as a function of driven frequency at selective wavelengths throughout the

exciton resonance. The spectra are displaced with equal spacing for clarity. Data was

taken with 𝑉𝑔 = 40𝑉 at 3.3K. The extracted (b) line width and (c) frequency of

mechanical resonance shown in (a) as a function of exciton detuning (bottom axis)

and incident wavelength (top axis). The green and blue solid lines in (b) and (c) are

extracted contributions from photothermal softening(symmetric) and backaction

(anti-symmetric). The red solid lines in (b) and (c) are fitting lines by summing up

green and blue lines. The data without symmetric contribution are shown in (d) and

(e), where the red solid lines now are fitting from Eqns. (4-1) and (4-13). ..................... 74

Figure 4-5: Gate dependence of optical damping across exciton resonance. (a) The shift of

mechanical linewidth induced by antisymmetric contribution as a function of exciton

detuning at selective gate voltages. Data is displaced with equal spacing for clarity.

The solid lines are fitting based on Eqns. (4-1) and (4-13). (b) The extracted amplitude

of the antisymmetric contribution of mechanical linewidth as a function of gate

voltages. The red dashed line is prediction based on Eqns (4-1) and (4-13). .................. 84

Figure 4-6: Power dependence of optical damping across exciton resonance. (a) The shift

of mechanical linewidth induced by antisymmetric contribution as a function of

exciton detuning at selective incident power. Data is displaced with equal spacing for

clarity. The solid lines are fitting based on Eqns. (4-1) and (4-13). (b) The extracted

amplitude of the antisymmetric contribution of mechanical linewidth as a function of

incident power. The red dashed line is prediction based on Eqns (4-1) and (4-13). ........ 85

Figure 4-7: Thermal motions of monolayer MoSe2 resonator. (a) Optical reflection

spectrum of sample #2 at 𝑉𝑔 = −60𝑉. The blue and green dashed lines are aligned

with 752 nm and 759 nm, which used for measurements of thermal noise in (b) and

(c). The red dash lines in (b) and (c) are Lorentzian fittings. ........................................... 87

Figure 5-1: Valley magnetoelectric effect in strained single-layer MoS2. (a) Piezoelectric

field 𝓔𝒑𝒛 is produced (right) as a result of stress-induced distortion of the hexagonal

lattice (left) and the associated ion charge polarization. Blue and yellow balls

represent Mo and S atoms, respectively. (b) Out-of-plane valley magnetization 𝑴𝑽 is

induced by an in-plane current density 𝑱 in the presence of the piezoelectric field 𝓔𝒑𝒛.

(c) Electronic band structure of n-doped single-layer MoS2 around the K and K’

valleys of the Brillouin zone. The band extrema are shifted from the K/K’ point (black

dashed lines) to opposite directions by a uniaxial strain, which also do not coincide

with the extrema of the Berry curvature distribution (blue dashed lines). The Fermi

level is tilted under an in-plane bias electric field. ........................................................... 92

Figure 5-2: Basic optical characterization in single-layer MoS2. (a) Intensity of the second-

harmonic component parallel to the excitation polarization (𝐼) as a function of the

excitation polarization angle measured from the armchair direction (𝜃) (symbols).

Solid line is a fit to 𝐼 = 𝐼0cos2(3𝜃) with 𝐼0 denoting the maximum intensity. (b)

Photoluminescence (PL) spectrum of single-layer MoS2 on a flexible substrate with

0% (blue), 0.1% (cyan) and 0.2% (green) strain along the armchair direction. Red line

Page 14: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

xiv

is the PL spectrum of the flake after being transferred onto a silicon substrate. The

spectra are normalized to the peak intensity and vertically shifted for clarity. ................ 94

Figure 5-3: Magneto-optic KR microscopy setup for strained single-layer MoS2. (a)

Optical image of strained MoS2 device. Black arrow indicates the direction of

electrical current. (b) Schematics of a MoS2 FET with bias voltage (Vsd) applied on

the source–drain electrodes and gate voltage (Vg) applied through the

Si/SiO2 substrate. For a strained single-layer MoS2 FET, a longitudinal electrical

current density J (black arrow) is optimally flowed in the transverse direction of

the strained-induced piezoelectric field Epz (orange arrow). The valley

magnetoelectric effect is detected by focusing a linearly polarized probe beam onto

the device under normal incidence and measuring the Kerr rotation angle θKR of the

reflected beam. ................................................................................................................ 97

Figure 5-4: Valley Hall effect and valley magnetoelectric effect in single-layer MoS2. Kerr

rotation image of an unstrained (a) and strained (b) single-layer MoS2 device under

two opposite bias directions (black arrows). Boundaries of the electrodes and the

device channel are marked in dashed black and green lines, respectively. The

unstrained device was measured with 𝑉𝑑𝑠= 2.5 V and Vg = 0 V (𝐽 = 22 A/m). KR is

observed at channel edges only. The strained device was measured with 𝑉𝑑𝑠 = 2.5 V

and Vg = 20 V (𝐽 = 13 A/m). KR is observed over the entire channel. ............................ 98

Figure 5-5: Dependences of KR angle on magnitude of bias current and in-plane magnetic

field. (a) Gate dependence of the KR at a fixed location on the strained device

(symbols) and gate dependence of channel current density 𝐽 measured with 𝑉𝑑𝑠 = 2.5

V (red solid line). Inset shows the corresponding bias dependences with Vg = 20 V. (d)

KR at a fixed location on the strained device as a function of in-plane magnetic field

B. Inset illustrates the Hanle effect for spin magnetization 𝑀𝑆 and the absence of it for

valley magnetization 𝑀𝑉. ................................................................................................. 99

Figure 5-6: Dependence on current direction of the valley magnetoelectric effect. a,

Piezoelectric field 𝓔𝒑𝒛 is produced along the armchair direction when single-layer

MoS2 is stressed along the armchair direction; b, KR image of sample (a) measured

in the Corbino disk geometry with 𝑉𝑑𝑠 = 1 V and Vg = 25 V (𝐽 = 3.9 A/m); c, Predicted

spatial distribution of valley magnetization by Eqn. (5-1). The direction of 𝓔𝒑𝒛 is

shown in orange lines. Boundaries of the electrodes and the device channel are marked

in dashed black and green lines, respectively. d to f are the same as a to c with single-

layer MoS2 stressed along the zigzag direction. The KR image was measured under

𝑉𝑑𝑠 = 6.9 V and Vg = 20 V (𝐽 = 2.2 A/m). ........................................................................ 103

Figure 5-7: Temperature dependence of the valley magnetoelectric effect. Gate

dependence of the two-point conductance (a) and KR at a fixed location on a strained

device (b) measured with 𝑉𝑑𝑠 = 2 V at four selected temperatures 10, 50, 90, and 110

K. Inset of b displays the gate dependence of the KR (symbols) and conductance (red

solid lines) at room temperature. ..................................................................................... 105

Figure 5-8: Valley ME effect at room temperature. a, Kerr rotation image of the device

shown in Fig. 5-6(b) at room temperature measured with Vds = 2.1 V and Vg = 20 V.

Page 15: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

xv

b, The background noise level measured with Vds = 0 and Vg = 0. c, Bias dependence

of the KR at a fixed location of the device channel (symbols) and the corresponding

current density (solid red line) measured with Vg = 20 V. .............................................. 106

Figure 5-9: Kerr rotation as a function of probe light polarization. KR was measured on

the strained device in Corbino disk geometry shown in Fig. 5-6(b) at two locations of

the device channel corresponding to two different bias fields (red and blue lines). The

linear polarization of the probe beam was varied by a half-wave plate and the sample

was fixed. ........................................................................................................................ 107

Figure 5-10: Dependence of KR on probe photon energy. Photoluminescence spectrum

shows the A exciton resonance in single-layer MoS2. The peak energy is marked by

the dashed line. KR was measured as a function of probe photon energy, the range of

which is limited by the tuning range of our laser. ........................................................... 108

Figure 5-11: Kerr rotation image under gate modulation. KR image of the unstrained

device shown in Fig. 5-4(a) measured with bias voltage Vds = 0 and gate voltage Vg

modulated at 4.11 kHz with a peak-to-peak amplitude of 2.5 V. Black and green

dashed lines indicate the electrode and sample boundaries, respectively. No

identifiable signal can be observed anywhere on the device channel. ............................. 109

Figure 5-12: Electrostatic simulation of Corbino disk devices. a, Microscope image of a

monolayer MoS2 device with Corbino disk electrodes: electrodes in gold, Si substrate

in grey and monolayer MoS2 in light blue. b, c, Contour plot of the simulated electric

potential (b) and electric-field strength (c) of the device. Inner and outer electrode are

fixed at 5 and 0 V, respectively. Black arrows in c indicate the local bias field direction

(arrow direction) and amplitude (arrow length). .............................................................. 111

Figure 5-13: Simulation on Corbino disk devices. a, Piezoelectric field ℇ𝑝𝑧 is produced

along the armchair direction when single-layer MoS2 is stretched along the armchair

direction. b, KR image of sample (a) measured in the Corbino disk geometry with 𝑉𝑑𝑐 = 1 V and Vg = 25 V. c, Predicted spatial distribution of valley magnetization based

on the valley ME effect. The direction of ℇ𝑝𝑧 is shown in orange lines. Boundaries of

the electrodes and the device channel are marked in dashed black and green lines,

respectively. d, Simulation of current density J including the conductivity anisotropy

induced by strain (0.5%, left column; 50%, right column). e to h are the same as a to

d with single-layer MoS2 stretched along the zigzag direction. The KR image was

measured under 𝑉𝑑𝑐 = 6.9 V and Vg = 20 V. .................................................................... 112

Figure 5-14: KR mapping of Corbino disk device under reversed bias. KR image with

current flowing radially outward (a) and inward (b). The KR signal flips sign under

reversed bias. .................................................................................................................... 113

Figure 5-15: Valley Hall effect in single-layer MoS2 at 30 K. a, Kerr rotation microscope

image of an unstrained monolayer MoS2 device under Vsd = 2.5 V and Vg = 0 V (J =

22 A/m). The boundaries of the electrodes and the device channel are shown in black

and green dashed lines, respectively. b, Conductivity as a function of gate voltage (Vg).

Inset: current as a function of bias voltage (Vds) with Vg = 20 V and a microscope

Page 16: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

xvi

image of the device. c, Gate dependence of the KR at three different locations on the

device channel indicated by circles in (a) measured with Vds = 2.5 V. Red, black and

blue line correspond to KR at the lower edge, center, and top edge of the channel,

respectively. d, Bias dependence of the KR at selected gate voltages near the lower

edge of the channel. ......................................................................................................... 115

Figure 5-16: Image of normalized KR signal by the channel current density for unstrained

(left) and strained (right) monolayer MoS2. ..................................................................... 116

Figure 5-17: Band structure of monolayer MoS2 at the K and K’ point. The shaded regions

illustrate the occupied states in the conduction band. a, Unstrained monolayer MoS2

under a finite bias field, giving rise to a tilted Fermi surface. b, Strained monolayer

MoS2 under bias. Net valley magnetization is generated in this configuration. The

dashed black lines indicate the K/K’ point and dashed blue lines indicate the extrema

of the Berry curvature distribution. .................................................................................. 123

Figure 5-18: Optical absorbance of monolayer MoS2 near the A exciton resonance at 10

K. ...................................................................................................................................... 126

Page 17: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

xvii

LIST OF TABLES

Table 4-1: Comparison of backaction effects in cavity optomechanics and this work. .......... 77

Table 5-1: Comparison of the KR measurements and the model based on conductivity

anisotropy in the Corbino disk devices. ........................................................................... 112

Table 5-2: Comparison of the effect reported in Ref. [185] and our work ............................ 114

Page 18: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

xviii

ACKNOWLEDGEMENTS

When looking back to my life in graduate school, it is great honor for me to work with lots

of wonderful people. Here I want to thank all the help and inspiration I got in past five years to

bring this dissertation to light.

First and foremost, I would like to acknowledge my academic advisor, Prof. Kin Fai Mak,

for giving me a chance to work on the exciting frontiers of two-dimensional physics. Fai has

incredible insight of physics and unceasing enthusiasm to explore the unknowns. He is always

very creative to come up with research topics and immediately find a smart way to solve the

problem. Along with his knowledgeable background of physics, Fai’s hands-on experimental skills

also impressed me much. When discussing with Fai on my research, I always do my utmost to

digest some of his wisdom and passion for science. I come to realize how generous and patient he

is before my experiments finally work. I would never forget the basic principles of research Fai

used to emphasize, “First, you know what to do. Then you know what you are doing. At last, you

understand what you have seen.” He had shown me how to be a great physicist with constant pursuit

of new ideas, solid understanding of physical picture, and critical attention to experimental details

that I am trying with greatest efforts to approach in the coming years. I am very fortunate to work

with Fai and indebted to him with training academically and personally.

I would like to thank Prof. Jie Shan for mentoring me as a co-advisor. Her uncompromising

questions in group meeting and in private discussion illuminate me to better understand my research.

Jie always explained those tough concepts to me with the greatest patience I can never find

elsewhere. Also, she kindly provided me with numerous suggestions on career development and

professionalism, which I particularly appreciate with. And her persistence in research and great

passion for science are invaluable personalities that I admired much. Without the support and

cultivation from Fai and Jie, I cannot imagine how far I can go on these dissertation works.

Page 19: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

xix

I would also like to thank Prof. Chaoxing Liu for serving as chair of my dissertation

committee and helping me finish lots of paperwork. As a student enrolled in Prof. Liu’s two courses:

statistical physics and semiconductor physics, I have benefited a lot from his lectures and achieved

a better understanding of my research problems. I am very thankful to Prof. Mikael C. Rechtsman

and Prof. Venkatraman Gopalan for their careful reading and valuable suggestions on my

dissertation.

I am extremely fortunate to be able to work in two amazing universities within my graduate

life – Penn State and Cornell. As one of the ‘oldest’ members in Mak and Shan group, it would be

a unique experience to witness the ‘never-off’ light shifted from labs on 3rd floor of Davey Lab at

Penn State to those on 2nd basement of Physical Science Building at Cornell. My thanks go to all

the former and current members of Mak and Shan group: Prof. Xiaoxiang Xi, Prof. Jieun Lee, Dr.

Liang Zhao, Prof. Xiuhong Zhu, Dr. Yanyuan Zhao, Dr. Zefang Wang, Egon Sohn, Dr. Yi-Hsin

Chiu, Dr. Shengwei Jiang, , Kaifei Kang, Lizhong Li, Dr. Yanhao Tang, Dr. Xiao-Xiao Zhang, Dr.

Tingxin Li, Tong Zhou, Dr. Yang Xu, Dr. Greg Stiehl, Dr. Chenhao Jin, Zui Tao, Avi Shragai and

all the young members.

Several people deserve my special thanks. I am extremely indebted to Jieun for her selfless

and patient mentoring during my first two years in Mak and Shan lab. Her knowledge background,

work ethic and first-rated skills set a high career model for me. Special thanks to Xiaoxiang, for

constantly helping me with both academical and personal stuff. I really appreciate all the knowledge

about superconductors from him. Another special thanks to Shengwei as my senior collaborator,

for giving me opportunities to work with him on several projects, and for teaching me how to work

smartly and efficiently. Thanks to my ‘long-term’ officemates: Egon and Zefang at Penn State and

Yanhao at Cornell for all the discussion of physics, music, films we had, just to name a few. I also

thank Yang for his experience on sample fabrication and discussion on magnetic materials. I also

Page 20: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

xx

thank Lizhong and Tingxin to be my ‘dinnermates’ occasionally.

I would also like to express my thanks to my classmates for discussion and company in

courses: Jiabin Yu, Yang Ge, Jiho Noh, Tsung-yao Wu, Ya-wen Chuang, Yu Jiang, Boyang Zheng,

Tao Wang, Dr. Jianyun Zhao, Yixuan Chen, Tomy Granzier-Nakajima, Sean Thomas, Justin

Rodriguez, Pavlo Bulanchuk, Krishnanand Mallayya, Tim Phillsbury, Umar Mauroof, Eugene

Tupikov, Daniel Schussheim, Leonardo E. Schendes Tavares, Ssohrab Borhanian, Benjamin Katz,

Neel Malvania, Patrick Godwin, Monica Tatiana Ricon Ramirez.

I also want to express my special thanks for friends I met throughout my PhD: Dr. Rui-

Xing Zhang, Prof. Bing Yao, Prof. Cui-Zu Chang, Dr. Zhong Lin, Dr. Yuhe Zhang, Dr. James

Kally, Dr. Gang Yang, Dr. Yufei Shen, Dr. Yixin Yan, Dr. Yu Pan, Dr. Qingze Wang, Dr. Di Xiao,

Dr. Jue Jiang, Dr. Jianxiao Zhang, Dr. Xiao Gan, Yifan Zhao, Songyang Pu, Yicheng Zhang, Run

Xiao, Jiali Lu, Tongzhou Zhao, Jiawei Wen, Xiaoche Wang (at Penn State) and Yanxin Ji, Ruofan

Li, Vishakha Gupta (at Cornell).

It would be a valuable experience to work in two of the best academic cleanrooms with

amazing staffs that include: Michael Labella, Kathy Gehoski, Guy Lavallee (at Penn State Nanofab)

and Garry Bordonaro, Aaron Windsor, Christopher Alpha (at Cornell CNF).

I am also grateful to Prof. Yuanbo Zhang at Fudan University for giving me a chance to do

undergraduate summer research intern in his lab and triggered my interest of two-dimensional

physics. I would also to thank Prof. Kostya Novoselov for his advising and encouragement on my

MPhys degree research at University of Manchester.

Finally, I would like to show my particular thanks to my beloved parents for their

unconditional love, understanding and support throughout my life. And my warmest thanks go to

Rui. With her tacit understanding and witty ideas, my life has been filled with joy and excitement.

Thank you and love you!

Page 21: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

xxi

To my parents

Page 22: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

Chapter 1

Introduction

Size and restricted geometry play decisive roles in physical properties of materials. In

reduced dimensionality, quantum physics teaches us that the energy of the spatially confined

electrons becomes discrete. In particular for two-dimensional (2D) space, in a strong magnetic field

and at ultra-low temperatures, the electron orbit can be completely quantized. In this case, the

resistance along the direction of the current vanishes completely, indicating a non-dissipating flow

of electrons, just like in superconductors. Additionally, the Hall effect — characterized by the

resistance measured perpendicular to the current demonstrates discrete steps with defined plateaus

over large intervals of magnetic fields. This quantization is called the quantum Hall effect [1].

Discovery of integer and fractional multiples of 𝑒2 ℎ⁄ (where 𝑒 denotes elementary charge of

electron, ℎ denotes Planck constant) in quasi-two-dimensional quantum wells lead to two Nobel

prizes for the study of these magic quantum mechanical effect [1]–[3]. Since then, people never

cease to hunt for realistic systems that can completely confine electrons in 2D and expect to exhibit

even more bizarre quantum effects.

Graphene, with only single layer of carbon atoms, is the first truly 2D system which had

been experimentally demonstrated by Andre Geim and Kostya Novoselov in 2004 [4]. This later

leads to their winning of the 2010 Nobel prize in physics. Date back to 1946, Canadian physicist

Philip Wallace pioneeringly hypothesized the existence of graphene and predicted the linear energy

dispersion without a band gap around the Fermi level [5]. This single hexagonal-structured layer

of carbon atoms has demonstrated exceptional properties more than prediction in versatile fields,

including ballistic transport [4], [6]–[8], quantum hall effect at room temperature [9], excellent

thermal conductivity [10], [11], high mechanical tolerance and flexibility [12], [13], high gate

Page 23: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

2

tunability [4], [7], [8]. Thus, graphene was initially expected to revolutionize the high-speed

electronic application [14], [15]. However, as a double-edged sword, the gapless electronic nature

of graphene make it challenged to fulfill applicable on/off ratio [14], [16] for modern electronic

device like transistors. Numerous efforts were applied to induce a band gap in graphene that include

breaking graphene superlattice’s symmetry by substrate [17], chemical doping [18] and electric

field [19]–[21]. However, the ‘gapped’ graphene cannot gain sufficient gap without sacrificing its

high mobility as a trade-off. In addition, the weak spin orbit coupling in graphene [22] hinders

novel quantum phenomena from emerging, which can be useful for further optoelectronic

application. Consequentially, much interest has been stimulated to other 2D materials beyond

graphene for both fundamental science and practical application [23], [24]. And transition metal

dichalcogenides (TMDs) is such a new class of 2D materials that attracts numerous interests

following the ‘gold rush’ of graphene.

Transition metal dichalcogenides (TMDs) is a class of II-VI semiconductors with the type

MX2, where M labels a transition metal atom (e.g. Mo, W, etc.) and X represents a chalcogen atom

(e.g. S, Se, Te). Most of known TMDs are formed with a layered structure, from which we can

isolate monolayer TMD from bulk crystal by either mechanical exfoliation or chemical growth.

Distinctly different from graphene, intrinsic semiconductor gaps with magnitude of 1~2 eV [23],

[24] naturally reside within monolayer TMD crystals. Ranging from visible to near-infrared, such

band gaps offer highly interesting optoelectronic application based on 2D materials. In addition,

restricted geometry in low dimensionality triggers a series of novel electronic and optical properties

in 2D TMDs. Most particular is the robust formation of a quasiparticle, namely as exciton, and

resulted occurrence of strong light-matter interaction in monolayer TMDs [23], [24]. This will

come as the core topic of our discussion in following chapters.

Nanoelectromechanical systems (NEMS) is a family of devices coupling mechanical

motion with electrical or optical functionalities in nanoscale. In NEMS devices, the movable

Page 24: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

3

ingredients can be actuated in static displacement or driven into dynamical vibration. Thus, the

mechanical motion in NEMS can be transduced into electrical or optical signals with high

sensitivity [25], [26]. Moreover, external perturbation can easily influence mechanical properties

of NEMS due to the miniscule nature built in such systems. Therefore, it would be highly interesting

to explore the probe-induced back-action effect on the mechanical performance of NEMS. When

integrated with 2D materials, this NEMS structure provides an ideal platform to study intrinsic

properties of atomically thin TMDs with mechanical knob. More importantly, by suspending

single-layered TMDs, it opens up tempting opportunities to study the interplay between mechanical

properties and rich electronic and optical features within this class of 2D materials. Our prior

interest will focus on coupling between mechanical mode and optical pumping through strong light-

matter interaction in suspended monolayer TMD semiconductor.

The main focus of this dissertation is about the NEMS based on TMD monolayers. By

suspension, both novel optical phenomena and interplay between mechanics and strong light-matter

interaction in 2D semiconductors were systematically studied and discussed in the following

sequence. First, we will discuss the emerging optical nonlinearity and its optical control in

monolayer WSe2 in Chapter 3. Second, we will demonstrate dynamical control of suspended

monolayer MoSe2 by light through the material’s strong excitonic resonance in Chapter 4. In

Chapter 5, along with flexible substrate, we will present the study of magneto-electricity effect in

strained single-layered MoS2. Here, we will begin our discussion with the introduction of

monolayer TMDs.

1.1 Monolayer transition metal dichalcogenides (TMDs)

The study of TMDs has a long history which can date back to 1963 for obtaining ultra-thin

MoS2 samples with use of adhesive tape [27]. Over the past decade, the ‘gold rush’ of graphene

Page 25: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

4

stimulated the resurgence of interest for this class of van der Waals materials. In general, there are

several prototypes of lattice structure for TMDs including 1T, 1T’, 2H and 3R [28]. Throughout

this dissertation, we only focus on TMDs stacked in 2H phase since the rich physics that we will

discuss only exist in this structure phase. In 2H phase, single-layered TMDs form honeycomb

lattice from top view, from which a layer of transition metal atom M is sandwiched by two layers

of chalcogen atoms X to form trigonal prismatic structure [23]. Similar to graphene, monolayer

layered TMDs also possess K and K’ valleys in momentum space. When thinned down to

monolayer limit, though its atomic structure remains unchanged, the reduced dimensionality

induces renormalization of electronic structure and new crystalline symmetry in 2D TMDs [23].

Thus, emerging electronic and optical features distinguish monolayer TMDs from their bulk

counterparts as we discussed below.

Figure 1-1: Trigonal prismatic structure of monolayer TMDs in 2H phase. The honeycomb lattice

structure with broken inversion symmetry and the first Brillouin zone with the high symmetry

points are shown. Figure is obtained from Ref. [23].

Page 26: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

5

1.1.1 Crossover from indirect to direct band gap

Here, we take MoS2 as a sample and discuss layer-dependent evolution of semiconductor

gaps in atomically thin TMDs. The bulk MoS2 is well-known as an indirect band gap semiconductor

with the gap size of 1.29 eV [29]. And this indirect band transition occurred between the valence

band maximum at Γ point and the conduction minimum along the Γ and Κ points of Brillouin zone.

Meanwhile, the smallest direct band gap resides at the K point with a gap size of 1.8 eV [29]. In

addition, the electronic bands associated with band gaps above have different orbital origins. At Κ

point, the direct band gap is mainly contributed from the 𝑑 orbitals of Mo atoms (𝑑𝑧2 for the

conduction band minimum and 𝑑𝑥2−𝑦2 for the valence band maximum, respectively) [29]–[31]. In

contrast, both 𝑑𝑧2 orbital of Mo atoms and 𝑝𝑧 orbital of S atoms contributed mostly to the indirect

band gap along the Γ and Κ points [29]–[31]. Note that orbitals from S atoms not only interact with

atoms within the same layer but also ‘talk’ to orbitals originated from atoms at neighboring layer.

Since Mo atoms sandwiched between two layer of S atoms, the direct bandgap states from Mo-𝑑

orbitals are more localized and less influenced by interlayer coupling in out-of-plane direction

compared to the indirect bandgap states which include S- 𝑝 orbitals. As the thickness reduces, the

indirect band gap gradually opens up while the direct bandgap remains almost unchanged due to

increasing quantum confinement effect. When thinned down to monolayer limit, the indirect band

gap along the Γ and Κ points exceed the direct band gap at Κ point. Thus, MoS2 undergoes a

crossover from an indirect band gap to direct band gap semiconductor in monolayer as depicted in

Fig. 1-2. Mak et al. [29] and Splendiani et al. [30] pioneeringly investigated this layer-dependent

evolution of electronic structure in MoS2 through optical measurements. Through systematical

characterizations, particularly for the work of Ref. [29], optical feature from the indirect band gap

consistently blue shifted down to bilayer and completely disappeared in monolayer limit.

Meanwhile, the feature from direct band gap transition changed little down to monolayer with at

Page 27: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

6

least two order of magnitude stronger in scale than the counterpart from indirect band gap. All the

experimental observation demonstrated that there was a crossover between indirect to direct band

gap occurring in monolayer MoS2 [29]–[31]. The indirect to direct band gap transition is essentially

the same for other group VI semiconductor monolayers described above using MoS2.

Figure 1-2: Calculated electronic band structure of atomically thin MoS2. (a) Atomic orbital

projection of band structures for monolayer MoS2 from first-principle predictions, without spin-

orbit coupling. Fermi energy is set to zero. Symbol size is proportional to its contribution to

corresponding state. (1) Contributions from Mo-d orbital. (2) Total contributions from p orbitals,

dominated by S atoms. (3) Total s orbitals. (b) The simulated band structure of bulk, quadri-layer,

bilayer and single-layer MoS2 (from left to right). A transition from indirect band gap to direct band

gap semiconductor occurs at monolayer limit. Figures are obtained from Ref [30], [31].

Page 28: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

7

1.1.2 Electronic band structure

In this section, we illustrate the basic properties of electronic band structure near the direct

band gap in monolayer MoS2 and the general description can also be applied to other group VI

TMDs monolayers. The trigonal prismatic structure of one MoS2 unit cell splits the d orbitals of

Mo atom into three subgroups: 𝐴1(𝑑𝑧2), 𝐸(𝑑𝑥𝑦, 𝑑𝑥2−𝑦2) and 𝐸′(𝑑𝑥𝑧, 𝑑𝑦𝑧) [31], [32]. As discussed

in section 1.1.1, the in-plane mirror symmetry allows the hybridization only between 𝐴1 and 𝐸

from the conduction band and valence band [31], [32], respectively, which opens the direct band

gap at K or K’ (namely, K’) point in Brillouin zone. Thus, the wavefunctions that describe states

in the adjacency of K/K’ points are given as the following symmetry-adapted basis, [32]

|𝜙𝑐⟩ = |𝑑𝑧2⟩ |𝜙𝑣𝜏⟩ =

1

√2(|𝑑𝑥2−𝑦2⟩ + 𝑖𝜏|𝑑𝑥𝑦⟩) (1-1)

where |𝜙𝑐⟩ is the state basis for conduction band, |𝜙𝑣⟩ is for valence band, and τ denotes ± with

upper and lower signs for the K (K’) valley index related by time reversal operation. To lowest

energy approximation, two-band k ∙ p theory gives out an effective Hamiltonian as [32]

��0 = 𝑎𝑡(𝜏𝑘𝑥��𝑥 + 𝑘𝑦��𝑦) +Δ

2��𝑧 (1-2)

where �� denotes the Pauli matrices for the tow basis functions above, 𝑎 is the lattice constant, 𝑡 is

the effective hopping integral, and Δ is the energy gap. This effective Hamiltonian of monolayer

MoS2 is akin to the ‘gapped’ graphene since they share the similar symmetry properties. And the

nature of the heavy metal d orbitals provides strong spin orbit coupling (SOC) to lift the spin

degeneracy within electronic bands. In addition, the reflection symmetry within a layer only allows

SOC to have out-of-plane component [31], [32]. Namely, under first order approximation, orbitals

of 𝐸(𝑑𝑥𝑦, 𝑑𝑥2−𝑦2) have non-zero SOC component while SOC from 𝐴1(𝑑𝑧2) vanishes.

Consequently, the top valence bands at K and K’ points are spin-polarized and separated by

hundreds of meV while the conduction bands remain spin degenerate [32]. This spin-polarized top

Page 29: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

8

valence bands give rise to two optical transitions from upper and lower valence bands to the

conduction bands at K and K’ points, which denoted as resonance A and B, respectively [29], [30].

The resonance A with lowest transition energy is our prior interest in this dissertation since major

population of carrier involves within this transition process. This induces prominently strong light-

matter interaction and related novel phenomena which we will present more related details in

Chapter 3 and 4. In reality, finite spin splitting of conduction bands at K and K’ points [23] can be

generated by SOC in TMDs with heavier atoms include MoSe2, WS2 and WSe2 monolayers. It

plays an important role on formation of richer quasiparticle complexes [23] and long lifetime of

valley carriers [23], [33]. But all these interesting physics go beyond the interest of this dissertation

and we will not discuss the phenomena induced by spin-polarized conduction bands. The 𝑘 ∙ 𝑝

model-based Hamiltonian above neglects SOC and now we modify ��0 with SOC approximated by

intra-atomic contribution 𝑳 ∙ 𝑺. The total Hamiltonian is then transformed as [32]

�� = 𝑎𝑡(𝜏𝑘𝑥��𝑥 + 𝑘𝑦��𝑦) +Δ

2��𝑧 − 𝜆𝜏

��𝑧−1

2��𝑧 (1-3)

where 2𝜆 is the spin splitting generated by SOC at the valence band edge and ��𝑧 is the Pauli matrix

for spin. It is noteworthy to emphasis that the spin orientation of upper valence band is opposite at

K and K’ points due to time reversal symmetry. This ‘spin-valley coupling’ effect gives rise to

unique valley-contrasting optical selection rules [23], [33] and further magnetic control [23], [33]

of valley index as a new degree of freedom. We will cover these two novel phenomena in next two

sections.

This spin splitting is not limited to specific materials and directly resulted from inversion

symmetry breaking. In monolayer MoS2, the inversion symmetry of lattice is intrinsically broken

due to the trigonal prismatic arrangement of S atoms [31], [32]. Namely, if taking the Mo atom as

the inversion center, an S atom will map to a vacuum position without identical counterparts.

Except for bandgap opening at K and K’ points, another explicit consequence of inversion

Page 30: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

9

symmetry breaking is valley-contrasting orbital magnetic moment 𝝁𝑉 [33], [34] in TMD

monolayers, which originated from self-rotation of electron wavepackets. This valley magnetic

moment can impose an anomalous velocity into a bias-driven electron in the vicinity of K valley.

The direction of anomalous velocity [33], [34] is switched at K’ valley to satisfy time reversal

symmetry. Intuitively, this effect can be viewed as the drifted electron experienced a Lorentz force

induced by a magnetic field in momentum space, namely the Berry curvature Ω𝐵 [33], [34]. Note

that the valley magnetic moment 𝝁𝑉 is proportional to the Berry curvature Ω𝐵 since these two

physical quantities share the same geometric origin [33], [34]. We will focus more on the valley

orbital magnetic moment due to its direct response to external fields. For 2D massive Dirac system

as monolayer TMDs, the valley magnetic moment has the simple form 𝝁𝑉 = 𝜏𝑒ℏ

2𝑚∗ �� at K and K’

points [33], [34], where 𝑚∗ is the effective mass of the Bloch band and �� is the unit vector pointing

out-of-plane. It mimics a spin magnetic moment with an effective Bohr magneton 𝜇𝐵∗ =

𝑒ℏ

2𝑚∗. In

addition, the conduction and valence bands carry valley magnetic moments with the same sign [33].

At the presence of an external magnetic field, the valley magnetic moment can generate a Zeeman-

type response as −𝝁𝑉 ∙ 𝐵. This coupling between valley magnetic moment and external magnetic

field, together with the strong SOC introduced above, allows the control of the valley degree of

freedom in TMD monolayers by circularly polarized light [35]–[37] and out-of-plane magnetic

field [38]–[43] as we discuss in section 1.13 and 1.14 for more details below.

Page 31: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

10

1.1.3 Valley-dependent optical selection rules

Th opposite valley magnetic moment explicitly gives rise to valley-dependent optical

selection rules in single-layer TMDs. Originated mostly from 𝑑𝑧2 orbitals, the conduction band

edges denote a zero magnetic quantum number, whereas the valence band edges with

|𝑑𝑥2−𝑦2⟩ + 𝑖𝜏|𝑑𝑥𝑦⟩ orbitals carry a magnetic quantum number 2𝝉 [33]. Due to the canceled spin

contribution, inter-band transition is thus forbidden at K and K’ points if only considering atomic

orbital angular moment under electric dipole approximation ∆𝑚 = ±1 [33]. However, as we

discussed in previous section, the valley magnetic moment at K and K’ points carries a magnetic

Figure 1-3: Optical selection rules and valley magnetic response. (a)Electronic bands (here for W

compounds) around K and K’ points, which are spin-split by the spin-orbit interactions. The spin

(up and down arrows) and valley (K and K’) degree of freedom are locked together. Azimuthal

quantum number (m) for each band is shown, and the valley-dependent optical selection rules are

depicted. (b) Zeeman shift of the K and K’ valleys under an out-of-plane magnetic field. The dashed

and solid lines represent bands of opposite spin (split by spin-orbit coupling). 𝜇𝑠, 𝜇𝑉 and 𝜇𝑂 are the

spin, valley and intra-atomic orbital magnetic moments, respectively. Figures are obtained from

Ref.[23], [33].

Page 32: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

11

quantum number −𝝉, which naturally lead to the total angular momentum change ±𝝉 at K and K’

valleys, respectively [33]. This allows the occurrence of direct optical transition at K and K’ valleys

exclusively coupled to the incident light with opposite helicity, namely the left- and the right-

circularly polarized light [33], [34]. The valley-contrasting optical selection rules have been

experimentally observed in single-layered TMDs by several groups independently [38]–[42]. Due

to SOC-induced spin valley locking, together with large momentum separation of K and K’ valleys,

the emitted photoluminescent (PL) preserves a large PL handedness ρ that suggests well-

maintained valley polarization upon excitation near the transition energy [33], [34].

