118
Symmetry Breaking and Harmonic Generation in Metasurfaces and 2-Dimensional Materials Jared S. Ginsberg Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy under the Executive Committee of the Graduate School of Arts and Sciences COLUMBIA UNIVERSITY 2021

Symmetry Breaking and Harmonic Generation in Metasurfaces

  • Upload
    others

  • View
    9

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Symmetry Breaking and Harmonic Generation in Metasurfaces

Symmetry Breaking and Harmonic Generation in Metasurfaces and 2-Dimensional Materials

Jared S. Ginsberg

Submitted in partial fulfillment of the

requirements for the degree of

Doctor of Philosophy

under the Executive Committee

of the Graduate School of Arts and Sciences

COLUMBIA UNIVERSITY

2021

Page 2: Symmetry Breaking and Harmonic Generation in Metasurfaces

© 2021

Jared S. Ginsberg

All Rights Reserved

Page 3: Symmetry Breaking and Harmonic Generation in Metasurfaces

Abstract

Symmetry Breaking and Harmonic Generation in Metasurfaces and 2-Dimensional Materials

Jared S. Ginsberg

A strong argument can be made that physics is, at its core, the study of symmetries.

Nonlinear optics is certainly no exception, with an enormous number of distinct processes each

depending in its own way on the underlying symmetries of the physical system, the light, or of

nature itself. Restricting ourselves to optical harmonic generation, we will explore three unique

physical systems as well as three symmetries. In each case, the controlled breaking of that

symmetry will lead to optical enhancements, novel nonlinearities, or deep physical insights.

Beginning with silicon metasurfaces, we will explore the effects of even and odd spatial

symmetries in optical systems. The periodic breaking of this symmetry will lead us to the highly

engineerable physics of bound states in the continuum. By studying the harmonic emission from

an atomic gas in the volume surrounding the metasurface, we will come to understand that

significant nonlinear optical enhancements can be engineered with any linewidth and at any

wavelength.

In the context of the two-dimensional material hexagonal boron nitride, we will

investigate and break crystal inversion symmetries. Using an intense laser tuned to the phonon

resonance of hexagonal boron nitride, large amplitude anharmonic ionic motions will provide us

a powerful degree of control over the internal symmetries of the system at an atomic level.

Breaking this symmetry, we measure short-lived even-order nonlinearities that would otherwise

be forbidden in equilibrium. Our observations for second- and third- harmonic generation are

Page 4: Symmetry Breaking and Harmonic Generation in Metasurfaces

confirmed by time-dependent density functional theory. Those simulations further extend the

understanding of this symmetry-breaking effect to even higher-order processes.

Lastly, single-crystal graphene and graphite provide an ideal platform through which to

explore time-reversal symmetry. Chiral photons, or optical beams with ellipticity and

handedness, are well known to break time-reversal symmetry. While applying high-power, chiral

light to graphene, the breaking of time-reversal lifts a degeneracy of the K and K’ valleys in the

momentum space Brillouin zone. Lifting this degeneracy, we unveil underlying spatial symmetry

properties of graphene in odd-order third- and fifth- harmonic generation which should otherwise

be unobservable. We also show experimentally, for the first time, that valley polarization and

population can be extracted using our technique.

Page 5: Symmetry Breaking and Harmonic Generation in Metasurfaces

i

Table of Contents

List of Figures ................................................................................................................................ iv

Acknowledgments......................................................................................................................... vii

Dedication ...................................................................................................................................... ix

Introduction ..................................................................................................................................... 1

Chapter 1: Low- and High- Order Harmonic Generation ............................................................... 4

Section 1.1: The Perturbative Regime of Harmonic Generation ................................................ 4

Section 1.2: The Semi-Classical Recollision Model for High-Order Harmonics ....................... 7

Step 1: Ionization ................................................................................................................... 8

Step 2: Acceleration ................................................................................................................ 9

Step 3: Recollision ................................................................................................................ 11

Section 1.3: Extension of the Semi-classical Model to Solids.................................................. 12

Chapter 2: Enhanced Harmonic Generation in Gases Using an All-Dielectric Metasurface ....... 15

Section 2.1: Quasi- Bound States in the Continuum................................................................. 16

Fundamental Principles:........................................................................................................ 16

Symmetry- Protected BICs: .................................................................................................. 18

Generalized Asymmetry Parameter and Quality Factor: ...................................................... 19

Section 2.2: Dimerized High- Contrast Gratings ...................................................................... 22

Dimerization and Spatial Mode Control: .............................................................................. 22

Final Device Design:............................................................................................................. 27

Section 2.3: Enhanced Nonlinearities in Gases ........................................................................ 28

Page 6: Symmetry Breaking and Harmonic Generation in Metasurfaces

ii

Wavelength Dependence: ..................................................................................................... 28

Polarization Dependence: ..................................................................................................... 30

Pressure Dependence: ........................................................................................................... 32

Section 2.4: Final Remarks ....................................................................................................... 33

Chapter 3: Nonlinear Phononics ................................................................................................... 37

Section 3.1: The Charged Harmonic Oscillator Model and Linear Optical Responses............ 37

Section 3.2: Lattice Anharmonicity .......................................................................................... 40

Chapter 4: Enhanced Harmonics and Phonon-Induced Symmetry Breaking in HBN ................. 44

Section 4.1: The Hyperbolic Phonon-Polaritons of Hexagonal Boron Nitride ........................ 45

Section 4.2: Frequency Dependent Third Harmonic Enhancement From Phonons ................. 52

Section 4.3: Time- Resolved Even-Order Nonlinearity in hBN ............................................... 59

Section 4.4: Polarization Dependence of Second Harmonic Generation ................................. 63

Section 4.5: Outlook on the Future of Nonlinear Phononics in hBN ....................................... 66

Chapter 5: Graphene Harmonic Generation and Time- Reversal Symmetry Breaking................ 68

Section 5.1: An Introduction to Graphene ................................................................................ 69

Electronic Properties ............................................................................................................. 69

Optical Properties.................................................................................................................. 72

Section 5.2: Elliptically Polarized Excitation of Graphene ...................................................... 75

Section 5.3: A Brief Review of the Role of Point-Groups of Symmetry in Harmonic

Generation from Solids ............................................................................................................. 82

Section 5.4: Time- Reversal Symmetry Breaking and Valley Polarization Effects ................. 86

Experimental Setup ............................................................................................................... 86

Page 7: Symmetry Breaking and Harmonic Generation in Metasurfaces

iii

Measurements ....................................................................................................................... 87

Theoretical Explanation of Emergent Flower Symmetry ..................................................... 89

Conclusions ................................................................................................................................... 96

References ..................................................................................................................................... 99

Page 8: Symmetry Breaking and Harmonic Generation in Metasurfaces

iv

List of Figures

1.1.1: Potential energy landscapes of anharmonic potentials…………….……..………….……...6

1.2.1: Trajectories of electrons vs tunneling times…………………………..……………..………8

1.2.2: Ponderomotive kinetic energy as a function of creation phase…………………….……….9

1.2.3: Lack of even harmonics in a centrosymmetric medium……………………………..……..10

1.3.1: Three-step high harmonic model for solids………………………………………….…….12

2.1.1: Depiction of bound state in the continuum………………………………………….……..17

2.1.2: Quality factor of BICs vs asymmetry parameter….…………………………………..……21

2.2.1: SEM image of dimerized high contrast grating in SOI……………………………….…...23

2.2.2: Band diagrams of HCG and DHCG…………………………………………………..……24

2.2.3: FDTD simulation of BIC mode in DHCG…………………………………………….…...25

2.2.4: Simulated and measured unpolarized reflectance from DHCGs……………………….….26

2.3.1: Wavelength dependence of THG on and off DHCG……………………………………....29

2.3.2: Polarization dependence of THG on DHCGs……………………………………………...30

2.3.3: Argon pressure dependence of THG and 5HG from DHCG…………………...………….32

2.4.1: Simulations of field and reflectance from 2D DHCG proposed design…………………. ..34

3.1.1: Optical and acoustic phonons in a harmonic potential……………………………………..38

3.1.2: Field and phonon response to harmonic potential……………………………………….....39

3.2.1: Field and phonon response with FT to anharmonic potential………………………………41

4.1.1: Toy model dispersion relations deriving Reststrahlen band……………………………….46

4.1.2: Image of hBN flake on silicon substrate…………………………………………………...47

4.1.3: Cartoon of hBN lattice with TO phonon changes also shown……………….…………….48

4.1.4: Phonon band diagram of hBN……………………………………………………………...49

Page 9: Symmetry Breaking and Harmonic Generation in Metasurfaces

v

4.1.5: Real permittivity of hBN both in plane and in z direction………………………………....50

4.1.6: Reflectance measurement on hBN at upper Reststrahlen band…………………………....51

4.2.1: Simulated atomic motion of B and N under phonon resonant excitation………………….52

4.2.2: Wavelength and power dependence of THG on resonance with hBN phonon………….....54

4.2.3: Experimental setup for THG in hBN……………………………………………………....55

4.2.4: Simulated HHG spectra for TO and LO phonon pumping…………………………………56

4.2.5: Pump intensity dependence simulations of resonant HHG in hBN………………………..57

4.3.1: TDDFT simulation of even order harmonics when pumping phonons……………………58

4.3.2: Experimental setup for pump-probe SHG in hBN…………………………………………59

4.3.3: Measured SHG time delay from hbn pump probe plus FT………………………………..61

4.3.4: Pump and probe power dependence of hBN SHG…………………………………………62

4.4.1: Theoretical and measured polarization dependence of hBN SHG…………………………63

4.4.2: Linecuts for polarization dependent SHG in hBN…………………………………………64

4.4.3: Simulated polarization dependence of SHG from hBN……………………………………65

5.1.1: Calculated electronic band dispersion of graphene in 2D…………………………………70

5.1.2: Dirac cones of graphene with various filling levels/ Fermi energies……………………….71