1.1.4 Valley magnetic response

The presence of valley magnetic moment opposite at K and K’ valleys also lead to

selectively addressing two valleys by a magnetic field perpendicular to TMD monolayers. The total

magnetic moment 𝝁𝑡𝑜𝑡 imposed on an electron in the vicinity of K and K’ valleys is largely

attributed by summing up three independent contributions: spin magnetic moment 𝝁𝑠, intra-atomic

orbital magnetic moment 𝝁𝑂 and valley (inter-atomic) magnetic moment 𝝁𝑉 [33], [34]. Through

Zeeman interaction in finite magnetic field, both conduction and valence bands experience energy

shifts under the influence of these three magnetic moments [33], [34]. Within those allowed inter-

band transition, the spin-induced Zeeman shifts cancel each other for conduction and valence bands

at K and K’ valleys. The intra-atomic orbital magnetic moments contribute energy shift of 2𝜏𝝁𝐵𝐵

owing to different nature of d orbitals for conduction and valence band edges [33]. Similarly, the

Zeeman contribution from valley magnetic moments is given as 𝜏𝝁𝐵𝐵(𝑚0

𝑚𝑐∗ −

𝑚0

𝑚𝑣∗), which stems

from different effective mass of conduction and valence bands [33], [34]. In the vicinity of K and

K’ points, the effective masses of conduction and valence band edges are approximately equal [31],

Page 33: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

12

[32]. Thus, the total Zeeman splitting between K and K’ valleys can further reduce to −4𝝁𝐵𝐵

purely due to the contribution of intra-atomic orbital moment. This valley Zeeman splitting matches

the experimental results with −(3.8 ± 0.2)𝝁𝐵𝐵 surprisingly well [38], [39]. In addition, it lifts the

degeneracy of left-and right circularly polarized emission in energy scale and allows the further

control of K and K’ valleys individually, together with valley-contrasting optical selection rules.

1.1.5 Exciton

Exciton is a quasi-particle made of an electron and hole pair bound by Coulomb interaction.

With optical illumination in gapped materials, an electron can be excited to the conduction band

and left a hole in the valence band by absorbing an incident photon. As the explicit consequence,

such electron-hole pair is correlated by Coulomb interaction and forming a hydrogen-like

compound, namely exciton. Though their different physical origins, it is still helpful to make an

analogy between exciton and well-studied hydrogen atom. First, the binding energy of exciton is

similar to the ionization energy of hydrogen atom, which lowers the total energy from the energy

sum of individual electron and hole. In addition, exciton’s binding energy is easily influenced by

dielectric environment due to its origin from Coulomb interaction. Second, as in hydrogen atom,

Bohr radius is an important physical quantity to characterize the wavefunction of exciton. It

classifies different types of exciton in length scale stemming from competition between Coulomb

interaction and Bloch lattice potential. Like hydrogen atom, excitons demonstrated various of

excited states with different energy and symmetries.

Typically, excitons can be categorized as Frenkel type and Wannier-Mott (Wannier for

simplicity) type. Frenkel excitons generally exist in weak dielectric systems with strong Coulomb

interaction and large binding energy (~ 0.1eV). As the explicit consequence, these tightly-bound

excitons localize within Bohr radius smaller than lattice spacing, of which the Coulomb coupling

Page 34: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

13

dominates over lattice potential. In contrast, Wannier-typed excitons reside in material systems

with large dielectric constant. Since the wavefunction of Wannier exciton spread over several unit

cell, the lattice potential can be ‘felt’ by Wannier excitons and be incorporated into the effective

masses of the electron and hole as Bloch particles. As result, the binding energy of Wannier

excitons is around ~ 0.01eV much smaller than localized Frenkel counterparts. All the excitons

discussed in this dissertation are Wannier-typed excitons.

Dimensionality, in addition to intrinsic dielectric capability of materials, is another influential

element to engineer properties of excitons. With the development of nanotechnology, material

systems in reduced dimensionality was fabricated and intensively studied, some of which show

promising potential as platform to study excitons. From 0D quantum dot, 1D carbon nanotube to

quasi-2D quantum wells, geometric restriction enhances the quantum confinement effect and also

further reduces dielectric screening by allowing a large portion of electric fields from electron-hole

pair to extend into the vacuum [44]. Thus, excitonic features are more pronounced due to the

enhanced Coulomb interaction in low dimensionality. And excitons had been intensively studied

in quantum wells system [45] and manifested lots of exciting physics include the strong signature

of exciton Bose-Einstein condensation [46], [47]. Monolayer TMD semiconductor is true 2D

system which basically is surface everywhere. Consequentially, both theoretical calculation and

experimental measurements have showed huge exciton binding energy ranging from 0.3 to 1 eV

for different TMDs [33], which is roughly one orders of magnitude larger than those in quasi-2D

quantum wells [23]. The significantly high binding energy of excitons in TMD monolayers gives

rise to the robust exciton resonance up to room temperature [23]. More interestingly, together with

strong SOC and valley dependent optical selection rules, excitons with contrasting valley index can

be selectively excited and controlled by different handedness of circularly polarized light [35]–

[37]. The valley-dependent optical selection rules also become spin-dependent due to spin-valley

coupling [32] and thus allow excitons being further distinguished by external magnetic field [38]–

Page 35: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

14

[42]. In addition, rich exciton species in TMD monolayers had been pronouncedly identified and

characterized at higher temperature [23], [33], which are usually elusive in other systems. These

excitons beyond fundamental excitation include charged exciton (trion), exciton pair (bi-exciton)

and even higher order excitonic complexes [23], [33].

Throughout this dissertation, we only focus on the 1s typed fundamental exciton state in

single-layered TMD semiconductors. Near this fundamental exciton resonance, high reflectivity

(>80%) with narrow linewidth (~ 2 meV) from TMD monolayers had been experimentally

demonstrated [48]–[50], which indicated the occurrence of extremely strong light-matter

interaction. This ‘cavity-like’ optical response originated purely from monolayer semiconductors.

Without substrate screening, the binding energy of excitons in suspended TMD monolayers can be

further enhanced [51]. In addition to the intrinsic properties of excitons, we are more interested in

Figure 1-4: Strong excitonic effect in 2D TMDs. (a) Schematics of electrons and holes bound into

excitons for the three-dimensional bulk and a quasi-two-dimensional monolayer. The changes in

the dielectric environment are indicated schematically by different dielectric constants 휀3𝐷 and 휀2𝐷

and by the vacuum permittivity 휀0.(b) Absorption spectrum of monolayer MoS2 at 10K (solid green

line). A and B are exciton resonances corresponding to transitions from the two spin-split valence

bands to the conduction bands. The blue dashed line shows the absorbance (arbitrary units) if

excitonic effects were absent. The inset illustrates the Coulomb bounding between an optically

generated electron-hole pair, forming a bound exciton. Figures are obtained from Ref. [23], [44]

Page 36: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

15

optically tuning physical properties of suspended TMD monolayers by using exciton resonance as

a cavity-less ‘cavity resonance’. Thus, with large on-resonance excitonic reflectivity and narrow

linewidth, suspended TMD monolayer stands out as an ideal platform to study numerous optical

phenomena include optical bistability that conventionally required delicate cavity design and

operation [52]. Meanwhile, it is particularly compelling to study the exciton-optomechanical

coupling with such cavity-less structure. In Chapter 3, we will show that the suspended WSe2

monolayers exhibit power- and wavelength-dependent nonlinearities. Together with unique valley-

contrasting optical rules and magnetic response, this exciton-induced bistability can help to control

light not only with light intensity but also with its polarization using monolayer materials. In

Chapter 4, we will demonstrate the light-induced damping and anti-damping of mechanical

vibrations and modulation of the mechanical spring constant at the μW level near the exciton

resonance with variable detuning in a suspended MoSe2 monolayer. Owing to its strain tunability

of exciton resonance energy, the monolayer MoSe2 resonator generates a dynamical strain as

vibrating and manifested different mechanical responses that stems from contrast photothermal

back-action at opposite detunings of exciton resonance. These two works above will open up

exciting opportunities to study and engineer optical nonlinearity and cavity-less optomechanics

using a monolayer semiconductor.

1.1.6 Mechanical properties

Intrinsically, the stability of the sp2 bounds gives rise to graphene’s exceptional mechanical

properties. Measurements from AFM nano-indentation show that the in-plane elastic stiffness of

graphene yields of 340 N/m [12], which corresponds to a bulk Young’s modulus of 1TPa. The

measurements also reveal the ultra-strength of graphene with a breaking strain of 25% [12]. Such

Page 37: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

16

large mechanical strength and strain sustainability greatly exceed those for any of conventional

materials used for NEMS application.

Atomically thin TMDs include MoS2, WS2, and WSe2 also share similar mechanical

properties as their graphene cousin. Through the similar indentation measurements, either

exfoliated or CVD grown MoS2 monolayers were found to have a Young’s modulus around 270

GPa [53], [54] with applied strain of 10% before breaking [53]. The mechanical properties of

tungsten-based TMDs have also been identified. For few-layered (4-15 layers) WSe2 by exfoliation

is found to have a Young’s modulus of 167 GPa [55] with a breaking strain of 7% [55]. Also, the

CVD-grown WS2 monolayers demonstrate a Young’s modulus with 272 GPa [54], similar to that

of single-layered MoS2. With external strain applied, both theoretical calculation and experimental

results show a redshift of band gap energy in both direct and indirect transition of TMDs that is

measured to be linear with strain. And the correspond redshift at a rate of ~70 meV per percent

applied strain for direction gap transitions [56] has been determined by optical and electronical

means for monolayer TMD semiconductors [57]. Such large strain sustainability introduced above,

together with pronounced strain-tunability in band gap, allows the mechanical tuning of electronic

properties of TMD monolayers [57]. Simultaneously, it opens up the possibility of coupling novel

optical properties with mechanical modes in suspended TMD monolayers as we will discussed

below.

1.2 Nanoelectromechanical Systems (NEMS)

1.2.1 2D NEMS

Due to the nature of two-dimensional materials, the emerging properties of TMDs monolayers

illustrated in the above sections can be greatly influenced by the substrate they are residing on. Any

Page 38: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

17

external perturbation particularly includes substrate effects (roughness, dielectric screening,

contamination etc.) can limit the study of intrinsic properties in TMD monolayers. The natural

thought to solve this problem is to suspended these atomically thin flakes to get rid of the substrate

effect. Through this way, the ultra-high carrier mobility [58], [59], thermal conductivity[10], [11]

and mechanical properties [12], [13] of graphene, as well as fundamental properties [53]–[55],

[60]–[64]of 2D TMDs, had been studied, which is challenged to investigate in the substrate-

supported samples.

Meanwhile, we can fabricate nanoelectromechanical system (NEMS) within atomic scale

using these suspended 2D membranes. And atomically thin monolayer materials have attracted

considerable interest in the studies of NEMS devices [65]–[86] because of their extremely

lightweight, mechanical flexibility, gate-tuneability and strong inter-mode couplings[75], [81],

[84], [86]. They also hold appealing potential in versatile sensing applications [80], [87]–[98]. In

turns, the high sensitivity of 2D NEMS is a benefit to not only study these intrinsic properties of

materials, but also to study interplay between thin film mechanics and other properties of 2D

materials[60], [72], [74], [99]–[105], especially for using mechanical resonance to probe the charge

density wave[102]–[104], superconductivity[72], [103] in 2D electron-correlated materials.

For 2D semiconductors, although the mechanical aspects of NEMS devices based on

monolayer materials have been extensively studied in the literature [60], [65], [66], [80], [85], the

dynamical coupling between mechanical vibrations and light (i.e. optomechanical coupling)

remains as an open subject. In particular, the presence of strong excitonic resonances in monolayer

TMD semiconductors[23], [33], which strongly interact with light (can reflect ~ 90 % of the

incident light) [48]–[50], presents a very interesting avenue to explore optomechanical coupling

effects in atomically thin materials without the need of an optical cavity.

Page 39: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

18

1.2.2 Mechanical model of membrane

In this section, we will review the static and dynamical mechanics of a suspended membrane.

The mechanical model developed below is aimed to practically extract physical quantities (e.g.

displacement of membrane, Young’s modulus, initial strain etc.) from our optomechanical

experiments using gate voltages.

The differential equation of motion for a circular membrane with clamped edges under

uniform tension is given as [106]

𝜌𝜕2𝑧

𝜕𝑡2= −𝐷∇4𝑧 + 𝜎∇2𝑧 + 𝑃 (1-4)

Figure 1-5: NEMS devices made of 2D materials. By fabricating suspended samples over a trench

structure, intrinsic properties include (a) electrical transport (e.g. carrier mobility), (b) optical

response (PL& linear absorbance) (c) mechanical properties (e.g. Young’s modulus) and (d)

thermal transport (thermal conductivity) of 2D materials were systematically studied, which would

be challenged for samples on substrate. Also, 2D NEMS is a good platform to study interplay

between nanomechanics and other interesting DOF (e.g. charge, spin, valley etc.) in real 2D limit.

Figures are obtained from Ref. [10], [12], [29], [58].

Page 40: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

19

where z is the displacement of the membrane, 𝜌 is the area density, 𝐷 is the 2D bending stiffness,

𝜎 is the 2D tension, and 𝑃 is the pressure acting on the membrane. The above equation is adapted

from Foppl-von Karman model which is used for describing the deflection of thin flat plates[106],

[107]. Before further derivation, we can just intuitively analysis the contributions from each term

that related to mechanical properties. The 2D bending stiffness is defined as 𝐷 =𝐸𝑌𝑡

2

12(1−𝜈2) [106],

where 𝐸𝑌 is the Young’s modulus of the membrane material, 𝑡 is the thickness of the plate, and 𝜈

is the Poisson ratio. If we plug in the typical numbers for TMD monolayers include 𝐸𝑌 ~ 200 𝐺𝑃𝑎,

𝑡 ~ 0.7 𝑛𝑚 and 𝜈 ~ 0.18 [60], then the area bending stiffness is 𝐷 ~ 10−8 . For the tension

contribution, by taking a typical tension 𝜎 ~ 0.1𝑁/𝑚 and the diameter 𝐿 ~10 𝜇𝑚 of our

monolayer NEMS, the tension term is of typical value around 𝜎𝐿 ~ 10−6 . From the simple

estimation above, we can reasonably conclude that the tension is dominant within our monolayer

sample and the bending term can safely drop out for further discussion. For equilibrium condition,

after setting the motion term to zero at the left-hand side of equation, the equation of motion for

the clamped membrane is rewritten as

𝜎∇2𝑧 + 𝑃 = 0 (1-5)

For simplicity, we neglect the angle contribution and the Laplace notation only has radial

dependence as ∇2=1

𝑟

𝜕

𝜕𝑟+

𝜕2

𝜕𝑟2 in cylindrical coordinate. With the boundary conditions both at the

center and edge of membrane, then the displacement profile 𝑧(𝑟) can be described in a parabola

shape [107]

𝑧(𝑟) = 𝑧0(1 −𝑟2

𝑅2) (1-6)

where 𝑧0 = 𝑃𝑅2/4𝜎 is the displacement at the center of membrane, thus the force acting at the

center is given as 𝐹 = 4𝜋𝜎𝑧.

Page 41: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

20

In our experiments, through doping extra charges into the membrane, the suspended flake can

be pulled down and the deflected displacement is well controlled by an electrostatic force that stems

from a DC gate voltage 𝑉𝑔. Thus, the whole suspended device forms a parallel-plate capacitor

structure and the corresponding electrostatic force is described as

𝑃(𝑧) =1

2

𝜕𝐶

𝜕𝑧𝑉𝑔2 =

𝜖0𝑉𝑔2

2(𝑑−𝑧(𝑟))2 (1-7)

where 𝐶 is the unit capacitance of the device, 𝜖0 is the vacuum dielectric constant, 𝑑 is the vacuum

gap between the suspended membrane and the bottom surface of trench. Typically for our

experiments, the maximal vertical displacement of membrane is of ~ 50 nm, which is much

smaller than the vacuum gap of ~ 600 nm. Thus, we can further simplify the electrostatic pressure

above to 1st order approximation as

𝑃 = 𝑃0 (1 −𝑧

𝑑)−2≅ 𝑃0 (1 + 2

𝑧

𝑑) (1-8)

where 𝑃0 =𝜖0𝑉𝑔

2

2𝑑2 is the displacement at the center of membrane. After plugging the electrostatic

pressure back to the equilibrium equation of net force as

𝜎∇2𝑧 + 𝑃0 (1 + 2𝑧

𝑑) = 0 (1-9)

We can reformulate the 2nd order differential equation in differential order of radius

variable as

𝑟2𝑧′′ + 𝑟𝑧′ + Λ2𝑟2(𝑧 +𝑑

2) = 0 (1-10)

where Λ = √2𝑃0

𝜎𝑑 and one can solve this zeroth order Bessel’s equation and obtain the general

solution by considering the boundary condition 𝑧(𝑅) = 0

𝑧(𝑟) =𝑑

2(𝐽0(Λ𝑟)

𝐽0(Λ𝑅)− 1) (1-11)

where 𝐽0 is the Bessel function with first kind. We note that the calculated profile of displacement

based on this electrostatic approximation is similar to that derived from parabola model but ‘bended’

Page 42: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

21

deeper at the center of membrane since the force will increase as the membrane pulled close to the

back-gate surface. Since Λ is determined by applied gate voltage 𝑉𝑔, this gate dependence provides

us a very convenient knob to tune and calibrate the static deflection of membrane together with the

interferometry technique introduced in section 2.2.4.

Similar analysis applies to dynamical description for the motion of a clamped membrane. We

start with replacing 𝑧 ⟶ 𝑧 + �� and 𝑃 ⟶ 𝑃 + �� , where the tilde notation represents the time-

dependent oscillation component. Assuming mechanical properties of the membrane are time-

independent, we can split the equation of motion Eqn. (1-5) into the static and dynamic parts

𝜎∇2𝑧 + 𝑃 = 0

𝜎∇2�� + �� = 𝜌𝜕2��

𝜕𝑡2 (1-12)

After setting �� = 0 for the convenience of discussion and using the conventional

separation of variables method, we can obtain the solution for dynamic displacement �� as

��(𝑟, 𝑡) = ∑ 𝐴𝑚𝐽0 (𝜁𝑚

𝑅𝑟) 𝑒𝑖Ω𝑚𝑡∞

𝑚=0 (1-13)

where the 𝐴𝑚 is the complex amplitude of m𝑡ℎ vibration mode and Ω𝑚 is the angular resonant

frequency of the m𝑡ℎ vibration mode as written blow

Ω𝑚 =𝜁𝑚

𝑅√𝜎

𝜌 (1-14)

where 휁𝑚 is the m𝑡ℎ positive root of Bessel function 𝐽0, and 휁0 ≅ 2.405 for the fundamental mode

(of our interest). To obtain the m𝑡ℎ complex amplitude 𝐴𝑚 of mechanical mode, we need to plug

the specific expression of external force into the equation of oscillation above. To determine the

resonance frequency of resonators, an electrostatic force with both DC and AC components is

applied to drive the suspended membrane into vibration. From previous discussion for electrostatic

pressure 𝑃 ≅𝜖0𝑉𝑔

2

2𝑑2(1 + 2

𝑧

𝑑), we can explicitly write down the AC component of the driving force

Page 43: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

22

by letting 𝑉𝑔⟶ 𝑉𝑔 + ��𝑔. Only keeping 1st order approximation term, we have the expression for

dynamical electrostatic pressure �� as

�� ≅𝜖0𝑉𝑔

2𝑑2��𝑔 (1-15)

To solve the dynamics part of motion, we can also rewrite the above �� in terms of Fourier

Bessel series to match the format of dynamical displacement as �� = ∑ 𝑃𝑚𝐽0 (𝜁𝑚

𝑅𝑟) 𝑒𝑖Ω𝑚𝑡∞

𝑚=0 .

Plugging the dynamical displacement �� to the dynamics equation of motion, now we are able to

solve for the complex amplitude as 𝐴𝑚 =𝑃𝑚

𝜌(Ω𝑚2 −Ω2)

, where the series coefficient for electrostatic

pressure is defined as

𝑃𝑚 = ∫ 𝑟��𝑅

0𝐽0(

𝜁𝑚

𝑅𝑟)𝑑𝑟

𝑅2

2(𝐽1(휁𝑚))

2⁄ =2��

𝜁𝑚𝐽1(𝜁𝑚) (1-16)

where the electrostatic pressure effectively projects itself onto different modes of motion scaled by

a factor of 2

𝜁𝑚𝐽1(𝜁𝑚). Thus, by adding a damping term ργ

∂z

∂t, the complex amplitude 𝐴𝑚 can be

modified as

𝐴𝑚 =𝐼𝑚

(Ω𝑚2 −Ω2+𝑖Ωγ)

(1-17)

where 𝐼𝑚 =2��

𝜁𝑚𝐽1(𝜁𝑚)𝜌=

𝜖0𝑉𝑔��𝑔

𝜌𝑑2𝜁𝑚𝐽1(𝜁𝑚) is directly related to both DC and AC voltages applied onto

the sample. Namely, the magnitude of membrane motion is a Lorentzian function in the frequency

domain and the maximal amplitude is determined by applied gate voltages (both DC and AC gates).

We emphasis here that the gate-tunability of complex amplitude 𝐴𝑚 allows us to estimate the

dynamical strain induced by vibration purely from mechanical resonance measurement. And the

resultant dynamical shift of exciton resonance opens up possibilities to dynamically control the

motion of suspended TMD monolayers through its exciton resonance as we will introduce next

section.

Page 44: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

23

1.3 Optomechanics

Optomechanics is an interdisciplinary field with utilizing laser light to control mechanical

objects, usually for resonators in meso- or nano-scale [108], [109]. In addition to trapping and

manipulating small particles, light-induced forces can also transduce and modulate

micromechanical motion [108], [109]. Specially, those NEMS devices with lightweight exhibit

noticeable advantages for optomechanical study and applications, and the resultant low stiffness in

turn endows this NEMS structure with higher sensitivity to be probed or controlled by feeble forces

induced by light. As consequence, although optomechanical effects have been utilized extensively

in macroscales include the gravitational detection in kilometer length and a gram-scale cavity

mirror [110], the expanding interests of optomechanics growingly stem from the progress of NEMS

technology.

Basically, dynamical control of mechanical motion by optomechanics depends on the use

of a ‘back-action’ force [108], [109]. Back-action can be intuitively understood as the perturbation

induced by a probe (usually laser beam) is no longer negligible for the sample and in turn it can

influence properties of the sample. And in our case, when optically measuring the motion of a

resonator, an extra damping was added by feedback from optical probe and thus modify the

oscillating states of the resonator. Numerous studies and application for these types of effect have

been reported extensively, which includes force sensing [80], non-volatile mechanical storage [111]

and photonic signal conversion [112], [113]. Nevertheless, an overwhelming majority of

optomechanics researches arise from the possible cooling down to the quantum ground state by

optomechanical coupling, where the average phonon occupation is less than one [109]. Meanwhile,

a system that resides around the ground state is a fundamental requirement for testing of quantum

theory and implementation of quantum computing [108], [109]. We will first demonstrate the

backaction dynamics by modeling a movable mirror driven by light-induced force with retardation,

Page 45: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

24

then discuss optomechanical backaction effects (cooling and self-oscillation) with derivation, and

our motivation for excitonic optomechanics using TMD monolayers.

1.3.1 Modeling optomechanical coupling

In this section, we will review the classical picture of optomechanical coupling by formulism.

Here, we start with a driven damped oscillator model and demonstrate how the light-induced force

that acts on oscillator with retardation can generates a backaction to modify the mechanical

oscillation. The dynamical motion of a damped harmonic oscillator driven by a light-induced force

is described as the following equation:

𝑚�� +𝑚𝛤𝑀�� + 𝑚Ω𝑀2 𝑧 = 𝐹𝑝ℎ(𝑧(𝑡)) (1-18)

where 𝑚 is the effective mass of the oscillator, Ω𝑀 is the mechanical resonance frequency, 𝛤𝑀 is

the intrinsic damping rate, and 𝐹𝑝ℎ is the driven force induced by light. And this driven force can

Figure 1-6: Schematic of the cavity optomechanical system. A mechanical oscillator comprises

one end of a Fabry-Perot cavity, coupling mechanical motion to the optical cavity length. Figure is

obtained from Ref.[108]

Page 46: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

25

be modeled by a delayed force that tends to reach its proper value in terms of position coordinate

𝑧 with some retardation 𝜏

𝐹𝑝ℎ(𝑧(𝑡)) = 𝐹(𝑧0) + ∫𝑑𝐹(𝑧(𝑡′))

𝑑𝑡′𝑇(𝑡 − 𝑡′)𝑑𝑡′

𝑡

0 (1-19)

where 𝑧0 is the mechanical equilibrium position, and the function 𝑇(𝑡) describes the time delay.

Thus, the equation of motion we need to solve is reformulated as:

𝑚�� +𝑚𝛤𝑀�� + 𝑚Ω𝑀2 𝑧 = 𝐹(𝑧0) + ∫

𝑑𝐹(𝑧(𝑡′))

𝑑𝑡′𝑇(𝑡 − 𝑡′)𝑑𝑡′

𝑡

0 (1-20)

We notice that the above equation can be solved easily in Fourier space by Laplace

transformation by adapting the Fourier transformation between frequency and time domains:

z(Ω) = ∫ 𝑧(𝑡)𝑒−𝑖Ω𝑡𝑑𝑡∞

0 (1-21)

The constant term 𝐹(𝑧0) has no time-dependence and just changes the equilibrium position

statically, which can be dropped out by shifting the displacement coordinate to this new equilibrium

position. Then, the equation of motion yields the following form

−𝑚Ω2𝑧(Ω) + 𝑖Ω𝑚𝛤𝑀𝑧(Ω) + 𝑚Ω𝑀2 𝑧(Ω) = ∫ 𝑒−𝑖Ω𝑡𝑑𝑡 ∫

𝑑𝐹(𝑧(𝑡′))

𝑑𝑡′𝑇(𝑡 − 𝑡′)𝑑𝑡′

𝑡

0

0 (1-22)

As 𝐹(𝑧(𝑡′)) implicitly depends on time and its derivative in terms of time can be rewritten as

𝑑𝐹(𝑧(𝑡′))

𝑑𝑡′=𝜕𝐹(𝑧(𝑡′))

𝜕𝑧

𝜕𝑧(𝑡′)

𝜕𝑡′ (1-23)

Under small amplitude of oscillation, 𝐹(𝑧(𝑡′)) is expanded around 𝑧(𝑡0) as: 𝐹(𝑧(𝑡′)) ≈

𝐹(𝑧(𝑡0)) + [𝑧(𝑡′) − 𝑧(𝑡0)]∇𝐹, where we adapt the abbreviation 𝜕𝐹(𝑧(𝑡′)) 𝜕𝑧⁄ |

𝑧=𝑧(𝑡0)= ∇𝐹. And

we can safely drop the subscript note with the small amplitude approximation as 𝜕𝐹(𝑧(𝑡′)) 𝜕𝑧⁄ =

∇𝐹. As consequence, the equation of motion is rewritten as

−𝑚Ω2𝑧(Ω) + 𝑖Ω𝑚𝛤𝑀𝑧(Ω) + 𝑚Ω𝑀2 𝑧(Ω) = ∫ 𝑒−𝑖Ω𝑡𝑑𝑡 ∫ ∇𝐹

𝜕𝑧(𝑡′)

𝜕𝑡′𝑇(𝑡 − 𝑡′)𝑑𝑡′

𝑡

0

0 (1-24)

With the helpful properties of the Laplace transform for convolutions as

𝐹1(Ω)𝐹2(Ω) = ∫ 𝑒−𝑖Ω𝑡𝑑𝑡 ∫ 𝐹1(𝑡′) 𝐹2(𝑡 − 𝑡

′)𝑑𝑡′𝑡

0

0 (1-25)

Page 47: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

26

where 𝐹1 and 𝐹2 are general functions used for discussion here. The equation of motion is

reformulated as

−𝑚Ω2𝑧(Ω) + 𝑚Ω𝑀2 𝑧(Ω) + 𝑖Ω𝑚𝛤𝑀𝑧(Ω) = 𝑖Ω∇𝐹𝑇(Ω)𝑧(Ω) (1-26)

And the light-induced force mostly stems from two components as radiation pressure and

photothermal effect. Both these two photo-induced processes react with an exponential behavior.

Thus, we can reasonably assume the profile of retardation function 𝑇 is of exponential type,

𝑇(𝑡) = 1 − 𝑒−𝑡/𝜏 (1-27)

Consequentially, the Fourier counterpart of the retardation function 𝑇(𝑡) is given by

𝑇(Ω) =1

𝑖Ω(1+𝑖Ω𝜏) (1-28)

After plugging the frequency-dependent retardation function back to the equation of

motion, we reformulated this oscillation equation in the power of Ω as

−𝑚Ω2𝑧(Ω) + 𝑖Ω𝑚𝛤eff𝑧(Ω) +𝑚Ωeff2 𝑧(Ω) = 0 (1-29)

with an effective resonance frequency (namely, spring constant) and an effective damping as [114]

Ωeff2 = Ω𝑀

2 (1 −1

(1+Ω2𝜏2)

∇𝐹

𝐾) ,

𝛤𝑒ff = 𝛤𝑀(1 + 𝑄𝑀Ω𝑀𝜏

(1+Ω2𝜏2)

∇𝐹

𝐾) (1-30)

where 𝐾 = 𝑚Ω𝑀2 is spring constant of the oscillator, and 𝑄𝑀 =

Ω𝑀

𝛤𝑀 is the mechanical quality factor.

From the expressions of effective resonance frequency and damping above, several interesting

phenomena can be explicitly concluded. The frequency dependence of the effective mechanical

quantities above indicates the propriate frequency window of oscillation required to observe the

pronounced changes induced by light. At very high mechanical frequency, the light-induced

contribution becomes negligible as the oscillator vibrates in the dark. In the contrast, as the

oscillator vibrates with small frequency, the light-induced terms are constants and no sufficient

damping change generates, especially for mechanical damping. With all other parameters fixed, the

Page 48: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

27

maximal optomechanical damping change occurs when Ω𝑀τ ~ 1. This highlights the importance

of τ to estimate the effectiveness of optomechanical modification. For a typical mechanical

resonance of our samples with Ω𝑀

2𝜋 ~ 40 MHz, the optimal time constant 𝜏 would then be 4 ns.

Second, the oscillator systems with low mass (namely, small rigidity) are more preferable to

observe such light-mechanics interplay. Most importantly, sitting at different slopes of light-

induced force (namely, gradient of force ∇𝐹 ) can give rise to completely opposite responses

corresponding to damping (red-detuned) or anti-damping (blue-detuned) of mechanical motion as

we will discuss in details below.

1.3.2 Optomechanical backaction

In optomechanical systems, the damping of mechanical resonator can be optically modified

by introducing a backaction-induced damping [108], [109], [115]. The physics behind this extra

damping can be understood by a simple Fabry-Perot cavity as depicted in Fig. 1-6 with

displacement-dependent resonance frequency 𝜔𝑐𝑎𝑣(𝑧): the displacement z increases as the cavity

elongates, and thus the cavity resonance frequency (of interest) decreases. And the light-induced

force is proportional to the profile of cavity resonance as a function of displacement. Both radiation

pressure and photothermal forces give rise to the same physics as we will illustrate below. Here,

let us start with the movable mirror residing at the rising slope of the resonance in displacement

coordinate, which sustains the optical mode frequency still higher than the laser frequency. Namely,

the incident beam is red-detuned with respect to the cavity resonance.

Page 49: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

28

Now we vibrate the movable mirror forwards and backwards in a small cycle. If the mirror

moves around with infinitely slow rate (adiabatically), the light-induced force can be ‘experienced’

by the moving mirror without doing additional work. Thus, no mechanical quantities will be

modified by the light-induced force under such condition. However, as the mirror moves with finite

sweep rate, the mirror ‘feels’ the light-induced force with time delay. Such retardation originates

from the finite decay time that incident photons or photothermal deformations take to build up

equilibrium. Thus, if the time delay is much shorter than one cycle of oscillation (namely, the

system is at equilibrium all the time), the light-induced force acts instantaneously on the mirror and

follows the profile of cavity resonance without any retardation. However, if the time delay is

comparable to one cycle of oscillation, the light-induced force that actually acts on the moving

Figure 1-7: Schematics of back-action effects in cavity optomechanics. (a) Schematic diagram of

light-induced force experienced by the end mirror as a function of displacement in Fabry-Perot

system. The finite decay rate of confined photons induces a delayed response to mirror motion. (b)

“P-V” style schematic depicting the work done by the radiation force during one cycle of

oscillation. The work is given by the enclosed area swept in the force-displacement diagram, which

is due to the retardation of the force (finite cavity decay rate). The work is negative or positive,

depending on whether one is on the red-detuned or blue-detuned side of resonance. This thus gives

rise to damping or amplification, respectively. Figure is obtained and modified from Ref.[109]

Page 50: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

29

mirror will be different from the adiabatic counterpart at the same location (see Fig 1-7(a)). In the

first half circle, as the mirror moves towards the resonance with non-adiabatic speed, the light-

induced force remains smaller than it would be in the adiabatic motion. Conversely, as the mirror

moves back towards the starting point, the light-induced force exceeds that in infinitely slow

motion. Overall, the work done by light-induced force (the net enclosed area in force-displacement

plot in Fig. 1-7(b)) is negative at red-detuned pump. Namely, the light-induced force extracts

mechanical energy from the mirror.

Similarly, if we put the movable mirror into the falling slope of the resonance in displacement

coordinate and play the same ‘game’, the net work done by light-induced force is positive and the

mechanical motion is amplified. Intuitively, the optomechanical backaction is like a ‘guide’ that

drives the energy flowing between mechanical motion and light. As the explicit consequence, the

linewidth of mechanical motion is broadened (narrowed) when the work done by light is positive

(negative). Namely, a broadened linewidth of mechanical modes means that the mechanical energy

now decays faster than before, and thus the effective temperature of this mechanical mode is

lowered. This process above is optomechanical ‘cooling’ of mechanical motion. Conversely,

optomechanical ‘amplification’ can be realized by introducing a feedback that effectively reduce

damping in a resonator. This can ultimately lead to self-oscillation of a resonator, in which the

optomechanical backaction reduces the resonator’s damping to zero and the resonator can sustain

its vibration only by thermal fluctuation. Owing to its rich and compelling potential for novel

phenomena in both fundamental science and application, numerous races had been setting off to

either cool a NEMS resonator to its quantum state [116]–[124] or pump a NEMS resonator into

self-oscillation [116], [118], [125]. Both photothermal force [116], [118] and radiation pressure

force[117], [119]–[125] had been exploited to demonstrate sufficient optomechanical cooling or

gain in NEMS systems.

Page 51: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

30

All these optomechanical systems above that require delicate cavity design and operation

[108], [109], [126], [127], including taper fiber access, coupling adjustment, and cavity stability,

impedes their potential application to cutting-edge opto-electronic devices, such as photonic

transistors and sensors. It would be particularly compelling to realize such optomechanical effects

in a mechanical resonator with a cavity-less structure. Hence, optical control of micro-resonators

with cavity-less structure is a promising alternative to achieve applicable integration for on-chip

micromechanical systems. Also, higher sensitivity of optical sensing can be reached in a micro-

resonator with low stiffness Thus, to realize the cavity-less optomechanics in mesoscopic

mechanics systems, an optical resonator with materials that possess both good mechanical

properties and strong light-matter coupling is highly demanded. Together with the excellent

mechanical strength and flexibility, the strong excitonic effects in monolayer TMD semiconductors

makes this type of materials as a favorable candidate to study and develop cavity-less

optomechanics using atomically thin materials. In Chapter 4, we will demonstrate our observation

of light-induced damping and anti-damping of mechanical modes and modification of mechanical

stiffness by optical excitation near exciton resonance in a monolayer MoSe2 resonator.