5.1.3: Real and imaginary parts of optical conductivity of graphene vs Fermi level…………….73

5.2.1: Reproduction of ellipticity dependence of harmonics in graphene………………………..74

5.2.2: Gating mechanism experimental design for graphene harmonics…………………………77

5.2.3: Experimental setup for gated graphene HHG..……………………………………………78

5.2.4: Gate dependent and ellipticity dependent THG and 5HG of graphene……………………79

5.2.5: Parallel and perpendicular components of THG in gated graphene…………….…………80

5.3.1: Comparison of CVD grown and exfoliated graphene crystals…………………………….81

Page 10: Symmetry Breaking and Harmonic Generation in Metasurfaces

vi

5.3.2: Theoretical polarization dependence of copolarized SHG in TMDs………………………83

5.3.3: Theoretical polarization dependence of copolarized THG in graphene……………………84

5.4.1: Experimental measurements of orientation dependent 5HG in graphite…………………..86

5.4.2: Flower pattern modulation depth vs pump ellipticity……………………………………..87

5.4.3: Orientation dependent 5HG for a thicker graphite flake…………………………………..88

5.4.4: Orientation dependent 5HG for a range of pump powers…………………………………88

5.5.1: Theory of 5HG from graphene as the sum of K and K’ valley contributions……..………90

5.5.2: Rotation of SHG flower pattern due to valley polarization……………………………..…91

5.5.3: Emergence of harmonic modulations resulting from valley polarization…………………92

5.5.4: 5HG orientation dependence (theoretical) for different valley populations……………….93

Page 11: Symmetry Breaking and Harmonic Generation in Metasurfaces

vii

Acknowledgments

There is not one project presented within this dissertation that I could have completed

without extensive help from my colleagues, collaborators, friends, and family. So many people

have contributed to this work, directly and indirectly. I hope they all enjoyed working with me as

much as I have with them.

First my advisor, Professor Alex Gaeta, who from the very beginning gave me the

freedom to forge my own research path, always trusting that we’d arrive at a place that was

exciting for us and impactful for others. I had not imagined before joining his group that a

graduate student could be given the resources and responsibility to drive their own research so

early on. It made for a truly unforgettable experience.

Everything I know about laboratory techniques, setups, lasers, the whole lot of it, I either

learned from or learned with Gauri Patwardhan. I think back to my first year in the office,

completely clueless and often times totally terrified, and I am so grateful to have had a mentor

like Gauri around to build my confidence in the lab. There’s nobody I’d rather flip an OPA

upside down with.

Our Ultrafast subgroup has been a tight-knit group of only a few people these past six

years. I sincerely hope that I’ve been as useful of a mentor to Cecilia Chen as Gauri was to me.

I’ve never worked with someone so willing to explore new techniques, new devices, or whacky

ideas as Cecilia. In working with her I’ve probably had more fun and unexpected days in the lab

than with anyone else. I can feel confident that the future of our little team is very safe with her.

I would be the first to admit that a new colleague showing up and flipping everything we

know and do on its head can be a painful experience. The opposite was true when Mehdi Jadidi

Page 12: Symmetry Breaking and Harmonic Generation in Metasurfaces

viii

came to Alex’s group. Mehdi took my research in directions that I may have never been exposed

to otherwise. I think it’s fair to say that our projects together have been the most rewarding of all.

Those of us who work with incredible devices and samples often take for granted our

collaborators who live in cleanrooms for our benefit. From Prof. Nanfang Yu’s group I would

especially like to thank Adam Overvig, who saw the potential in one of my earliest ideas and

worked with me to see it through to the very end. From James Hone’s group I’d like to mention

Sang Hoon Chae, who was unreasonably patient with me each time a laser annihilated one of his

beautiful samples.

Jin Zhang and Angel Rubio must be the most patient collaborators I have ever known.

Aside from his vital role as the primary theorist for my favorite paper, Jin went on all manner of

wild goose-chases looking for fascinating physics with me these past two years. I hope that it

was as worthwhile for their team as it was for me, and I hope to keep the collaboration going for

some time.

As proud as I am to have completed my degree at Columbia University, I am far prouder

to have done the work in such an incredible group. Yoshi, Yun, Bok, Chaitali, Chaitanya, Jae,

Miri, Yair, Andrew, Ricky, MJ, Alessandro, you’ve all contributed to a work environment that

I’ve truly loved every minute of.

Lastly, my friends and my family. I don’t know how I ended up so lucky to be

surrounded by people who support my every decision. Thank you all so much for supporting me

these last 6 years.

Page 13: Symmetry Breaking and Harmonic Generation in Metasurfaces

ix

Dedication

To my family.

Page 14: Symmetry Breaking and Harmonic Generation in Metasurfaces

1

Introduction

Strong field physics and attosecond science comprise a many decades long journey built

on first understanding, and then applying, phenomena of electron tunneling and recombination.

Beginning with noble gas experiments in the earliest days and only just being applied to solid-

state systems some ten years ago, strong field physics is now explored at the frontier of every

conceivable material system. From nanostructured arrays to the most exotic two-dimensional

materials, attosecond science now provides as much fundamental physics about the systems

themselves as those systems add to our ability to tailor beams of light.

One of the pillars of ultrafast optics, and the common thread that will link the various

explorations in this dissertation, is harmonic generation. In atomic systems, the ionization and

recollision of electrons and their parent ions by intense near- and mid-infrared lasers has led to

truly profound technological advancements. These culminated in the ability to generate

unbelievable optical bandwidths covering as many as hundreds of octaves and reaching energies

in the x-ray regime. These broad pulses allow scientists to perform some of the fastest

measurements possible, with time-resolutions of a few 10-18 seconds, and all with table-top laser

systems.

Though it may be many years before solid-state systems reach the performance abilities

of their atomic counterparts, interest in harmonic generation from solids is at an all-time high.

The large conversion efficiencies made possible by the densely packed lattices of solids was just

the first of what is now a plethora of advantages. The ability to (1) manufacture resonances, (2)

explore band-structures, and vibrational properties of matter, or (3) study symmetries and

generate previously impossible optical polarizations, makes solids a clear favorite platform for

future research.

Page 15: Symmetry Breaking and Harmonic Generation in Metasurfaces

2

The goal of this thesis is therefore to provide some of the earliest demonstrations of each

of these three avenues of study in solid-state devices. In doing so, I hope to prove to the reader

that solid and hybrid solid-gas systems are the future of ultrafast nonlinear optics and harmonic

generation. For each of the three areas I’ve listed above, we will need to utilize a unique material

system, specifically tailored to our goals. As such, this dissertation will be broken up into the

following main sections:

• The first chapter is where we will establish foundational knowledge of harmonic

generation and nonlinear optics. We’ll begin with the well-known perturbative regime

and its application to lower-order processes before exploring in more detail the semi-

classical theory of higher-order harmonic generation in gases. A beautiful analogy

between the atomic process and what takes place in bulk solids will cover the full scope

of systems for the remainder of the chapters.

• In Chapter 2 we will dive into our first specific system and symmetry, a hybrid solid-gas

platform for nonlinear optics that aims to apply the biggest advantages of nano-structured

devices to the well-established technique of gas-phase high-harmonic generation. In

doing so, we will take a detour into the physics of bound states in the continuum a

powerful concept in which spatial symmetry breaking can be leveraged to design

semiconductor Metasurfaces at any operating wavelength and with any quality factor.

Our unique insight will be to add spatial control to the long list of advantages we get

from BICs.

• Chapters 3 and 4 introduce an entirely new type of material resonance, one which is

much more rare in nature, yet every bit as applicable to our pursuit of advancing

nonlinear optics. Here we will be focusing on phonon-polaritonics in the van Der Waals

Page 16: Symmetry Breaking and Harmonic Generation in Metasurfaces

3

material hexagonal-boron nitride. Chapter 3 will establish the theoretical basis for

nonlinear phononics, building up from a harmonic oscillator model straight through to a

robust theory of lattice anharmonicity. Chapter 4 then applies that lattice anharmonicity

to boron nitride in two ways. First, we show enhanced nonlinearities due to phonon

coupling. Second, strong optical pumping is utilized to break the inherent inversion

symmetry of a material, enabling entirely new nonlinearities.

• Finally, we will touch on some of the most curious topics to have emerged in harmonic-

generation related experiments in the last 3 years, all relating to graphene and its thicker

cousin graphite. As a material without a bandgap, graphene exists in a unique space,

seemingly having optical properties of a semi-metal at certain times and those of an

atomic system at others. We will come to understand the role of graphene’s fermi level

and how it relates to this fascinating dichotomy. After that, we will explore ways in

which graphene’s unique relationship to time-reversal symmetry allows us to determine

material characteristics that should be entirely impossible to measure optically in a

system with its rotational symmetry properties.