Page 52: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

Chapter 2

Experimental reviews

In this chapter, we will introduce experimental methods to fabricate 2D

monolayer/heterostructure devices, especially for suspended MEMS structure. We will also review

basic optical techniques that applied to characterize the optical/mechanical properties of atomically

thin crystals. Experimental detection of static displacement and dynamic resonance for monolayer

resonators will be discussed in detail individually.

2.1 Device fabrication

2.1.1 Sample preparation

Due to the nature of layered materials, external perturbation can easily break down the weakly-

bound van der Waals interactions among layers [128], [129]. Starting from this simple idea,

intensive efforts were employed to obtain atomically thin flakes, or ultimately monolayers till the

first successful isolation and identification of graphene by Andre Geim and Kostya Novoselov [4].

In this groundbreaking work, single-layer graphene was exfoliated using a surprisingly easy method:

repeatedly peeling flakes off bulk crystal by scotch tape [4]. Soon the magic of this ‘scotch tape’

method was applied to a wider range of layered materials, including the family of transition metal

dichalcogenides [6], [128], [129]. Nowadays, mechanical exfoliation still demonstrates its power

on efficient access to atomically thin crystals with high sample quality. And all the TMD

monolayers we studied in this dissertation are mechanically exfoliated from bulk crystals.

Page 53: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

32

Here we will briefly illustrate how to prepare atomically thin samples from scratch. First, a

bulk crystal was placed onto the adhesive face of scotch tape served as ‘parent’ tape. And we took

a second scotch tape with similar length and width and peeled a piece of thin flake from the bulk

crystal. This scotch tape with thin flake is ‘daughter’ tape and will be used for further exfoliation.

Then we repeated this peel-exfoliate process between two ‘daughter’ tapes for several times till

exfoliated thin flakes showed desired optical contrast. During this tape preparation, we avoided the

overlapping of thin flakes and tried to cover most of tape with thin-down flakes. Then tapes with

considerable thin flakes were gradually placed onto targeted substrates with some titled angle and

gently massaged with a cotton tip for several minutes. This step is aimed to avoid bubbles forming

between tape and substrate. One should also be careful about the mechanical strength for massage

process. Some external force is needed to form good contact between flakes and the substrate. This

will strengthen the van der Waals force between substrate to pull flakes down from bulk on the

scotch tape. Meanwhile, thin flakes can easily be broken into messy clusters under unnecessarily

strong shear force. After tape massage, we gentle lifted the tape with small tiled angle and left

noticeable thin flakes onto the targeted substrate for further search and identification. Note that one

may want to adjust the type of scotch tapes for different layered materials onto different targeted

substrates. To increase the yield when exfoliating thin flakes, we usually chose white ‘magic tape’

for silicon substrate with oxide layer coating atop whereas blue Nitto tape was used for sample

exfoliation onto flexible substrate include PDMS films. In practice, we noticed that the yield of

obtaining thin flakes significantly enhanced using silicon substrates with oxygen plasma treatment

beforehand. This is a key step to harvest high sample quality monolayers/few-layers with relatively

larger size, up to ~ 100 𝜇m.

To identify exfoliated monolayers, optical contrast varies for materials on different

substrates. Both graphene and TMDs show noticeable optical contrast on 280 nm oxide-coated

silicon substrate [130]–[132], whereas monolayer hBN is almost unidentifiable under such

Page 54: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

33

condition [133]. This is caused by interference effect between monolayer material and silicon oxide

layer. Silicon substrate with different oxide thickness can possibly be applied to search of extra thin

hBN flakes. Another commonly-used substrate for exfoliation is polymer PDMS film. The

‘transparency’ of PDMS challenged the search of monolayer graphene on it. Unlike graphene,

monolayer TMDs especially for TMD semiconductors show enough ‘visibility’ on PDMS film due

to strong light-matter interaction. Also, adhesion of PDMS film is controlled by temperature and

beneficial for exfoliating larger monolayer TMD crystals. Thus, one can directly identify and

transfer desired monolayer TMDs onto targeted chips all by one PDMS film [134]. This efficient

transfer method avoids extra lithography or solvent exposure and was mainly used for fabricating

purely monolayer and suspended monolayer devices, which will be discussed in section 2.1.3. In

the section 2.2, I will discuss in details about how to reliably identify monolayers with not only

optical contrast but also other optical spectroscopy methods.

2.1.2 Van der Waals assembly

In this section, we will illustrate a facile and efficient method to fabricate 2D material-based

FET with high sample quality. The basic idea of this fabrication method is as following: after

Figure 2-1: Optical images of monolayer TMDs exfoliated on (a) silicon substrate with 285 nm

oxide layer atop and (b) PDMS substrate doubly-capsulated with thin hBN flakes.

Page 55: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

34

identifying monolayers ideal for further study, we utilized a polymer stamp on a glass slide to pick

up and stack these monolayers in a desirable sequence. Once finished stacking layer by layer, the

sandwich-like van der Waals heterostructure was released onto the preprinted substrates and thus

the optoelectronic or optomechanical devices based on van der Waals materials were fabricated.

Thus, as well as probing the materials’ intrinsic properties, the heterostructure stacked by this ‘dry-

transfer’ method also allows us to design and engineer artificial material structure with favorable

sequence or stacking order for novel phenomena and application. This polymer-assisted technique

was first exploited in fabricating ultra-clean graphene field-effect devices [135]. Within the

invention of this intelligent fabrication method, a big band gap (namely insulator) layered materials

– hexagonal boron nitride (hBN) was introduced [135]. This graphene cousin later becomes an

important building block to fabricate 2D heterostructure devices with high quality.

Now we briefly review this ‘dry-transfer’ method firstly introduced in Ref.[136]. First, a piece

of sticky polymer film made from polypropylene carbonate (PPC) was put onto another soft

polymer polydimethylsiloxane (PDMS) with greater thickness to form a stamp. And this polymer

stamp can generate greater adhesion with exfoliated hBN flakes and peel them up from oxide-

coated silicon substrates by controlling the temperature of substrates. Due to the nature of van der

Waals materials, the picked hBN flake can form stronger bound with the subsequent atomically

thin flakes and thus easily pick up these flakes from silicon substrates, which is usually challenged

by directly using PPC film include graphene. Moreover, the hBN-encapsulated graphene was

protected from ‘dirty’ environment and demonstrated high-quality performance in transport

measurement due to atomically flatness and clean interface provided from hBN. In some sense, this

‘dry-transfer’ technique, or more professional terminology as ‘van der Waals assembly’, is a

cornerstone that paves paths towards numerous possibilities of observing interesting physics only

emerged with high sample quality.

Page 56: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

35

Inspired by Ref [137], we developed this van der Waals assembly with several modification

for our research interests. First, we replace the PPC film with another polymer with greater

stickiness and higher melting point, named as polycarbonate (PC). The explicit consequence is that

new stamp structure can directly pick up almost all the layered materials studied in our research

group, which greatly enrich the diversities of devices we can fabricate. Second, we add one more

PPC layer between PC film and PDMS substrate as a soft buffer layer, which prevents the falling

or stretching of PC film from PDMS. This PPC film can naturally form a curvature and thus allow

the smooth occurrence of layer stacking with finite tilted angles instead of putting two layers

together instantaneously that caused the formation of bubbles.

Figure 2-2: Process flow for heterostructure fabrication using PC as adhesive polymer. This

process flow can be used to fabricate heterostructure FET or NEMS device. The final

heterostructure maintains the stacking order. Step 1: Attach a PC/PPC/PDMS stamp on the glass

slide and place it on micro-manipulator of transfer stage; Step 2: First layer pick-up; Step 3: Unload

the bottom substrate and load the other exfoliated substrate; Step 4: Second layer pick-up, Step 5:

Place a new prepatterned substrate with electrodes for final transfer; Step 6: The new prepatterned

substrate makes a contact with heterostructure and PC stamp at 150 C; Step 7: PC and

heterostructure detached from handle substrate and transferred to the new prepatterned substrate.

Figure is obtained from Ref.[137].

Page 57: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

36

2.1.3 Suspend atomically thin membrane

In this dissertation, most of interesting phenomena originated from suspended TMD

monolayers. Conventional fabrication of a suspended 2D device inevitably exposes the material to

lithography solvents and aggressive acid liquids followed by a delicate critical point drying. Such

low-yielded process usually brings additional contamination and seriously degrades the sample

quality especially for TMD materials. Thus, a lithography-free method is undoubtedly critical to

fabricate a suspended TMD device that sustains high sample quality. Ideally, it would be highly

favorable to transfer the monolayers onto the prepattern trench directly from tapes or substrates that

used for exfoliation. In principle, the high flexibility and large breaking strength of TMD

monolayers make this one-step fabrication possible in practice, which can sufficiently limit extra

physical damage or chemical contamination. Consequentially, two different routines to acquire

suspended TMD devices were conducted in our research. Together with pre-deposited metal

electrodes in substrates, both two fabrication methods required the prepattern trenches etched by

reactive ion etching (RIE). The first strategy we tried to suspend TMD monolayers is to directly

exfoliate TMD monolayers onto silicon substrates prepatterned with dense trench arrays. As normal

exfoliation on silicon substrates, there would be some chance to suspend monolayers over a trench

during this exfoliation process. Obviously, this method can produce high-quality suspended TMD

samples, but its low yield greatly jeopardizes the efficiency of device fabrication. In practice, we

found that the polymer PDMS was a high-yielded exfoliation ‘tape’ to obtain TMD monolayers in

large size (usually > 30 μ𝑚). In addition, we come up with the second fabrication method that

monolayers were exfoliated and transferred onto the target trench all by one piece of PDMS. And

this PDMS deterministic transfer method [134] had been exploited to fabricate free-standing

devices. The working mechanism behind this technique depends on the viscoelasticity of PDMS.

Namely, PDMS behaves like an elastic solid within a short period while it moves like a liquid at

Page 58: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

37

much longer timescales [134]. The monolayer flakes were easily ‘peeled’ from its bulk due to the

intimate contact can be formed between PDMS and the flakes. By controlling the temperature

during this transfer, the contact between PDMS and flakes can be weakened. This can explicitly

lead to the detachment of flakes from PDMS, which adhere favorably to the silicon surface. Due to

the absence of capillary force within this PDMS-based transfer, the transfer of monolayer flakes

over a trench is straightforward with surprisingly high yield (> 90%). Note that partial cover of the

trench would likely give rise to tear or damage of flakes. By peeling off very slowly, the suspended

region of monolayers can completely sustain without noticeable wrinkles or mechanical cracks. All

the suspended monolayer samples were fabricated through this PDMS deterministic transfer

technique.

2.2 Optical spectroscopy of monolayer semiconductors

Due to the intrinsically strong excitonic effects, optical spectroscopy turns out to be an

efficient and powerful experimental technique to unveil the properties of monolayer TMD

Figure 2-3: PDMS deterministic dry transfer method for 2D NEMS.

Page 59: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

38

semiconductors. Meanwhile, the interesting light-valley interactions in such atomically thin

materials endows versatile knobs (e.g. polarization of light) to resolve and control the internal DOF

(e.g. spin and valley) of these monolayers with optical excitation. In this section, we will mainly

review the photoluminescence spectroscopy and optical reflection spectroscopy as two routine

spectroscopy methods employed in the studies within this dissertation. In addition, the basic

principle of magnetic circular dichroism spectroscopy will be briefly introduced. And we also will

discuss how to transduce the mechanical performance of suspended membranes into optical signals

using interferometry techniques.

2.2.1 Photoluminescence spectroscopy

Photoluminescence (PL) is a light emission process that occurs after absorbing photons in

materials. In general, electrons within a material can be excited to permissible excited states by

gaining excess energy from incident photons and the vacuums (or holes) are correspondingly left

by the excited electrons in valence band. Then the excited electrons will rapidly relax to the

conduction band edge and stay there for a certain time duration (~ femtosecond to picosecond). It

will be energetically favored for the excited electrons relaxing back to the valence bands. As

consequence, the recombined electron-hole pairs can emit photons (PL) that are of lower energy

than those are absorbed. This is a radiative recombination process, which, at the same time,

competes with other possible non-radiative recombination process include Auger process and

defect trapping. Note that electronic transition that occurs at indirect band gaps needs the assistance

of phonons due to the conservation of momentum, which can significantly reduce the quantum

efficiency of PL. Thus, the PL features in monolayer TMD semiconductors are of order of

magnitudes higher than those in bilayer or few-layered samples, which provides a fast and reliable

way to determine monolayer TMD flakes.

Page 60: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

39

The experimental setup of PL measurements is shown in Figure 2-4. A monochrome

excitation beam was guided through a beam splitter and focally projected onto the monolayer

samples with ~ 1𝜇m2 spot size using an optical objective. The emitted PL signals was collected

with the same objective and reflected by the beam splitter for spectrum detection. The

corresponding PLD spectrum was measured by a dispersive spectrometer with liquid nitrogen-

chilled charge-coupled device (CCD). Usually, to optimize the signal-to-noise ratio (SNR) in PL

measurements, a long pass filter was employed to filter out the excitation photons for delicate

spectrum features and a spatial filter (pin hole) was used to eliminate the scattered light from

substrates and only allow the signals from sample pass through.

2.2.2 Optical reflection spectroscopy

Optical absorption spectroscopy can unveil rich physical properties in materials include band

gap, excitonic resonance and spin texture of band structure etc. The ideal way of obtaining

Figure 2-4: Schematics of microscopic confocal system in reflection geometry.

Page 61: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

40

absorption of a material is to perform optical reflection and transmission measurements to make

use of the following relation as

A(𝜔) = 1 − R(𝜔) − T(𝜔) (2-1)

As a complimentary technique to PL spectroscopy, optical absorption spectroscopy is

employed to directly probe the band-to-band transition of materials from photon absorption in

contrast to detection of light emission in PL spectroscopy. Meanwhile, the oscillation strength of

electronic band-to-band transition is much stronger than the counterpart of band-to-defect transition.

As consequence, optical absorption exhibits clear features of transition between band edges with

least influence from defect states. In thin film limit, the sample absorbance can be directly

calculated from the reflection/transmission contrast [56], [138]:

A(𝜔) =𝑛𝑠2−1

4

𝑅−𝑅0

𝑅0=𝑛𝑠+1

2

𝑇−𝑇0

𝑇0 (2-2)

which indicates the more evident contrast from reflection measurements than that in transmission.

However, one cannot directly extract the sample absorbance from its reflection contrast spectrum

on SiO2/Si substrate due to the distortion from multilayer reflection. Instead, we practically obtain

the sample absorbance from Krames-Kronig constrained variable fitting as the case shown below.

For suspended atomically thin TMD materials studied in this dissertation, we can obtain

the corresponding optical reflectance 𝑅/𝑅0 which is characterizing by the ratio of the reflected

power from the suspended sample 𝑅 to that from the bare trench 𝑅0 under normal incidence. These

two quantities can be expressed through the parameters of the TMD monolayers/vacuum/Si

multilayer structure and the optical properties of its constitutes as [50], [139]

𝑅(𝜔) = |−𝐶−𝐷[1−��(𝜔)]

𝐶+𝐷[1+��(𝜔)]|2

, 𝑅0(𝜔) = |−𝐶−𝐷

𝐶+𝐷|2. (2-3)

Here we have assumed the incident power to be unit. The coefficients C and D are related to the

refractive index 𝑛𝜁 , thickness 𝑑𝜁 , and phase shift 𝛿𝜁 = 𝑛𝜁𝜔𝑑𝜁/𝑐 in the vacuum (휁 = 𝑣) and Si

(휁 = 𝑆𝑖) layers [50]:

Page 62: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

41

(𝐶𝐷) = (

𝑠𝑖𝑛𝛿𝑣 𝑖𝑛𝑆𝑖𝑐𝑜𝑠𝛿𝑣𝑖𝑐𝑜𝑠𝛿𝑣 𝑛𝑆𝑖𝑠𝑖𝑛𝛿𝑣

) (𝑠𝑖𝑛𝛿𝑆𝑖 + 𝑖𝑛𝑆𝑖𝑐𝑜𝑠𝛿𝑆𝑖𝑖𝑐𝑜𝑠𝛿𝑆𝑖 + 𝑛𝑆𝑖𝑠𝑖𝑛𝛿𝑆𝑖

). (2-4)

And the complex absorbance ��(𝜔) of monolayer TMD semiconductor is modeled by a series of

Lorentzian arisen from different exciton resonances [50](A, B exciton resonance or charged exciton)

��(𝜔) = ∑𝐴𝑗

1−𝑖ℏ(𝜔−𝜔𝑗)/𝛾𝑗𝑗 . (2-5)

Here 𝐴𝑗 is the area of the jth Lorentzian, 𝜔𝑗 is the jth resonance frequency, and 𝛾𝑗 is the jth total

exciton linewidth.

The setup of optical reflection spectroscopy is similar to the one introduced in the PL

measurement. Here we use a broadband beam (supercontinuum laser or warm color lamp) as the

excitation source instead of a monochromic laser. And the dispersive CCD spectrometer we used

is reliable to measure the reflection spectrum from 400 to 850 nm with 0.1nm resolution.

2.2.3 Magnetic circular dichroism (MCD) spectroscopy

Magnetic circular dichroism (MCD) stems from the different absorption of materials between

left- and right-handed circularly polarized light at the presence of a magnetic field perpendicular to

the sample surface. The intensity of MCD signals is defined as ρ𝑀𝐶𝐷 =𝐴𝐿−𝐴𝑅

𝐴𝐿+𝐴𝑅, where 𝐴𝐿(𝑅) denotes

the sample’s absorbance of left (right) handed circularly polarized light with L(R).

To enhance the signal to noise ratio (SNR), the optical excitation was modulated

between left and right circular polarization by a photo-elastic modulator at certain modulating

frequency. The reflected beam was collected by the same objective and detected by a photodiode.

The MCD was determined as the ratio of the AC component at modulating frequency (measured

by a lock-in amplifier) and the DC component (measured by a multimeter) of the reflected light

intensity.

Page 63: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

42

2.2.4 Characterization of mechanics in 2D NEMS

The optical spectroscopy methods illustrated above are employed to characterize the

physical properties of a material by measuring the corresponding spectrum. For the suspended

TMD monolayers, we not only care about the intrinsic properties of the atomically thin materials

but also aim to investigate the mechanical motions of these free-standing membranes. Hence, the

interferometry technique was introduced to calibrate and measure both static displacement and

dynamical vibration of suspended monolayers [65]. In 2D monolayer-based NEMS device, both

suspended membrane and back trench surface can reflect the incident laser beam and the two

reflected beams can form interference pattern with each other. As the membrane position change,

the total intensity from two reflected beams also change due to the modulated interference condition.

Hence, the mechanical motion of suspended membranes can be transduced to optical response

through this interferometry technique.

The suspended TMD monolayers can be charged and actuated vertically by applying a gate

voltage due to capacitive nature of the resonator. Meanwhile, the reflectivity R of suspended

monolayers is dependent on the vertical displacement Δz towards the silicon back surface. Thus,

one can extract the vertical displacement of membrane by measuring the gate-dependent reflectance.

Modeling the gate dependence of 𝑅/𝑅0 at a given wavelength can unveil the gate dependence of

Δz, the separation of the center of the suspended membrane from the underneath silicon substrate.

By balancing the electrostatic force 𝐹𝑔 ≈𝜀0𝜋𝑅

2𝑉𝑔2

2𝑑𝑣2 due to gating with the force due to tension, i.e.

4𝜋𝜎Δ𝑧 − 𝐹𝑔 = 0, (2-6)

the vertical displacement Δz of the center of the membrane as a function of gate voltage 𝑉𝑔 can be

calculated. Here 휀0 is the vacuum permittivity, 𝑅 is the radius of the suspended membrane, and

𝜎 = 𝜎0 + 𝜎′is the total tension in the membrane with 𝜎0 and 𝜎′ being the initial built-in tension and

Page 64: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

43

the gate-induced tension, respectively. Furthermore, 𝜎′ can be written in terms of the material’s 2D

Young’s modulus 𝐸𝑌, the Poisson ratio 𝜈, membrane thickness 𝑡 and the strain 𝜖 as 𝜎′ =𝐸𝑌𝑡

1−𝜈2𝜖.

The strain is related to the vertical displacement Δz as 𝜖 =2

3(Δ𝑧

𝑅)2

. The corresponding gate

dependence of the vertical displacement Δz and the strain 𝜖 is numerically shown in Chapter 4.

Furthermore, the exciton peak position as a function of the gate voltage 𝑉𝑔 due to strain can also be

fitted by the model by using a peak shift rate of 0.1 eV/% strain (see Figure 4-2(a)), which agrees

very well with reported values [60].

Thus, we monitor mechanical resonance of our 2D NEMS devices by periodically driving the

resonator capacitively and measure the corresponding reflectance response in terms of driving

frequency. The driven vibration of monolayer membrane reaches the maximum as the AC driving

frequency resides at the natural frequency of the resonator. This on-resonance driven vibration

Figure 2-5: Interferometry technique to characterize mechanics of 2D NEMS. (a) Schematic of

reflected beams from both suspended membrane and silicon surface. (b) The normalized reflectance

as a function of displacement of membrane. The blue arrows indicate the transduction between

displacement of membrane and reflectance. The oscillation of membrane can periodically modulate

the sample reflectance (brown dash lines). Inset: gate dependence of normalized reflectance of

suspended sample with 450 nm probe wavelength. The red solid line is prediction with Eqns. (2-3)

to (2-6).

Page 65: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

44

leads to a Lorentzian-shaped reflectivity change of the suspended MoSe2 at natural frequency. To

make the optical response of reflectivity change more evident, a DC gate voltage is needed to break

the out-of-plane mirror symmetry of mechanical motion. This is also consistent with the derivation

of the complex amplitude in mechanical vibration shown in Eqn. (1-17).

Page 66: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

Chapter 3

Valley selective exciton bistability in suspended monolayer WSe2

During the past few decades people have witnessed at least two major innovations in

science which have had substantial impact on technology as well as science itself, even our daily

lives. We refer, of course, these two to be transistor and laser. Transistor is a semiconductor device

that used to amplify or switch electrical signals by employing electrons in an effective two-

dimensional region that residing at the interface between the silicon surface and its oxide layer.

Some charges, or current flow, can inevitably convert to heat dissipation during the operation. As

transistors are getting smaller and denser in logic circuit nowadays, the accumulated heat within

nanoscale can cause lots of troubles and degrade the device performance.

Alternatively, a transistor based on photons [140], [141] is proposed to solve this problem.

In principle, it can realize the switching times at the speed of light, and lower the power

consumption during the signal transmission and has promising potential to greatly enhance the

efficiency in current optical communication systems [142]. The most essential building block could

be implemented optical transistor is the concept of optical bistability [52]. In this chapter, we will

demonstrate the realization of optical bistability with strong excitonic effect in monolayer materials,

as well as the repeatable switching of light not only by light intensity also purely by its helicity

using monolayer TMDs semiconductors.

This chapter is based on published work in which I am the first author [50]. This chapter

is reprinted with permission from H. Xie, S. Jiang, J. Shan, K. F. Mak, Nano. Lett. 2018 (5), 3213-

3220 (https://pubs.acs.org/doi/abs/10.1021/acs.nanolett.8b00987).

Page 67: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

46

3.1 Optical bistability

Optical bistability, the phenomenon of two well-discriminated stable states depending

upon the history of the optical input, is the key concept for all-optical transistors, switches, logical

gates, and memory [52]. Namely, we can control the light with the intensity of light through the

optical transistor device. So far, optical bistability had been realized in photonic crystals or optical

cavity systems with nonlinear medium, which requires the delicate device configuration and well-

selective nonlinear materials [52]. Monolayer TMD semiconductors, with extremely strong

excitonic effect [44], [139], [143]–[147], provide an ideal playground to study the optical

nonlinearities stem from light-matter interaction. The emerging valley degree of freedom (DOF)

[32] and its coupling to light helicity [35]–[37], [148], [149] and magnetic field [38]–[43], gives us

extra knobs to play with such nonlinearity.

Figure 3-1: Schematic of optical bistability. 𝐼𝑓 and 𝐼𝑏 are the critical intensities for transmittance

bistability under forward and backward intensity sweeps, respectively. Figure is obtained from

Ref.[150].

Page 68: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

47

3.2 Experimental setup and basic characterization of suspended WSe2 monolayer

Suspended monolayer WSe2 was fabricated by transferring monolayer WSe2 flakes onto

trenched SiO2/Si substrates. Figure 3-2(b) shows the optical image and schematic side view of a

representative monolayer sample suspended over a circular trench of 8 μm in diameter and 600 nm

in depth. Suspended WSe2 samples were measured in an optical cryostat as shown in Figure 3-2(a).

Light was coupled into and out of it using a high numerical aperture microscope objective under

normal incidence. The beam size on the sample is typically about 1 μm. For reflectance spectrum

measurements, broadband radiation from a super-continuum light source and a spectrometer

equipped with a liquid nitrogen cooled charge coupled device (CCD) were used. The power on the

sample was limited to < 1 μW to avoid any nonlinear effects. For the cw excitation measurement,

a tunable cw Ti-sapphire laser and a Si avalanche photodiode were employed. The laser power on

the sample was controlled by rotating a half waveplate sandwiched between two linear polarizers.

It was limited to 250 μW to avoid sample damage. In the pump-probe measurement, the reflected

pump beam was filtered by a crossed linear polarizer. In the helicity-resolved measurement, a liquid

crystal phase plate was used to change the incident light polarization from linear to circular. For

the MCD measurement and the reflectance switching experiment, the light helicity was modulated

by electrically modulating the liquid crystal phase plate at several Hz’s and the reflected light was

detected by an avalanche photodiode. The MCD signal was recorded with a lock-in amplifier and

the time trace of the reflected light was recorded with an oscilloscope.

Page 69: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

48

Figure 3-3(a) is the reflectance spectrum of a suspended WSe2 monolayer at temperatures

ranging from 4 K to 330 K. The reflectance 𝑅/𝑅0 at any given wavelength was determined as the

ratio of the reflected power from the suspended sample 𝑅 to the reflected power from the bare

trench 𝑅0 under normal incidence. It exhibits a sharp exciton resonance with over 80% reflectance

contrast at low temperatures (Figure 3-3(a)). The spectra can be simulated by considering light

propagation in the WSe2/vacuum/Si multilayer structure with the WSe2 optical response described

by a single Lorenzian that arises from the exciton resonance (Figure 3-3(b)). The small discrepancy

around 720 nm at 4 K is due to the contribution of the charged exciton to the optical response of

monolayer WSe2 [149], [151], [152], which has been neglected in the simulation for simplicity.

The exciton peak wavelength and linewidth employed in the simulation are shown in Figure 3-3(c)

as a function of temperature. In the low temperature limit, the exciton is peaked at 709 nm (1.75

eV) with a linewidth of ~ 7 meV. With increasing temperature, the exciton resonance redshifts and

broadens. This behavior is well described by the exciton-phonon interaction (solid lines) with an

Figure 3-2: Experimental setup and suspended WSe2 device. (a) Sketch of experimental setup. The

suspended samples are placed in an optical cryostat. Optical excitation, either broadband radiation

or monochromatic beam, is illuminated onto sample. For single-colored source, both the incident

and reflected laser light are monitored by a pair of photodiodes. The reflected broadband spectrum

from the sample is recorded by a dispersive spectrometer. (b) Schematic side view of a suspended

monolayer WSe2 sample over a drumhead trench of 8 μm in diameter and 600 nm in depth on a Si

substrate. The substrate has a 600-nm oxide layer and pre-patterned Ti/Au electrodes. The inset is

an optical image of a sample illuminated by white light. Gold bars are electrodes and the dark ring

is the edge of the trench.

Page 70: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

49

average phonon energy of ~ 16 meV and LO-phonon energy of ~ 30 meV [153], [154]. These

values are in good agreement with the values reported for WSe2 [153].

Figure 3-3: Temperature dependence of exciton resonance in suspended WSe2 monolayer. (a)

Reflectance spectrum measured at 4 - 330 K. A sharp exciton resonance is observed at low

temperatures. The resonance redshifts and broadens with increasing temperature. (b) Simulated

reflectance spectrum at 4 - 330 K. (c) Exciton peak wavelength and linewidth (symbols) as a

function of temperature used in the simulation of (b). Solid lines are fits to Eqns. (3-2) and (3-3).

Page 71: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

50

3.3 Observation of optical bistability in suspended monolayer WSe2

In this section, we present the systematical study of the optical bistability through both

monochromatic and broadband probe under optical pump near exciton resonance of suspended

WSe2 monolayers.

3.3.1 Power dependence of monochromatic reflectance spectra across exciton resonance

We investigate the optical nonlinearity in monolayer WSe2 near the exciton resonance at 4

K. Figure 3-4 shows the power dependence of the reflectance for cw excitation tuned through the

exciton resonance in sample #1. The beam size on the sample is typically about 1 μm. The results

are similar up to ~ 100 K, for which thermal broadening is insignificant. They are also reproducible

in all samples studied here except small variations in the exciton resonance wavelengths and

linewidths. Results at selected wavelengths, marked as vertical dashed lines on the experimental

reflectance spectrum (Fig. 3-4(a)) and on the corresponding simulated absorbance spectrum (Fig.

3-4(b)), are shown in Fig. 3-4(c). At 708 nm (blue-detuned from the exciton resonance), the

reflectance increases with increasing power. Onset of nonlinearity occurs at a power level as low

as 1 μW. At all other wavelengths (red-detuned from the exciton resonance), the reflectance

exhibits jumps at critical powers and a hysteretic behavior with power sweeps. These are signatures

of optical bistability. Moreover, both the critical powers and the size of the bistability region

increase for excitation at longer wavelengths. We note that several μW is sufficient to generate

optical bistability in suspended monolayers here, in contrast to 10 mW in monolayer TMDs

encapsulated in hBN under similar experimental conditions [48].

Page 72: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

51

Figure 3-4: Optical bistability in suspended monolayer WSe2. (a) Reflectance spectrum of sample

#1 at 4 K. (b) Absorbance spectrum described by a single Lorentzian that best fits the reflectance

spectrum of (a). (c, d) Reflectance at representative wavelengths (708, 712, 720, 725 and 728 nm)

under forward and backward power sweeps from experiment (c) and simulation (d). The

wavelengths are marked by vertical dashed lines in (a, b).

70 0 71 0 72 0 73 00.0

0.6

1.2

0.0

0.6

1.2

0.0

0.6

1.2

0.0

0.6

1.2

0.0

0.6

1.2

0 100 200 300 4000.0

0.6

1.2

Re

flec

tance

Wavelength (nm)

(a)

(c)

708 nm

Reflecta

nce

712 nm

720 nm

725 nm

Power (mW)

728 nm

exp

70 0 71 0 72 0 73 00.0

0.5

1.0

0.0

0.6

1.2

0.0

0.6

1.2

0.0

0.6

1.2

0.0

0.6

1.2

0 100 200 300 4000.0

0.6

1.2

Wavelength (nm)

Abs

orb

ance

(b)

(d)

708 nm

712 nm

Reflecta

nce

720 nm

725 nm

Power (mW)

728 nm

sim

Page 73: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

52

3.3.2 Wavelength dependence of broadband reflectance spectra across exciton resonance

To understand the origin of the observed optical bistability, we monitor the reflectance

spectrum with a low-power super-continuum probe of suspended samples that are pumped by cw

excitation at a fixed power of 100 μW. Figure 3-5(a) and (b) are the contour plots of the reflectance

as a function of the pump (left axis) and probe (bottom axis) wavelengths of sample #2. Figure 3-

5(a) corresponds to a forward sweep of the pump wavelength (i.e. increasing wavelength), and Fig.

3-5(b), a backward sweep (i.e. decreasing wavelength). The bright features along the diagonal

dashed lines are the residual of the pump intensity. For the forward sweep, a clear exciton resonance

is initially observed at 715 nm with a width of 10 meV. As the pump wavelength approaches the

exciton resonance, a significant redshift of the exciton by over 20 nm accompanied with a

broadening by ~ 100 meV is observed. A sudden “recovery” of the exciton resonance follows when

the pump wavelength increases beyond a critical value of ~ 745 nm. In contrast, for the backward

sweep of the pump wavelength, the exciton resonance suddenly redshifts to ~ 738 nm at a lower

critical pump wavelength of ~ 735 nm. The exciton gradually returns to the original position as the

pump wavelength further decreases. We summarize the pump wavelength dependence of the

exciton peak and linewidth in Fig. 3-5(c) and (d), respectively. These values were extracted from

the analysis of the reflectance spectra (Methods). Clear hysteresis in both the peak wavelength and

the linewidth can be observed. Figure 3-5(a) and (b) also show that reflectance at probe wavelengths

off the exciton resonance (or far above the resonance at 450 nm, not shown) remains independent

of pump. Since the reflectance from the multilayer structure is very sensitive to the vacuum cavity

length due to optical interference, our result indicates a negligible pump-induced vertical movement

of the suspended membrane.

Page 74: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

53

The significant pump-induced exciton redshift and broadening near the bistability region

suggests the importance of a photo-thermal mechanism. We estimate that a temperature rise of ~

300 K is required to explain the observed exciton resonance shift under near-resonance (740 nm)

excitation at 100 μW. Such a large temperature rise, as we discuss below, is plausible due to the

low thermal conductance of suspended samples. As the sample temperature rises under optical

pumping in the photo-thermal effect, the exciton resonance “runs” away from the pump wavelength

for a blue-detuned pump, which in turn decreases the sample absorbance at the pump wavelength

and reduces the temperature rise. No cumulative effect can build up. In contrast, the exciton

resonance “runs” towards the pump wavelength for a red-detuned pump, which increases the

sample absorbance at the pump wavelength and, in turn, causes a further increase in the sample

temperature and more redshift towards the pump wavelength. This process is cumulative, i.e. a

positive feedback, and can cause the frequency of exciton resonance overshooting that of optical

Figure 3-5: Bistable excitonic reflection spectra under monochromatic pump. (a, b) Contour plots

of reflectance of sample #2 at 4 K as a function of probe wavelength (bottom axis) and pump

wavelength (left axis). The pump wavelength is swept forward (i.e. increases) in (a) and backward

(i.e. decreases) in (b). The pump power is fixed at 100 μW. The diagonal dashed lines are the

residual pumps. (c, d) The extracted exciton peak wavelength (c) and linewidth (d) from the

experimental reflectance spectra. (e, f) Simulation of the parameters shown in (c, d) based on the

photo-thermal effect as described in the main text.

720

740

760

0.2150

1.600

(a)

700 720 740 760

720

740

760

(b)

Pu

mp

wa

vele

ng

th (

nm

)

Probe wavelength (nm)

720

740

760

700 720 740 7600

50

100

Re

so

nan

ce

wa

ve

len

gth

(nm

)

exp

(c)

(d)

Lin

ew

idth

(m

eV

)

Pump wavelength (nm)

720

740

760

700 720 740 7600

50

100

Re

so

nan

ce

wa

ve

len

gth

(nm

) (e)

(f)

Lin

ew

idth

(m

eV

)

Pump wavelength (nm)

sim

Page 75: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

54

pump. This internal passive feedback, together with optical nonlinearity, leads to optical bistability

in monolayer WSe2 (Fig. 3-6(a)). For a red-detuned pump under a forward power sweep, the system

becomes unstable and switches suddenly from a low-temperature state to a high-temperature state

due to the rapid shift of the exciton resonance, especially when the pump power 𝑃 exceeds a critical

value 𝑃𝑓. Under a backward power sweep, the system starts from the high-temperature state, and

can remain there even below 𝑃𝑓. Only when the pump power decreases below a much lower critical

value 𝑃𝑏 , optical pumping can no longer keep the system in the high-temperature state, and it

switches suddenly to the low-temperature state. This is mainly due to the accumulated heat in

suspended WSe2 monolayers that own low thermal conductivity. As consequence, a hysteretic

bistable behavior occurs.