Page 17: Symmetry Breaking and Harmonic Generation in Metasurfaces

4

Chapter 1: Low- and High- Order Harmonic Generation

Section 1.1: The Perturbative Regime of Harmonic Generation

Around the time of the completion of this dissertation, the first demonstration of optical

harmonic generation by Franken, et. al [1] will be celebrating its 60th anniversary. Using a ruby

laser emitting at 694 nm and quartz as a nonlinear medium, Franken and his team observed

second harmonic generation (SHG) in what we now understand to be the perturbative (or weak-

field) nonlinear regime. The years since this pioneering experiment have seen the field of

ultrafast nonlinear optics and harmonic generation extended far beyond the weak regimes and

simple bulk systems. And yet now, decades later, some of the most interesting harmonic

generation experiments are aimed at bending and manipulating foundational concepts of

harmonic generation. So that is where we will begin.

Linear optics supposes that the polarization in a material, induced by an applied electric

field 𝐸(𝑡), is given by 𝑃(𝑡) = 𝜖0𝜒(1)𝐸(𝑡) for a linear susceptibility 𝜒(1), which acts as a

material dependent proportionality constant. We allow for the possibility of nonlinear optical

interactions by generalizing this relationship to:

𝑷(𝒕) = 𝝐𝟎[𝝌(𝟏)𝑬(𝒕) + 𝝌(𝟐)𝑬𝟐(𝒕) + 𝝌(𝟑)𝑬𝟑(𝒕) + ⋯ ] ( 1)

which has been expanded as a power series in the electric field and now includes higher order

(nonlinear) susceptibilities, 𝜒(𝑛). Supposing now that the applied field is monochromatic with

frequency 𝜔0, amplitude 𝐸0, and of the form 𝐸(𝑡) = 𝐸0𝑒−𝑖𝜔0𝑡 + 𝑐. 𝑐, it is clear how the

nonlinear polarization of Equation 1 leads to the creation of new frequency components:

𝑷(𝒕) = 𝝐𝟎[𝝌(𝟏)𝑬𝟎𝒆𝟎

−𝒊𝝎𝟎𝒕 + 𝝌(𝟐)𝑬𝟎𝟐𝒆𝟎

−𝟐𝒊𝝎𝟎𝒕 + 𝝌(𝟑)𝑬𝟎𝟑𝒆𝟎

−𝟑𝒊𝝎𝟎𝒕 +⋯] ( 2)

This highly simplified expression fails to consider the possibility of multiple frequencies, but

nonetheless illustrates some important features. First, the integer multiples of the applied

Page 18: Symmetry Breaking and Harmonic Generation in Metasurfaces

5

frequency 𝜔0 are the harmonics we will be concerned with in this thesis. Second, we can read off

the field scaling of the different harmonic components, observing that the nth polarization order

scales as 𝐸0𝑛, and by extension the measured power of the nth harmonic scales as 𝐸0

2𝑛. The

perturbative definition of the nonlinear polarization given in Equation 1 is precisely why these

field dependencies constitute the “perturbative regime” of harmonic generation. Deviations from

that scaling, such as those we will see ourselves in Chapter 2, are considered to be the domain of

“nonperturbative” high-harmonic generation (HHG) which will be described in detail in the

following Section 1.2.

Much of the nonlinear behavior in a perturbative harmonic generation system can be

captured by an anharmonic oscillator model [2], describing the trajectories of electrons in a

potential well such as those provided by their atoms. An introduction to this technique is

included for the sake of preparing us for Chapter 3, in which the anharmonic behavior will

become the role of the lattice ions of a crystal, as opposed to the relatively freer electrons.

However, for brevity we will truncate this derivation at the point where we can start to extract

some useful conclusions, particularly about symmetries, and from there we will generalize.

Take the position of an electron to be x. Subjected to an applied electric field 𝐸(𝑡) with

frequency 𝜔0 and a potential well with damping constant 𝛾, the equation of motion for the

electron is:

𝒅𝟐𝒙

𝒅𝒕𝟐+ 𝟐𝜸

𝒅𝒙

𝒅𝒕+𝝎𝟎

𝟐𝒙 + 𝒂𝒙𝒏 = −𝒆𝑬(𝒕)

𝒎𝒆. ( 3)

I have deliberately chosen not to specify the exponent of the restoring force proportional to a.

The value of n defines the symmetry of the potential well, with even n leading to non-symmetric

potentials, and odd n restoring that symmetry (Figure 1.1.1). To solve, we expand the position x

Page 19: Symmetry Breaking and Harmonic Generation in Metasurfaces

6

as a power series, 𝑥 = 𝜖𝑥(1) + 𝜖2𝑥(2) + 𝜖3𝑥(3) +⋯, for small 𝜖. Substituting in and separating

the different perturbation orders, we find:

𝑑2𝑥(1)

𝑑𝑡2+ 2𝛾

𝑑𝑥(1)

𝑑𝑡+ 𝜔0

2𝑥(1) = −𝑒𝐸(𝑡)

𝑚𝑒

{

𝒅𝟐𝒙(𝟐)

𝒅𝒕𝟐+ 𝟐𝜸

𝒅𝒙(𝟐)

𝒅𝒕+𝝎𝟎

𝟐𝒙(𝟐) + 𝒂𝒙(𝟏)𝟐 = 𝟎, 𝒏 𝒆𝒗𝒆𝒏

𝒅𝟐𝒙(𝟐)

𝒅𝒕𝟐+ 𝟐𝜸

𝒅𝒙(𝟐)

𝒅𝒕+𝝎𝟎

𝟐𝒙(𝟐) = 𝟎, 𝒏 𝒐𝒅𝒅 ( 4)

𝑑2𝑥(3)

𝑑𝑡2+ 2𝛾

𝑑𝑥(3)

𝑑𝑡+ 𝜔0

2𝑥(3) + 2𝑎𝑥(1)𝑥(2) = 0

Let’s consider the steady state solutions of the middle pair of equations describing the second-

order in perturbation theory. With all time derivatives neatly set to 0 we are left with:

{ 𝒙(𝟐) = −

𝒂(𝒆𝑬𝟎𝒎)𝟐𝒆−𝟐𝒊𝝎𝟎𝒕

(𝟐𝒊𝝎𝟎𝜸)𝟐 , 𝒏 𝒆𝒗𝒆𝒏

𝒙(𝟐) = 𝟎, 𝒏 𝒐𝒅𝒅

( 5)

Some steps involving the solution of the first order component and its substitution into the next

step of the perturbation have been omitted. Carrying these two expressions forward, one finds

that the second order susceptibility for even n for a monochromatic field is given by 𝜒(2) =

𝑁(𝑒3

𝑚2)𝑎

𝜖0(2𝑖𝜔0𝛾)2[𝜔02−(2𝜔0)2−2𝑖𝜔0𝛾 ]

, whereas for odd n, the value of 𝜒(2) is identically 0. By considering

Figure 1.1.1: The potential energy landscape for anharmonic centrosymmetric media (red) and non-

centrosymmetric media (blue) compared to a perfectly harmonic potential (black). The symmetry of the

material, and hence the symmetry of the potential energy, places selection rules on the allowed harmonic orders.

Page 20: Symmetry Breaking and Harmonic Generation in Metasurfaces

7

an anharmonic oscillator model of nonlinear optics we have arrived at a key result, namely that

centrosymmetric media (odd n) are forbidden to have even order nonlinearities, while symmetry-

broken media have no such restriction placed on them. In Chapters 3 and 4 this distinction will

become our motivation to design a system in which the inversion symmetry can be changed

optically and at will.

In writing the nonlinear susceptibilities for a single frequency component and in the form

of a proportionality constant, the true tensor nature of the various 𝜒(𝑛) was lost. In Chapter 5 we

will return to extend the above discussion of inversion symmetries to include spatial symmetries

of crystalline solids, giving us a tool to understand the symmetries of emitted light as a function

of the arrangements of atoms.

Section 1.2: The Semi-Classical Recollision Model for High-Order Harmonics

While there are many different theories for describing the emission of higher- order

harmonics (HHG) from atomic systems [3–5], none have been so widely accepted and applied

with such success as the semi-classical model of Paul Corkum [6,7]. As luck would have it, this

most useful explanation also happens to be the simplest, consisting of three distinct steps, each of

which we will explore in some detail. Broadly speaking these steps include: (1) The tunneling of

an electron away from its parent atom, a type of ionization process resulting in a free electron

with little or no kinetic energy. (2) Acceleration of the electron by the electric field of the laser

that released it. (3) Recollision of the electron with its ion, caused by the changing sign of the

electric field, and resulting in the conversion of the electron’s kinetic energy to a high energy

Page 21: Symmetry Breaking and Harmonic Generation in Metasurfaces

8

photon. Each of these three stages is sensitive to different field and laser parameters, and by

treating each one in detail, we can establish for ourselves some benchmarks and goals that will

be revisited in future sections.

Step 1: Ionization

Of the three stages involved in the model, the first is the one which most heavily relies on

a quantum mechanical description, while providing the least information about the high energy

harmonics that will ultimately result from the full process. However, by considering the

combined atomic potential and applied electric field, we are able to write the ionization rate as:

𝑾𝒅𝒄 =𝑬𝒔𝟎

ℏ|𝑪𝒏∗𝒍∗|

𝟐𝑮𝒍𝒎 (𝟒𝑬𝒔𝟎

𝒆𝑬(𝟐𝒎𝒆𝑬𝒔𝟎)

−𝟏𝟐

)

𝟐𝒏−𝒎−𝟏

𝒆−𝟒𝑬𝒔𝟎

ℏ/𝟑𝒆𝑬(𝟐𝒎𝒆𝑬𝒔

𝟎)−𝟏𝟐, ( 6)

where 𝐸𝑠0 and 𝐸𝑠

ℎ are the ionization potential of the atom and the ionization potential of

hydrogen, respectively. 𝐺𝑙𝑚 and 𝐶𝑛∗𝑙∗ are uniquely determined by azimuthal and magnetic

quantum numbers, and E is the amplitude of the applied electric field. That is all to say, that the

Figure 1.2.1: Colors represent a range of one-dimensional electron trajectories for different tunneling times

(phases). The black curve is included as a guide, representing the period of the monochromatic field that

initiates the electron tunneling. The horizontal dotted line represents the atom’s position, illustrating the

range of different recollision times. Some electron trajectories, such as the top brown one, never return to

the atom.

Page 22: Symmetry Breaking and Harmonic Generation in Metasurfaces

9

probability of ionization 𝑊𝑑𝑐(𝐸(𝑡))𝑑𝑡, within the time interval dt, is purely a function of the

atomic potential, field strength, quantum numbers, and fundamental constants. The key

takeaway then is that the tunneling is a fundamentally probabilistic process, and yet the exact

moment of ionization is significant. As we will see in the following steps, the driving field phase

at the free electron’s creation time will lead to a strongly varying trajectory (Figure 1.2.1) and

final kinetic energy (Figure 1.2.2).

Step 2: Acceleration

We can now formulate the second step in the following simple way: we have an

effectively free electron in vacuum to which we are applying an oscillating electric field

𝐸 = 𝐸0 cos𝜔0𝑡 . The initial position x and momentum 𝑚𝑒𝑥 at time t0 are both 0. Then as a

function of time and carrier frequency, the position of the electron may be expressed as:

𝒙(𝒕) =𝑬𝟎

𝝎𝟎𝟐 [(𝒄𝒐𝒔𝒕 𝝎𝟎𝒕 − 𝒄𝒐𝒔𝝎𝟎𝒕𝟎) + (𝝎𝟎𝒕 − 𝝎𝟎𝒕𝟎)𝒔𝒊𝒏𝝎𝟎𝒕𝟎], ( 7)

Figure 1.2.2: Bottom: the kinetic energy imparted to the electron as a function of the phase of the laser at

the time of tunneling. Above: the laser amplitude at that time (phase). It is found that the maximum kinetic

energy that can be imparted is 3.17 Up, which occurs at a tunneling phase of 18o.

Page 23: Symmetry Breaking and Harmonic Generation in Metasurfaces

10

and it therefore has a kinetic energy of

𝑬𝒌𝒊𝒏𝒆𝒕𝒊𝒄(𝜽) = 𝟐𝑼𝒑(𝒔𝒊𝒏𝜽 − 𝒔𝒊𝒏𝜽𝒊)𝟐, ( 8)

where we have made two simplifications. First, we have replaced 𝜔𝑡 with the relevant phase 𝜃,

and introduced a new quantity, the ponderomotive energy 𝑈𝑝 =𝑒2𝐸0

2

4𝑚𝑒𝜔2, which represents the

average energy of a free electron in a monochromatic field. There are two key observations to be

made already, which will be revisited again and again in the coming chapters. That is, the

ponderomotive energy is linearly proportional to laser power, and scales quadratically with

wavelength. Given Eq (2) and (3), the final kinetic energy of the electron can be computed

numerically as a function of the driving laser’s phase, as shown in Figure 1.2.2. We find that at a

driving phase of 18o, a maximum kinetic energy of 3.17 times the very same ponderomotive

energy is possible. This maximum energy plus the atom’s potential energy is the highest energy

possible in an HHG process and fundamentally limits the short wavelength cut-off of a high-

harmonic spectrum [8,9].

Figure 1.2.3: (a) When performing harmonic generation with a perfectly symmetry atom such as Argon, only

odd order harmonics can be generated. This is due to the destructive interference of even orders from successive

high harmonic pulses, illustrated in the inset. (b) In atomic systems, elliptically polarized light can prevent the

proper recollision of the electron and ion, see here as a steep monotonic fall-off in harmonic generation power.

Page 24: Symmetry Breaking and Harmonic Generation in Metasurfaces

11

Step 3: Recollision

Having established that the maximum energy of an emitted photon during the recollision

process is predetermined to be equal to the ionization energy plus the kinetic energy calculated in

the previous section, we can turn our attention to two important subtleties of the HHG recollision

process. The first is that experimental harmonic spectra cannot be explained by a single

tunneling and recollision process. Specifically, it is well established that in HHG experiments

that utilize noble gases, only odd-order harmonics of the driving frequency are observed [10].

We will see this behavior in the following chapter in our own projects utilizing Argon, an

example of which is reproduced in Figure 1.2.3 (a) for a 3 μm pump beam. The full 3 step

process established in this chapter occurs every half-cycle of the driving laser, and to understand

the full picture of experimental observations, we need to consider not one, but a series of

tunneling, acceleration, and recollision processes.

Consider the emission of the nth harmonic in the first recollision process to be a simple

oscillating field 𝐶𝑜𝑠(𝑛𝜔0𝑡), and from the subsequent emission one half-cycle later to be

𝐶𝑜𝑠(𝑛(𝜔0𝑡 + 𝜋)). For combined emission we find that constructive interference occurs for all

odd orders of n, while destructive interference prevents any even-order harmonics from being

experimentally observed. A careful reader will realize that the requirement we’ve placed on the

harmonic orders is not so different from that which we derived for centrosymmetric media in

Section 1.1.1. In fact, the perfect spherical symmetry of a noble gas atom is an equally valid way

of arriving at the same conclusion.

The second important subtlety relates to whether the recollision takes place at all. Up

until now, we have considered the simplest situation of a one dimensional free-electron

trajectory. However, as we will see in later chapters the introduction of ellipticity to the driving

Page 25: Symmetry Breaking and Harmonic Generation in Metasurfaces

12

field can have profound effects on harmonic emission. In the case of an atomic gas, the

ellipticity causes a deviation of the electron from its straight path [11], increasing the likelihood

that the electron will fail to recollide with its original ion. Figure 1.2.3 (b) quantifies this picture,

showing the reduction of harmonic yield as the pump ellipticity is increased- a result of a failed

recollision step.

Section 1.3: Extension of the Semi-classical Model to Solids

Much of what we’ve learned in the previous section assumed a single electron tunneling

away from a single atom, and therefore feels as though it should fail to describe the interaction of

intense light with a crystalline solid consisting of many neatly arranged atoms. The trick to

making a tidy analogy between the atomic and solid cases is to recognize that the most powerful

description of electrons in a lattice is not a spatial picture, but an energy-momentum one [12,13].

Rather than tracking the trajectory of an electron in real-space, we will understand the 3-step

model of HHG in solids by their propagation along valence and conduction bands in a so-called

band diagram.

Figure 1.3.1: Illustration of the band

diagram for the one-dimensional chain

of atoms with two different masses.

Also shown are the three steps of the

solid HHG process. First, a large

number of fundamental photons excite

an electron from a lower band up to the

conduction band. The electron (hole)

accelerates in the conduction (valence)

band. The acceleration of a charge

contributes optical emission known as

intra-band HHG. Finally the electron

and hole recombine, emitting a high

energy photon in the process.

Page 26: Symmetry Breaking and Harmonic Generation in Metasurfaces

13

The calculation of a band diagram for a real material can be a substantial undertaking,

requiring powerful computation techniques such as Density Functional Theory (DFT) [14]. We

will instead derive the simplest two-band model of a one-dimensional diatomic chain of atoms,

and use that to illustrate the three steps. Consider two species of atoms with masses m and M,

connected by identical springs of spring constant 𝜆. The equations of motion for the positions

𝑢 of the two species are given by

{ 𝒎�̈�𝟐𝒏 = −𝝀(𝟐𝒖𝟐𝒏 − 𝟐𝟐𝒏−𝟏 − 𝒖𝟐𝒏+𝟏) 𝑴�̈�𝟐𝒏+𝟏 = −𝝀(𝟐𝒖𝟐𝒏+𝟏 − 𝟐𝟐𝒏 − 𝒖𝟐𝒏+𝟐)

. ( 9)

We make the standard ansatz that the solutions for 𝑢 oscillate in space and time according to

𝑢2𝑛 = 𝐴𝑒−𝑖𝜔𝑡−2𝑖𝑘𝑛𝑎 and 𝑢2𝑛+1 = 𝐵𝑒−𝑖𝜔𝑡−2𝑖𝑘𝑛𝑎. Plugging these in and solving for the

frequencies we find two solutions to the equations of motion:

𝝎± =𝝀

𝒎𝑴[𝒎 +𝑴±√(𝒎−𝑴)𝟐 + 𝟒𝒎𝑴 𝑪𝒐𝒔𝟐(𝒌𝒂) ] ( 10)

These two solutions are referred to as the dispersion curves for the one-dimensional diatomic

chain, and together make up the band diagram in Figure 1.3.1. An electron that exists in a lower

energy band, hereafter referred to as the valence band, may be promoted up to the higher energy

collection of states (the conduction band) by the absorption of one or more photons from an

applied laser. This takes the role of the tunneling step in analogy to the atomic system. The

difference now is that the role of an ionization potential is played by the bandgap.

The second step is likewise very similar to the free electron case. Now, a promoted

electron (as well as the hole that it left in the valence band) is accelerated within its new band,

acquiring kinetic energy. An important distinguishing detail here is that the oftentimes

anharmonic shape of the band leads to radiation in this intermediate step. This additional

contribution to the overall observed radiation is referred to as intraband harmonic

generation [13,15,16]. Finally, the electron and hole recombine, emitting a new photon at an

Page 27: Symmetry Breaking and Harmonic Generation in Metasurfaces

14

energy equal to the bandgap plus the accumulated energy that the electron gained during its

acceleration. This final process is known as interband HHG, and completes the analogy.