3.4 Modelling mechanism of the observed optical bistability

We would like to emphasis again that the occurrence of optical bistability mainly

attributing to the nonlinear optical response and internal positive feedback from external excitation.

The sharp exciton resonance and its passive feedback (redshift under illumination) endow WSe2

monolayers with feasibility to realize optical bistability near exciton resonance. Optical pump near

exciton resonance can redshift exciton resonance through either relaxing the existed strain or

heating up the sample. In Section 3.3.2, significant broadening of exciton line width under pump

near exciton resonance implies the overwhelming role of photothermal effect from laser light. This

is also verified by the comparison between temperature dependence of exciton resonance in Figure

3-3(c) and the pump-induced exciton redshift in Figure 3-5(c). In this section, we build up a

photothermal model to simulate the temperature profile of WSe2 membrane that optically pumped

near exciton resonance and predict the corresponding optical bistability.

Page 76: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

55

3.4.1 Simulation of photothermal model

Now we turn to a more quantitative discussion of optical bistability from the photo-thermal

effect. The exciton reflection contrast of monolayer WSe2 can be modeled by Eqn. (S1) and (S2)

by a Lorentzian for the exciton resonance

��(𝜔) =𝐴0

1−2𝑖ℏ(𝜔−𝜔0)/𝛾. (3-1)

Here 𝐴0 is the area of the Lorentzian and 𝛾 is the total exciton linewidth. As shown in Fig. 3-3(b),

the reflectance spectrum at 4 K can be fitted very well by choosing 𝐴0 ≈ 0.85, ℏ𝜔0 ≈ 1.748 eV

and 𝛾 ≈ 7 meV.The reflection contrast at higher temperatures can also be simulated by taking into

account the temperature dependent exciton peak position 𝜔0(𝑇) and linewidth 𝛾(𝑇) [153], [154],

𝜔0(𝑇) = 𝜔0(0) − 𝑆⟨𝜔𝐷⟩[coth (⟨ℏ𝜔𝐷⟩

2𝑘𝐵𝑇) − 1], (3-2)

𝛾(𝑇) ≈ 𝛾(0) +𝛾0

𝑒𝑥𝑝(ℏ𝜔𝐷/𝑘𝐵𝑇)−1. (3-3)

Here ℏ𝜔0(0) ≈ 1.75 eV is the exciton resonance at zero temperature, ⟨ℏ𝜔𝐷⟩ ≈ 16 meV and

ℏ𝜔𝐷 ≈ 30 meV are the average phonon energy and LO-phonon energy of WSe2, respectively, 𝑆 ≈

2.25 is the coupling parameter, 𝑘𝐵 is the Boltzmann constant,𝛾(0) ≈ 7 meV is the linewidth at

zero temperature, and 𝛾0 ≈ 76 meV is the characteristic linewidth corresponding to exciton-

phonon scattering [153]. Eqns. (3-2) and (3-3) describe the experimental results in Fig. 3-3(c) very

well. In Fig. 3-3(b) we show the simulated exciton reflection contrast spectra at varying

temperatures. The results agree well with the experimental data. Now we demonstrate a thermal

mechanism to show how the incident laser induces change in optical response of suspended 2D

semiconductor. The temperature profile 𝑇(𝑟) of the membrane induced by central laser spot when

using the heat diffusion equation is as

𝜕𝑇

𝜕𝑡=

𝜅

𝜌𝑐𝑝∇2𝑇 (3-4)

Page 77: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

56

where 𝜅 is the sheet thermal conductance of the sample, 𝜌 is the area mass density of the sample,

and 𝑐𝑝 is the specific heat capacity. In steady state, the above heat diffusion equation just

transforms as Laplace’s equation as

0 = ∇2𝑇 =1

𝑟

𝜕𝑇

𝜕𝑟+𝜕2𝑇

𝜕𝑟2 (3-5)

Note that the boundary condition for the above temperature profile is the temperature at the edge

of membrane 𝑇(𝑅) = 𝑇0 = 300 K. Another boundary condition comes from the heat flow 𝑞 across

the edge of laser spot 𝑟𝑠 as 𝜕𝑇

𝜕𝑟|𝑟=𝑟𝑠

= −𝑞

𝜅, where 𝑞 is in unit of W/m. And the equilibrium condition

of membrane can only be satisfied when the heat flow into the membrane is equal to the absorbed

power of incident laser as

𝐴𝐿𝑃 = 2π𝑟𝑠𝑞 (3-6)

where 𝐴 is the real part of complex absorbance �� in the suspended WSe2, 𝐿 is the local field factor

and with the laser power 𝑃. And solving the steady thermal equation with the boundary conditions

gives rise to the temperature profiled 𝑇′(𝑟) as

𝑇′(𝑟) = 𝑇0 +𝐴𝐿𝑃

2𝜋𝜅ln(

𝑅

𝑟) (3-7)

Thus, we can express the average sample temperature 𝑇 under optical excitation at wavelength 𝜆

as

𝑇(𝑃, 𝜆) =2𝜋

𝜋(𝑅2−𝑟𝑠2)∫ 𝑇′(𝑟)𝑟𝑑𝑟𝑅

𝑟𝑠= 𝑇0 +

𝐴𝐿𝑃

4𝜋𝜅(1 −

2𝑟𝑠2 ln

𝑅

𝑟𝑠

𝑅2−𝑟𝑠2 ) ≅ 𝑇0 +

𝐴(𝑇,𝜆)𝐿(𝑇,𝜆)𝑃

4𝜋𝜅(𝑇) (3-8)

where the average temperature of membrane is a good approximation with small spot size applied

to our case (𝑟𝑠 ~0.5 𝜇𝑚). Optical bistability requires the existence of more than one solution to 𝑇

for a given 𝑃 (Fig. 3-6(b)). Therefore, in addition to normal regions of monotonically increasing 𝑇

with 𝑃, there must exist a region of decreasing 𝑇 with 𝑃, i.e. 𝑑𝑇

𝑑𝑃≤ 0. Carrying out the derivatives

gives us the condition

Page 78: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

57

𝑑𝐴

𝑑𝑇

𝑇−𝑇0

𝐴+𝑑𝐿

𝑑𝑇

𝑇−𝑇0

𝐿≥ 1 . (3-9)

Meanwhile, the first term in Eqn. (3-9) is significantly bigger than the second term because of the

weak temperature dependence of the local field factor. In fact, the local field factor can be well

approximated by 𝐿 = |1 −𝐶−𝐷[1−𝐴(𝑇,𝜆)]

𝐶+𝐷[1+𝐴(𝑇,𝜆)]|2 ≈ |1 −

𝐶−𝐷

𝐶+𝐷|2

. Ignoring the weak temperature

dependence of the local field factor (which is mainly determined by the substrate properties) and

the sample thermal conductance for simplicity, we obtain from Eqn. (3-9) a criterion for optical

bistability

𝑑𝐴

𝑑𝑇≳

𝐴

𝑇−𝑇0 . (3-10)

Since 𝑑𝐴

𝑑𝑇 is negative (positive) on the blue (red) side of the exciton resonance (Fig.3-3(a), (b)) and

𝐴

𝑇−𝑇0 is always positive, optical bistability can only occur for sufficiently red-detuned optical

pumping. This is fully consistent with our experimental observation and the above intuitive picture

for the phenomenon.

Figure 3-6: (a) Schematic of the photo-thermal effect. The exciton resonance redshifts under

optical excitation for both blue- and red-detuned optical excitation from the exciton resonance

(shown by blue and red vertical arrows, respectively). For the red-detuned optical excitation, the

exciton resonance runs towards the pump and the effect is cumulative. This internal passive

feedback together with nonlinearity can lead to optical bistability. (b) Simulated average sample

temperature as a function of excitation power at 708 nm (blue-detuned) and 728 nm (red-detuned).

𝑃𝑓 and 𝑃𝑏 are the critical powers for temperature instability under forward and backward power

sweeps, respectively. (c) Simulated average sample temperature as a function of wavelength at 1,

10 and 100 μW. 𝜆𝑓 and 𝜆𝑏 are the critical wavelengths for temperature instability under forward

and backward wavelength sweeps, respectively. The region between the critical powers

(wavelengths) is bistable.

700 735 7700

150

300

lb

Te

mp

era

ture

(K

)

Wavelength (nm)

1mW

10mW

100mW

lf

(c)

0 250 5000

150

300

Te

mp

era

ture

(K

)

Power (mW)

708 nm

728 nm

Pf

Pb

(b)

700 720 7400.0

0.5

1.0

Abso

rban

ce

Wavelength (nm)

(a)

Page 79: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

58

3.4.2 Comparison to the observed bistable exciton resonance

In this section, we compare in details the predictions of the photo-thermal effect and the

experimental observations. To this end, we solve Eqn. (3-8) numerically for 𝑇(𝑃, 𝜆). For simplicity

we have used a constant thermal conductance (𝜅 ≈ 5 × 10−9 W/K). Figure 3-6(b) shows 𝑇(𝑃) for

two representative wavelengths. At 708 nm (blue-detuned from the exciton resonance), 𝑇 increases

monotonically with 𝑃. In contrast, at 728 nm (red-detuned from the exciton resonance), temperature

instability occurs at critical power 𝑃𝑓 and 𝑃𝑏 for forward and backward power sweeps, respectively.

The system is bistable between these two powers. Figure 3-6(c) is 𝑇(𝜆) for three representative

powers (1, 10 and 100 μW). Temperature instability is observed only at 10 and 100 μW. Again, the

system is bistable between the critical wavelength 𝜆𝑓 and 𝜆𝑏 , at which temperature instability

occurs for forward and backward wavelength sweeps, respectively. Moreover, the bistability region

shifts towards more red-detuned wavelengths for larger powers.

Now equipped with the sample temperature, we can simulate its reflectance using the

measured temperature dependent absorbance. Figure 3-4(d) is the simulated power dependence of

the sample reflectance. The simulation reproduces the overall trend of the experiment (Fig. 3-4(c))

for all wavelengths with critical powers within a factor of 2 – 3 of the measured values. Figure 3-

7(d) – (f) are the simulated wavelength dependence of the sample reflectance at a fixed power of 1,

10 and 100 μW, respectively. The agreement between simulation and experiment (Fig. 3-7(a) – (c))

is also very good including the critical wavelengths. We have also included in Fig. 3-5(e-f)

simulation for the pump-probe experiment (Fig. 3-5(c), (d)). Again, good agreement is seen for the

exciton characteristics as a function of pump wavelength. We note that there is only one free

parameter (the sample thermal conductance) in the simulation. The value of 5 × 10−9 W/K was

chosen to fit the entire set of the experimental observations. It is consistent with the measured [155],

[156] and calculated [157] thermal conductivity of WSe2. The low thermal conductance of

Page 80: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

59

suspended WSe2 is primarily responsible for the low critical powers required for optical bistability

here. We conclude that the photo-thermal effect captures the main features of optical bistability in

suspended monolayer WSe2 despite the crude assumptions made in the simulation, including the

temperature-independent thermal conductance and spatially uniform heating.

3.5 Valley contrasting bi-exciton resonance and control of light by helicity

Finally, we demonstrate the unique valley-selective optical bistability and control of light by

its helicity in suspended monolayer WSe2. The idea relies on the valley-dependent optical response

of monolayer TMDs [23], [32]–[43], [148], [151], in which the two degenerate direct gaps at the K

and K’ valleys of the Brillouin zone couple exclusively to the left (𝜎+) and right (𝜎−) circularly

polarized light, respectively. Since the K and K’ valleys are time-reversal copies of each other, the

Figure 3-7: Wavelength dependence of the sample reflectance for forward (blue) and backward

(red) wavelength sweeps. (a, b, c) are experiment and (d, e, f) are simulation. The excitation power

is fixed at 1 μW (a, d), 10 μW (b, e) and 100 μW (c, f).

700 710 720 730

0.0

0.5

1.0

1.5

Reflecta

nce

Wavelength (nm)

1 mW

(d)

700 710 720 730

0.0

0.5

1.0

1.5

Reflecta

nce

Wavelength (nm)

10 mW

(e)

700 720 740 760

0.0

0.5

1.0

1.5

Reflecta

nce

Wavelength (nm)

100 mW

(f)

700 710 720 7300.0

0.5

1.0

1.5

Refle

cta

nce

Wavelength (nm)

1 mW

(a)

700 720 740 7600.0

0.5

1.0

1.5

Re

flecta

nce

Wavelength (nm)

100 mW

(c)

700 710 720 7300.0

0.5

1.0

1.5

Re

fle

cta

nce

Wavelength (nm)

10 mW

(b)

Page 81: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

60

valley degeneracy can be lifted by applying an out-of-plane magnetic field (i.e. the exciton valley

Zeeman effect [38]–[43]). This is shown in Fig. 3-8(a) for the helicity-resolved reflectance

spectrum measured under a magnetic field of 8 T. The exciton resonance is split by ~ 1.8 meV, in

agreement with the reported magnitude of the exciton valley Zeeman effect [38]–[43]. We study in

Fig. 3-8(b) the optical nonlinearity for the 𝜎+ and 𝜎− excitation at a red-detuned wavelength (710

nm, marked by the dashed vertical line in Fig. 3-8(a)). Optical bistability is observed for both

circular polarizations with larger critical powers and a wider hysteresis loop for the 𝜎+ excitation.

This is consistent with the fact that 710 nm corresponds to a larger redshift from the exciton

resonance for the 𝜎+ excitation.

Moreover, there exists a finite range of power between the two bistability regions (marked by

two dotted vertical lines). This can be utilized for repeatable switching of the sample reflectance

by light helicity, as illustrated in the power dependence of the magnetic circular dichroism (MCD)

(Fig. 3-8(c)). This observation can be understood in terms of an energy barrier for optical switching.

For the hysteresis loop corresponding to left-handed optical pumping under 8 T, the lowest energy

state is the high-temperature (low-reflectance) state. During the forward scan, switching the

incident light handedness (e.g. by the modulator) from left to right within the hysteretic power

range changes the energy barrier for optical switching, and can cause a one-time only switch to the

high-temperature (low-reflectance) state. Subsequent switching of the light handedness back to left

merely changes the energy barrier, and cannot overcome the energy barrier to bring the state back

to the higher energy low-temperature (high-reflectance) state. As a result, the state is trapped to the

high-temperature (low-reflectance) state and no MCD signal (averaged over many modulation

cycles) can be observed. The same mechanism works for the hysteresis loop corresponding to right-

handed optical pumping, in which the lowest energy state is now the low-temperature (high-

reflectance) state. Rather than merely changing the energy barrier for optical switching, switching

the incident light handedness for the narrow power range in between the left- and right-handed

Page 82: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

61

hysteresis loops (shaded) actually changes the relative energy minima of the system between the

low-temperature (high-reflectance) state and the high-temperature (low-reflectance) state. Similar

argument can be made for the backward power scan. This is consistent with the result of Fig. 3-8(c)

if we take into account the broadened hysteresis loops. Repeatable switching of the sample

reflectance by pure helicity modulation is therefore possible in this narrow power range.

Figure 3-8: Valley-dependent optical bistability and control of light by its helicity. (a) Helicity-

resolved reflectance of sample #3 at 4 K under an out-of-plane magnetic field of 8 T. The K and

K’ valley excitons are split by ~ 2 meV. The vertical dashed line corresponds to 710 nm. (b)

Helicity-resolved reflectance under power sweeps at 710 nm. There exists a finite power range

between two bistability regions (between the vertical dotted lines). (c) Magnetic circular dichroism

(MCD) under forward and backward power sweeps at 710 nm. (d) Helicity-controlled optical

‘switching’ in real time at 710 nm and 39 μW. The helicity of light is modulated at 1 Hz by a liquid

crystal modulator (upper panel) and the sample reflectance follows the helicity modulation with a

change of reflectance by ~ 30% (lower panel).

Page 83: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

62

We thus choose a power between the bistability regions (39 μW) at 710 nm to demonstrate

repeatable light switching by pure helicity in Fig. 3-8(d). In the upper panel, the light helicity is

modulated by a liquid crystal modulator at 1 Hz. Repeatable switching of the sample reflectance

by ~ 30% is achieved (lower panel). (Note the delayed response in the reflectance when switching

from 𝜎+ to 𝜎− is due to the slow response of our liquid crystal modulator rather than limited by

the intrinsic response of the material. Using an electro-optic modulator, we have observed a

switching frequency up to 50 kHz, which again is limited by the speed of our light modulator.)

Although the unique valley-dependent optical bistability and helicity-controlled optical switching

can only be achieved under a finite magnetic field here, it is, in principle, also feasible to achieve

such a phenomenon under zero magnetic field utilizing the magnetic proximity effect. Recent

experimental studies have shown strong magnetic proximity coupling of monolayer TMDs to

magnetic insulators with a proximity exchange field exceeding 10 T [158], [159]! Our study has

thus opened up exciting opportunities in controlling light with light, including wavelength, power,

and helicity, using monolayer materials.

3.6 Conclusion

In this chapter, we have demonstrated robust optical bistability, the phenomenon of two

well-discriminated stable states depending upon the history of the optical output, in fully suspended

monolayers of WSe2 at low temperatures near the exciton resonance. Optical bistability has been

achieved under continuous-wave optical illumination that is red-detuned from the exciton

resonance at an intensity level of 103 W/cm2. The observed bistability is originated from a

photothermal mechanism, which provides both optical nonlinearity and passive positive feedback,

two essential elements for optical bistability. The low thermal conductance of suspended samples

is primarily responsible for the low excitation intensities required for optical bistability. At the

Page 84: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

63

presence of external magnetic field, the exciton bistability becomes helicity dependent due to

exciton valley Zeeman effect, which enables repeatable switching of the sample reflectance by light

polarization (left- or right-handedness). Our study has opened up exciting opportunities in

controlling light with light, including its wavelength, power, and polarization, using monolayer

semiconductors.

Page 85: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

Chapter 4

Tunable exciton-optomechanical coupling in suspended monolayer MoSe2

The extremely strong excitonic interaction in monolayer transition metal dichalcogenide

(TMD) semiconductors has enabled many fascinating light-matter interaction phenomena [23],

[33], [34], such as strongly coupled exciton-polaritons [160]–[162] and nearly perfect monolayer

mirrors [48], [49]. The strong light-matter interaction also opens up the possibility to dynamically

control the mechanical motion of a monolayer semiconductor through its exciton resonance [163].

Such exciton-optomechanical coupling in monolayer TMDs has not been studied so far. In principle,

the strong excitonic resonance in atomically thin semiconductor is reminiscent to the cavity

resonance of an optical cavity. Thus, we expect to observe the light-induced damping and anti-

damping of mechanical vibrations and modulation of the mechanical spring constant by moderate

optical pumping near the exciton resonance with variable detuning.

In this chapter, we discuss the first demonstration of exciton-optomechanical coupling in

monolayer MoSe2 resonators, as well as the its tunability using atomically thin semiconductor for

potential cavity-less optomechanics. The observation of gate-tunable exciton-optomechanical

coupling in a monolayer semiconductor may find novel applications in nanoelectromechanical

systems (NEMS) and in exciton-optomechanics.

This chapter is based on unpublished work in which I am the first author [164]. This chapter

is reprinted with permission from a manuscript in preparation to be published with the same title.

Page 86: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

65

4.1 Exciton-optomechanical backaction through strain

Atomically thin monolayer materials have attracted tremendous interests in the studies of

NEMS devices [65], [66], [75]–[84], [67], [85], [86], [68]–[74] because of their extremely

lightweight, mechanical flexibility, gate-tuneability and strong inter-mode couplings [75], [81],

[84], [86]. They also hold great promise in many sensing applications [80], [87], [96]–[98], [88]–

[95]. Although the mechanical aspects of NEMS devices based on monolayer materials have been

extensively studied in the literature [60], [65], [66], [68]–[71], [75], [76], [78]–[80], [83], [85], the

dynamical coupling between mechanical vibrations and light (i.e. optomechanical coupling)

remains as an open subject. In particular, the presence of strong excitonic resonances in monolayer

TMD semiconductors [23], [33], [34], which strongly interact with light (can reflect ~ 90 % of the

incident light) [48]–[50], presents a very interesting avenue to explore optomechanical coupling

effects in atomically thin materials without the need of an optical cavity. Although such cavity-less

optomechanical coupling has been demonstrated in quantum well heterostructures [163], the unique

mechanical properties and the extremely strong excitonic effects in monolayer TMDs can enable

fundamental studies and applications into new regimes.

We study in this work gate-tunable exciton-optomechanical coupling in suspended

monolayer MoSe2 (Fig. 4-1). The silicon back gate can induce charge in the MoSe2 membrane and

pull it down, which breaks the mirror symmetry of the mechanical device. Mechanical vibration of

the membrane can then create an oscillating strain in the sample, which in turn produces a

dynamical spectral shift in the exciton resonance due to its linear dependence on strain [56], [165].

As a result, illumination of the sample with light near the exciton resonance can generate a

photothermal force with dynamical backaction, producing interesting exciton-optomechanical

coupling effects we will explore below. Note that exciton-phonon parametric coupling cannot

generate efficient optomechanical backaction acting onto our suspended 2D systems. Under such

Page 87: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

66

mechanism, sufficient optomechanical feedback can only be achieved as the frequency magnitude

of mechanical modes is comparable to line width of exciton resonance. In our case, full line width

of exciton resonance (~ 3 meV ≈ 0.73 THz, see Figure 4-2(a)) is five order of magnitudes larger

than the vibration frequency of the suspended membrane (~ 40 MHz, see Figure 4-2(b)). Instead,

such an effect can be straightforwardly understood as exciton analogue of the standard

optomechanics. Similar to cavity resonance with position dependence, strain-tunable band gaps of

TMD semiconductors can periodically modulate exciton resonance by simply vibrating the

suspended membrane. Tensile strain causes redshift of exciton resonance [56], [60], [165] and thus

leads to asymmetrical photothermal deformation for optical excitation at opposite detuning of

exciton resonance. Additional strain is induced onto the sample by this photothermal deformation

potential. And the photon-induced strain takes time to accumulate (~ 10 ns for single-layer MoSe2

in cryogenic temperature) [105], which is determined by thermal properties of the suspended TMD

flakes. Thus, in principle, a backaction effect can build up via strain that mediates exciton resonance

with external optical excitation.

4.2 Experimental setup

Polydimethylsiloxane (PDMS) was employed to exfoliate monolayer MoSe2 flakes from

bulk crystals. After being identified and confirmed by optical contrast and photoluminescence, the

desired monolayers of MoSe2 were transferred onto the prepatterned trench substrates with metal

electrodes nearby. Circular drumhead trenches with diameter ~ 5 μm were structured adjacent to

electrodes and etched to 600-nm depth by photolithography and subsequent reactive ion etching

(RIE).

The suspended samples were measured under normal incidence with a microscope coupled

to a cryostat at 3.3K, of which filled with helium exchange gas ~19 Torr. For the reflectance

Page 88: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

67

measurements, either a supercontinuum light source with broadband radiation or a tunable-

wavelength Ti-Sapphire laser were employed to illuminate the samples. Illumination was focused

onto samples with a 40× objective with long working distance and the reflected radiation was

collected with the same objective. With a supercontinuum light source, the reflection was detected

by a spectrometer equipped with a charge coupled device (CCD). For a tunable-wavelength Ti-

Sapphire laser, both incident and reflected beams were recorded by a pair of silicon Avalanche

photodetectors. The reflectance spectra were obtained by comparing the reflectance from the

suspended MoSe2 and bare trench regions.

The suspended MoSe2 can be charged and actuated vertically by applying a gate voltage

due to capacitive nature of the resonator [65], [66], [68], [85]. In addition, the reflectivity 𝑅 of the

suspended MoSe2 is dependent on the vertical displacement 𝑧 towards the silicon backplane [81].

Thus, we characterize the mechanical resonance of our MoSe2 device by driving the resonator

capacitively and monitoring the corresponding reflectivity response as a function of the driving

frequency with a fast photodetector. The driven vibration of MoSe2 membrane reaches the

maximum as the ac driving frequency resides at the natural frequency of the resonator. This on-

resonance driven vibration leads to a Lorentzian-shaped reflectivity change of the suspended

MoSe2 at natural frequency. To convincingly visualize optical response of reflectivity change for

measurement of mechanical resonance, a dc gate voltage is needed to break the out-of-plane mirror

symmetry of periodic oscillation.

To observe the optomechanical effect acting on the MoSe2 resonator, a tunable-wavelength

Ti-Sapphire laser was employed. The illumination beam near exciton resonance was served both

as a pump and probe for measuring mechanical response under different pumping conditions. The

pump power was continuously tuned by rotating the polarization angle of a half-waveplate

sandwiched between two linear polarizers in the incident beam path.

Page 89: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

68

4.3 Observation of exciton-optomechanical coupling in monolayer MoSe2 resonator

In this section, we present the experimental findings that exhibit mutual interplay between

exciton resonance and mechanical resonance of monolayer MoSe2 resonators via strain. And such

exciton-optomechanical coupling in turn can modify mechanical modes of monolayer MoSe2

resonator without assistance of a cavity structure.

Figure 4-1: Schematic illustration of the device geometry and experimental setup. The mechanics

of suspended membrane is characterized with optical interferometry technique. And the incident

and reflected beams are recorded by a normal and fast photodetector, respectively. The static

displacement 𝑧 and dynamical oscillation 𝛿𝑧 of membrane can transduce to reflection signals,

which are measured with a multimeter and a network analyzer correspondingly. Meanwhile, the

vibration of monolayer MoSe2 membrane generates a dynamical strain 𝛿𝜖 that give rise to an

oscillating spectral shift of exciton resonance δ𝜔𝑋. In addition, the illumination of the sample with

light near the exciton resonance can generate a photothermal force 𝐹𝑝ℎ. As a result, the oscillation

of monolayer semiconductor (oscillating exciton spectral shift) can periodically modulate the

photothermal force as δ𝐹𝑝ℎ.

Page 90: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

69

4.3.1 Optical response of strained MoSe2 resonator

Figure 4-2(a) shows the sample reflection contrast spectra at varying gate voltages (𝑉𝑔 = 0 −

40 V) in a contour plot from device #1. (We only focus on positive 𝑉𝑔 in this study. Results from

negative 𝑉𝑔 are similar.) Spectra at selected gate voltages are also shown. A strong dip in the

reflection contrast due to the neutral exciton resonance in monolayer MoSe2 can be clearly seen.

The sharp exciton linewidth (𝛾 ~ 3 meV) and the large on-resonance reflection contrast (~ 90 %)

are consistent with results reported for nearly perfect monolayer mirrors [48], [49], which illustrate

the high sample quality here. The exciton resonance 𝜔𝑋 redshifts with increasing gate voltage due

to build up of in-plane strain in the sample as the gate pulls it down. It provides the necessary

mechanism for exciton-optomechanical coupling [81], as we will investigate below. Here the

nonlinear gate-induced spectral shift ∆𝜔𝑋 is due to the nonlinear dependence of the sample strain

on 𝑉𝑔 [81], [166], [167]. The theoretical gate-dependence of the exciton resonance 𝜔𝑋 is also

plotted in the lower panel of Fig. 4-2(a), which shows good agreement with the experimental result.

Detailed simulation can be found in section 4.4.1.

To characterize the mechanical resonance of our suspended sample, we drive the MoSe2

into mechanical vibrations by a small ac voltage (δ𝑉𝑔~ 100 mV in amplitude) superimposed on top

of the dc gate voltage 𝑉𝑔. We illuminate the sample by a laser beam with center wavelength 765

nm near its exciton resonance, and detect the reflectivity change as a function of the driving

frequency (details see section 4.4.2). Figure 4-2(b) shows the gate dependence of the fundamental

mechanical mode in a contour plot. With increasing |𝑉𝑔|, the resonance frequency first decreases

slightly and then increases afterwards. The behavior is consistent with results reported in the

literature [66], [71], [76], [78], [166]–[168], which can be explained by gate modulation of the

Young’s modulus in the sample as shown by the solid line fit (Fitting model see section 4.4.2). At

small 𝑉𝑔, our mechanical resonator displays high quality factor Q ~ 20,000 as shown in inset of Fig.

Page 91: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

70

4-2(b). The quality factor decreases quickly at high 𝑉𝑔 due to increased mechanical losses,

consistent with results from the literature [68], [70].

4.3.2 Exciton-modified mechanical response of MoSe2 resonator

Next we investigate the spectral dependence of the mechanical response in order to

understand the origin of the detected vibration-induced reflectivity change. To do this we vary the

incident laser wavelength near the exciton resonance and keep a small incident power fixed at 𝑃 =

1 μW. Figure 4-3(a) shows the detected reflectivity change due to the fundamental mechanical

mode at 𝑉𝑔 = 40 V and at varying incident wavelength. The peak response as a function of the

incident wavelength and frequency detuning Δ𝜔 ≡ 𝜔 − 𝜔𝑋 from the exciton resonance 𝜔𝑋 is

Figure 4-2: Gate tuning of excitonic and mechanical resonances in suspended MoSe2 resonator.

(a) Optical reflectance spectra of sample #1 are shown at selective gate voltages (upper panel) and

in contour plot (lower panel). The orange dashed line in lower panel of (a) is prediction of Eqn. (4-

6). (b) Mechanical resonance shift as a function of gate voltages. The grey dashed line is the

prediction of Eqn. (4-8). Inset: A typical mechanical resonance of our resonator at 3.3K, and the

corresponding Lorentzian fit with Q ~ 20,000.

Page 92: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

71

shown in Fig. 4-3(b). The response changes sign at the exciton resonance, maximizes in amplitude

at a small detuning (Δ𝜔 ~ 3 THz), and diminishes with further increase in detuning. The vanishing

response exactly at the exciton resonance also provides a very reliable way to calibrate the

frequency detuning Δ𝜔 . The spectral response in Fig. 4-3(b) can be reproduced by taking a

frequency derivative of the measured reflection contrast 1

𝑅

𝑑𝑅

𝑑𝜔 (Fig. 4-2(a)), and multiplied by a

frequency shift δ𝜔𝑋 ≈ 0.1 THz. The spectral dependence of the mechanical response can therefore

be understood in terms of a dynamical shift in the exciton resonance (δ𝜔𝑋) caused by an oscillating

strain (δ𝜖) in the suspended MoSe2 sample undergoing mechanical vibrations (inset of Fig. 4-3(b)).

The dynamical shift δ𝜔𝑋 produces the reflectivity change as observed, which also provides an

effective method of detecting thermal vibrations of a monolayer semiconductor at cryogenic

temperatures (see section 4.6). Such picture can be further tested in Fig. 4-3(c) by comparing the

𝑉𝑔-dependence of the extracted dynamical shift δ𝜔𝑋 with the quantity 𝑑𝜔𝑋

𝑑𝑉𝑔δ𝑉𝑔 computed from the

exciton spectral shift in Fig. 4-2(a) (under ac driving voltage δ𝑉𝑔~ 70 mV). The good agreement

supports our interpretation. The result in Fig. 4-3(c) also suggests stronger exciton-mechanical

coupling at higher gate biases. We will come back to this point later.

Page 93: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

72

4.3.3 Photothermal backaction of exciton-optomechanical coupling

Equipped with an understanding of the excitonic and mechanical characters of our device, we

now proceed to examine effects due to exciton-optomechanical backaction. Figure 4-4(a), similar

to Fig. 4-3(a), shows the optically detected fundamental mechanical mode at 𝑉𝑔 = 40 V and at

Figure 4-3: Interaction between excitonic and mechanical resonances in monolayer MoSe2

resonator. (a) Normalized reflectivity change pumped with 1 μW as a function of driven frequency

at selective wavelengths throughout the exciton resonance. The spectra are displaced with equal

spacing for clarity. Data was taken with 𝑉𝑔 = 40𝑉 at 3.3K. The illumination power is 1 μW. (b)

The peak response of mechanical resonance across exciton resonance (symbols with line) as a

function of incident wavelength (bottom axis) and exciton detuning Δ𝜔 (top axis). Data was

extracted from (a). The dashed red line is the illustration of spectral shift due to dynamical strain.

Inset: A sketch of a dynamical shift in the exciton resonance (δ𝜔𝑋). (c) The extracted dynamic shift

of exciton resonance δ𝜔𝑋 at selective gate voltages (symbols). The blue solid line is obtained by

multiplying quantity 𝑑𝜔𝑋

𝑑𝑉𝑔 extracted from the lower panel of Fig.4-2 (a) with a dynamic voltage

δ𝑉𝑔~ 70 mV.

Page 94: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

73

varying incident wavelength. Compared to Fig. 4-3(a), the incident light power is increased from

𝑃 = 1 μW to 𝑃 = 10 μW in order to better show effects from dynamical backaction. Very

asymmetric dependence on optical detuning is seen. In particular, the blue-detuned mechanical

resonance is significantly broader than the red-detuned counterpart. The shift in the mechanical

resonance is also very asymmetric with detuning. These behaviors are summarized by the detuning

dependences of the mechanical linewidth and resonance frequency shown in Fig. 4-4(b) and (c),

respectively. For large detuning from the exciton resonance, the fundamental mechanical mode is

largely unperturbed by the incident light. Near the exciton resonance, however, the mechanical

resonance frequency redshifts and the mechanical linewidth broadens with very asymmetric

detuning dependences.

Page 95: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

74

Figure 4-4: Light induced damping and anti-damping of mechanical modes across exciton

resonance in monolayer MoSe2 resonator. (a) Normalized reflectivity change pumped with 10 μW

as a function of driven frequency at selective wavelengths throughout the exciton resonance. The

spectra are displaced with equal spacing for clarity. Data was taken with 𝑉𝑔 = 40𝑉 at 3.3K. The

extracted (b) line width and (c) frequency of mechanical resonance shown in (a) as a function of

exciton detuning (bottom axis) and incident wavelength (top axis). The green and blue solid lines

in (b) and (c) are extracted contributions from photothermal softening(symmetric) and backaction

(anti-symmetric). The red solid lines in (b) and (c) are fitting lines by summing up green and blue

lines. The data without symmetric contribution are shown in (d) and (e), where the red solid lines

now are fitting from Eqns. (4-1) and (4-13).

Page 96: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

75

The observations can be understood in a model of photothermal backaction and gate-induced

mirror symmetry breaking in the mechanical device. At 𝑉𝑔 = 40 V, the monolayer MoSe2

membrane is pulled down by the applied electrostatic force from the back gate, which lowers the

out-of-plane mirror symmetry of the mechanical device (Fig. 4-1). With the lowered symmetry, a

small driven mechanical vibration in the out-of-plane direction creates an oscillating strain δ𝜖 in

the membrane that produces a dynamical spectral shift in the exciton resonance δ𝜔𝑋. Meanwhile,

illumination of the sample by laser light near the exciton resonance produces a photothermal force

𝐹𝑝ℎ, whose magnitude depends on the detuning from the exciton resonance Δ𝜔 and the overall

vertical displacement 𝑧 of the membrane (Fig. 4-1). As a result, the oscillating membrane (and the

oscillating exciton resonance) periodically modulates the photothermal force 𝐹𝑝ℎ and creates a

dynamical backaction on its mechanical vibration, which changes its mechanical linewidth Γ and

resonance frequency Ω according to [105], [114]

Γ = Γ0 (1 + 𝑄Ω0𝜏

1+Ω02𝜏2

∇𝐹𝑝ℎ

𝛫),

Ω ≈ Ω0 (1 −1

1+Ω02𝜏2

∇𝐹𝑝ℎ

2𝛫). ` (4-1)

Here Γ0 and Ω0 are the unperturbed mechanical linewidth and resonance frequency, respectively,

𝜏 is the time delay for the photothermal force in response to the mechanical movement of the

membrane, ∇𝐹𝑝ℎ is the gradient of the photothermal force in the out-of-plane direction, and 𝛫 is

the unperturbed spring constant of the mechanical resonator, 𝑄 =Ω0

2πΓ0 is the mechanical quality

factor.