With this we have established much of the groundwork necessary to understand the

different examples of low- and high- harmonic generation throughout this thesis. Where

necessary, further theoretical frameworks will be presented to complete each story, such as

bound states in the continuum in Chapter 2 or phonon-polaritons following that. However, the

reader should always keep in mind the laser parameter dependencies, symmetry arguments, and

methods described thus far, as they will inform many of our motivations moving forward.

Page 28: Symmetry Breaking and Harmonic Generation in Metasurfaces

15

Chapter 2: Enhanced Harmonic Generation in Gases

Using an All-Dielectric Metasurface

In the previous chapter we derived the ponderomotive energy, a relationship between laser

power, wavelength, and the maximum energy photon that can be generated from a gas by HHG.

Though counterintuitive, the key to reaching ever shorter final wavelengths is to begin with the

longest wavelength possible! Next-generation studies of ultrafast atomic phenomena demand

higher repetition rates and shorter pulses with greater bandwidths. Thus, the quadratic

dependence of the maximum HHG bandwidth on pump wavelength [17] provides strong

motivation to push the wavelength of the pump field deeper into the mid-infrared. Strict intensity

requirements of nearly 100 terawatt/cm2 in order to access the non-perturbative regime of HHG

in gases has, however, mostly restricted the field to the use of chirped pulse, multi-pass, and

regeneratively amplified lasers. These large-footprint systems have low repetition rates, high

average power, and are limited to the near-visible regime, hindering progress towards mid-

infrared pumping. This chapter is motivated by a simultaneous pursuit of efficient HHG at longer

wavelengths and lower pump powers.

The lower intensities required to observe HHG from solids [18] has generated significant

interest and led to rapid innovation in the decade since its initial discovery. A promising

direction, and the topic of the current Chapter, has been the merging of nonlinear optics with

metasurfaces to create an efficient platform for HHG [19–24]. We will see how these

nanostructured devices allow for the engineering of far-field emission profiles, selective

wavelength enhancement, and exceptionally strong field-confinement. Sub-wavelength scale

plasmonics (made of metallic antennas) and dielectrics have both been demonstrated for these

Page 29: Symmetry Breaking and Harmonic Generation in Metasurfaces

16

purposes. While the strong field-confinement permitted by these two types of systems further

reduces the pump power requirements for harmonic generation in solid-state systems, enhancing

intensities within the device material poses its own challenges. Plasmonic devices are severely

hampered by ohmic losses and are susceptible to melting, even at modest intensities [25]. Both

plasmonics and dielectrics must also contend with the reabsorption of harmonics within an

opaque generation material [26], reducing the effective interaction length of the nonlinear

medium to only a few tens of nanometers. For these reasons we will focus on a hybrid approach

in this Chapter, combining the designable aspects of metasurfaces with the robustness of gases.

To do so, spatially selective confinement of pump energy within the regions just outside the

metasurface will be accomplished with bound states in the continuum (BIC), a class of optical

resonance with broadly engineerable mode profiles supported by all-dielectric metasurfaces [27–

30]. A variety of BIC-enhanced light-matter interactions including optical absorption [31], third-

harmonic generation (THG) in solids [32], and Raman scattering [33] were previously reported.

Section 2.1: Quasi- Bound States in the Continuum

Fundamental Principles:

In every discipline of physics from optics to quantum mechanics, the principle of a bound

state arises as a consequence of a potential well with discretized energy states. Look no further

than the quantum mechanical square well, with discrete, bound energy levels given by ℏ2𝑛2

2𝑚𝑎2 for

integer 𝑛, mass 𝑚, and well width 𝑎. In the limit that the potential well becomes finite, these

bound modes exist separate from the surrounding continuum of states, which are capable of

radiating energy out of the system to infinity. Though it may appear as if energy can be confined

within a continuum of states, it is often the case that there is still some coupling through which

Page 30: Symmetry Breaking and Harmonic Generation in Metasurfaces

17

energy is lost, and these states are referred to as resonances. Only in the limit where the lifetime

of the resonance goes to infinity and its linewidth goes to zero does the resonance truly become

bound. What could it mean then for a state to be bound despite existing in the continuum? Figure

2.1.1 is included as a graphical representation, but only serves to deepen the mystery.

Furthermore, for a mode to be entirely decoupled from all other channels and to have zero

linewidth is to say that it cannot be excited by any external fields, in addition to not radiating into

them. These two questions of existence and coupling will be tackled in turn, but not before a

brief history.

Bound states in the continuum were first proposed as a mathematical curiosity of

quantum mechanics by von Neumann [34], who painstakingly wrote out an infinite potential that

would support an electronic bound state. Though an experimental analog to von Neumann’s first

theoretical approach has never been realized, the universality of the potentials and Schrodinger

equation on which the concepts were formulated have motivated further searches for these

Figure 2.1.1: Various waveforms representing the continuum and bound states. The continuum states outside

of the potential well in any given system will have no spatial localization and tend to be a higher energy. By

comparison a leaky-mode, labeled here as a resonance, will have some spatialization but will ultimately radiate

away. Without any means of radiating, the state is perfectly localized, represented by the black curve.

Page 31: Symmetry Breaking and Harmonic Generation in Metasurfaces

18

elusive bound states. In the decades following, many such systems have been found, ranging

from entirely accidental bound states due to continuous parameter tuning [35,36] to

separability [37,38]. Perhaps the simplest such system, and the one we will describe in the next

section, is the bound state in the continuum protected by symmetries.

Symmetry- Protected BICs:

Our work in this chapter will primarily focus on one specific avenue for arriving at a

bound state in the continuum, and that is by exploiting symmetry. Those readers who have

experience in fiber optics or some mathematics may already be aware of the following fact: that

the modes belonging to two different symmetry classes naturally decouple from each other. This

can be understood as a trivial consequence of the mode overlap efficiency 𝜂, defined generally as

𝜼 =| ∫ 𝑬𝟏

∗𝑬𝟐𝒅𝑨|

∫|𝑬𝟏|𝟐𝒅𝑨 ∫|𝑬𝟐|

𝟐𝒅𝑨 ( 11)

for modes described by electric fields 𝐸1and 𝐸2 covering an area 𝑑𝐴. If we consider the simplest

case, in which our even mode is described by cos(𝑥), and our odd mode by sin (𝑥) in one

dimension, then clearly the mode coupling vanishes as ∫ sin(𝑥) cos(𝑥) 𝑑𝑥 = 0 over any integer

number of periods of oscillation.

There are as many different approaches to understanding symmetry protected BICs as

there are physical systems to explore. BICs have been realized in air pressure waves [39], water

waves [40], and a bit more recently quantum wires [41]. In the field of photonics there have even

been a number of nice demonstrations from which we are able to extract a key result: periodic

structures provide a natural starting place for symmetry- protected BIC design [42–44]. To be a

little more specific, planar photonic crystals fabricated on semiconductor wafers can be written

in a huge number of different ways. Let’s consider the likely scenario in which a planewave or

gaussian-shaped laser beam is incident on a structured photonic crystal. Then the symmetric

Page 32: Symmetry Breaking and Harmonic Generation in Metasurfaces

19

nature of the mode necessitates an anti-symmetric device mode in order for the resulting total

system to contain an odd state that is bound in the even-state continuum. This can be

accomplished when the periodic structure follows both time-reversal and inversion symmetry in

the plane. The resulting BIC occurs at one of the high symmetry points in momentum space. This

is an oversimplification of the full design considerations to come in later sections, but before

focusing in on any one specific design, let us tackle the looming question of how to access (and

make us of) an inaccessible resonance.

Generalized Asymmetry Parameter and Quality Factor:

The previous sections have established the principles behind the symmetry-protected

BIC, however it is important to realize that the issues of infinitesimal linewidth and hence zero

coupling make these bound states experimentally unviable. Without coming up with some

scheme that allows for the excitation of such states, we’ll have no hope of using them for any

impactful application. We are actually left with only two options in order to proceed: break one

of the two symmetries in the system- either that of the incident mode or the one you wish to

couple to- in such a way that the linewidth of the resonance becomes experimentally viable. The

applied perturbation, which can come in any number of forms depending on the device design

and BIC scheme, should not go so far as to make the BIC mode lifetime too short, so some care

should be taken.

Returning to our model from before, let’s imagine a planewave (even symmetry)

normally incident on a planar, periodic photonic crystal supporting a mode with odd symmetry.

The overlap integral of the two modes is zero and hence they cannot couple. To break the even

symmetry of the incident beam is actually a very simple task- it can be accomplished by slightly

altering the angle of incidence to be off the normal axis. In practice, this is not the preferred

Page 33: Symmetry Breaking and Harmonic Generation in Metasurfaces

20

approach, as the quality and wavelength of the resonance will become a sensitive function of the

incident beam, and not of the device design. The second option is considered the best practice-

to break the odd symmetry of the device mode [45] and allow coupling in and out. As we are

interested in periodic structures in this Chapter, and we wish to maintain periodicity in our total

system, the perturbation that we apply must also be periodic. For a series of semiconductor

pillars making up a metasurface, this can include anything from adding material, removing

material, shifting the position of, or tilting the constituent pillars. Regardless of which

perturbation approach is taken, we will now see that the coupling to a leaky mode is universal.