To quantitatively model our experiment, we need to calculate the gradient of the

photothermal force ∇𝐹𝑝ℎ.Thus, the contribution that gives rise to dynamical backaction is

𝛻𝐹𝑝ℎ

𝛫= 𝛼𝑃

4(𝛥𝜔)

[(𝛥𝜔)2+𝛾2]2∙ ∆𝜔𝑋 (4-2)

Page 97: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

76

where 𝛼 represents the overall amplitude of the photon-induced rigidity compared to the

unperturbed spring constant of the resonator. Here the overall amplitude 𝛼 depends on parameters

of the device such as the Young’s modulus, the thermal expansion coefficient and the thermal

conductivity of MoSe2 etc. Detailed expression of 𝛻𝐹𝑝ℎ

𝛫 will be derived and presented in the

following section 4.4.3. As expected, the result is linearly proportional to the incident power 𝑃 (see

Fig. 4-6(a),(b)) and the gate-induced shift in the exciton resonance ∆𝜔𝑋 (Fig. 4-2(a)), and exhibits

antisymmetric dependence on detuning Δ𝜔. It reflects the fact that the oscillating exciton resonance

dynamically modulates the photothermal force. In addition to the dynamical backaction

contribution, there is also a contribution from photothermal softening [68], [81], [105] of the

mechanical spring constant ∝𝑃

(Δ𝜔)2+𝛾2. Since it is originated from optical absorption near the

exciton resonance, it shows symmetric dependence on detuning Δ𝜔. It is also independent of ∆𝜔𝑋.

The results in Fig. 4-4(b) and (c) can be described by including both the symmetric and the

antisymmetric contributions with the corresponding amplitudes and the exciton linewidth 𝛾 as

fitting parameters. In order to isolate the antisymmetric contribution linear to ∆𝜔𝑋, we subtract the

symmetric contribution from Fig. 4-4(b) and (c), and show the results in Fig. 4-4(d) and (e). The

dependences on detuning Δ𝜔 are reminiscent to the optical spring effect and optical damping

reported in the literature on cavity optomechanics [108], [109], [115], [126]. The underlying

mechanism is in fact very similar. In cavity optomechanics, the mechanical vibrations of the end

mirror of an optical cavity dynamically change the cavity resonance frequency, which in turn

modulates the intracavity optical field and the radiation force on the mirror. The periodically

modulating radiation force creates a dynamical backaction on the mirror’s oscillations, and changes

its mechanical resonance frequency and linewidth. Here the mirror is replaced by the MoSe2

membrane and the cavity resonance is replaced by the exciton resonance (no optical cavity is

required). Because the exciton resonance redshifts (∆𝜔𝑋 < 0) with increasing strain (i.e. increasing

Page 98: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

77

downward displacement 𝑧 of the membrane), the sign of the detuning dependences in Fig. 4-4(d)

and (e) is opposite to that in cavity optomechanics. In particular, red-detuning (blue-detuning)

causes anti-damping (damping) and blueshift (redshift) on the mechanical resonance.

Cavity optomechanics

Measurement: optical spring effect and

optical damping of mechanical motion

through the cavity resonance

This work

Measurement: light induced damping and anti-

damping of mechanical motion through the

fundamental exciton resonance

Resonantly reflect or transmit the incident

light at cavity resonance frequency due to the

intracavity interference of trapped light

Resonantly reflect or transmit the incident light

at exciton resonance frequency due to the

radiative damping of oscillating exciton dipole

Cavity resonance with position dependence Exciton resonance with strain dependence

Mirror at movable end Suspended MoSe2 monolayer

Radiation force due to the momentum

transfer of confined photons

Photothermal force due to light induced

tension change

Time delay: decay lifetime of trapped photon

inside cavity.

Time delay: thermal equilibrium time of

photothermal deformation

4.4 Modeling strain induced exciton-optomechanical coupling in suspended MoSe2

We utilize mechanical model of thin film developed in Section 1.2.2 to model the gate-

induced strain effect of excitonic and mechanical resonance in our MoSe2 resonators. Particularly,

integrating with required mechanical and optical theories, we derive in details to show a dynamical

backaction can build up through a photothermal force, which stems from the light absorption near

exciton resonance and the resultant tension change. And such photothermal force acts on the

resonator with some time delay related to thermal diffusion and subsequent deformation.

Table 4-1: Comparison of backaction effects in cavity optomechanics and this work.

Page 99: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

78

4.4.1 Gate induced exciton spectral shift

Next, we find the exciton spectral shift as a function of gate voltage. The suspended MoSe2

monolayer was pulled down by electrostatic force and the strained flake redshifted its excitonic

resonance correspondingly. For a circular drumhead trench, the deflection of the membrane under

uniform-loaded force can be described with a parabola distribution 𝜉(𝑟) = 𝑧(1 −𝑟2

𝑅2), where z is

the vertical displacement at the center of the suspended membrane, and 𝑅 = 2.5 𝜇𝑚 is the radius

of the circular drumhead trench. The stored elastic energy within the strained membrane 𝑈𝑒𝑙 is

[166], [167]

𝑈𝑒𝑙 =𝜋𝐸𝑌𝑡

1−𝜈2∫ [(1 − 𝜈2)𝜖0 +

1

2(𝜕𝜉

𝜕𝑟)2]2

𝑅

0𝑟𝑑𝑟 =

𝜋𝐸𝑌𝑡

1−𝜈2[2𝑧4

3𝑅2+ (1 − 𝜈2) (𝜖0𝑧

2 +1

2𝜖02𝑅2)] (4-3)

where 𝐸𝑌 is the 2D Young’s modulus of MoSe2, 𝜈 is the Poisson ratio, 𝑡 = 0.67 𝑛𝑚 is the

thickness of monolayer MoSe2 flake and 𝜖0 is the initial strain that applied during device

fabrication. And the electrostatic energy of the gated resonator 𝑈𝑒𝑠 is

𝑈𝑒𝑠 = −1

2𝐶𝑔𝑉𝑔

2 (4-4)

where 𝐶𝑔 is the electrostatic capacitance of the resonator and 𝑉𝑔 is the gate voltage. Therefore, to

find the equilibrium position of the membrane, we set the total energy to be minimum, namely:

𝐹𝑡𝑜𝑡 = −𝜕(𝑈𝑒𝑙+𝑈𝑒𝑠)

𝜕𝑧= 0, (4-5)

Consequently, we obtain the vertical displacement of suspended flake 𝑧 ≈𝜀0

8𝐸𝑌𝑡𝜖0

𝑅2

𝑑2𝑉𝑔2. The

strain 𝜖 is related to the vertical displacement 𝑧 as 𝜖 =1

2𝑅∫ (

𝜕𝜉

𝜕𝑟)2𝑅

0𝑟𝑑𝑟 =

2

3(z

𝑅)2. Therefore, the

gate induced exciton shift is

∆𝜔𝑋 = 𝛽𝑋 𝜖 ≈𝛽𝑋𝜀0

2𝑅2

96𝐸𝑌2𝑡2𝜖0

2𝑑4𝑉𝑔4 (4-6)

Page 100: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

79

where 𝛽𝑋 represents the strain-induced excitonic spectral shift rate. And the extracted exciton peak

position as a function of the gate voltage 𝑉𝑔 can also be fitted by the model using 𝛽𝑋~ 0.1 eV/%

strain (lower panel of Fig. 4-2(a)), which agrees reasonably well with reported values [56], [60],

[165].

4.4.2 Gate induced shift in mechanical resonance

The above membrane mechanics can also illustrate the observed shift in mechanical resonance

of our resonator in terms of gate voltage. The spring constant of the resonator 𝑘𝑒𝑓𝑓 is given by

taking the second derivative of the total energy

𝑘𝑒𝑓𝑓 =𝜕2(𝑈𝑒𝑙+𝑈𝑒𝑠)

𝜕𝑧2=

8𝜋𝐸𝑌𝑡

(1−𝜈2)𝑧2 + 4.924𝐸𝑌𝑡𝜖0 −

1

2

𝜕2𝐶𝑔

𝜕𝑧2𝑉𝑔2 (4-7)

Thus, the corresponding frequency of mechanical vibration Ω0 is as following

Ω0 = √𝑘𝑒𝑓𝑓

𝑀𝑒𝑓𝑓= √

𝜋𝜀02

8(1−𝜈2)𝐸𝑌𝑡𝜖02𝑅2

𝑑2𝑉𝑔4−

𝜀0𝜋𝑅2

3𝑑3𝑉𝑔2+4.924𝐸𝑌𝑡𝜖0

𝑀𝑒𝑓𝑓 (4-8)

Here, 휀0 is vacuum dielectric constant, 𝑑 = 600 𝑛𝑚 is the trench depth between the top SiO2

surface and the back-gate silicon surface. By using young’s modulus 𝐸𝑌 = 310 GPa , effective

mass 𝑀𝑒𝑓𝑓 = 9 × 10−17 𝑘𝑔 and built-in strain 𝜖0 = 0.36% as free fitting parameters at 3.3K, the

measured gate-induced frequency can be well described by the above frequency equation in terms

of gate voltage 𝑉𝑔. The extracted effective mass 𝑀𝑒𝑓𝑓 and young’s modulus 𝐸𝑌, respectively, are

slightly larger than the reported values in room temperature [169].This discrepancy suggests

additional mass and initial strain imparted into the resonator by the polymer residues from

fabrication process. In addition, a stiffened modulus at low temperature was reported in tensioned

graphene resonator [170] likely due to influence of adsorbent molecules, which is similar to our

observation here. Symmetrical frequency shift by gate excludes the negligible doping effect on

Page 101: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

80

mechanical frequency (results at negative gates not shown). To minimize the optomechanical effect,

a probe wavelength away from exciton resonance (765 nm) was employed in gate-dependent

measurements with low power intensity (~ 5 𝜇W).

4.4.3 Modelling photothermal backaction in strained MoSe2 resonator

To better understand our experiment, we need to make quantitative predictions for the

photothermal coupling strength (namely, the gradient of photothermal force) ∇𝐹𝑝ℎ for our system.

The strain-mediated feedback mechanism is the following (see lower left panel of Fig. 4-1): the

monolayer MoSe2 membrane is initially deformed by a gate voltage, and it absorbs light

proportional to both the excitonic absorbance �� and the incident laser power P. The membrane

heats up locally near the laser spot, it takes some timescale 𝜏 for the temperature to equilibrate

across the membrane, the tension σ changes, which equivalently apply a photothermal force 𝐹𝑝ℎ

with time delay 𝜏 onto the membrane. This retarded force can exert nonzero work within one

oscillation cycle and cause the resonator to experience a change in both frequency and line width

of mechanical motion.

The quantitative formula for this light-induced change in frequency and line width of

mechanical modes have been given as Eqn. (4-1), which are all proportional to the gradient of

photothermal force 𝛻𝐹𝑝ℎ. Here, we derive the gradient of photothermal force 𝛻𝐹𝑝ℎ ≡ 𝑑𝐹𝑝ℎ/𝑑𝑧 for

our resonator system, where 𝐹𝑝ℎ = 4𝜋𝜎𝑝ℎ𝑧 is the photothermal force felt by the suspended

membrane and 𝜎𝑝ℎ is the photothermal-induced tension. And we have for photothermal force

gradient related to dynamical backaction as

𝛻𝐹𝑝ℎ = 4𝜋𝑧𝑑𝜎𝑝ℎ

𝑑𝑧 (4-9)

Page 102: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

81

If we write down the photothermal tension from laser heating onto our monolayer MoSe2 membrane,

then as [81]

𝜎𝑝ℎ = −𝐸𝑌𝑡

1−𝜈2𝛼𝑡ℎ𝐴𝐿𝑃

4𝜋𝜅 (4-10)

where 𝛼𝑡ℎ is the thermal expansion coefficient, 𝜅 is the sheet thermal conductivity, �� is the

complex absorbance of the monolayer MoSe2 flake, 𝐿 is local field factor, 𝑃 is the power of the

incident laser. Here the complex absorbance is modeled by a single Lorentzian that arises from the

exciton resonance ��(𝜔) =𝐴0

1−𝑖(𝜔−𝜔𝑋(𝑧))

(𝛾)

. Since the exciton resonance of monolayer MoSe2

membrane is linearly proportional to the applied strain, which is also a function of vertical

displacement of the suspended sample. Then the gradient of photothermal tension can be written

by substituting the complex absorbance.

𝑑𝜎𝑝ℎ

𝑑𝑧≈ −

𝐸𝑌𝑡

1−𝜈2𝛼𝑡ℎ𝐿

4𝜋𝜅

𝜕��

𝜕𝑧𝑃 = −

𝐸𝑌𝑡

1−𝜈2𝛼𝑡ℎ𝐿

4𝜋𝜅𝑃

𝐴0

(1+(Δ𝜔)2

𝛾2)2

2(Δ𝜔)

𝛾2∂𝜔𝑋

∂ϵ

∂ϵ

∂z (4-11)

Here we use β𝑋 =∂𝜔𝑋

∂ϵ and

∂ϵ

∂z=4

3

𝑧

𝑅2 from section 4.4.2. Then we can substitute the gradient of

absorbance due to exciton resonance into gradient of the photothermal force,

𝛻𝐹𝑝𝑡ℎ = −𝐸𝑌𝑡

1−𝜈2𝛼𝑡ℎ𝐿𝑃

𝜅(𝑧

𝜕��

𝜕𝑧) = −

𝐸𝑌𝑡

1−𝜈2𝛼𝑡ℎ𝐿

𝜅𝑃��(

4(Δ𝜔)

𝛾2+ℏ2(Δ𝜔)2Δ𝜔𝑋) (4-12)

Thus, the contribution that gives rise to dynamical backaction is

𝛻𝐹𝑝ℎ

𝛫= −

1

𝐾

𝐸𝑌𝑡

1−𝜈2𝛼𝑡ℎ𝐿𝐴0

𝜅

𝛾2

4𝑃

4(𝛥𝜔)

[(𝛥𝜔)2+𝛾2]2∙ ∆𝜔𝑋 = 𝛼𝑃

4(𝛥𝜔)

[(𝛥𝜔)2+𝛾2]2∙ ∆𝜔𝑋 (4-13)

where 𝛼 = −1

𝐾

𝐸𝑌𝑡

1−𝜈2𝛼𝑡ℎ𝐿𝐴0

𝜅

𝛾2

4 is the detailed expression for the overall magnitude of the light

induced stiffness compared to the natural spring constant of the resonator as shown in Eqn.(4-2).

Two significant implications can be concluded here. First, when ∆𝜔𝑋 = 0, namely no deflection

of membrane, 𝛻𝐹𝑝ℎ will vanish. In other word, there is no dynamical backaction from photothermal

force without breaking out-of-plane mirror symmetry by gate [81]. Second, the sign of 𝛻𝐹𝑝ℎ is

determined jointly by exciton detuning 𝛥𝜔 and excitonic spectral shift ∆𝜔𝑋 induced by strain.

Page 103: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

82

So far, we still need the explicit expression for the time delay 𝜏 that photothermal force in

response to the mechanical movement of the membrane to fulfill the quantitative model of

photothermal backaction. Since a photothermal model within membrane have been built up in

Chapter 3, we also start from the steady heat diffusion shown as Eqn. (3-4) and add a time-

dependent term ��(𝑟, 𝑡) with the following substitution 𝑇 → 𝑇(𝑟) + ��(𝑟, 𝑡) . After applying

separation of variables and boundary condition ��(𝑅, 𝑡) = 0, we acquire the solution for ��(𝑟, 𝑡)

similar to the dynamical displacement ��(𝑟, 𝑡) in section 1.2.2 as

��(𝑟, 𝑡) = ∑ 𝐴𝑚𝐽0 (𝑟√𝜌𝑐𝑝

𝜅𝜏𝑚)∞

𝑚=0 𝑒−𝑡/𝜏𝑚 (4-14)

where the 𝐴𝑚 is the amplitude and 𝜏𝑚 is the time constant of the mth phonon diffusion mode. Apart

from the oscillating counterpart in the dynamical displacement ��(𝑟, 𝑡), the time dependent part in

��(𝑟, 𝑡) is an exponential decay. Thus, the time delay 𝜏 is equivalent to the longest time constant of

the possible phonon diffusion process as

𝜏 =𝑅2𝜌𝑐𝑝

2.4052𝜅 (4-15)

In principle, we now have each piece of modelling photothermal backaction and can

quantitatively predict the light-induced damping and anti-damping acting on our resonators. We

can also estimate the overall amplitude of the photo-induced change in line width ∆Γ =Ω02𝜏

1+Ω02𝜏2

∇𝐹𝑝ℎ

2𝜋𝛫

by combining Eqn. (4-1) with Eqn. (4-13). The estimated ∆Γ = 0.5𝑀𝐻𝑧 (V𝑔 = 40 𝑉) is in good

agreement with the extracted value from the experiment (Fig. 4-4(d)) by using the measured

mechanical resonance Ω0 = 2𝜋 × 41 𝑀𝐻𝑧 (V𝑔 = 40 𝑉 ), spring constant 𝐾 = 4.8 𝑁𝑚−1 , the

extracted young’s modulus 𝐸𝑌 = 208 𝑁𝑚−1, thermal expansion coefficient 𝛼𝑡ℎ = 4 × 10

−6𝐾−1

[171], local field factor 𝐿 = 2.27, sheet thermal conductivity 𝜅 = 2 × 10−8𝑊𝐾−1 [62], [105],

[172], excitonic absorbance 𝐴0 = 0.85 , and exciton line width 𝛾 = 3.8 𝑇𝐻𝑧 , strain-induced

exciton shift ∆𝜔𝑋 = 3.8 𝑇𝐻𝑧, the time for heat travel out of drum 𝜏 =𝑅2𝜌𝐶𝑝

2.4052𝜅= 5 × 10−9s [105]

Page 104: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

83

(with specific heat capacity 𝐶𝑝 = 20 𝐽/(𝑘𝑔 ∙ 𝐾) agrees with values in the literature for monolayer

[105] and bulk [173] MoSe2). And we will demonstrate the numerical prediction of photothermal

backaction as developed above and compare the predictions to the extracted experimental findings

within next section.

4.5 Tunable exciton-optomechanical effects

We have performed multiple checks to further examine the observed damping and anti-

damping in mechanical modes through exciton resonance for our system. In this section, we present

gate- and power-dependence of light induced shift of mechanical line width ΔΓ. By comparing

experimental results to the model of photothermal backaction, we thus verify its exciton-

optomechanical origin of the observed modification of mechanical modes across the exciton

resonance in a monolayer MoSe2 resonator. Below we will focus on the mechanical linewidth

because optical damping is generally much more pronounced than the optical spring effect due to

the enhancement factor 𝑄 in Eqn. (4-1).

4.5.1 Gate dependence of exciton-optomechanical coupling

Finally, we note that Eqn. (4-13) also predicts the antisymmetric contribution is important

only when there is a significant gate-induced redshift in the exciton resonance (∆𝜔𝑋 ≳ 𝛾). To test

this prediction, we perform optical detuning studies at different applied gate voltages, which vary

the in-plane sample strain and thus the shift in the exciton resonance ∆𝜔𝑋. Figure 4-5(a) shows the

detuning dependences of the antisymmetric contribution of the mechanical linewidth at different

gate voltages 𝑉𝑔 (the incident power is fixed at 𝑃 = 3.3 μW). We have subtracted the symmetric

contributions from the total (symmetric plus antisymmetric) mechanical linewidth to obtain Fig. 4-

Page 105: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

84

5(a). We can see that the amplitude of the antisymmetric contribution increases significantly with

increasing 𝑉𝑔, as summarized in Fig. 4-5(b). In Fig. 4-5(b), we also show the prediction from Eqn.

(4-1) and (4-13) taking into account the 𝑉𝑔-dependence of the exciton resonance shift ∆𝜔𝑋 from

Fig. 4-2(a). The good agreement strongly supports our physical interpretations.

4.5.2 Power dependence of exciton-optomechanical coupling

To further understand the observation on the sample’s mechanical response as a function of

detunings, we perform the measurement on the power dependence of the mechanical resonance

under the same condition (3.3K, Vg ~ 40V). The magnitudes of light induced change of line width

are background-subtracted and summarized in Fig. 4-6(a) at several selected powers. For line width

Figure 4-5: Gate dependence of optical damping across exciton resonance. (a) The shift of

mechanical linewidth induced by antisymmetric contribution as a function of exciton detuning at

selective gate voltages. Data is displaced with equal spacing for clarity. The solid lines are fitting

based on Eqns. (4-1) and (4-13). (b) The extracted amplitude of the antisymmetric contribution of

mechanical linewidth as a function of gate voltages. The red dashed line is prediction based on

Eqns (4-1) and (4-13).

Page 106: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

85

of mechanical motion, a clear enhancement of optomechanical damping and anti-damping is

observed with increasing pump power. These amplitude of optical-induced mechanical response

clearly shows a linear dependence as a function of power. By using the young’s modulus 𝐸𝑌, initial

strain 𝜖0 and extracted parameters from fitting Eqn. (4-9), experimental results and theoretical

calculation reach a reasonably good agreement for photo-induced line width shown in Fig. 4-6(b).

Note that there is no sudden jump, namely optical bistability, occurring for the amplitudes of

mechanical resonance as a function of detuning for all our measurement. This indicates the incident

powers still remain at low level without generating any nonlinearity behavior. In addition, the

increased power can lead to a redshift of exciton resonance, which further enlargers the optically

induced change of line width. Thus, the power dependence of line width shifted by optical pump is

in agreement with the prediction from photothermal backaction model.

Figure 4-6: Power dependence of optical damping across exciton resonance. (a) The shift of

mechanical linewidth induced by antisymmetric contribution as a function of exciton detuning at

selective incident power. Data is displaced with equal spacing for clarity. The solid lines are fitting

based on Eqns. (4-1) and (4-13). (b) The extracted amplitude of the antisymmetric contribution of

mechanical linewidth as a function of incident power. The red dashed line is prediction based on

Eqns (4-1) and (4-13).

Page 107: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

86

4.6 Detection of thermal vibrations of monolayer MoSe2 at cryogenic temperatures

In driven vibration, the vibrated MoSe2 monolayer generates an oscillating strain (δ𝜖) and

corresponding dynamical shift of exciton resonance (δ𝜔𝑋) that can greatly modulate the reflectivity

change near exciton resonance. Namely, the mechanical response of the suspended MoSe2 sample

can be detected more ‘easily’ with probe wavelength near exciton resonance than those far from

exciton resonance. To verify this exciton-enhanced sensitivity of mechanical resonance, we

measure the thermal vibrations of another MoSe2 mechanical resonator (D2) with the same device

structure at 3.3K. After testing the spectral range of fundamental mechanical mode, the modulated

force is turned off and the resulting thermal vibration noise is recorded with a spectrum analyzer.

To clearly resolve the thermal vibration process at such cryogenic temperature, we used the probe

wavelength near exciton resonance with incident power P = 100 𝜇W.

Figure 4-7(a) shows the reflection contrast measurement of device #2 with gate voltage (~-

60V) at 3.3K, with exciton resonance around 745nm. The above-mentioned 100𝜇W probe power

for resolving thermal motion can significantly heat up the suspended membrane, which further

redshifts the exciton resonance of MoSe2 monolayer to around 750nm. In addition, we can see that

the reflectivity of sample only changes rapidly around the exciton resonance compared to the

regions outside the exciton resonance. In Fig. 4-7 (d) and (e), we demonstrated thermal noise

spectra at the same gate voltage for probe wavelength (b) near exciton resonance (752nm) and (c)

away from exciton resonance (759 nm). Mechanical resonance around 55.2MHz is clearly observed

by 752nm probe as the resonance peak with 759 nm probe remains elusive. This indicates that

dynamical shift of exciton resonance due to thermal motion can amplify the thermal noise signal

by modulating the reflectivity change near exciton resonance. Namely, one can utilize the exciton

resonance with appropriate probe wavelength to enhance the measurement sensitivity of

mechanical motion of suspended monolayers.

Page 108: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

87

4.7 Conclusion

In this chapter, we have demonstrated efficient dynamical control of the mechanical motion

of suspended monolayer MoSe2 by light through the material’s strong excitonic resonance. Gate-

tunable optomechanical damping and anti-damping of the mechanical vibrations due to

photothermal backaction have been observed. The damping and anti-damping effects are effective

Figure 4-7: Thermal motions of monolayer MoSe2 resonator. (a) Optical reflection spectrum of

sample #2 at 𝑉𝑔 = −60𝑉. The blue and green dashed lines are aligned with 752 nm and 759 nm,

which used for measurements of thermal noise in (b) and (c). The red dash lines in (b) and (c) are

Lorentzian fittings.

Page 109: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

88

(Fig. 4-4 b and d) in the sense that the photo-induced change in the mechanical linewidth at

moderate optical pumping is significantly larger than the unperturbed linewidth Γ0 . Further

optimization of the device that allows larger gate-induced shift in the exciton resonance ∆𝜔𝑋 ,

which favors dynamical backaction, could open up interesting applications in optomechanical

cooling and mechanical lasing in monolayer TMD semiconductors.

Page 110: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

Chapter 5

Valley magnetoelectricity in single-layer MoS2

Magnetoelectric (ME) effect, the phenomenon of inducing magnetization by application of

an electric field or vice versa, holds great promise for magnetic sensing and switching applications

[174]. Studies of the ME effect have so far focused on the control of the electron spin degree of

freedom (DOF) in materials such as multiferroics [175] and conventional semiconductors [176].

Meanwhile, the crystalline structure of solids can attribute electrons a valley degree of freedom

(DOF) [23], [32]–[34], which represents the degenerate energy extrema in momentum space. And

the appealing feasibility for dynamic control of valleys has led to the concept of valleytronics

[23], [33], [34], [177], which aims at harvesting valley, together with charge and spin, to store

and transduce information.

In this chapter, we will demonstrate a new form of the ME effect based on the valley DOF

in two-dimensional (2D) Dirac materials [32]–[34], [178]. By manipulating carriers’ momentum

distribution, the valley magnetization can arise from the joint effect of current and strain. And such

magnetoelectric effect from valley DOF that sustains up to room temperature offers a practical

mechanism for converting charge to magnetization in valleytronic devices.

This chapter is based on published work in which I am a co-author [179]. This chapter is

reprinted with permission from J. Lee, Z. Wang, H. Xie, K. F. Mak, J. Shan, Nat. Mater. 2017 16,

887–891. ( https://www.nature.com/articles/nmat4931).

Page 111: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

90

5.1 Magnetization from valley degree of freedom

Electrons in two-dimensional (2D) Dirac materials including gapped graphene and single-

layer TMDs possess a new two-fold valley degree of freedom (DOF) corresponding to the K and

K’ valleys of the Brillouin zone. The valley DOF carries orbital magnetic moment [32]–[34], [178].

We would like to emphasis that the magnitude of valley magnetic moment depends on carrier’s

momentum. Due to time reversal symmetry and the limit of 2D nature, valley magnetic moment

points out of 2D plane with opposite direction at K and K’ valleys. Thus, a nonzero valley

magnetization can arise from either a finite population imbalance between the valleys (i.e. a net

valley polarization) or a distribution difference between them without a population imbalance [178].

Such valley magnetization forms the basis for valley-based applications. Whereas the former

relaxes by intervalley scattering, the latter is largely limited by intravalley scattering. The presence

of valley contrasting Berry curvatures in 2D Dirac materials, which can couple to external

electromagnetic excitations, enables the control of valley magnetization [33], [34], [38]–[43],

[180]–[186]. Although the control of valley magnetization by circularly polarized light and by a

vertical magnetic field has now become routine [33], [34], [38]–[43], [181]–[183], [185], [186],

the development of practical valleytronic devices requires the pure electrical control of valley

magnetization. The valley magnetoelectric (ME) effect is an attractive approach for this purpose.

To realize the linear ME effect, a material has to possess broken time-reversal and spatial-

inversion symmetries [174], [187]. For an electrical conductor, time-reversal symmetry can be

broken naturally by application of a bias voltage, under which dissipation through carrier scattering

is caused by a charge current. The magnetoelectricity produced in this manner is known as the

kinematic ME effect [188], [189]. Single-layer TMDs such as MoS2, which are intrinsically non-

centrosymmetric, satisfy these symmetry requirements with finite doping and biasing. The

characteristic three-fold rotational symmetry of these materials, however, renders the relevant

Page 112: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

91

valley ME susceptibilities zero [187], [190]. In this work, we apply a uniaxial stress to lower the

crystal symmetry and demonstrate the generation of steady-state out-of-plane valley magnetization

by an in-plane charge current in single-layer MoS2. Such an effect can be understood intuitively as

the valley analog of the Rashba-Edelstein effect for spins [189], [191], [192](Fig. 5-1(b)). In the

presence of an in-plane electric field 𝓔𝒑𝒛 , an in-plane current density 𝑱 generates an effective

magnetic field ∝ 𝓔𝒑𝒛 × 𝑱, resulting in finite out-of-plane valley magnetization 𝑴𝑽. In place of

the Rashba field (an out-of-plane built-in electric field at the semiconductor heterostructure

interfaces) in the Rashba-Edelstein effect [189], [191], [192], 𝓔𝒑𝒛 here is the in-plane piezoelectric

field that arises from the strain-induced lattice distortion and the associated ion charge polarization

(Fig. 5-1(a)), a phenomenon that has been recently demonstrated in single-layer MoS2. The effect

provides a possible valley magnetization-current conversion mechanism for magnetization

switching and detection applications [191], [192].

Page 113: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

92

Figure 5-1: Valley magnetoelectric effect in strained single-layer MoS2. (a) Piezoelectric field 𝓔𝒑𝒛

is produced (right) as a result of stress-induced distortion of the hexagonal lattice (left) and the

associated ion charge polarization. Blue and yellow balls represent Mo and S atoms, respectively.

(b) Out-of-plane valley magnetization 𝑴𝑽 is induced by an in-plane current density 𝑱 in the

presence of the piezoelectric field 𝓔𝒑𝒛. (c) Electronic band structure of n-doped single-layer MoS2

around the K and K’ valleys of the Brillouin zone. The band extrema are shifted from the K/K’

point (black dashed lines) to opposite directions by a uniaxial strain, which also do not coincide

with the extrema of the Berry curvature distribution (blue dashed lines). The Fermi level is tilted

under an in-plane bias electric field.

Page 114: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

93

5.2 Experimental setup

In our experiment, single-layer MoS2 was mechanically exfoliated from bulk crystals onto

flexible polydimethylsiloxane (PDMS) substrates. The flakes were stressed along a particular high-

symmetry axis by mechanically stretching the substrate. The strained flakes were then transferred

onto SiO2/Si substrates with pre-patterned electrodes to form field-effect transistors with Si as the

back gate. Strain was fixed for all subsequent measurements. During the process, the strain level in

the MoS2 flakes was monitored by the photoluminescence (PL) energy shift based on the strain-

induced decrease of the band gap (~ 20 nm/% strain independent of direction [56], [165]) (Fig. 5-

2(b)). The crystallographic orientation of the flakes was determined by the second-harmonic

generation (SHG) under normal incidence [56]. Figure 5-2(a) shows the intensity of the second-

harmonic component parallel to the excitation as a function of the polarization angle, the maximum

of which corresponds to the armchair direction. Polar magneto-optic Kerr rotation (KR) microscopy

at variable temperatures was employed to image the spatial distribution of the magnetization in the

device channels [184]. To this end, linearly polarized light was focused onto the device under

normal incidence and its polarization rotation (i.e. KR angle) upon reflection was detected. The

polar configuration is sensitive to out-of-plane magnetization only. To enhance the detection

sensitivity, we have used a probe wavelength close to the MoS2 A exciton resonance and a lock-in

detection method with the bias voltage being modulated at 4.11 kHz. The KR sensitivity is about

0.5 μrad/Hz1/2 (details see section 5.2.2).

Page 115: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

94

5.2.1 Device fabrication

Atomically thin flakes of MoS2 were mechanically exfoliated from synthetic bulk crystals

(2D Semiconductors) onto flexible polydimethylsiloxane (PDMS) substrates. Single-layer flakes

were identified by the combination of their optical contrast and photoluminescence (PL) spectrum.

The crystallographic orientation of each single-layer flake was determined by performing optical

second-harmonic generation (SHG) on multiple locations of the flake to ensure that it contains a

single domain. The flakes were then stressed uniaxially along a high-symmetry axis by stretching

the PDMS substrates using a mechanical translation stage. The PL spectrum was simultaneously

measured to monitor the strain level. Strain up to about 0.2 - 0.3% (corresponding to a shift in the

PL peak wavelength by 4 - 6 nm) was typically achieved in monolayer MoS2 before the flake started

to slip or break. A glass slide was then employed to support the stretched PDMS for sample transfer

Figure 5-2: Basic optical characterization in single-layer MoS2. (a) Intensity of the second-

harmonic component parallel to the excitation polarization (𝐼) as a function of the excitation

polarization angle measured from the armchair direction (𝜃) (symbols). Solid line is a fit to 𝐼 =𝐼0cos

2(3𝜃) with 𝐼0 denoting the maximum intensity. (b) Photoluminescence (PL) spectrum of

single-layer MoS2 on a flexible substrate with 0% (blue), 0.1% (cyan) and 0.2% (green) strain along

the armchair direction. Red line is the PL spectrum of the flake after being transferred onto a silicon

substrate. The spectra are normalized to the peak intensity and vertically shifted for clarity.

Page 116: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

95

onto a Si substrate with a 100-nm SiO2 layer. During the transfer, the plane of the MoS2 flake and

the Si substrate were kept parallel so that strain was best maintained. The Si substrates were pre-

patterned with electrodes (Ti 3nm/Au 30nm) by either e-beam or photo-lithography followed by

metal evaporation so that the MoS2 devices are free of resist contamination.

5.2.2 Second-harmonic generation (SHG)

To determine the crystallographic orientation of single-layer MoS2 flakes, we employed

SHG measurements based on a femtosecond Ti:Sapphire oscillator centered at 820 nm with a

repetition rate of ~ 80 MHz and a pulse duration of ~ 100 fs. The laser beam was focused onto the

sample under normal incidence using a 40x microscope objective. The reflected second-harmonic

(SH) radiation peaked at 410 nm was filtered and detected by a spectrometer equipped with a

nitrogen-cooled charge-coupled device (CCD). The excitation power was kept below 1 mW on the

sample. The crystallographic orientation of the crystal was determined from the SH intensity as a

function of the excitation polarization with respect to the sample’s high-symmetry axes. We chose

to fix the sample while varying the linear polarization of the excitation beam using a broadband

half-wave plate. The reflected SH beam was passed through the same broadband half-wave plate

and an analyzer for the measurement of the SH component parallel to the excitation polarization.

The integration time for each SH spectrum was 5 seconds with binning applied to the CCD readout.

In Fig. 5-2(a), we show the integrated SH intensity as a function of excitation polarization. A six-

fold pattern was observed as expected from D3h symmetry of the crystal. The SH maxima

correspond to the armchair direction.

Page 117: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

96

5.2.3 Photoluminescence (PL) spectroscopy

PL spectroscopy was performed on MoS2 flakes to first identify their monolayer thickness

and then to characterize the strain level. A continuous-wave solid-state laser at 532 nm (up to 50

µW on the sample) was employed as the excitation source. PL spectrum was collected with a

confocal microscope setup, filtered, and recorded by a spectrometer equipped with a CCD.

5.2.4 Magneto-optic Kerr rotation (KR) microscopy

KR microscopy was performed on devices in an optical cryostat. The schematics of KR

microscopy setup is shown in Figure 5-3 (b). A linearly polarized probe laser was tuned to a

wavelength slightly red shifted from the PL peak wavelength and focused perpendicularly onto the

devices by a 40x microscope objective. The power on the sample was 200 - 350 µW. The reflected

probe was collected by the same objective, passed through a half-wave plate, split by a Wollaston

prism, and detected by a pair of balanced photodetectors. In order to increase the signal-to-noise

ratio, an oscillating bias voltage at 4.11 kHz was applied to the devices and the KR was detected

using a lock-in amplifier with a time constant of 30 msec. The noise level was about 0.5 µrad/Hz1/2

away from the boundaries of the gold electrodes. The noise level is typically higher near the

boundaries due to distortions of the light polarization upon reflection. For two-dimensional (2D)

spatial mapping, the devices were scanned by an XY piezo-stage while the probe beam was fixed.

The KR measurement in the presence of an in-plane magnetic field was performed in a cryostat

equipped with a superconducting magnet in the Voigt geometry.

Page 118: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

97

5.3 Observation of valley magnetoelectric effect in single-layer MoS2

In the following section 5.3 and 5.4, we investigate the valley associated magnetization

without external magnetic field in single-layer MoS2 FET devices by employing Kerr rotation

microscopy. Particularly, we focus more on the magnetization induced by a distribution difference

of valley magnetic moments between two valleys instead of a population imbalance.

5.3.1 Valley magnetization in single-layer MoS2

Here we present the KR image of two single-layer MoS2 devices at 30 K, one with

intentional strain and one without in Fig. 5-4 (a) and (b). Drastic differences are observed. In the

Figure 5-3: Magneto-optic KR microscopy setup for strained single-layer MoS2. (a) Optical image

of strained MoS2 device. Black arrow indicates the direction of electrical current. (b) Schematics

of a MoS2 FET with bias voltage (Vsd) applied on the source–drain electrodes and gate voltage

(Vg) applied through the Si/SiO2 substrate. For a strained single-layer MoS2 FET, a longitudinal

electrical current density J (black arrow) is optimally flowed in the transverse direction of the

strained-induced piezoelectric field Epz (orange arrow). The valley magnetoelectric effect is

detected by focusing a linearly polarized probe beam onto the device under normal incidence

and measuring the Kerr rotation angle θKR of the reflected beam.

Page 119: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

98

unstrained device (Fig. 5-4(a)), the KR is present only near the channel edges and is of opposite

sign on the two edges. When the current direction is reversed, the KR switches sign on the two

edges. (Detailed studies on the gate and bias dependences are shown in Section 5.4.7) Such a KR

signal is originated from valley polarization accumulated on the channel edges that is driven by the

valley Hall effect (VHE), as reported by recent experiments [180], [184], [186]. On the other hand,

in the strained device (Fig. 5-4(b)), a much stronger KR signal per unit current (in Fig. 5-16) is

observed nearly uniformly over the entire device channel. The signal also changes sign under a

reversed current direction. For this particular device, the sample was strained by ~ 0.2% along the

zigzag direction and biased along the same direction.

Figure 5-4: Valley Hall effect and valley magnetoelectric effect in single-layer MoS2. Kerr

rotation image of an unstrained (a) and strained (b) single-layer MoS2 device under two opposite

bias directions (black arrows). Boundaries of the electrodes and the device channel are marked in

dashed black and green lines, respectively. The unstrained device was measured with 𝑉𝑑𝑠= 2.5 V

and 𝑉𝑔 = 0 V (𝐽 = 22 A/m). KR is observed at channel edges only. The strained device was measured

with 𝑉𝑑𝑠 = 2.5 V and 𝑉𝑔 = 20 V (𝐽 = 13 A/m). KR is observed over the entire channel.

Page 120: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

99

5.3.2 Dependences of bias current and in-plane magnetic field

We performed multiple control measurements to further examine the observed KR signal

in the bulk of strained single-layer MoS2. First, we studied the dependence of the KR signal on the

channel current density (𝐽). Figure 5-5(a) shows the KR (symbols) and the current density (solid

lines) as a function of gate voltage (𝑉𝑔), as well as bias voltage (𝑉𝑑𝑠) (inset of Fig. 5-5(a)). A typical

n-type transport behavior is seen. The observed KR is linearly proportional to 𝐽. Second, we studied

the effect in the presence of an in-plane magnetic field in the Voigt geometry. No dependence of

the KR signal on the magnetic field is observed up to ± 0.5 T (Fig. 5-5(b)). The absence of the

Hanle effect (that is, Larmor precession of spin magnetization under an in-plane magnetic field

[193]) excludes the spin nature of the observed effect in case of weak spin-orbit coupling (SOC).

It is, however, fully compatible with valley magnetization, which does not couple with an in-plane

magnetic field [148]. Third, KR was measured as a function of the probe beam’s polarization

direction. The observed negligible dependence excludes strain-induced optical birefringence as a

possible origin of the observed KR (more details see Section 5.4).

Figure 5-5: Dependences of KR angle on magnitude of bias current and in-plane magnetic field.

(a) Gate dependence of the KR at a fixed location on the strained device (symbols) and gate

dependence of channel current density 𝐽 measured with 𝑉𝑑𝑠 = 2.5 V (red solid line). Inset shows

the corresponding bias dependences with 𝑉𝑔 = 20 V. (d) KR at a fixed location on the strained

device as a function of in-plane magnetic field B. Inset illustrates the Hanle effect for spin

magnetization 𝑀𝑆 and the absence of it for valley magnetization 𝑀𝑉.

Page 121: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

100

5.3.3 Hamiltonian model for valley magnetoelectric effect in strained monolayer MoS2

To understand the microscopic mechanism of the observation, we consider the effect of

strain on the Dirac Hamiltonian of single-layer MoS2 in the k·p approximation [32], [194], [195],

𝐻0 = ℏ𝑣𝐹(𝑘𝑥𝜎𝑥 + 𝑠𝑘𝑦𝜎𝑦) +∆

2𝜎𝑧. Here ℏ and 𝑣𝐹 denote, respectively, the Planck’s constant and

the Fermi velocity, ℏ𝒌 = ℏ(𝑘𝑥�� + 𝑘𝑦��) is the crystal momentum measured from the K (s = +1)

and K’ (s = -1) point of the Brillouin zone, 𝝈 is the Pauli matrix for the sublattice index, and ∆ is

the band gap. We choose unit vector 𝑥 and �� along the armchair and zigzag direction, respectively

(�� corresponds to the out-of-plane direction). The SOC has been ignored since we only consider n-

type MoS2 and the SOC is weak in the lowest-energy conduction bands of MoS2 [31]. Under

uniform uniaxial strain along high-symmetry axes, the Hamiltonian is modified as 𝐻 ≈

ℏ(𝑣𝐹𝑥𝑘𝑥′ 𝜎𝑥 + 𝑠𝑣𝐹𝑦𝑘𝑦

′ 𝜎𝑦) +∆′

2𝜎𝑧 + 𝑠𝛼(𝑢xx − 𝑢𝑦𝑦)ℏ𝑣𝐹𝑘𝑦

′ [194], [195]. Here the crystal

momentum ℏ𝒌 is replaced by the canonical momentum ℏ𝒌′ due to a strain-induced fictitious vector

potential, and a new (third) term that does not couple to the Pauli matrix appears with 𝛼 ~ 1 and

𝑢ij denoting, respectively, a dimensionless parameter and a strain tensor element [194], [195]. The

Hamiltonian also takes into account the anisotropy in the Fermi velocity due to broken three-fold

rotational symmetry (𝑣𝐹𝑥 ≠ 𝑣𝐹𝑦) and the shift of the energy gap from ∆ to ∆′. The major effect of

these changes is the shift of the band extrema and the extrema of the Berry curvature distribution

(blue dashed lines) illustrated schematically in Fig. 5-1(c). The band extrema are shifted from the

original K/K’ point (black dashed lines) to opposite directions, which also do not coincide with the

extrema of the Berry curvature distribution. Under a bias field, a current is formed and the Fermi

surface tilts to the direction of the bias field. Although the two valleys maintain an equal population,

the modified band structure and Berry curvature distribution, in combination with the non-

equilibrium carrier distribution, lead to a net valley magnetization:

Page 122: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

101

𝑴𝑽 ≈3𝛼𝜀ℏ𝑣𝐹

𝑒𝑥𝑥Δ′ ℇ𝒑𝒛 × 𝑱. (5-1)

Here 휀 and 𝑒𝑥𝑥 denote, respectively, the dielectric constant and the piezoelectric tensor element of

single-layer MoS2, and only the leading-order term in both the piezoelectric field ℇ𝒑𝒛 and the

current density 𝑱 is considered. Similarly, the maximum absorbance difference for left- and right-

handed light (which determines the KR angle) is calculated by ignoring the excitonic effect as 𝐴+ −

𝐴− ≈ 𝑀𝑉𝐴Φ0

3Γ with Φ0 , A and Γ denoting, respectively, the magnetic flux quantum, the peak

absorbance and the width of the optical transition (See Sections 5.5 for detailed derivations).

Equation (5-1) reproduces our intuitive result for the valley ME effect in single-layer

TMDs as illustrated in Fig. 5-1(b). It is also consistent with the symmetry requirements [187], [190].

We consider the ME susceptibility 𝛼𝑖𝑗 that relates the bias electric field ℰ𝑗 to magnetization 𝑀𝑖 =

𝛼𝑖𝑗ℰ𝑗. Since single-layer MoS2 does not spontaneously break time-reversal symmetry, a charge

current under bias that causes dissipation is required for finite 𝛼𝑖𝑗 [188], [189]. For unstrained

single-layer MoS2 with D3h symmetry, the ME susceptibilities 𝛼𝑧𝑥 and 𝛼𝑧𝑦 that are relevant for the

generation of out-of-plane magnetization by an in-plane bias field all vanish [187], [190]. However,

under a uniaxial stress along the armchair or the zigzag direction, the symmetry point group is

reduced to C2v with the mirror line along the 𝑥 (armchair) axis. The 𝛼𝑧𝑦 component becomes

nonzero, and finite magnetization 𝑀𝑧 = 𝛼𝑧𝑦ℰ𝑦 emerges for a bias field with nonzero component

perpendicular to the mirror line (that is, ℰ𝑦 ≠ 0) [187], [190].

5.3.4 Dependences of the current and strain directions with respect to the crystalline axis

Our major experimental findings shown in Fig. 5-4 can also be explained by Equation (5-

1), namely, the need of finite strain to generate valley magnetization and the linear dependence of

Page 123: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

102

the KR on the amplitude and sign of the current density. To further verify the vector relation of

𝑴𝑽 ∝ 𝓔𝒑𝒛 × 𝑱, we fabricated devices in the Corbino disk geometry, which allows current along

the radial direction over a wide range of angles relative to the applied stress. Two devices are shown

in Fig. 5-6 with a uniaxial stress applied along the armchair (upper row) and zigzag axes (lower

row). The piezoelectric field 𝓔𝒑𝒛 is along the 𝑥-axis with opposite direction in the two cases (solid

orange lines). Note the effect of tensile strain along the zigzag direction is similar to compressive

strain in the armchair direction. Both devices show positive KR in one half and negative KR in the

second half of the channel (Figs. 5-6(b) and (e)). In particular, the KR signal vanishes along the

orange line and reaches maximum approximately perpendicular to it. The observed patterns agree

reasonably well with the simulation result based on 𝑀𝑉 ∝ �� ∙ (𝓔𝒑𝒛 × 𝑱) for both devices (Figs. 5-

6(c) and (f)). (Details of simulation are provided in Section 5.4.4.) Deviations of the experimental

results from the predictions may originate from the presence of inhomogeneous strain in the

samples. Inhomogeneous doping and optical responses associated with it, as well as strain-induced

conductivity anisotropy may also play a role. However, the observed images with one-fold

symmetry cannot be explained by either the strain-induced conductivity anisotropy or the trigonal

warping effect of the band structure [185]. (See Sections 5.4.5 and 5.4.6 for more details.)

Page 124: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

103

Moreover, Eqn. (5-1) predicts a KR signal level that is comparable to the value observed

in experiment as discussed below. The detailed analysis that relates the KR angle 𝜃𝐾𝑅 to the optical

properties of the sample has been discussed in Supplementary Section 6 of Ref.[184]. In short, the

KR angle is proportional to the difference in absorbance of left- and right-handed light at the probe

energy [184]: 𝜃𝐾𝑅 = 𝐼𝑚[𝛽𝑙𝑜𝑐(𝐴+ − 𝐴−)] . Here Im[…] denotes the imaginary part and 𝛽𝑙𝑜𝑐

describes the combined properties of the substrate, oxide layer and the local field factors that

influence wave propagation in the multilayer structure and is on the order of unity (𝛽𝑙𝑜𝑐 ~ 1). For

a typical measurement, a KR angle of ~ 100 𝜇𝑟𝑎𝑑 was observed at 10 K in MoS2 strained by ~ 0.5%

and in the presence of a current density of 𝐽 ~ 10 A/m. Using Eqn. (5-1) and the experimental

parameters we estimate the circular dichroism to be 𝐴+ − 𝐴−~ (2 − 3) × 10−5 (KR angle 𝜃 ~ 20

Figure 5-6: Dependence on current direction of the valley magnetoelectric effect. a, Piezoelectric

field 𝓔𝒑𝒛 is produced along the armchair direction when single-layer MoS2 is stressed along the

armchair direction; b, KR image of sample (a) measured in the Corbino disk geometry with 𝑉𝑑𝑠 =

1 V and 𝑉𝑔 = 25 V (𝐽 = 3.9 A/m); c, Predicted spatial distribution of valley magnetization by Eqn.

(5-1). The direction of 𝓔𝒑𝒛 is shown in orange lines. Boundaries of the electrodes and the device

channel are marked in dashed black and green lines, respectively. d to f are the same as a to c with

single-layer MoS2 stressed along the zigzag direction. The KR image was measured under 𝑉𝑑𝑠 =

6.9 V and 𝑉𝑔 = 20 V (𝐽 = 2.2 A/m).

Page 125: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

104

– 30 𝜇𝑟𝑎𝑑) if we take Γ ~ 3.5 meV (thermal broadening), Δ′ ~ 2 eV, 𝑣𝐹 ≈ 0.6 × 106 ms-1 (ref.[32]),

𝛼 ~ 1 (ref.[194], [195]) and peak absorbance 𝐴 ~ 0.3 – 0.5 (in Fig. 5-18). We also estimate the

valley magnetization to be 𝑀𝑉 ~ 4 × 10−11 Amp. It corresponds to a volumetric converse

magnetoelectric coefficient 𝛼𝑧𝑦 ~ 𝜇0𝑀𝑉/ℇ𝑑𝑐𝑡 ~0.5 ps/m. Here 𝜇0 and 𝑡 (≈ 0.67 nm) denote the

vacuum permeability and the approximate thickness of monolayer MoS2, respectively. We thus

conclude that out-of-plane valley magnetization has been generated and observed by KR in strained

single-layer MoS2 through the valley ME effect.

5.3.5 Temperature dependence of valley magnetoelectric effect

Finally, we examine the temperature dependence of the effect. The gate dependence of the

two-point square conductance 𝜎 (Fig. 5-7(a)) and the KR (Fig. 5-7(b)) from the same device is

shown side by side at several selected temperatures. A clear metal-insulator crossover is observed

from the transport data around 𝑉𝑔 ≈ 12 V. The device channel shows a metallic behavior above

this gate voltage (that is, 𝜎 decreases with increasing temperature), and an insulating behavior

below it (that is, 𝜎 increases with temperature). The observed square conductance at the crossover

(𝜎 ≈ 20 μS) is smaller than the Ioffe-Regel criterion (𝜎 ≈ 40 μS) [196], which is likely due to an

underestimated experimental value by the two-point method with Schottky contacts. The KR of the

device under fixed strain shows a similar temperature dependence, as expected from 𝑴𝑽 ∝

𝓔𝒑𝒛 × 𝑱. The detailed temperature dependences of the two quantities, however, are not fully

identical since the parameters involved in these measurements (such as the optical conductivity at

the probe wavelength, the optical resonance width and the Schottky barrier height) are also

temperature dependent. Detailed studies of the temperature dependences are beyond the scope of

this work.

Page 126: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

105

However, we draw attention to an important finding here: valley ME effect persists at room

temperature. The gate dependence of the valley magnetization at room temperature is shown in the

inset of Fig. 5-7(b), and the spatial image and bias dependence in Fig. 5-8. The observation of

electric-field induced valley magnetization at room temperature opens up the possibility of valley-

based applications, such as magnetic switching [191], [192] and flexible devices [99], [197], by

engineering the Berry curvature effects.

Figure 5-7: Temperature dependence of the valley magnetoelectric effect. Gate dependence of the

two-point conductance (a) and KR at a fixed location on a strained device (b) measured with 𝑉𝑑𝑠 = 2 V at four selected temperatures 10, 50, 90, and 110 K. Inset of b displays the gate dependence

of the KR (symbols) and conductance (red solid lines) at room temperature.

Page 127: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

106

5.4 Extended support experiments for the observed valley magnetoelectricity

The nonzero Kerr rotation in the polar geometry observed in this experiment has been

interpreted as a result of out-of-plane magnetization (either near the channel edges in the VHE or

throughout the entire device channel in the ME effect). As we have argued in the main text and

supported by the in-plane magnetic field dependence (Fig. 5-5(b)), the magnetization is the valley

magnetization. We have also studied the dependence of the observed effect on strain, bias current,

and the direction of the current density with respect to the piezoelectric field through the Corbino

disk geometry. All of these results are consistent with the microscopic model of valley

magnetoelectricity. On the other hand, other effects could also give rise to the observed KR signal.

Figure 5-8: Valley ME effect at room temperature. a, Kerr rotation image of the device shown in

Fig. 5-6(b) at room temperature measured with Vds = 2.1 V and Vg = 20 V. b, The background noise

level measured with Vds = 0 and Vg = 0. c, Bias dependence of the KR at a fixed location of the

device channel (symbols) and the corresponding current density (solid red line) measured with Vg

= 20 V.

Page 128: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

107

We performed several control experiments on both strained and unstrained samples to rule out these

potential artifacts. These studies include the dependence of the effect on the probe wavelength,

probe polarization, sample temperature, bias and gate voltage, and gate modulation. The bias and

gate dependences, as well as the temperature dependence, have been discussed in the section 5.3

(Figs. 5-4, 5-5, 5-6 and 5-7). In this section, we present studies on the potential effects from sample

birefringence, sample conductivity anisotropy, and source-gate coupling. We also present results

on the photon energy dependence of the observed KR signal.

5.4.1 Dependence of Kerr rotation on probe polarization

Anisotropy in the equilibrium optical response due to strain can give rise to KR although

such an effect is not expected to be bias dependent and to be detected using the lock-in method

with the bias voltage being modulated. We examined the effect of strain-induced birefringence on

the KR by measuring the KR as a function of the probe beam’s incident polarization. Strain-induced

Figure 5-9: Kerr rotation as a function of probe light polarization. KR was measured on the

strained device in Corbino disk geometry shown in Fig. 5-6(b) at two locations of the device

channel corresponding to two different bias fields (red and blue lines). The linear polarization of

the probe beam was varied by a half-wave plate and the sample was fixed.

KR

rad

)

Page 129: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

108

birefringence would strongly depend on the relative angle between the light polarization and the

applied stress direction. In Fig. 5-9, we show the KR measured on a strained monolayer MoS2

device under two bias directions as a function of the probe beam polarization. No dependence is

found, confirming that strain-induced birefringence has a negligible effect on the measured KR.

5.4.2 Dependence of Kerr rotation angle on probe photon energy

The KR angle measures the left- and right-handed absorbance difference at the probe

photon energy (see sections 5.5.4 and 5.5.5). As a result the KR signal is expected to exhibit

significant photon energy dependence near the A exciton resonance. To verify this we have

measured the dependence of the KR signal on the probe photon energy, tuned by varying the base

temperature of our diode laser (range is thus limited). Figure 5-10 shows the dependence, which

exhibits a significant change near the A exciton resonance. The response is similar to Ref. [193].

Under this condition our probe is sensitive to the presence of a small valley magnetization.

Figure 5-10: Dependence of KR on probe photon energy. Photoluminescence spectrum shows the

A exciton resonance in single-layer MoS2. The peak energy is marked by the dashed line. KR was

measured as a function of probe photon energy, the range of which is limited by the tuning range

of our laser.

0

5000

10000

15000

PL r

eson

ance

(a.u

.)

1.85 1.90 1.95 2.00-40

-20

0

20

40

60

80

KR

(m

rad

)

KR

(m

rad

)

Photon energy (eV)

Page 130: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

109

5.4.3 Kerr rotation under pure gate modulation

We evaluated the potential contribution of the source-gate coupling to the KR by measuring

the KR under a pure gate modulation. Instead of applying a modulating bias source-drain voltage

as in the original measurement, we applied a modulating gate voltage at 2.5 V peak-to-peak under

zero bias. The observed KR image (Fig. 5-11) shows negligible signal over the entire device

channel, suggesting that the source-gate coupling effect is negligible in our measurements. Fig. 5-

11 shows the result for the same device as in Fig. 5-4(a) without intentional strain. Similar results

were observed for strained devices.

5.4.4 Simulation of valley ME effect in Corbino disk geometry

To study the valley ME effect as a function of current relative to the stress direction, we

employed devices in the shape of a Corbino disk. The KR image from experiment and simulation

for two devices stressed along two different high-symmetry axes is shown in Fig. 5-6. Here we

Figure 5-11: Kerr rotation image under gate modulation. KR image of the unstrained device shown

in Fig. 5-4(a) measured with bias voltage Vds = 0 and gate voltage Vg modulated at 4.11 kHz with

a peak-to-peak amplitude of 2.5 V. Black and green dashed lines indicate the electrode and sample

boundaries, respectively. No identifiable signal can be observed anywhere on the device channel.

2 µm

KR (µrad)

-120 1200

Page 131: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

110

show the details of the simulation. Figure 5-12(a) is the optical microscope image of one of the

devices with electrodes in gold, Si substrate in grey, and monolayer MoS2 in light blue. To find the

local bias electric field (and current density assuming an isotropic conductivity), we numerically

solved the Laplace equation with the boundary conditions given by the device geometry and known

electric potential on the electrodes using the Comsol package. Figures 5-12(b) and 3c show,

respectively, the electric potential and electric field amplitude (√ℰ𝑥′2 + ℰ𝑦′

2) distribution around

the Corbino disk when the inner and outer electrode potential are set at 5 V and 0 V, respectively.

The axes 𝑥′ and 𝑦′ are defined as shown in Fig. 5-12. The black arrows in Fig. 5-12(c) show the

direction and amplitude of the local bias electric field, which is mostly radial except near the edges

of the outer electrode. The result confirms the intuitive expectation that away from the edges, the

field distribution is radially uniform. Valley magnetization on the device channel was then

simulated using 𝑀𝑧 = 𝛼𝑧𝑦ℰ𝑦 by setting 𝛼𝑧𝑦 to be a constant and converting the calculated {ℰ𝑥′,

ℰ𝑦′} to {ℰ𝑥 , ℰ𝑦} in the basis of the high-symmetry axes of the crystal. The crystallographic

orientation of the sample was inferred from the second-harmonic generation (SHG) measurement.

The simulated spatial distribution of the valley magnetization is shown in Figs. 5-6 (c) and (f).

Page 132: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

111

5.4.5 Effects of conductivity anisotropy

In Fig. 5-13 we show the simulation result for the spatial distribution of current density in

the Corbino disk devices by including strain-induced anisotropy in conductivity. We have assumed

that the conductivity is larger perpendicular to the strain direction. In the first set of simulation we

have assumed a small conductivity anisotropy (0.5%) and in the second set we have greatly

exaggerated the anisotropy to 50% in order to make the effect clearly seen. For comparison, we

have also included in the figure results of our experiment and simulation based on our model of

Figure 5-12: Electrostatic simulation of Corbino disk devices. a, Microscope image of a

monolayer MoS2 device with Corbino disk electrodes: electrodes in gold, Si substrate in grey and

monolayer MoS2 in light blue. b, c, Contour plot of the simulated electric potential (b) and electric-

field strength (c) of the device. Inner and outer electrode are fixed at 5 and 0 V, respectively. Black

arrows in c indicate the local bias field direction (arrow direction) and amplitude (arrow length).

b c

x' (µm)

y'(µ

m)

Electric potential (V)

2

1.5

1

0.5

0

Field strength (V/µm)

x' (µm)

y'(µ

m)

a

5 µm

Page 133: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

112

valley ME effect (Fig. 5-6). The conductivity anisotropy cannot explain the experimental

observations. In Table S1 we compare their key differences.

Furthermore, we have performed two additional control experiments to support the above

conclusion. First, we note that any model that only concerns the magnitude of the conductivity will

not be able to explain the experimental observations since the effect also depends on the direction

Table 5-1: Comparison of the KR measurements and the model based on conductivity anisotropy

in the Corbino disk devices.

Experiment: KR near A

exciton resonance

Model: conductivity

anisotropy (larger

conductivity ⊥ strain)

Spatial distribution 1-fold symmetry 2-fold symmetry

Strain direction

- Armchair

- Zigzag

Maximum KR direction

- Zigzag

- Zigzag

Maximum current direction

- Zigzag

- Armchair

Figure 5-13: Simulation on Corbino disk devices. a, Piezoelectric field ℇ𝑝𝑧 is produced along the

armchair direction when single-layer MoS2 is stretched along the armchair direction. b, KR image

of sample (a) measured in the Corbino disk geometry with 𝑉𝑑𝑐 = 1 V and 𝑉𝑔 = 25 V. c, Predicted

spatial distribution of valley magnetization based on the valley ME effect. The direction of ℇ𝑝𝑧 is

shown in orange lines. Boundaries of the electrodes and the device channel are marked in dashed

black and green lines, respectively. d, Simulation of current density J including the conductivity

anisotropy induced by strain (0.5%, left column; 50%, right column). e to h are the same as a to d

with single-layer MoS2 stretched along the zigzag direction. The KR image was measured under

𝑉𝑑𝑐 = 6.9 V and 𝑉𝑔 = 20 V.

Page 134: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

113

of the current. This has been seen above in Table 5-1 in the 1-fold symmetry of the KR image. This

can also be seen in Fig. 5-14, where the KR changes sign when the current direction is reversed.

Second, we performed KR mapping of the Corbino disk devices with different probe polarization

orientations with respect to the strain direction and obtained identical results (also see Fig. 5-9). If

the dominant mechanism were strain-induced conductivity anisotropy, the result would be

dependent on the polarization orientation with respect to the strain direction.

Based on these evidences we summarize that strain-induced conductivity anisotropy cannot

explain our experimental results using the Corbino disk geometry. However, it is conceivable that

the strain-induced conductivity anisotropy contributes to the experimental observation. It may, for

example, distort the ideal KR image (Fig. 5-6 or Fig. 5-13) in addition to doping and strain

inhomogeneities that are present in our devices.

5.4.6 Comparison to the effect of trigonal warping of the band structure

In Table 5-2 we compare the major differences between the effect observed in this work

and the effect reported in Ref. [185]. Ref. [185] reported an observation of polarized

Figure 5-14: KR mapping of Corbino disk device under reversed bias. KR image with current

flowing radially outward (a) and inward (b). The KR signal flips sign under reversed bias.

a b

200 -200 0

KR (µrad)

Page 135: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

114

electroluminescence in forward biased p-n junctions in few-layer WSe2. The effect was modeled

as different electron-hole overlap at the K and K’ valleys under a bias field due to trigonal warping.

In our experiment, we observed KR in biased and strained monolayer MoS2 and modeled the effect

as strain- and current-induced valley magnetization, in analogue to the Rashba-Edelstein effect in

semiconductor quantum wells. The effect observed in this work arises from a completely different

mechanism and cannot be explained by the trigonal warping effect of the band structure.

Science, 344, 725 (2014)

Measurement: degree of circular polarization of

electroluminescence

Model: trigonal warping

Our work

Measurement: Kerr rotation near the A

exciton resonance

Model: strain- and current-induced valley

magnetization

(p) 3-fold symmetry as a function of current

direction: ∝ 𝑠𝑖𝑛3𝜑 (𝜑 is the azimuthal angle)

(o & p) 1-fold symmetry as a function of

current direction: ∝ 𝑠𝑖𝑛𝜑

(p) Peaked for current along the armchair

direction

(o & p) Peaked for current along the zigzag

direction

(o) Signal shows no to weak dependence on the

amplitude of the current.

(o & p) Signal depends linearly on the

amplitude of the current.

5.4.7 Valley Hall effect in single-layer MoS2

We show extra measurement results of the valley Hall effect (VHE) at 30 K in the

unstrained single-layer MoS2 device (Fig. 5-4(a)). Figure 5-15(a) (same as Fig. 5-4(a)) shows the

two-dimensional (2D) Kerr rotation (KR) image of the device. A clear KR signal is present only

near the two channel edges, as expected for the VHE [184]. An optical microscope image of the

device is shown in the lower inset of Fig. 5-15(b). The gate dependence of the two-point

conductivity (Fig. 5-15(b)) shows a typical n-type behavior. The upper inset shows the bias (Vds)

Table 5-2: Comparison of the effect reported in Ref. [185] and our work

(p-predicted property; o-observed property)

Page 136: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

115

dependence of the current (I) with Vg = 20 V. The two-point mobility of the device was estimated

to be about 120 cm2/Vs. The gate dependence of the KR near the two edges (red and blue lines in

Fig. 5-15(c)) resembles that of the channel conductivity or the carrier density. Negligible KR signal

was measured at the center of the device channel (black line) for all gate voltages. In Fig. 5-15(d),

KR near one of the edges was shown as a function of the bias voltage Vds at several selected gate

voltages.

Figure 5-15: Valley Hall effect in single-layer MoS2 at 30 K. a, Kerr rotation microscope image

of an unstrained monolayer MoS2 device under Vsd = 2.5 V and Vg = 0 V (J = 22 A/m). The

boundaries of the electrodes and the device channel are shown in black and green dashed lines,

respectively. b, Conductivity as a function of gate voltage (Vg). Inset: current as a function of bias

voltage (Vds) with Vg = 20 V and a microscope image of the device. c, Gate dependence of the KR

at three different locations on the device channel indicated by circles in (a) measured with Vds = 2.5

V. Red, black and blue line correspond to KR at the lower edge, center, and top edge of the channel,

respectively. d, Bias dependence of the KR at selected gate voltages near the lower edge of the

channel.

b

c

a

2 µm

KR (µrad)

-120 1200

d

-20 -10 0 10 20

0

50

100

0 1 2 30

20

40

(S

)

Vg (V)

I (

A)

Vds

(V)

0 1 2 3

0

50

100

150

0 V

-20 V

KR

(ra

d)

Vds

(V)

Vg = 20 V

-20 -10 0 10 20-100

-50

0

50

100

KR

(ra

d)

Vg (V)

Page 137: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

116

For comparison we also show in Fig. 5-16 the images of normalized KR signals by the

channel current density of unstrained (left) and strained (right) monolayer MoS2. The KR signal

from the VHE is much smaller than that from the magnetoelectric (ME) effect under the same

current density.

5.5 Microscopic model of valley magnetization in strained and biased MoS2

5.5.1 Derivation of the piezoelectric field in single-layer MoS2 under a uniaxial stress

In monolayer MoS2, which lacks inversion symmetry, uniaxial stress results in a nonzero

piezoelectric field 𝓔𝒑𝒛 as a result of strain-induced lattice distortion and the associated ion charge

polarization [99], [197]. In this section, we show the relation between strain and 𝓔𝒑𝒛 by considering

crystal’s symmetry.

The linear piezoelectric effect can be described as the first-order coupling between surface

polarization vector 𝑷 and mechanical strain tensor 𝑢𝑗𝑘: 𝑃𝑖 = 𝑒𝑖𝑗𝑘𝑢𝑗𝑘 with 𝑒𝑖𝑗𝑘 denoting the third-

Figure 5-16: Image of normalized KR signal by the channel current density for unstrained (left)

and strained (right) monolayer MoS2.

2 µm

KR per current density (mrad·m/A)

-250

250 0

Page 138: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

117

rank piezoelectric tensor 34. For monolayer MoS2, which belongs to D3h point group, 𝑒𝑖𝑗𝑘 has only

one nonzero independent coefficient. By employing the Voigt notation to reduce the number of

indices, the piezoelectric tensor can be written as [99], [197]–[199]

[𝑒] = (𝑒𝑥𝑥 −𝑒𝑥𝑥 00 0 00 0 0

0 0 00 0 −𝑒𝑥𝑥0 0 0

), (5-2)

where x, y and z are along the armchair, zigzag, and out-of-plane direction. By employing [𝑢𝑇] =

(𝑢𝑥𝑥, 𝑢𝑦𝑦, 𝑢𝑧𝑧, 2𝑢𝑦𝑧, 2𝑢𝑥𝑧, 2𝑢𝑥𝑦), the resulting surface polarization becomes

[𝑃] = [𝑒𝑢] = 𝑒𝑥𝑥 (

𝑢𝑥𝑥 − 𝑢𝑦𝑦−2𝑢𝑥𝑦0

). (5-3)

In our devices, stress 𝑡𝑘𝑙 was applied to monolayer flakes to induce strain 𝑢𝑖𝑗 = 𝑆𝑖𝑗𝑘𝑙𝑡𝑘𝑙,

where 𝑆𝑖𝑗𝑘𝑙 is the symmetry-constrained compliance. For D3h point group, it has the following form

[198]:

[𝑆] =

(

𝑆11 𝑆12 𝑆13𝑆12 𝑆11 𝑆13𝑆13 𝑆13 𝑆33

0 0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0 0

𝑆44 0 0 0 𝑆44 0 0 0 2(𝑆11 − 𝑆12))

. (5-4)

When uniaxial stress is applied along the x-axis (armchair direction), i.e. only 𝑡𝑥𝑥 ≠ 0, the

resulting strain tensor becomes [𝑢𝑇] = 𝑡𝑥𝑥(𝑆11, 𝑆12, 𝑆13, 0,0,0), giving rise to a piezoelectric

polarization 𝑷 = 𝑒𝑥𝑥𝑡𝑥𝑥(𝑆11 − 𝑆12, 0,0) pointing in the armchair direction. When stress is

applied along the y-axis (zigzag direction), however, only tyy ≠ 0 , which induces a piezoelectric

field 𝑷 = 𝑒𝑥𝑥𝑡𝑦𝑦(𝑆12 − 𝑆11, 0,0) of opposite sign.

Page 139: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

118

5.5.2 Electronic band structure of strained single-layer MoS2

The 𝑘 ∙ 𝑝 Hamiltonian of single-layer MoS2 ignoring the spin-orbit coupling (SOC) can be

written as [32], [178]

𝐻0 = ℏ𝑣𝐹(𝑘𝑥𝜎𝑥 + 𝑠𝑘𝑦𝜎𝑦) +∆

2𝜎𝑧. (5-5)

Here ℏ = ℎ/(2𝜋) is the Planck’s constant, 𝑣𝐹 is the Fermi velocity, ℏ𝒌 = ℏ(𝑘𝑥�� + 𝑘𝑦��) is the

crystal momentum counted from the K (s = +1) and K’ (s = -1) point of the Brillouin zone, 𝝈 is the

Pauli matrix accounting for the sublattice index, and ∆ is the band gap. The corresponding basis

functions for the conduction band |𝑐⟩ and the valence band |𝑣⟩ are

|𝑐⟩ = |𝑑𝑧2⟩, |𝑣⟩ =1

√2(|𝑑𝑥2−𝑦2⟩ + 𝑖𝑠|𝑑𝑥𝑦⟩), (5-6)

where |𝑑𝑧2⟩ , |𝑑𝑥2−𝑦2⟩ and |𝑑𝑥𝑦⟩ are the d-orbitals of the Mo atoms. The valence band

wavefunction at the two valleys is related by the time reversal operation. The SOC is weak in the

conduction bands of MoS2 and is ignored for n-doped samples that are studied in this work. The

above simple two-band model with electron and hole symmetry is thus a good description of n-

doped single-layer MoS2.