We start with a brief derivation of the inverse lifetime 𝛾𝑟𝑎𝑑 of the mode following [45], because

if we can observe a change from an infinite lifetime to a finite lifetime, we will have succeeded

in our goal. Consider a total resonance electric field 𝐸𝑟 = 𝐸0 + Δ𝐸, where 𝐸0 solves the Maxwell

equations and Δ is a small perturbation.

Each resonant state 𝐸𝑟 with complex frequency 𝜔𝑟 = 𝜔0 −𝑖𝛾

2, must come with a

corresponding complex conjugate state 𝐸𝑟∗. The complementarity of the complex conjugate states

actually gives us an elegant way to solve for the inverse lifetime,

[𝝎𝒓𝟐 −𝝎𝒓

∗𝟐] ∫ 𝒅𝑽 𝝐𝑬𝒓𝑬𝒓∗ − 𝒄𝟐𝑺𝟎∑ 𝒔(

𝑬𝒓𝜹𝑬𝒓∗

𝒅𝒛−

𝑬𝒓∗𝜹𝑬𝒓

𝒅𝒛)𝒔=+𝟏,−𝟏𝑽 ( 12)

Which can be simplified and estimated to first order using perturbation theory to actually be

𝜸

𝒄= 𝑺𝟎∑ 𝑺𝟎(|𝑬𝒓,𝒙|

𝟐+ |𝑬𝒓,𝒚 |

𝟐)𝒔=+𝟏,−𝟏 ( 13)

This sum can be computed with a few tricks that are beyond the scope of this section. To

summarize the result though, we can find that

𝜸

𝒄= |

𝒌𝟎

√𝟐𝑺𝟎[𝒑𝒙 −

𝒎𝒚

𝒄+

𝒊𝒌𝟎

𝟔𝑸𝒛𝒙]|

𝟐

+ |𝒌𝟎

√𝟐𝑺𝟎[𝒑𝒚 +

𝒎𝒙

𝒄+

𝒊𝒌𝟎

𝟔𝑸𝒚𝒛]|

𝟐 ( 14)

Page 34: Symmetry Breaking and Harmonic Generation in Metasurfaces

21

Where 𝑝 and 𝑚 are the electric and magnetic dipole and 𝑄 are the components of the electric

quadrupole. 𝑘0 is the wavevector and 𝑆0is the surface area of the integration. 𝑐 as always is the

speed of light. Now we can return to considering our periodic perturbations, and observe the

impact that this has on the lifetime we’ve computed. We proceed with the derivation for any

perturbation operation that maintains the mirror symmetry of the unit cell (x) goes to (-x). The

perpendicular components of the total 2D electric field can be described according to how they

transform under this symmetry operation, noting that 𝑥 and 𝑦 components will always be of

opposite parity. Let’s take the case where 𝑥 is the even mode direction and 𝑦 is odd. The second

term in Equation (14) vanishes for this scenario and the inverse lifetime simplifies dramatically

to

𝜸

𝒄=

𝒌𝟎𝟐

𝟐𝑺𝟎|𝒑𝒙|

𝟐. ( 15)

Let’s now transform our language from lifetime to quality factor, defined to be proportional to

that lifetime and resonance frequency according to 𝑄 =𝜔0

𝛾. The key to finishing the derivation

Figure 2.1.2: The generalized asymmetry parameter, a relevant example of which is given in the inset,

universally scales the quality factor according to 𝑸 ∝ 𝜶−𝟐. As the asymmetry parameter approaches zero, the

BIC lifetime goes to infinity.

Page 35: Symmetry Breaking and Harmonic Generation in Metasurfaces

22

lies in the fact that the dipole moment is proportional to the magnitude of the perturbation, and so

in total, 𝑄 =𝑆0

2𝑘0

1

Δ2|𝑝0|2 , where I have brought back Δ as the perturbation magnitude and 𝑝0 as

the dipole moment amplitude. We have quickly derived an extremely important result, that once

it is made accessible, the quality factor of a BIC scales inversely as the square of the size of the

perturbation.

All that remains is to generalize what we mean when we say “size of the perturbation”.

Let’s define a new perturbation parameter called the asymmetry, which is normalized according

to the device parameters in such a way that it scales from 0 to 1. 0 indicates no perturbation,

while 1 represents the extreme. As an example, if the perturbation is to the length of a lattice

“atom”, then the asymmetry parameter would be defined as the change in length divided by

original length, etc. This generalized result is depicted in Figure 2.1.2.

What we’ve just described is the BIC which is no longer completely bound, and because

of this it is generally referred to as a quasi-bound state in the continuum. The discussions of

bound states and asymmetry parameters have been kept as general as possible so far, but in the

following Section we will finally elaborate on the specific system for our experiments.

Section 2.2: Dimerized High- Contrast Gratings

Dimerization and Spatial Mode Control:

In Chapter 1.3.1 we derived the energy-momentum dispersion relationship for a simple

system of masses joined by identical springs. What we neglected to discuss at the time was that

this toy system, which I claimed to be the simplest, was quite clearly a step up in complexity

compared to a system with all identical masses. By alternating between two different “atoms” of

masses 𝑚 and 𝑀, we replaced the monomer unit cell of the system with a two atom “dimer”. At

Page 36: Symmetry Breaking and Harmonic Generation in Metasurfaces

23

the time, this was a necessary step to ensure that our model system would have a bandgap. As we

have now seen, the disruption of the periodicity can have more profound effects (such as creating

leaky modes) depending on the perturbed system.

An ideal platform to study these periodic perturbations (and the one we will perform our

experiments on) is a high-contrast grating [46–50]. In this context, “contrast” refers to the

difference in refractive index between the active dielectric medium and the surrounding

environment. Originally, lower contrast periodic gratings grew to prominence because the

lifetime of the guided mode resonance that they supported could be controlled as a function of

the depth of the etch surrounding the grating fingers [51]. High contrast gratings are then the

limit of dielectric gratings as the contrast becomes maximal, that is, the surrounding index

approaches unity. Deep etching solved an issue of scale in dielectric gratings, which is outside

the realm of this Chapter. But to put it simply, shallow corrugation necessitated fabrication of

very large gratings [52].

Figure 2.2.1: SEM of a dimerized high contrast grating used in this work. The effectively infinite extent of the

grating fingers in the y-direction makes this, in essence, a one-dimensional system. The dimerizing perturbation

can clearly be seen to be applied to the gaps between bars as opposed to their widths. The reasons for this choice

will become evident in the next section.

Page 37: Symmetry Breaking and Harmonic Generation in Metasurfaces

24

Following the recipe developed in the previous section of introducing a periodic

perturbation to a regular array of dielectric bars or resonators, we will now discuss a subset of

high contrast gratings known as dimerized high contrast gratings (DHCGs) pioneered by

collaborators to this work Nanfang Yu and Adam Overvig [46]. In one periodic dimension, let’s

consider each resonator to be an infinitely long silicon bar (sometimes referred to as a grating

finger here) in the y-direction, and periodic in x. Reducing the system to effectively only one

dimension also limits the total number of dimerizing perturbation classes to just two. On one

hand, every other grating finger can have its width adjusted in the x-direction. Alternatively, the

widths can be held constant and every bar can be shifted from its original location along x in

alternating directions. An SEM image of a gap-perturbed dimerized high-contrast grating is

shown in Figure 2.2.1. As discussed above, the change in width or position divided by their

respective starting dimensions specifies our asymmetry parameter, and hence the bandwidth and

Q-factor of the system.

Figure 2.2.2: HCG vs DHCG band structures. (a) The band structure of a 1D HCG. The arrow points to a mode

of interest to us that exists at the edge of the first Brillouin zone (FBZ). This mode is below the light line,

illustrated by the shaded region, making it inaccessible. Inset: a sketch of the corresponding grating. (b)

Dimerization of the HCG folds the FBZ. The arrow points to the same mode as before, now located at k=0. It

can be experimentally accessed.

Page 38: Symmetry Breaking and Harmonic Generation in Metasurfaces

25

The dimerization of a high-contrast grating alters the physics of the system in ways

beyond the perturbation to the mode symmetry described in Section 2.1. In the momentum space

picture, the doubling of the spatial period of the structure implies a halving in k-space, known as

Brillouin zone folding [53]. This “folds” modes at the edges of the first Brillouin zone (FBZ) to

the gamma point, making them accessible to normal incidence excitation. Such modes had

previously only existed in an unusable state below the light-line. The Brillouin zone folding of a

high contrast grating is illustrated in Figure 2.2.2.

Brillouin zone folding comes with one more inherent advantage: the modes that exist at

the Brillouin zone boundaries are typically much flatter than modes found at the gamma point,

and the slope of the band is directly proportional to the group velocity of light in the mode [54].

Therefore, high degrees of confinement in small devices benefit from the flattest possible bands.

Furthermore, the combination of small slopes and a reduced FBZ means DHCGs can exhibit

sharper spectral features with narrower linewidths than either unperturbed or low-contrast

gratings.

Figure 2.2.3: Finite-Difference Time-Domain simulation of the gap-centered mode originating from a gap-

perturbation to a 1D DHCG. Though the mode extends into the silicon and substrate, the mode volume in the

gaps is more than sufficient to enhance the optical intensity felt by argon atoms by orders of magnitude.

Page 39: Symmetry Breaking and Harmonic Generation in Metasurfaces

26

What remains is to choose between the width and gap varieties of perturbation. This

choice is ultimately tied to the intended application of the final metasurface, so let’s quickly

review the motivation presented at the start of the Chapter. We are envisioning a system in which

the resonance provided by the DHCG can enhance the electric field strength being used to

perform high-harmonic generation in gases. We must therefore choose a periodic perturbation

which provides the quasi-BIC mode most localized in the gaps between grating fingers. Those

gaps will be the perfect place to flow our atomic sample.

We choose to modulate the spacing (i.e., the gap-perturbation) because this folds the

unique mode depicted in Figure 2.2.3 to the gamma point. Mode and device simulations are

performed with the commercial finite-difference time-domain (FDTD) solver licensed by

Lumerical. The mode is centered in the gap between adjacent grating fingers, where it provides a

uniform two orders of magnitude enhancement in the local field intensity for a modest quality

factor Q = 1500. This mode affords us the ability to inject our gaseous nonlinear medium (e.g.,

Figure 2.2.4: (a) Simulated unpolarized reflectance spectrum of DHCGs with alternating gaps of 190 nm ± δ.

The reflectance peaks at the BIC Fano-resonance design wavelength near 1550 nm. Smaller perturbations

correspond to larger Q-factors and narrower linewidths. (b) FTIR measured reflectance of unpolarized near-

infrared light. Losses lead to the observed variation in reflectance magnitude.

Page 40: Symmetry Breaking and Harmonic Generation in Metasurfaces

27

Argon atoms) into the gaps. Alternatively, a mode centered purely within the fingers could have

been achieved with a width perturbation and fixed gaps.

Final Device Design:

We designed and fabricated silicon-on-insulator (SOI) DHCGs on a standard wafer with

a 250-nm thick silicon device layer, 1-μm of oxide, and a 500-μm silicon substrate. Figure 2.2.1

shows a top view SEM image of that very device, fabricated by e-beam lithography, which

consists of a series of silicon fingers 270-nm wide with a total grating period of 920-nm. The

large grating period is a result of the dimerization of an unperturbed high contrast grating with a

period of 460-nm. We also fabricated a second grating with approximately twice the width and

spacing, allowing us to test two very different operating wavelengths of 1550 nm and 2900 nm.

By scaling the period and duty cycle, the system can be made to operate at any wavelength and is

limited only by fabrication sensitivity.

We both simulated and experimentally tested a variety of perturbation sizes to the silicon

gratings. Figures 2.2.4 (a) and (b) show the simulated and measured unpolarized reflectance

spectra of a set of DHCGs designed with a central wavelength of 1550 nm and a range of Q-

factors as high as 1000. We also include the symmetry-protected case of unperturbed gap widths

of 190 nm and confirm that no resonance is observed (this is a true BIC, not Quasi). In our

experiments we choose to increase the size of the perturbation in order to accommodate the

bandwidth of our laser pulses, while simultaneously making it easier for the field to couple into

the mode.

Only when the incident laser polarization of the pump is oriented along the finger

direction can the required BIC mode be efficiently excited. The total number of modes in the

Page 41: Symmetry Breaking and Harmonic Generation in Metasurfaces

28

system is also a function of the grating depth. By restricting the device to 250 nm in height we

thereby reduce the likelihood of a parasitic resonance being located near our gap-centered mode.

Section 2.3: Enhanced Nonlinearities in Gases

Wavelength Dependence:

At last, our ideal metasurface is designed and fabricated, capable of providing field

enhancements throughout the infrared wavelength range and with any Q-factor we desire. The

experimental setup used to test the devices for enhanced harmonic generation is as follows:

An optical parametric amplifier (Light Conversion HE Topas Prime) pumped by an

amplified Titanium-sapphire laser system (Coherent Legend Elite) operating at a 1-kHz

repetition rate with 6 mJ of pulse energy to generate tunable, 60-fs duration signal pulses with

center wavelengths from 1485 nm to 1580 nm. This parametric amplifier output is used to pump

an additional difference frequency generation module for our mid-infrared measurements at 2.9

μm with a pulse duration of 70 fs. The pump-beam polarization is controlled with a broadband

zero-order half wave plate before being focused on the sample by a 5-cm focal length CaF2 lens.

Any unconverted seed light is removed with the appropriate long-pass filters. Reflected

harmonics are collected and measured on a fiber-coupled ultraviolet-visible spectrometer (Ocean

Optics) with a 350 nm to 1100 nm detection range. The devices are mounted in a high-pressure

stainless-steel gas cell with a 1-inch diameter sapphire window. For sufficiently large beam

diameters, the intensity on the sapphire window is low enough that it does not contribute to the

nonlinear signal in our experiments. Argon pressures in excess of 200 psi could be held on the

devices within the cell.

The first thing we wish to test is whether enhanced nonlinearity is restricted to the design

wavelength of our DHCG. The third-harmonic-generation (THG) signals generated on and off

Page 42: Symmetry Breaking and Harmonic Generation in Metasurfaces

29

the 1550 nm- resonant DHCG are shown in Figure 2.3.1. Off the metasurface structure, the pump

wavelength was swept from 1485 nm to 1580 nm, and the emitted third harmonic signal was

recorded in Figure 2.3.1 (a). We observe the expected behavior of harmonic emission that tracks

the pump wavelength without any additional spectral features. This picture changes when we

repeat the same wavelength scan on the device itself. The harmonic yield dramatically increases,

requiring over an order of magnitude lower power to generate, but within a limited bandwidth.

At pump wavelengths below 1530 nm or above 1580 nm no measurable increase is observed in

the third-harmonic signals plotted in Figure 2.3.1 (b), and only a small signal, far off-resonance

due to the peak of the 1485 nm pump is measured (blue arrow). Compared to this small signal,

the 1550-nm pump beam displays the greatest harmonic yield enhancement of a factor of 45,

implying the greatest overlap with the BIC resonance. This concept is illustrated in Figure 2.3.1

Figure 2.3.1: (a) Normalized third-harmonic signal generated from the bare substrate for pump wavelengths

in the range of 1485 nm to 1580 nm. Pump powers are 10X greater than in (b). (b) Third-harmonic spectra for

the same five pump wavelengths shown in (a), now on the DHCG. The large field enhancement on resonance

pins the third harmonic to 513 nm. The small signal at 495 nm (blue arrow) corresponds to the third-harmonic

signal from the peak of the 1485-nm pump that is generated off-resonance. (c) Schematic of the pump field

and resonance overlap resulting in the harmonics measured in B. The shaded overlap regions contribute to the

nonlinear signal.

Page 43: Symmetry Breaking and Harmonic Generation in Metasurfaces

30

(c), in which the overlap of the Gaussian pump spectra and Lorentzian BIC resonance determines

efficient harmonic generation. This explains the observed pinning of the THG to roughly 516 nm

and the reduced linewidths of detuned harmonics. The cut-off shape of the short wavelength

harmonics gives the clearest indication of the resonance edge.

Polarization Dependence:

We also examine the polarization dependence of our fabricated near-IR metasurfaces. As

described in an earlier Section, the grating mode under study is excited by TE polarized light

parallel to the long direction of the grating fingers. We hereafter refer to this polarization

direction as 0o. The theoretical dependence of the peak reflectance from the grating is studied in

a series of FDTD simulations and plotted as a blue line in Figure 2.3.2 (b). For clarity, we

Figure 2.3.2: (a) (A) Normalized third-harmonic spectra of a 1550 nm pump, corresponding to the red points

in B. Also included is the signal generated off the device, which is unmeasurable. 1 atmosphere of Argon

pressure was flowed over the sample. Contributions to the third harmonic from both gas and silicon are

present. (B) Simulated peak reflectance of the ideal grating (blue line), and integrated experimental third

harmonic yield (red dots), as a function of the pump polarization. 0o is the polarization parallel to the grating

fingers. The red dashed line indicates 𝑪𝒐𝒔 𝟔𝜽, corresponding to the expected scaling of perturbative THG.

The mismatch between the red theoretical curve and experimental dots indicates that the measured THG is

instead following a non-perturbative scaling typical of higher pump intensities.

Page 44: Symmetry Breaking and Harmonic Generation in Metasurfaces

31

subtract any contribution to the reflectance due to etalon effects of the thin device and oxide

layers, hence the reflectance has a minimum of zero. As the polarization is rotated by an angle 𝜃

beginning from 0o, the peak power reflectance due to the BIC follows precisely the expected

cos2 𝜃 behavior. The component of the field oriented along 90o therefore makes no contribution,

and the cos2 𝜃 behavior is a result of the projection of the pump beam onto the 0o axis. We

compare to this reflectance curve the integrated (measured) THG yield of 1550-nm pulses on

resonance, shown as red dots in Figure 2.3.2 (b), corresponding to the spectra in Figure 2.3.2 (a).

The experimental data follow a similar trend, but with some important differences. Were it a

perturbative third-order nonlinearity with intensity 𝐼 ∝ cos2 𝜃, a THG efficiency proportional to

cos6 𝜃 may be expected, given by the red line in (b). The measured harmonics follow neither the

exact cos2 𝜃 behavior of the resonance or the perturbative cos6 𝜃 dependence.

That observed saturation of the intensity dependence of the Nth order harmonic below 𝐼𝑁

is a signature of non-perturbative harmonic generation. Thus, we conclude that in these

experiments the generated signal corresponds to the non-perturbative regime with more than 32

times larger emission at 0o than at 80o. Within the nonperturbative regime for THG pumped in

the mid-IR, a conversion efficiency upper-bound of about 10-7 is found for low-pressure Argon

gas [55]. Fifth harmonic conversion efficiencies saturate about one order of magnitude lower at

10-8. While the intensity enhancement provided by the DHCGs demonstrated here does not raise

this efficiency limit, it does successfully lower the input power at which the maximum

conversion efficiency is reached. At 90o polarization, the harmonic yield converges back to the

signal off-resonance.