In general, the effects of uniform uniaxial strain on the Hamiltonian include 30,31: 1)

modification of the Fermi velocity because of the broken 3-fold rotational symmetry (the Fermi

velocity along the armchair and the zigzag directions 𝑣𝐹𝑥 ≠ 𝑣𝐹𝑦 ); 2) change of the crystal

momentum ℏ𝒌 to the canonical momentum ℏ𝒌′ = ℏ𝒌 −𝑒

𝑐𝑨, where 𝑨 is a strain-induced fictitious

vector potential, e and c denote the elementary charge and speed of light in vacuum, respectively;

3) modification of the energy gap ∆→ ∆′; and 4) addition of a new valley-dependent term that does

not couple to the sublattice pseudospin Pauli matrices 𝑠𝛼(𝑢𝑥𝑥 − 𝑢𝑦𝑦)ℏ𝑣𝐹𝑘𝑦′ . Here 𝛼 is a

Page 140: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

119

dimensionless parameter on the order of unity and 𝑢𝑖𝑗 is a strain tensor element [194], [195]. The

Hamiltonian for single-layer MoS2 strained along high-symmetry axes can be written as

𝐻 ≈ ℏ𝑣𝐹𝑥𝑘𝑥′ 𝜎𝑥 + 𝑠ℏ𝑣𝐹𝑦𝑘𝑦

′ 𝜎𝑦 +∆′

2𝜎𝑧 + 𝑠𝛼(𝑢𝑥𝑥 − 𝑢𝑦𝑦)ℏ𝑣𝐹𝑘𝑦

′ , (5-7)

where higher-order terms in strain 𝑢𝑖𝑗 have been ignored. The eigen-energies of the Hamiltonian

can be found as

𝜖𝑐,𝑣(𝒌′) = ±𝜖(𝒌′) + 𝑠𝛼(𝑢𝑥𝑥 − 𝑢𝑦𝑦)ℏ𝑣𝐹𝑘𝑦′ ,

(5-8)

𝜖(𝒌′) =1

2√∆′2 + 4ℏ2(𝑣𝐹𝑥

2 𝑘𝑥′ 2 + 𝑣𝐹𝑦

2 𝑘𝑦′ 2).

Here the “+” and “-“ sign are for the energy dispersion of the conduction band (𝜖𝑐) and the valence

band (𝜖𝑣 ), respectively. Similarly we can find the eigenstates of the Hamiltonian. The Berry

curvature of the bands, defined through the periodic part of the Bloch wavefunction 𝑢(𝒌) as

𝛀(𝒌) = 𝛁𝒌 × ⟨𝑢(𝒌)|𝑖𝛁𝒌|𝑢(𝒌)⟩, can be shown for the conduction band as [31], [32]

𝛀(𝒌′) = sΩ(𝒌′)�� = −sℏ2𝑣𝐹

2∆′

4[𝜖(𝒌′)]3��. (5-9)

For relatively small strain levels, we can approximate 𝑣𝐹𝑥 ≈ 𝑣𝐹𝑦 ≈ 𝑣𝐹 and the second line

of Eq. (5-8) becomes 𝜖(𝒌′) ≈∆′

2+ℏ2𝑘′

2

2𝑚, where 𝑚 =

∆′

2𝑣𝐹2 is the band mass. With this approximation

the energy dispersion can be expressed as

𝜖𝑐(𝒌′) ≈∆′

2+ℏ2(𝑘′+𝑠𝑘0)

2

2𝑚, 𝜖𝑣(𝒌′) ≈ −

∆′

2−ℏ2(𝑘′−𝑠𝑘0)

2

2𝑚 , (5-10)

where the shift of the band extrema with respect to 𝒌′ = 0 is given by 𝒌𝟎 = 𝑘0�� ≡𝛼(𝑢𝑥𝑥−𝑢𝑦𝑦)Δ

2ℏ𝑣𝐹��.

On the other hand, the extrema of the Berry curvature distribution is not shifted from 𝒌′ = 0 since

it is a property of the wavefunctions and the wavefunctions are not affected by the last term in the

Page 141: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

120

Hamiltonian that does not couple to the sublattice pseudospin Pauli matrices [190]. As we show

below that this leads to nonzero valley magnetization in the presence of a bias current. We note that

the strained Hamiltonian 𝐻 of Eq. (5-7) is very similar to that for a strained topological crystalline

insulator analyzed in Ref. [190]. The derived results of Eq. 5-8 and 5-9 are also consistent with the

results for the strained topological crystalline insulator [190].

5.5.3 Valley magnetization

It has been shown by Xiao et al. that the valley magnetization 𝑴 of an electron gas in a 2D

Dirac material (n-doped MoS2 in this case) can be calculated as [178]

𝑴 ≈2𝑒

ℏ∫

𝑑𝒌

(2𝜋)2𝑓(𝒌)(

Δ

2+ 𝜇)𝛀(𝒌). (5-11)

Here the integration is over the entire Brillouin zone, 𝑓(𝒌) is the electron distribution function, and

𝜇 is the chemical potential counted from the bottom of the conduction band. Note that 𝑴 is out-of-

plane since the Berry curvature is out-of-plane. We consider strained monolayer MoS2 under a

constant bias field ℇ𝒅𝒄 with both the sample and bias field being spatially uniform. In the absence

of the bias field, the electrons at temperature T are described by the Fermi-Dirac distribution 𝑓𝑐0(𝐸,

Δ

2+ 𝜇, 𝑇) with 𝐸 = 𝜖𝑐(𝒌). It can be shown that the valley magnetization vanishes since the Berry

curvature has opposite values in the two valleys (Eq. 5-9). Under the bias field, a non-equilibrium

correction to the distribution function emerges and can be approximated as

𝑓𝑐1 =𝑒𝜏

ℏℇ𝒅𝒄 ⋅ 𝛁𝒌𝑓𝑐0, (5-12)

where 𝜏 is the electron scattering time. This correction can be obtained under the relaxation time

approximation from the Boltzmann equation for the distribution of electrons.

Page 142: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

121

Below we calculate Eq. 5-11 under the non-equilibrium carrier distribution. The integration

can be broken into the contribution from the two valleys explicitly. We note that the energy

dispersion is centered at momentum 𝒌′ = −𝑠𝒌𝟎, whereas the Berry curvature is centered at 𝒌′ =

0. We change variables so that the energy is centered at 0 momentum and the Berry curvature at

𝑠𝒌𝟎 for simplicity of the calculation. Equation 5-11 now can be rewritten for strained MoS2 as

𝑀 ≈𝑒

ℏΔ′ ∫

𝑑𝒌

(2𝜋)2𝑓𝑐1(𝒌)[Ω(𝒌 + 𝒌𝟎) − Ω(𝒌 − 𝒌𝟎)], (5-13)

where the integration is for one valley only and 𝑓𝑐1(𝒌) = 𝑓𝑐1(𝐸,Δ′

2+ 𝜇, 𝑇) with 𝐸 ≈

∆′

2+ℏ2𝑘2

2𝑚. In

Eq. 5-13 we have also used the fact that the band gap energy (~ 2 eV) far exceeds the chemical

potential (~ 10’s meV) so that Δ′

2+ 𝜇 ≈

Δ′

2. Next, we plug Eq. 5-12 into Eq. 5-13 and do the

integration by parts

𝑀 = −Δ′𝑒2𝜏

ℏ2ℇ𝒅𝒄 ⋅ ∫

𝑑𝒌

(2𝜋)2𝑓𝑐0(𝒌)[𝛁𝒌Ω(𝒌 + 𝒌𝟎) − 𝛁𝒌Ω(𝒌 − 𝒌𝟎)]

(5-14)

≈12𝑒2𝜏

𝑚Δ′ℇ𝒅𝒄 ⋅ 𝒌𝟎 ∫

𝑑𝒌

(2𝜋)2𝑓𝑐0(𝒌).

We have used 𝜕Ω

𝜕𝜖≈ −6

ℏ2

𝑚Δ′2. Expressing the remaining integration through the chemical

potential, ∫𝑑𝒌

(2𝜋)2𝑓𝑐0(𝒌) =

𝑚𝜇

2𝜋ℏ2, we finally obtain

𝑀 = 12𝑒

ℎ(𝜇

Δ′)(𝑒𝜏ℇ𝒅𝒄)⋅(ℏ𝒌𝟎)

𝑚=

6

Φ0(𝜇

Δ′)(𝑒𝜏ℇ𝒅𝒄)⋅(ℏ𝒌𝟎)

𝑚. (5-15)

The second half of the equation expresses the valley magnetization in terms of the magnetic

flux quantum Φ0 =ℎ

2𝑒 . The 2D valley magnetization has the unit of [energy]/[magnetic flux

quantum] = Ampere. It is out-of-plane. Its magnitude is determined by ℇ𝒅𝒄 ⋅ 𝒌𝟎. Since 𝒌𝟎 = 𝑘0��,

Page 143: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

122

it’s most efficient to generate valley magnetization by biasing the sample in the same direction as

𝒌𝟎, i.e. along the zigzag direction.

We further note that the shift 𝒌𝟎 can also be expressed in terms of the piezoelectric field

ℇ𝒑𝒛 [194], [195], which is related to strain as εℇ𝒑𝒛 = [𝑒𝑢] with 휀 and [𝑒] denoting the dielectric

constant and the piezoelectric tensor. For uniform uniaxial in-plane strain, 𝒌𝟎 and ℇ𝒑𝒛 are related

as

𝒌𝟎 =𝛼𝜀Δ′

2𝑒𝑥𝑥ℏ𝑣𝐹�� × ℇ𝒑𝒛. (5-16)

The piezoelectric field is along the armchair direction, i.e. ℇ𝒑𝒛 = ℇ𝑝𝑧𝑥 for strain along the high

symmetry axes. We rewrite Eq. 5-15 as

𝑴 =3𝛼𝜀ℏ𝑣𝐹

𝑒𝑥𝑥Δ′ ℇ𝒑𝒛 × 𝑱, (5-17)

where 𝑱 is the current density and defined by the bias field, 𝑱 =2𝜇

𝜋ℏ2 𝑒2𝜏ℇ𝒅𝒄. We have thus obtained

the intuitive result of the main text that 𝑴 ∝ ℇ𝒑𝒛 × 𝑱. Although there is no population imbalance

between the two valleys, the net valley magnetization is nonzero in strained and biased monolayer

MoS2 due to the distinct carrier configuration of the two valleys. From the symmetry point of view,

unstrained single-layer MoS2 has D3h symmetry. The uniaxial in-plane strain along high symmetry

axes breaks the rotational symmetry and leaves only one mirror symmetry line along the armchair

direction. The resultant piezoelectric field is along the symmetry line. A net valley magnetization

emerges as long as there is a nonzero current component perpendicular to the symmetry line.

Our results can be intuitively understood based on Fig. 5-17. Figure 5-17(a) shows the

conduction bands (orange) and valence bands (blue) of unstrained monolayer MoS2 at the K and

K’ valley. The Fermi surface tilts at each valley under a bias current, but the extrema of both the

Page 144: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

123

bands and the Berry curvature distribution are centered at the K and K’ point (black dashed lines).

The contribution to the valley magnetization from the two valleys cancels exactly since the Berry

curvature is of opposite sign at the two valleys. When monolayer MoS2 is strained, both the bands

and the Berry curvature distribution shift either closer to or further away from each other in

momentum (depending on whether it’s tensile or compressive strain). The dashed blue lines

illustrate the extrema of the Berry curvature distribution, which do not coincide with the band

extrema. The contribution to the valley magnetization from the two valleys no longer cancels

although the carrier population of the two valleys remains equal.

Figure 5-17: Band structure of monolayer MoS2 at the K and K’ point. The shaded regions

illustrate the occupied states in the conduction band. a, Unstrained monolayer MoS2 under a finite

bias field, giving rise to a tilted Fermi surface. b, Strained monolayer MoS2 under bias. Net valley

magnetization is generated in this configuration. The dashed black lines indicate the K/K’ point

and dashed blue lines indicate the extrema of the Berry curvature distribution.

Page 145: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

124

5.5.4 Circular dichroism

A material’s response to an oscillating electric field at angular frequency 𝜔, ℇ𝒐𝒑(𝑡) =

𝑅𝑒{ℇ𝒐𝒑𝑒𝑖𝜔𝑡}, can be divided into interband and intraband transitions. For photon energies ℏ𝜔 ~ Δ,

the latter contribution is insignificant and can be ignored. In monolayer MoS2 the interband

transitions are modified by the strong Coulomb interactions to form excitonic states. For simplicity,

below we analyze circular dichroism in strained and biased monolayer MoS2 by ignoring the

excitonic interactions. The optical conductivity tensor element (𝜎𝛼𝛽) for interband transitions under

a two-band approximation can be evaluated as

𝜎𝛼𝛽(𝜔) =𝑖ℏ𝑒2

𝑚2Δ∫

𝑑𝒌

(2𝜋)2(𝑓𝑣 − 𝑓𝑐)𝑝𝑐𝑣,𝛼𝑝𝑣𝑐,𝛽𝛿(𝜖𝑐 − 𝜖𝑣 − ℏ𝜔). (5-18)

Here the integration is over the entire Brillouin zone, 𝑓𝑐 and 𝑓𝑣 (≈ 1) are the conduction band and

valence band carrier distribution functions, respectively, 𝒑𝒄𝒗 = ⟨𝑐|𝒑|𝑣⟩ is the transition matrix, and

𝛿(𝑥) is the Dirac delta function. It has been shown by Xiao et al. [32] that the transition matrix

element near the K (s = +1) and K’ (s = -1) point for left-handed (+) and right-handed (-) light can

be expressed as

|𝒑±(𝒌)|𝟐 ∝ (1 ± 𝑠

Δ

𝟐𝜖(𝒌))𝟐. (5-19)

For strained samples, the transition matrix keeps the same form with Δ → Δ′ and 𝒌 → 𝒌′

since only the wavefunctions are involved. We compute the normalized circular dichroism 𝜌(𝜔) =

𝜎+(𝜔)−𝜎−(𝜔)

𝜎+(𝜔)+𝜎−(𝜔) [or equivalently, 𝜌(𝜔) =

𝐴+(𝜔)−𝐴−(𝜔)

𝐴+(𝜔)+𝐴−(𝜔), defined through the absorbance for the left- and

right-handed light 𝐴+(𝜔) and 𝐴−(𝜔), respectively) following the same approach as for valley

magnetization:

𝜎+(𝜔) − 𝜎−(𝜔) ≈ −𝐶4𝑒𝜏ℇ𝒅𝒄⋅𝒌𝟎

𝜋Δ′ℏ(𝜖

𝜕𝑓𝑐0

𝜕𝜖)|𝜖=(ℏ𝜔−Δ′) 2⁄

(5-20)

Page 146: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

125

𝜎+(𝜔) + 𝜎−(𝜔) ≈ 𝐶2𝑚

𝜋ℏ2Θ(ℏ𝜔 − Δ′ − 2μ),

where constant 𝐶 is a combination of the materials parameters. We note that in the above

derivations we have assumed the thermal/lifetime broadening of the energy levels is larger than the

shift of the transition energies due to strain and bias. This is justified since the momentum shift due

to strain 𝑘0 or bias 𝑒𝜏

ℏℇ𝑑𝑐 is a small fraction of the Fermi wavevector (10-2 – 10-1)√2𝑚𝜇 ℏ⁄ under

typical experimental conditions and the corresponding energy shift is <1 meV.

The circular dichroism is dependent on the photon energy. We estimate its peak value by

using ℏ𝜔 ≈ Δ′ + 2μ and (𝜖𝜕𝑓𝑐0

𝜕𝜖)|𝜖=(ℏ𝜔−Δ′) 2⁄

≃ −𝜇

Γ, where the width Γ is given by 4𝑘𝐵𝑇 if the

transition is primarily thermally broadened, or by the experimental transition linewidth if the

transition is primarily disorder broadened. The normalized circular dichroism is thus estimated as

𝜌 =2

Γ(𝜇

Δ′)(𝑒𝜏ℇ𝒅𝒄)⋅(ℏ𝒌𝟎)

𝑚= 𝑀

Φ0

3Γ∝ 𝑀. (5-21)

Our result shows that the circular dichroism is linearly proportional to the net out-of-plane valley

magnetization, which is non-vanishing in strained and biased monolayer MoS2.

5.5.5 Comparison of microscopic theory with experiment

We compare the result of the above microscopic theory (Eq. 5-21) with typical

experimental results: a KR angle 𝜃𝐾𝑅 of ~ 100 𝜇𝑟𝑎𝑑 was observed at 10 K in monolayer MoS2

strained by ~ 0.5% along one of the high symmetry axes and in the presence of a current density of

𝐽 ~ 10 A/m. We estimate the normalized circular dichroism using Eq. 5-21 to be 𝜌 ~ 5 × 10−5 if

we take Γ ~ 3.5 meV (thermal broadening at 10 K), Δ′ ~ 2 eV, 𝑣𝐹 = 0.6 × 106 ms-1 (Ref. [32]) and

𝛼 ~ 1 (Ref. [194], [195]) (note that 𝛼 for single-layer MoS2 has not been well calibrated although

it is expected to be on the order of unity). Furthermore, by combining with the low-temperature

Page 147: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

126

peak absorbance of ~ 0.3 – 0.5 (depending on the transition width of the sample) from experiment

(Fig. 5-18), we obtain 𝐴+ − 𝐴−~2− 3 × 10−5.

On the other hand, in the polar Kerr rotation microscopy as employed in this study, linearly

polarized light impinges on a sample under normal incidence and the polarization of the reflected

light is analyzed. The KR angle 𝜃𝐾𝑅 is defined as the rotation of the polarization upon reflection.

We have shown previously that the KR angle 𝜃𝐾𝑅 can be related to the absorbance difference for

left- and right-handed light through a local field factor 𝛽𝑙𝑜𝑐 due to interference of the probe in the

multiple thin film structure of our devices (MoS2 on SiO2/Si substrates) [184]:

𝜃𝐾𝑅(𝜔) = 𝐼𝑚{𝛽𝑙𝑜𝑐[𝐴+(𝜔) − 𝐴−(𝜔)]}. (5-22)

The local field factor is on the order of unity (𝛽𝑙𝑜𝑐 ~ 1). The microscopic theory thus

predicts that the KR angle 𝜃𝐾𝑅 ~ 20 – 30 𝜇𝑟𝑎𝑑. Despite of the approximations employed in the

model (for instance, ignoring the excitonic effects), it predicts a KR level that is comparable to the

measured one.

The theory also predicts valley magnetization 𝑀 ~ 4 × 10−11 Amp. It corresponds to a

volumetric converse magnetoelectric coefficient of 𝜇0𝑀/ℇ𝑑𝑐𝑡 ~ 0.5 ps/m. Here 𝜇0 and 𝑡 (≈

Figure 5-18: Optical absorbance of monolayer MoS2 near the A exciton resonance at 10 K.

Page 148: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

127

0.67 nm) denote the vacuum permeability and the approximate thickness of monolayer MoS2,

respectively. Note that the value is comparable to that of Cr2O3, one of the known single-phase

materials with large converse ME coefficients [200].

5.6 Conclusion

In this chapter, by breaking the three-fold rotational symmetry in single-layer MoS2 via a

uniaxial stress, we have demonstrated the pure electrical generation of valley magnetization in this

material, and its direct imaging by Kerr rotation microscopy. The observed out-of-plane

magnetization is independent of in-plane magnetic field, linearly proportional to the in-plane

current density, and optimized when the current is orthogonal to the strain-induced piezoelectric

field. These results are fully consistent with a theoretical model of valley magnetoelectricity driven

by Berry curvature effects. Furthermore, the effect persists at room temperature, opening

possibilities for practical valleytronic devices.

Page 149: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

Chapter 6

Summary and future perspectives

In summary, this dissertation has demonstrated the interaction of nanomechanics with

photonics and valley-based optoelectronics using a monolayer TMD semiconductor. In particular,

we have been focusing on the role of exciton resonance with valley DOF in optomechanical

coupling. In Chapter 3, we have shown the realization of a robust optical bistability when optically

pumped with moderate power near exciton resonance of a suspended WSe2 monolayer. A

photothermal mechanism provides both optical nonlinearity and internal positive feedback that are

essential prerequisites of the observed bistable exciton resonance. The presence of an external

magnetic field further allows control of the sample reflectivity by light helicity due to its internal

valley DOF.

Moreover, in Chapter 4, we have presented the first dynamical control of mechanical

motion of a suspended MoSe2 monolayer through its exciton resonance without an optical cavity

structure. The oscillating membrane can periodically modulate a photothermal force that originated

from excitonic absorbance and build up a dynamical backaction on its mechanical motion. In

addition, gate tunable optomechanical damping and anti-damping of the mechanical vibrations due

to photothermal backaction have been observed in real two-dimensional limit.

Another important aspect of this dissertation is the generation of magnetization through the

control of electrons’ valley magnetic moment, which originated from Berry phase effect. Unlike

the magnetization induced from unbalanced population between valleys, valley magnetization can

also arise from unpolarized valleys by manipulating carriers’ momentum distribution. In Chapter

5, we have reported the realization of such magnetization without valley polarization through the

joint effect of current and strain in a monolayer MoS2. By lowering the symmetry of the MoS2

Page 150: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

129

lattice with uniaxial strain, the distributions of valley magnetic moment at valleys are displaced

with respect to the corresponding valley centers. In addition, the tilting Fermi pockets at K and K’

valleys, when flowing a charge current, can enclose inequivalent magnitude of magnetic moment.

This valley-unpolarized non-equilibrium momentum distribution thus gives rise to a net

magnetization, namely, valley magnetoelectric effect. Our finding that can persist up to room

temperature implies a practical possibility for integrating the valleytronics with magnetic

information processing.

For future perspectives, it would be still interesting to investigate some unexplored aspects

from the demonstrated studies in this dissertation. In principle, together with extreme lightweight,

the strong excitonic effect in monolayer TMD semiconductor make the material as a promising

platform to study cavity-less optomechanics. In our device structure, the photothermal-based

optomechanical effect overwhelmingly dominates over the radiation pressure-based counterpart.

However, there are several limitations that prevents us from better resolving the effects from

dynamical backactions and obtaining more convincing data. The first issue is that high laser power

could not be applied due to the low thermal conductance of monolayer TMD, which would greatly

limit the induced optomechanical effect. Otherwise, the significant photothermal heating can easily

lead to the bistability shown in Chapter 3 instead of enhanced optomechanical effect. For a larger

photothermal coupling, a high symmetry-breaking force (by gate voltage) would be desired but

simultaneously a large gate voltage could lower the quality factor of mechanical mode due to the

Joule dissipation of displacement current through the oscillation membrane. In low temperature,

exciton resonance in monolayer TMD semiconductor would become sharp with great contrast. But

it also requires much higher gate voltage to pull down the suspended monolayer since the

membrane becomes very stiff at cryogenic condition. Namely, there is a tradeoff between

symmetry-breaking force and quality factor to improve the photothermal coupling with applying

gate voltage. Also, shrinking the area of suspended membrane could help to avoid the generation

Page 151: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

130

of bistability but it also requires much higher gate voltage to pull the membrane down. To observe

much cleaner data for optical spring effect and optical damping, one need to do more thoughtful

simulation and optimization in the geometry of device to obtain pronounce effects from

optomechanical backaction.

Although, the optomechanical contribution from radiation pressure was almost negligible,

it would be still interesting to investigate the cavity optomechanics using excitonic effect in

monolayer TMD semiconductor. Instead of directly forming resonator with monolayers, it would

be desirable to put the monolayer semiconductor into a high-finesse cavity or place it very close to

an optical microcavity to enhance the radiation pressure felt by the membrane. Thus, lots of

interesting physics could arise from interfacing such exciton-polariton schematics with

nanomechanics. Compared to photothermal optomechanics, the interaction between exciton and

confined photons paves attractive routes towards cooling to the quantum ground state. In addition,

it would be highly interesting to investigate the generation of frequency comb by an optical micro-

resonator integrated with monolayer TMD semiconductors.

Within the past two years, the discovery of atomically thin magnetic materials opens a

window to a richer and largely unexplored regime in low dimensional physics, as well as studies

done in this dissertation. As indicating in Chapter 3, it would be encouraging to realize the helicity

control of optical switching without an external magnetic field in monolayer TMD semiconductor,

which can be magnetically proximitized by a neighboring 2D magnet (e.g. CrBr3). Meanwhile,

strain can play a very important role in the intrinsic magnetization of 2D magnetic materials. The

phase transition between anti-ferromagnetism (AFM) and ferromagnetism (FM) via strain have

been predicted in bilayer 2D magnetic materials without solid experimental evidence to date. In

addition, it would also be interesting to explore the strain-induced change in internal magnetization

of 2D magnet, where strain bridges the internal spin macroscopic mechanical DOF. The strong

magneto-elastic coupling within magnetic 2D atomic crystals (e.g. CrI3) would be another highly

Page 152: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

131

interesting topic that deserves thorough investigation. Meanwhile, as analogy to exciton-

optomechanics in Chapter 4, it would be highly interesting to study the coupling between

ferromagnetic resonance (FMR) and mechanics in a 2D magnet resonator, which can transduce

spin information into mechanical state. Such transduction between quantum mechanics and macro-

mechanics can also serve as a frequency transducer between GHz and MHz, which are normal

frequency range for FMR and mechanical modes in our device, respectively. And these novel spin-

mechanical devices, like above-mentioned exciton-optomechanical counterparts, may enable new

classes of measurements and find more interesting applications in the near future.

Page 153: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

Bibliography

[1] K. von Klitzing, G. Dorda, and M. Pepper, “New Method for High-Accuracy

Determination of the Fine Structure Constant Based on Quantized Hall resistance,” Phys.

Rev. Lett., vol. 45, no. 6, pp. 494–497, 1980.

[2] R. B. Laughlin, “Quantized Hall conductivity in two dimensions,” Phys. Rev. B, vol. 23,

no. 10, pp. 5632–5633, 1981.

[3] D. C. Tsui, H. L. Stormer, and A. C. Gossard, “Two-Dimensional Magnetotransport in the

Extreme Quantum Limit,” Phys. Rev. Lett., vol. 48, no. 22, pp. 1559–1562, 1982.

[4] K. S. Novoselov et al., “Electric Field Effect in Atomically Thin Carbon Films,” Science

(80-. )., vol. 306, no. 5696, pp. 666–669, 2004.

[5] P. R. Wallace, “The Band Theory of Graphite,” Phys. Rev., vol. 71, no. 9, pp. 622–634,

1947.

[6] K. S. Novoselov et al., “Two-dimensional atomic crystals,” Proc. Natl. Acad. Sci. United

States Am., vol. 102, no. 30, pp. 10451–10453, 2005.

[7] K. S. Novoselov et al., “Two-dimensional gas of massless Dirac fermions in graphene,”

Nature, vol. 438, no. 7065, pp. 197–200, 2005.

[8] Y. Zhang, Y.-W. Tan, H. L. Stormer, and P. Kim, “Experimental observation of the

quantum Hall effect and Berry’s phase in graphene,” Nature, vol. 438, no. 7065, pp. 201–

204, 2005.

[9] K. S. Novoselov et al., “Room-Temperature Quantum Hall effect in Graphene,” Science

(80-. )., vol. 315, no. 5817, p. 1379, 2007.

[10] S. Ghosh et al., “Extremely high thermal conductivity of graphene : Prospects for thermal

management applications in nanoelectronic circuits,” Appl. Phys. Lett., vol. 92, no. 15, p.

151911, 2008.

[11] A. A. Balandin et al., “Superior Thermal Conductivity of Single-Layer Graphene,” Nano

Lett., vol. 8, no. 3, pp. 902–907, 2008.

[12] C. Lee, X. Wei, J. W. Kysar, and J. Hone, “Measurement of the Elastic properties and

intrinsic strength of Monolayer Graphene,” Science (80-. )., vol. 321, no. 5887, pp. 385–

388, 2008.

Page 154: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

133

[13] I. W. Frank, D. M. Tanenbaum, A. M. Van Der Zande, and P. L. McEuen, “Mechanical

properties of suspended graphene sheets,” J. Vac. Sci. Technol. B Microelectron. Nanom.

Struct. Process. Meas. Phenom., vol. 25, no. 6, pp. 2558–2561, 2007.

[14] A. K. Geim, “Graphene : Status and Prospects,” Science (80-. )., vol. 324, no. 5934, pp.

1530–1534, 2009.

[15] K. S. Novoselov, V. I. Fal, L. Colombo, P. R. Gellert, M. G. Schwab, and K. Kim, “A

roadmap for graphene,” Nature, vol. 490, no. 7419, pp. 192–200, 2012.

[16] F. Schwierz, “Graphene transistors,” Nat. Nanotechnol., vol. 5, no. 7, pp. 487–496, 2010.

[17] S. Y. Zhou et al., “Substrate-induced bandgap opening in epitaxial graphene,” Nat.

Mater., vol. 6, no. 10, pp. 770–775, 2007.

[18] X. Wang et al., “N-Doping of Graphene Through Electrothermal Reactions with

Ammonia,” Science (80-. )., vol. 324, no. 5928, pp. 768–771, 2009.

[19] K. F. Mak, C. H. Lui, J. Shan, and T. F. Heinz, “Observation of an Electric-Field-Induced

Band Gap in Bilayer Graphene by Infrared Spectroscopy,” Phys. Rev. Lett., vol. 102, no.

25, p. 256405, 2009.

[20] Y. Zhang et al., “Direct observation of a widely tunable bandgap in bilayer graphene,”

Nature, vol. 459, no. 7248, pp. 820–823, 2009.

[21] F. Xia, D. B. Farmer, Y. Lin, and P. Avouris, “Graphene Field-Effect Transistors with

High On / Off Current Ratio and Large Transport Band Gap at Room Temperature,” Nano

Lett., vol. 10, no. 2, pp. 715–718, 2010.

[22] H. Min, J. E. Hill, N. A. Sinitsyn, B. R. Sahu, L. Kleinman, and A. H. Macdonald,

“Intrinsic and Rashba spin-orbit interactions in graphene sheets,” Phys. Rev. B, vol. 74,

no. 16, p. 165310, 2006.

[23] K. F. Mak and J. Shan, “Photonics and optoelectronics of 2D semiconductor transition

metal dichalcogenides,” Nat. Photonics, vol. 10, no. 4, pp. 216–226, 2016.

[24] Q. H. Wang, K. Kalantar-zadeh, A. Kis, J. N. Coleman, and M. S. Strano, “Electronics

and optoelectronics of two-dimensional transition metal dichalcogenides,” Nat.

Nanotechnol., vol. 7, no. 11, pp. 699–712, 2012.

[25] H. G. Craighead, “Nanoelectromechanical Systems,” Science (80-. )., vol. 290, no. 5496,

pp. 1532–1535, 2000.

[26] M. Roukes, “Nanoelectromechanical systems face the future,” Phys. World, vol. 14, no. 2,

pp. 25–31, 2001.

Page 155: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

134

[27] S. Manzeli, D. Ovchinnikov, D. Pasquier, O. V Yazyev, and A. Kis, “2D transition metal

dichalcogenides,” Nat. Rev. Mater., vol. 2, no. 17033, pp. 1–15, 2017.

[28] D. Voiry, A. Mohite, and M. Chhowalla, “Phase engineering of transition metal

dichalcogenides,” Chem. Soc. Rev., vol. 44, pp. 2702–2712, 2015.

[29] K. F. Mak, C. Lee, J. Hone, J. Shan, and T. F. Heinz, “Atomically Thin MoS2: A New

Direct-Gap Semiconductor,” Phys. Rev. Lett., vol. 105, no. 13, p. 136805, 2010.

[30] A. Splendiani et al., “Emerging Photoluminescence in Monolayer MoS2,” Nano Lett., vol.

10, no. 4, pp. 1271–1275, 2010.

[31] G. Liu, W. Shan, Y. Yao, W. Yao, and D. Xiao, “Three-band tight-binding model for

monolayers of group-VIB transition metal dichalcogenides,” Phys. Rev. B, vol. 88, no. 08,

p. 085433, 2013.

[32] D. Xiao, G.-B. Liu, W. Feng, X. Xu, and W. Yao, “Coupled Spin and Valley Physics in

Monolayers of MoS2 and Other Group-VI Dichalcogenides,” Phys. Rev. Lett., vol. 108,

no. 19, p. 196802, 2012.

[33] K. F. Mak, D. Xiao, and J. Shan, “Light – valley interactions in 2D semiconductors,” Nat.

Photonics, vol. 12, no. 8, pp. 451–460, 2018.

[34] X. Xu, W. Yao, D. Xiao, and T. F. Heinz, “Spin and pseudospins in layered transition

metal dichalcogenides,” Nat. Phys., vol. 10, no. 5, pp. 343–350, 2014.

[35] K. F. Mak, K. He, J. Shan, and T. F. Heinz, “Control of valley polarization in monolayer

MoS2 by optical helicity,” Nat. Nanotechnol., vol. 7, no. 8, pp. 494–498, 2012.

[36] H. Zeng, J. Dai, W. Yao, D. Xiao, and X. Cui, “Valley polarization in MoS2 monolayers

by optical pumping,” Nat. Nanotechnol., vol. 7, no. 8, pp. 490–493, 2012.

[37] T. Cao et al., “Valley-selective circular dichroism of monolayer molybdenum disulphide,”

Nat. Commun., vol. 3, no. 887, pp. 1–5, 2012.

[38] D. Macneill et al., “Breaking of Valley Degeneracy by Magnetic Field in Monolayer

MoSe2,” Phys. Rev. Lett., vol. 114, no. 03, p. 037401, 2015.

[39] Y. Li et al., “Valley Splitting and Polarization by the Zeeman Effect in Monolayer

MoSe2,” Phys. Rev. Lett., vol. 113, no. 26, p. 266804, 2014.

[40] G. Aivazian et al., “Magnetic control of valley pseudospin in monolayer WSe2,” Nat.

Phys., vol. 11, no. 02, pp. 148–152, 2015.

[41] A. Srivastava, M. Sidler, A. V Allain, D. S. Lembke, A. Kis, and A. Imamoğlu, “Valley

Zeeman effect in elementary optical excitations of monolayer WSe2,” Nat. Phys., vol. 11,

no. 02, pp. 141–147, 2015.

Page 156: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

135

[42] G. Wang et al., “Magneto-optics in transition metal diselenide monolayers,” 2D Mater.,

vol. 2, no. 3, p. 034002, 2015.

[43] A. V Stier, K. M. Mccreary, B. T. Jonker, J. Kono, and S. A. Crooker, “Exciton

diamagnetic shifts and valley Zeeman effects in monolayer WS2 and MoS2 to 65 Tesla,”

Nat. Commun., vol. 7, no. 10643, pp. 1–8, 2016.

[44] A. Chernikov et al., “Exciton Binding Energy and Nonhydrogenic Rydberg Series in

Monolayer WS2,” Phys. Rev. Lett., vol. 113, no. 07, p. 076802, 2014.

[45] S. Schmitt-Rink, D. S. Chemla, and D. A. B. Miller, “Linear and nonlinear optical

properties of semiconductor quantum wells,” Adv. Phys., vol. 38, no. 2, pp. 89–188, 1989.

[46] D. Snoke, “Spontaneous Bose Coherence of Excitons and Polaritons,” Science (80-. ).,

vol. 298, no. 5597, pp. 1368–1373, 2002.

[47] L. V Butov, C. W. Lai, A. L. Ivanov, A. C. Gossard, and D. S. Chemla, “Towards Bose-

Einstein condensation of excitons in potential traps,” Nature, vol. 417, no. 6884, pp. 47–

52, 2002.

[48] G. Scuri et al., “Large Excitonic Reflectivity of Monolayer MoSe2 Encapsulated in

Hexagonal Boron Nitride,” Phys. Rev. Lett., vol. 120, no. 3, p. 037402, 2018.

[49] P. Back, S. Zeytinoglu, A. Ijaz, M. Kroner, and A. Imamoglu, “Realization of an

Electrically Tunable Narrow-Bandwidth Atomically Thin Mirror Using Monolayer

MoSe2,” Phys. Rev. Lett., vol. 120, no. 3, p. 037401, 2018.

[50] H. Xie, S. Jiang, J. Shan, and K. F. Mak, “Valley-Selective Exciton Bistability in a

Suspended Monolayer Semiconductor,” Nano Lett., vol. 18, no. 5, pp. 3213–3220, 2018.

[51] D. Van Tuan, M. Yang, and H. Dery, “Coulomb interaction in monolayer transition-metal

dichalcogenides,” Phys. Rev. B, vol. 98, no. 12, p. 125308, 2018.

[52] H. M. Gibbs, Optical Bistability : Controlling Light by Light. Academic Press, Inc.:

Orlando, FL, 1985.

[53] S. Bertolazzi, J. Brivio, and A. Kis, “Stretching and Breaking of Ultrathin MoS2,” ACS

Nano, vol. 5, no. 12, pp. 9703–9709, 2011.

[54] K. Liu et al., “Elastic Properties of Chemical-Vapor-Deposited Monolayer MoS2 , WS2 ,

and Their Bilayer Heterostructures,” Nano Lett., vol. 14, no. 9, pp. 5097–5103, 2014.

[55] Rui Zhang et al., “Elastic properties of suspended multilayer WSe2,” Appl. Phys. Lett., vol.

108, no. 4, p. 042104, 2016.

Page 157: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

136

[56] K. He, C. Poole, K. F. Mak, and J. Shan, “Experimental demonstration of continuous

electronic structure tuning via strain in atomically thin MoS2,” Nano Lett., vol. 13, no. 6,

pp. 2931–2936, 2013.

[57] R. Roldan, A. Castellanos-Gomez, E. Cappelluti, and F. Guinea, “Strain engineering in

semiconducting two-dimensional crystals,” J. Phys. Condens. Matter, vol. 27, no. 31, p.

313201, 2015.

[58] K. I. Bolotin et al., “Ultrahigh electron mobility in suspended graphene,” Solid State

Commun., vol. 146, no. 9–10, pp. 351–355, 2008.

[59] X. Du, I. Skachko, A. Barker, and E. Y. Andrei, “Approaching ballistic transport in

suspended graphene,” Nat. Nanotechnol., vol. 3, no. 8, pp. 491–495, 2008.

[60] A. Castellanos-Gomez, V. Singh, H. S. J. van der Zant, and G. A. Steele, “Mechanics of

freely-suspended ultrathin layered materials,” Ann. Phys., vol. 527, no. 1–2, pp. 27–44,

2015.

[61] J. J. Bae et al., “Thickness-dependent in-plane thermal conductivity of suspended MoS2

grown by chemical vapor deposition,” Nanoscale, vol. 9, no. 7, pp. 2541–2547, 2017.

[62] X. Zhang et al., “Measurement of Lateral and Interfacial Thermal Conductivity of Single-

and Bilayer MoS2 and MoSe2 Using Refined Optothermal Raman Technique,” ACS Appl.

Mater. Interfaces, vol. 7, no. 46, pp. 25923–25929, 2015.

[63] Y. Li et al., “Measurement of the optical dielectric function of monolayer transition-metal

dichalcogenides : MoS2, MoSe2, WS2 and WSe2,” Phys. Rev. B, vol. 90, no. 20, p. 205422,

2014.

[64] N. Peimyoo, J. Shang, W. Yang, Y. Wang, C. Cong, and T. Yu, “Thermal conductivity

determination of suspended mono- and bilayer WS2 by Raman spectroscopy,” Nano Res.,

vol. 8, no. 4, pp. 1210–1221, 2015.

[65] J. S. Bunch et al., “Electromechanical resonators from graphene sheets,” Science (80-. ).,

vol. 315, no. 5811, pp. 490–493, 2007.

[66] C. Chen et al., “Performance of monolayer graphene nanomechanical resonators with

electrical readout,” Nat. Nanotechnol., vol. 4, no. 12, pp. 861–867, 2009.

[67] R. M. Cole et al., “Evanescent-Field Optical Readout of Graphene Mechanical Motion at

Room Temperature,” Phys. Rev. Appl., vol. 3, no. 2, p. 024004, 2015.

[68] N. Morell, A. Reserbat-plantey, I. Tsioutsios, K. G. Scha, F. H. L. Koppens, and A.

Bachtold, “High Quality Factor Mechanical Resonators Based on WSe2 Monolayers,”

Nano Lett., vol. 16, no. 8, pp. 5102–5108, 2016.

Page 158: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

137

[69] A. Castellanos-gomez, R. Van Leeuwen, M. Buscema, H. S. J. Van Der Zant, G. A.

Steele, and W. J. Venstra, “Single-Layer MoS2 Mechanical Resonators,” Adv. Mater., vol.

25, no. 46, pp. 6719–6723, 2013.

[70] J. Lee, Z. Wang, K. He, R. Yang, J. Shan, and P. X. L. Feng, “Electrically tunable single-

and few-layer MoS2 nanoelectromechanical systems with broad dynamic range,” Sci. Adv.,

vol. 4, no. 3, pp. 27–35, 2018.

[71] T. Miao, S. Yeom, P. Wang, B. Standley, and M. Bockrath, “Graphene

Nanoelectromechanical Systems as Stochastic-Frequency Oscillators,” Nano Lett., vol. 14,

no. 6, pp. 2982–2987, 2014.

[72] V. Singh, S. J. Bosman, B. H. Schneider, Y. M. Blanter, and G. A. Steele,

“Optomechanical coupling between a multilayer graphene mechanical resonator and a

superconducting microwave cavity,” Nat. Nanotechnol., vol. 9, no. 10, pp. 820–824, 2014.

[73] X. Song, M. Oksanen, J. Li, P. J. Hakonen, and M. A. Sillanpaa, “Graphene

Optomechanics Realized at Microwave Frequencies,” Phys. Rev. Lett., vol. 113, no. 2, p.

027404, 2014.

[74] P. Weber, I. Tsioutsios, D. E. Chang, and A. Bachtold, “Coupling Graphene Mechanical

Resonators to Superconducting Microwave Cavities,” Nano Lett., vol. 14, no. 5, pp. 2854–

2860, 2014.

[75] R. De Alba et al., “Tunable phonon-cavity coupling in graphene membranes,” Nat.

Nanotechnol., vol. 11, no. 9, pp. 741–746, 2016.

[76] J. P. Mathew, R. N. Patel, A. Borah, R. Vijay, and M. M. Deshmukh, “Dynamical strong

coupling and parametric amplification of mechanical modes of graphene drums,” Nat.

Nanotechnol., vol. 11, no. 9, pp. 747–751, 2016.

[77] V. Singh et al., “Probing thermal expansion of graphene and modal dispersion at low-

temperature using graphene nanoelectromechanical systems resonators,” Nanotechnology,

vol. 21, no. 16, p. 165204, 2010.

[78] J. Güttinger et al., “Energy-dependent path of dissipation in nanomechanical resonators,”

Nat. Nanotechnol., vol. 12, no. 7, pp. 631–636, 2017.

[79] M. Will et al., “High Quality Factor Graphene-Based Two-Dimensional Heterostructure

Mechanical Resonator,” Nano Lett., vol. 17, no. 10, pp. 5950–5955, 2017.

[80] J. Chaste, A. Eichler, J. Moser, G. Ceballos, R. Rurali, and A. Bachtold, “A

nanomechanical mass sensor with yoctogram resolution,” Nat. Nanotechnol., vol. 7, no. 5,

pp. 301–304, 2012.

Page 159: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

138

[81] R. A. Barton et al., “Photothermal self-oscillation and laser cooling of graphene

optomechanical systems,” Nano Lett., vol. 12, no. 9, pp. 4681–4686, 2012.

[82] X. Song, M. Oksanen, M. A. Sillanpaa, H. G. Craighead, J. M. Parpia, and P. J. Hakonen,

“Stamp Transferred Suspended Graphene Mechanical Resonators for Radio Frequency

Electrical Readout,” Nano Lett., vol. 12, no. 1, pp. 198–202, 2012.

[83] A. Eichler, J. Moser, J. Chaste, M. Zdrojek, and A. Bachtold, “Nonlinear damping in

mechanical resonators made from carbon nanotubes and graphene,” Nat. Nanotechnol.,

vol. 6, no. 6, pp. 339–342, 2011.

[84] A. Eichler, M. del A. Ruiz, J. A. Plaza, and A. Bachtold, “Strong Coupling between

Mechanical Modes in a Nanotube Resonator,” Phys. Rev. Lett., vol. 109, no. 2, p. 025503,

2012.

[85] J. Lee, Z. Wang, K. He, J. Shan, and P. X. Feng, “High Frequency MoS2 Nanomechanical

Resonators,” ACS Nano, vol. 7, no. 7, pp. 6086–6091, 2013.

[86] C. Liu, I. S. Kim, and L. J. Lauhon, “Optical Control of Mechanical Mode-Coupling

within a MoS2 Resonator in the Strong-Coupling Regime,” Nano Lett., vol. 15, no. 10, pp.

6727–6731, 2015.

[87] O. Arcizet, T. Briant, A. Heidmann, and M. Pinard, “Beating quantum limits in an

optomechanical sensor by cavity detuning,” Phys. Rev. A, vol. 73, no. 3, p. 033819, 2006.

[88] A. Tavernarakis et al., “Atomic Monolayer Deposition on the Surface of Nanotube

Mechanical Resonators,” Phys. Rev. Lett., vol. 112, no. 19, p. 196103, 2014.

[89] A. Gloppe et al., “Bidimensional nano-optomechanics and topological backaction in a

non-conservative radiation force field,” Nat. Nanotechnol., vol. 9, no. 11, pp. 920–926,

2014.

[90] A. W. Barnard, M. Zhang, G. S. Wiederhecker, M. Lipson, and P. L. Mceuen, “Real-time

vibrations of a carbon nanotube,” Nature, vol. 566, no. 7742, pp. 89–93, 2019.

[91] J.Moser;, A. Eichler, M. J. Esplandiu, D. E. Liu, M. I. Dykman, and A. Bachtold,

“Ultrasensitive force detection with a nanotube mechanical resonator,” Nat. Nanotechnol.,

vol. 8, no. 7, pp. 493–496, 2013.

[92] S. L. De Bonis et al., “Ultrasensitive Displacement Noise Measurement of Carbon

Nanotube Mechanical Resonators,” Nano Lett., vol. 18, no. 8, pp. 5324–5328, 2018.

[93] M. S. Haney et al., “Single-protein nanomechanical mass spectrometry in real time,” Nat.

Nanotechnol., vol. 7, no. 9, pp. 602–608, 2012.

Page 160: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

139

[94] J. E. Losby et al., “Torque-mixing magnetic resonance spectroscopy,” Science (80-. ).,

vol. 350, no. 6262, pp. 798–801, 2015.

[95] A. Mehlin et al., “Stabilized Skyrmion Phase Detected in MnSi Nanowires by Dynamic

Cantilever Magnetometry,” Nano Lett., vol. 15, no. 7, pp. 4839–4844, 2015.

[96] N. Rossi et al., “Vectorial scanning force microscopy using a nanowire sensor,” Nat.

Nanotechnol., vol. 12, no. 2, pp. 150–155, 2016.

[97] L. M. De Lépinay, B. Pigeau, B. Besga, P. Vincent, P. Poncharal, and O. Arcizet, “A

universal and ultrasensitive vertorial nanomechanical sensor for imaging 2D force fields,”

Nat. Nanotechnol., vol. 12, no. 2, pp. 156–162, 2016.

[98] Z. Wang et al., “Phase Transitions of Adsorbed Atoms on the Surface of a Carbon

Nanotube,” Science (80-. )., vol. 327, no. 5965, pp. 552–555, 2012.

[99] H. Zhu et al., “Observation of piezoelectricity in free-standing monolayer MoS2,” Nat.

Nanotechnol., vol. 10, no. 2, pp. 151–155, 2014.

[100] D. Lloyd et al., “Band Gap Engineering with Ultralarge Biaxial Strains in Suspended

Monolayer MoS2,” Nano Lett., vol. 16, no. 9, pp. 5836–5841, 2016.

[101] N. N. Klimov et al., “Electromechanical Properties of Graphene drumheads,” Science

(80-. )., vol. 336, no. 6088, pp. 1557–1561, 2012.

[102] S. Sengupta, H. S. Solanki, V. Singh, S. Dhara, and M. M. Deshmukh, “Electromechanical

resonators as probes of the charge density wave transition at the nanoscale in NbSe2,”

Phys. Rev. B, vol. 82, no. 15, p. 155432, 2010.

[103] M. Barmatz, L. R. Testardi, and F. J. Di Salvo, “Elastic measurements in the layered

dichalcogenides TaSe2 and NbSe2,” Phys. Rev. B, vol. 12, no. 10, pp. 4367–4376, 1975.

[104] S. Sengupta et al., “Plasmon Mode Modifies the Elastic Response of a Nanoscale Charge

Density Wave System,” Phys. Rev. Lett., vol. 110, no. 16, p. 166403, 2013.

[105] N. Morell et al., “Optomechanical Measurement of Thermal Transport in Two-

Dimensional MoSe2 Lattices,” Nano Lett., vol. 19, no. 5, pp. 3143–3150, 2019.

[106] S. Timoshenko, Theory of plates and shells. New York: McGraw-Hill, 1959.

[107] L. D. Landau and E. M. Lifshitz, Theory of Elasticity. Pergamon Press, 1970.

[108] T. J. Kippenberg and K. J. Vahala, “Cavity Optomechanics :Back-Action at the

Mesocale,” Science (80-. )., vol. 321, no. 5893, pp. 1172–1176, 2008.

[109] M. Aspelmeyer, T. J. Kippenberg, and F. Marquardt, “Cavity optomechanics,” Rev. Mod.

Phys., vol. 86, no. 4, pp. 1391–1452, 2014.

Page 161: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

140

[110] T. Corbitt et al., “An All-Optical Trap for a Gram-Scale Mirror,” Phys. Rev. Lett., vol. 98,

no. 15, p. 150802, 2007.

[111] M. Bagheri, M. Poot, M. Li, W. P. H. Pernice, and H. X. Tang, “Dynamic manipulation of

nanomechanical resonators in the high-amplitude regime and non-volatile mechanical

memory operation,” Nat. Nanotechnol., vol. 6, no. 10, pp. 726–732, 2011.

[112] M. Hossein-Zadeh and K. J. Vahala, “Photonic RF Down-Converter Based on

Optomechanical Oscillation,” IEEE Photonics Technol. Lett., vol. 20, no. 4, pp. 234–236,

2008.

[113] G. S. Wiederhecker, L. Chen, A. Gondarenko, and M. Lipson, “Controlling photonic

structures using optical forces,” Nature, vol. 462, no. 7273, pp. 633–636, 2009.

[114] C. Metzger, I. Favero, A. Ortlieb, and K. Karrai, “Optical self cooling of a deformable

Fabry-Perot cavity in the classical limit,” Phys. Rev. B, vol. 78, no. 3, p. 035309, 2008.

[115] I. Favero and K. Karrai, “Optomechanics of deformable optical cavities,” Nat. Photonics,

vol. 3, no. 4, pp. 201–205, 2009.

[116] M. Zalalutdinov et al., “Autoparametric optical drive for micromechanical oscillators,”

Appl. Phys. Lett., vol. 79, no. 5, pp. 695–697, 2005.

[117] M. Vogel, C. Mooser, K. Karrai, and R. J. Warburton, “Optically tunable mechanics of

microlevers,” Appl. Phys. Lett., vol. 83, no. 7, pp. 1337–1339, 2012.

[118] C. H. Metzger and K. Karrai, “Cavity cooling of a microlever,” Nature, vol. 432, no.

7020, pp. 1002–1005, 2004.

[119] O. Arcizet, P.-F. Cohadon, T. Briant, M. Pinard, and A. Heidmann, “Radiation-pressure

cooling and optomechanical instability of a micromirror,” Nature, vol. 444, no. 7115, pp.

71–74, 2006.

[120] A. Schliesser, P. Del’Haye, N. Nooshi, K. J. Vahala, and T. J. Kippenberg, “Radiation

Pressure Cooling of a Micromechanical Oscillator Using Dynamical Backaction,” Phys.

Rev. Lett., vol. 97, no. 24, p. 243905, 2006.

[121] S. Gigan et al., “Self-cooling of a micromirror by radiation pressure,” Nature, vol. 444,

no. 7115, pp. 67–70, 2006.

[122] D. Kleckner and D. Bouwmeester, “Sub-kelvin optical cooling of a micromechanical

resonator,” Nature, vol. 444, no. 7115, pp. 75–78, 2006.

[123] J. D. Teufel et al., “Sideband cooling of micromechanical motion to the quantum ground

state,” Nature, vol. 475, no. 7356, pp. 359–363, 2011.

Page 162: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

141

[124] J. Chan et al., “Laser cooling of a nanomechanical oscillator into its quantum ground

state,” Nature, vol. 478, no. 7367, pp. 89–92, 2011.

[125] T. J. Kippenberg, H. Rokhsari, T. Carmon, A. Scherer, and K. J. Vahala, “Analysis of

Radiation-Pressure Induced Mechanical Oscillation of an Optical Microcavity,” Phys.

Rev. Lett., vol. 95, no. 3, p. 033901, 2005.

[126] F. Marquardt and S. M. Girvin, “Optomechanics,” Physics (College. Park. Md)., vol. 2,

no. 40, 2009.

[127] L. V Butov, “Condensation and pattern formation in cold exciton gases in coupled

quantum wells Condensation and pattern formation in cold exciton gases in coupled

quantum wells,” J. Phys. Condens. Matter, vol. 16, pp. R1577–R1611, 2004.

[128] A. K. Geim and I. V Grigorieva, “Van der Waals heterostructures,” Nature, vol. 499, no.

7459, pp. 419–425, Jul. 2013.

[129] K. S. Novoselov, A. Mishchenko, A. Carvalho, and A. H. C. Neto, “2D materials and van

der Waals heterostructures,” Science (80-. )., vol. 353, no. 6298, p. aac9439, 2016.

[130] P. Blake et al., “Making graphene visible,” Appl. Phys. Lett., vol. 91, no. 6, p. 063124,

2007.

[131] M. M. Benameur, B. Radisavljevic, J. S. Heron, S. Sahoo, H. Berger, and A. Kis,

“Visibility of dichalcogenide nanolayers,” Nanotechnology, vol. 22, no. 12, p. 125706,

2011.

[132] A. Castellanos-Gomez, N. Agraït, and G. Rubio-Bollinger, “Optical identification of

atomically thin dichalcogenide crystals,” Appl. Phys. Lett., vol. 96, no. 21, p. 213116,

2010.

[133] R. V Gorbachev et al., “Hunting for Monolayer Boron Nitride : Optical and Raman

Signatures,” Small, vol. 7, no. 4, pp. 465–468, 2011.

[134] A. Castellanos-gomez et al., “Deterministic transfer of two-dimensional materials by all-

dry viscoelastic stamping,” 2D Mater., vol. 1, no. 1, p. 011002, 2014.

[135] C. R. Dean et al., “Boron nitride substrates for high-quality graphene electronics,” Nat.

Nanotechnol., vol. 5, no. 10, pp. 722–726, 2010.

[136] L. Wang et al., “One-Dimensional Electrical Contact to Two-Dimensional Material,”

Science (80-. )., vol. 342, no. 6158, pp. 614–618, 2013.

[137] K. Kim et al., “Van der Waals Heterostructures with High Accuracy Rotational

Alignment,” Nano Lett., vol. 16, no. 3, pp. 1989–1995, 2016.

Page 163: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

142

[138] K. F. Mak, M. Y. Sfeir, Y. Wu, C. H. Lui, J. A. Misewich, and T. F. Heinz, “Measurement

of the Optical Conductivity of Graphene,” Phys. Rev. Lett., vol. 101, no. 19, p. 196405,

2008.

[139] K. F. Mak, K. He, C. Lee, G. H. Lee, J. Hone, and T. F. Heinz, “Tightly bound trions in

monolayer MoS2,” Nat. Mater., vol. 12, no. 3, pp. 207–211, 2013.

[140] F. Hong and S. Xiong, “Single-photon transistor using microtoroidal resonators,” Phys.

Rev. A, vol. 78, no. 1, p. 013812, 2008.

[141] L. Neumeier, M. Leib, and M. J. Hartmann, “Single-Photon Transistor in Circuit Quantum

Electrodynamics,” Phys. Rev. Lett., vol. 111, no. 6, p. 063601, 2013.

[142] D. A. B. Miller, “Are optical transistors the logical next step ?,” Nat. Photonics, vol. 4, no.

1, pp. 3–5, 2010.

[143] K. He et al., “Tightly Bound Excitons in Monolayer WSe2,” Phys. Rev. Lett., vol. 113, no.

2, p. 026803, 2014.

[144] G. Wang et al., “Giant Enhancement of the Optical Second-Harmonic Emission of WSe2

Monolayers by Laser Excitation at Exciton Resonances,” Phys. Rev. Lett., vol. 114, no. 9,

p. 097403, 2015.

[145] Z. Ye et al., “Probing excitonic dark states in single-layer tungsten disulphide,” Nature,

vol. 513, no. 7517, pp. 214–218, 2014.

[146] D. Y. Qiu, F. H. Jornada, and S. G. Louie, “Optical Spectrum of MoS2 : Many-Body

Effects and Diversity of Exciton States,” Phys. Rev. Lett., vol. 111, no. 21, p. 216805,

2013.

[147] A. Ramasubramaniam, “Large excitonic effects in monolayers of molybdenum and

tungsten dichalcogenides,” Phys. Rev. B, vol. 86, no. 11, p. 115409, 2012.

[148] G. Sallen et al., “Robust optical emission polarization in MoS2 monolayers through

selective valley excitation,” Phys. Rev. B, vol. 86, no. 8, p. 081301(R), 2012.

[149] A. M. Jones et al., “Optical generation of excitonic valley coherence in monolayer WSe2,”

Nat. Nanotechnol., vol. 8, no. 9, pp. 634–638, 2013.

[150] G. He and S. H. Liu, Physics of Nonlinear Optics. World Scientific, 1999.

[151] Z. Wang, J. Shan, and K. F. Mak, “Valley- and spin-polarized Landau levels in monolayer

WSe2,” Nat. Nanotechnol., vol. 12, no. 10, pp. 144–149, Oct. 2017.

[152] Z. Wang, L. Zhao, K. F. Mak, and J. Shan, “Probing the Spin-Polarized Electronic Band

Structure in Monolayer Transition Metal Dichalcogenides by Optical Spectroscopy,”

Nano Lett., vol. 17, no. 2, pp. 740–746, 2017.

Page 164: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

143

[153] A. Arora, K. Nogajewski, M. Molas, M. Koperski, and M. Potemski, “Exciton band

structure in layered MoSe2: from a monolayer to the bulk limit,” Nanoscale, vol. 7, no. 48,

pp. 20769–20775, 2015.

[154] J. S. Ross et al., “Electrical control of neutral and charged excitons in a monolayer

semiconductor,” Nat. Commun., vol. 4, no. 1474, pp. 1–6, 2013.

[155] C. Chiritescu et al., “Ultralow Thermal Conductivity in Disordered, Layered WSe2

Crystals,” Science (80-. )., vol. 315, no. 5810, pp. 351–353, 2007.

[156] A. Mavrokefalos, N. T. Nguyen, M. T. Pettes, D. C. Johnson, and S. Li, “In-plane thermal

conductivity of disordered layered and superlattice films,” Appl. Phys. Lett., vol. 91, no.

17, p. 171912, 2007.

[157] W.-X. Zhou and K.-Q. Chen, “First-Principles Determination of Ultralow Thermal

Conductivity of monolayer WSe2,” Sci. Rep., vol. 5, no. 15070, pp. 1–8, 2015.

[158] C. Zhao et al., “Enhanced valley splitting in monolayer WSe2 due to magnetic exchange

field,” Nat. Nanotechnol., vol. 12, no. 8, pp. 757–762, 2017.

[159] D. Zhong et al., “Van der Waals engineering of ferromagnetic semiconductor

heterostructures for spin and valleytronics,” Sci. Adv., vol. 3, no. 5, p. e1603113, 2017.

[160] X. Liu et al., “Strong light-matter coupling in two-dimensional atomic crystals,” Nat.

Photonics, vol. 9, no. 1, pp. 30–34, 2014.

[161] S. Dufferwiel et al., “Exciton–polaritons in van der Waals heterostructures embedded in

tunable microcavities,” Nat. Commun., vol. 6, no. 8579, pp. 1–7, 2015.

[162] Y. Chen, J. D. Cain, T. K. Stanev, V. P. Dravid, and N. P. Stern, “Valley-polarized

exciton-polaritons in a monolayer semiconductor,” Nat. Photonics, vol. 11, no. 7, pp. 431–

435, 2017.

[163] K. Onomitsu et al., “Cavity-less on-chip optomechanics using excitonic transitions in

semiconductor heterostructures,” Nat. Commun., vol. 6, no. 8478, pp. 1–6, 2015.

[164] H. Xie, J. Shan, and K. F. Mak, “Tunable exciton-optomechanical coupling in suspended

monolayer MoSe2,” unpublished.

[165] H. J. Conley, B. Wang, J. I. Ziegler, R. F. Haglund, S. T. Pantelides, and K. I. Bolotin,

“Bandgap Engineering of Strained Monolayer and Bilayer MoS2,” Nano Lett., vol. 13, no.

8, pp. 3626–3630, 2013.

[166] C. Chen and J. Hone, “Graphene Nanoelectromechanical Systems,” Proc. IEEE, vol. 101,

no. 7, pp. 1766–1779, 2013.

Page 165: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

144

[167] J. Lee, Z. Wang, K. He, R. Yang, J. Shan, and P. X. L. Feng, “Electrically tunable single-

and few-layer MoS2 nanoelectromechanical systems with broad dynamic range,” Sci. Adv.,

vol. 4, no. 3, p. eaao6653, 2018.

[168] C. Chen et al., “Modulation of mechanical resonance by chemical potential oscillation in

graphene,” Nat. Phys., vol. 12, no. 3, pp. 240–244, 2016.

[169] Y. Yang et al., “Brittle Fracture of 2D MoSe2,” Adv. Mater., vol. 29, no. 1604201, pp. 1–

7, 2017.

[170] I. R. Storch et al., “Young ’ s modulus and thermal expansion of tensioned graphene

membranes,” Phys. Rev. B, vol. 98, no. 8, p. 085408, 2018.

[171] D. Cakir, F. M. Peeters, and C. Sevik, “Mechanical and thermal properties of h -MX2 ( M

= Cr , Mo , W ; X = O , S , Se , Te ) monolayers : A comparative study,” Appl. Phys. Lett.,

vol. 104, no. 20, p. 203110, 2014.

[172] M. Yarali et al., “Effect of Metal Doping and Vacancies on the Thermal Conductivity of

Monolayer Molybdenum Diselenide,” ACS Appl. Mater. Interfaces, vol. 10, no. 5, pp.

4921–4928, 2018.

[173] H. L. Kiwia and E. F. Westrum Jr, “Low-temperature heat capacities of molybdenum

diselenide and ditelluride,” J. Chem. Thermodyn., vol. 7, no. 7, pp. 683–691, 1975.

[174] M. Fiebig, “Revival of the magnetoelectric effect,” J. Phys. D. Appl. Phys., vol. 38, no. 8,

pp. R123–R152, 2005.

[175] W. Eerenstein, N. D. Mathur, and J. F. Scott, “Multiferroic and magnetoelectric

materials,” Nature, vol. 442, no. 7104, pp. 759–765, 2006.

[176] D. D. Awschalom and M. E. Flatté, “Challenges for semiconductor spintronics,” Nat.

Phys., vol. 3, no. 3, pp. 153–159, 2007.

[177] J. R. Schaibley et al., “Valleytronics in 2D materials,” Nat. Rev. Mater., vol. 1, no. 16055,

pp. 1–15, 2016.

[178] D. Xiao, W. Yao, and Q. Niu, “Valley-Contrasting Physics in Graphene : Magnetic

Moment and Topological Transport,” Phys. Rev. Lett., vol. 99, no. 23, p. 236809, 2007.

[179] J. Lee, Z. Wang, H. Xie, K. F. Mak, and J. Shan, “Valley magnetoelectricity in single-

layer MoS2,” Nat. Mater., vol. 16, no. 9, pp. 887–892, 2017.

[180] R. V. Gorbachev et al., “Detecting topological currents in graphene superlattices,” Science

(80-. )., vol. 346, no. 6208, pp. 448–451, 2014.

[181] J. Kim et al., “Ultrafast generation of pseudo-magnetic field for valley excitons in WSe2

monolayers,” Science (80-. )., vol. 346, no. 6214, pp. 1205–1208, 2014.

Page 166: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

145

[182] E. J. Sie, J. W. Mciver, Y. Lee, L. Fu, J. Kong, and N. Gedik, “Valley-selective optical

Stark e ect in monolayer WS2,” Nat. Mater., vol. 14, no. 3, pp. 290–294, 2015.

[183] Y. Ye et al., “Electrical generation and control of the valley carriers in a monolayer

transition metal dichalcogenide,” Nat. Nanotechnol., vol. 11, no. 7, pp. 598–602, 2016.

[184] J. Lee, K. F. Mak, and J. Shan, “Electrical control of the valley Hall effect in bilayer MoS2

transistors,” Nat. Nanotechnol., vol. 11, no. 5, pp. 421–425, 2016.

[185] Y. J. Zhang, T. Oka, R. Suzuki, J. T. Ye, and Y. Iwasa, “Electrically Switchable Chiral

Light-Emitting Transistor,” Science (80-. )., vol. 344, no. 6185, pp. 725–728, 2014.

[186] K. F. Mak, K. L. Mcgill, J. Park, and P. L. Mceuen, “The valley Hall effect in MoS2

transistors,” Science (80-. )., vol. 344, no. 6191, pp. 1489–1492, 2014.

[187] A. P. Cracknell, Magnetism in crystalline materials: Applications of the theory of groups

of cambiant symmetry. Pergamon Press, New York, 1975.

[188] L. S. Levitov, Y. V Nazarov, and G. M. Eliashberg, “Magnetoelectric effects in

conductors with mirror isomer symmetry,” Zhurnal Eksp. I Teor. Fiz., vol. 88, pp. 229–

236, 1985.

[189] V. M. Edelstein, “Spin polarization of conduction electrons induced by electric current in

two-dimensional asymmetric electron systems,” Solid State Commun., vol. 73, no. 3, pp.

233–235, 1990.

[190] I. Sodemann and L. Fu, “Quantum Nonlinear Hall Effect Induced by Berry Curvature

Dipole in Time-Reversal Invariant Materials,” Phys. Rev. Lett., vol. 115, no. 21, p.

216806, 2015.

[191] A. Chernyshov, M. Overby, X. Liu, J. K. Furdyna, Y. Lyanda-geller, and L. P. Rokhinson,

“Evidence for reversible control of magnetization in a ferromagnetic material by means of

spin – orbit magnetic field,” Nat. Phys., vol. 5, no. 9, pp. 656–659, 2009.

[192] I. M. Miron et al., “Current-driven spin torque induced by the Rashba effect in a

ferromagnetic metal layer,” Nat. Mater., vol. 9, no. 3, pp. 230–234, 2010.

[193] L. Yang et al., “Long-lived nanosecond spin relaxation and spin coherence of electrons in

monolayer MoS2,” Nat. Phys., vol. 11, no. 10, pp. 830–835, 2015.

[194] T. L. Linnik, “Effective Hamiltonian of strained graphene,” J. Phys. Condens. Matter, vol.

24, no. 20, p. 205302, 2012.

[195] H. Rostami, R. Roldan, E. Cappelluti, R. Asgari, and F. Guinea, “Theory of strain in

single-layer transition metal dichalcogenides,” Phys. Rev. B, vol. 92, no. 19, p. 195402,

2015.

Page 167: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

146

[196] M. Gurvitch, “Ioffe-Regel criterion and resistivity of metals,” Phys. Rev. B, vol. 24, no.

12, pp. 7404–7407, 1981.

[197] W. Wu et al., “Piezoelectricity of single-atomic-layer MoS2 for energy conversion and

piezotronics,” Nature, vol. 514, no. 7523, pp. 470–474, 2014.

[198] J. Tichy, J., Erhart, J., Kittinger, E., Privratska, Fundamentals of piezoelectrics sensorics:

Mechanical, dielectric and thermodynamical properties of piezoelectric materials.

Springer-Verlag, Berlin, Heidelberg, 2010.

[199] K.-A. N. Duerloo, M. T. Ong, and E. J. Reed, “Intrinsic Piezoelectricity in Two-

Dimensional Materials,” J. Phys. Chem. Lett., vol. 3, no. 19, pp. 2871–2876, 2012.

[200] V. J. Folen, G. T. Rado, and E. W. Stalder, “Anisotropy of the magnetoelectric effect in

Cr2O3,” Phys. Rev. Lett., vol. 6, no. 11, pp. 607–608, 1961.

Page 168: PROBING EXCITONIC MECHANICS IN SUSPENDED AND …

VITA

Hongchao Xie

EDUCATION

Visiting Ph.D. student 01/2018-07/2019

Cornell University, Ithaca, NY Advisor: Prof. Kin Fai Mak

Ph.D. in Physics 08/2014-08/2019

The Pennsylvania State University, University Park, PA Advisor: Prof. Kin Fai Mak

MPhys (Hons) Physics (1st class) 09/2012-06/2014

The University of Manchester, Manchester, UK Advisor: Prof. Kostya Novoselov

B.S. in Physics 09/2009-06/2013

Shandong University, Jinan, China

AWARDS

Homer F. Braddock Graduate Fellow Scholarship, The Pennsylvania State University, 2014

College Scholarship, Shandong University, 2010 & 2011

PUBLICATIONS

• H. Xie, J. Shan, K. F. Mak, “Tunable exciton-optomechanical coupling in suspended

monolayer MoSe2”, in preparation.

• H. Xie, S. Jiang, J. Shan, K. F. Mak, “Valley-selective exciton bistability in a suspended

monolayer semiconductor”, Nano Lett. 2018 (5), 3213-3220.

• J. Lee, Z. Wang, H. Xie, J. Shan, K. F. Mak, “Valley magnetoelectricity in single-layer MoS2”,

Nat. Mater. 2017, 16, 887-891

• G. M. Stiehl, R. Li, V. Gupta, I. El Baggari, S. Jiang, H. Xie, L. F. Kourkoutis, K. F. Mak, J.

Shan, R. A. Buhrman, D. C. Ralph, “Layer-dependent spin-orbit torques generated by the

centrosymmetric transition metal dichalcogenide 𝛽-MoTe2”, arXiv:1906.01068.