Page 45: Symmetry Breaking and Harmonic Generation in Metasurfaces

32

Pressure Dependence:

A final, crucial experimental check should be performed when working in a hybrid solid-

gas system. We need to be certain that the measured harmonics originate from the gas sample

and not the substrate or silicon device. To do so we also characterize the performance using the

scaled-up mid-infrared devices, allowing us to explore the third and fifth harmonics, both of

which fall within our detection range for a 2.9 μm pump wavelength. Figure 2.3.3 shows the

third- and fifth-harmonic spectra of these structures for a range of Argon gas pressures. For both

harmonic orders, the harmonic yields scale quadratically with gas pressure (see (b)) confirming

that the emission is due to the atoms in the gap regions. At a pump intensity of 2 x 1012 W/cm2,

this emission occurs at field strengths where efficient harmonic generation in Argon is not

expected to take place. In the absence of Argon, we still observe a non-zero third- and fifth-

harmonic signal, which we attribute to the silicon finger surfaces. This is further supported by

Figure 2.3.3: (a) Third and fifth harmonics of a 2.9 μm pump beam on a 2.9 μm resonant DHCG. Argon

pressures are varied from 100 psi to 200 psi in a high-pressure gas cell. The fifth harmonic has been scaled up

for visibility. (b) Integrated harmonic yields from the spectra in (A), matched by color. Pressure dependence

of both harmonic orders are fit to the square of the pressure added to a pressure independent contribution

that we assume is due to the silicon fingers. The fifth harmonic is similarly scaled up.

Page 46: Symmetry Breaking and Harmonic Generation in Metasurfaces

33

the expected overlap of the mode and grating predicted in our simulation (Figure 2.2.3). At 200

psi of Argon pressure, the harmonic yields from Argon significantly exceed that of the device

material.

Section 2.4: Final Remarks

Over the course of this Chapter, we have developed an understanding of quasi-bound

states in the continuum as well as a scheme for utilizing them for enhancing harmonic generation

in argon atoms. We have observed a greater than 10-fold enhancement of the third and fifth

harmonic yields in the presence of a BIC- supporting DHCG, with the yield enhancement for

higher-order harmonics being greater due to the higher order dependence on pump intensity. We

have evidence based on the polarization dependence of THG depicted in Figure 2.3.2 that we

have achieved the strong-field, non-perturbative regime, as the integrated THG does not follow a

cos6 𝜃 dependence on the polarization angle. From this observation we note that the hybrid gas-

DHCG technique lowers the required pump input intensity to saturate the THG conversion

efficiency at about 10-7. Furthermore, we have determined that for pressures between 100 psi and

200 psi, the contributions to our total signal from enhanced harmonics within the grating material

and Argon are of the same order of magnitude. With this we confirm the mode profile produced

by our FDTD simulations which predicts a gap-centered mode with leakage into the silicon

fingers.

Further improvement of our hybrid metasurface-gas harmonic scheme requires the ability

to restrict the strong field-confinement mode to only the gas-filled region. Ideally, this would be

accomplished while maintaining relatively large mode areas and all of the tunable wavelength

and quality factor properties afforded to us by the DHCG platform. We propose a device in

Figure 2.4.1, based on a 2-dimensional realization of a DHCG, that meets all of these stated

Page 47: Symmetry Breaking and Harmonic Generation in Metasurfaces

34

requirements. Figure (a) shows a representation of the device, which consists of silicon dimers

containing one circular and one elliptical nanopillar. The perturbation of the unit cell is to the

ellipse minor axis. For a target resonance wavelength in the mid-infrared, the disk major axes

should be 500 nm, and the grating period should be 2.1 μm. Figure (b) shows the BIC mode of

interest in the system, chosen at a wavelength of 3.15 μm. In this system, the strongest field

enhancement no longer spatially overlaps with the silicon. This offers tremendous flexibility to

the proposed system in two forms. First, the contribution to the harmonic signal from any non-

gas sources would be greatly diminished. Second, the quality-factors could be raised to their

fabrication limits, greatly increasing the local powers experienced by the atomic targets without

increasing the likelihood of device damage. Quality-factor tuning of the 2D DHCG observes the

Figure 2.4.1: (a) Proposed 2D DHCG. The height of the grating is 250 nm, disk radii 500 nm, and the period is

2.1 μm. The dimerizing perturbation is to every other disk’s minor axis. (a) FDTD simulation showing the

relative power enhancement in the region marked by a red square in (a) (top view). The strong field-

confinement mode is entirely contained in the Argon gas region, with no leakage into the walls of the grating

or substrate. (c) Simulated polarized reflectance spectrum of 2D DHCGs with alternating minor axes of 500

nm - δ for four magnitudes of the perturbation as well as the unperturbed case. The reflectance peaks at the

BIC Fano-resonance wavelength near 3.15 μm. Smaller perturbations correspond to larger Q-factors and

narrower linewidths.

Page 48: Symmetry Breaking and Harmonic Generation in Metasurfaces

35

required 𝛼−2 dependence, but now the lower-Q peaks in the reflectance curve tend toward the

short wavelength side as in Figure (c). Smaller ellipse minor axes corresponding to larger

perturbations result in less silicon in each unit cell, and hence a lower effective index, which

leads to a blueshift in the reflectance peak.

The potential for greatly amplified field strengths within gas-filled gaps gives renewed

relevance to high-harmonic generation from gases in the presence of metasurfaces. Since a

critical factor in the generation of high energy harmonics is, among other things, a sufficiently

large intensity, the approach we have developed in this Chapter can be extended readily to

higher-order harmonic generation. The possibility to tailor mode profiles to exclude leakage into

the grating, and therefore eliminate the damage that has so far limited plasmonic and dielectric

devices alike, deserves further study. Using a transparent nonlinear medium such as Argon

extends the interaction length of the harmonic generation process well beyond the 10’s of

nanometers available in either metal or silicon-based sources, while dramatically reducing

harmonic absorption when compared with solid-state media. The 250-nm tall gratings used in

our experiments, for example, are tall enough to allow greater coherent buildup of the harmonic

signal, while being sufficiently subwavelength in extent to sidestep phase-matching restrictions.

Lower power requirements open up the possibility to use higher repetition rate sources to initiate

HHG in gases and have the potential to make smaller footprint table-top attosecond sources more

viable. The ease with which Q-factors can be adjusted in the fabrication process means the

DHCG devices shown here can be made to match the bandwidth of a given pump laser.

Alternatively, the resonance linewidth can be chosen to select for the desired linewidth of the

generated harmonics. Moreover, the control provided by DHCGs can be further extended to

simultaneously include both wavefront and polarization shaping.

Page 49: Symmetry Breaking and Harmonic Generation in Metasurfaces

36

Looking to the future, our hybrid gas-DHCG scheme could be applied to the entire suite

of nonlinear optical studies being conducted in atomic systems. Beyond the gas phase, two-

dimensional materials are emerging as a promising nonlinear optical platform moving

forward [56–58] and will become the focus in the next few Chapters. The single- to few- atom

thickness of graphene and transition metal dichalcogenides limits their nonlinear interaction

length, making them prime candidates for integration with field-confining metasurfaces.

Furthermore, the all-dielectric metasurface is a leading platform for meta-lenses, capable of

providing the same level of phase-control and integrability with 2D materials as seen in previous

plasmonic nonlinear metalenses [59]. This is due to the high degree of engineerability of modes

provided by all-dielectric DHCGs, but they also bring the added benefit of eliminating potential

field hotspots in sensitive 2D materials, especially at higher harmonic orders and greater field

intensities.

Page 50: Symmetry Breaking and Harmonic Generation in Metasurfaces

37

Chapter 3: Nonlinear Phononics

Just as quickly as the development of metasurfaces led to their immediate application for

nonlinear optical enhancements, the discovery of two-dimensional materials has led to intense

interest in their electronic and nonlinear optical properties. We will see in this chapter and the

next that lower orders of harmonic generation can be used to gain valuable insights into the

behavior of special two-dimensional materials when excited by intense resonant radiation. We’ll

be especially focused on hexagonal boron nitride (hBN), a thin insulating material with one of

the highest energy infrared-active phonons known. After describing phonons in general, we will

develop a straightforward but robust theory of the nonlinear driving of phonons, referring back to

the harmonic oscillator picture of Chapter 1. One of the major consequences we will observe

from this theory is the opportunity to control the structural characteristics of hBN. Only then in

Chapter 4 will we go on to show that the nonlinear driving of the phonons leads to enhanced

odd-order harmonic generation on resonance, as well as induced even-order harmonic generation

which should, as we’ll soon understand, be forbidden.

Section 3.1: The Charged Harmonic Oscillator Model and Linear Optical Responses

In Chapter 1 we considered a model of electrons oscillating in an anharmonic potential

and from there derived the even and odd order harmonics based on symmetry. Returning to the

same picture, let’s now consider a variation of the situation, where the electron motion can be

neglected and instead the entire atom moves in a harmonic potential. This situation describes the

excitation of a phonon, a collective vibrational mode of all of the different sites in a lattice. An

illustration of two flavors of phonon in a 1-dimensional chain of atoms is given in Figure 3.1.1.

Page 51: Symmetry Breaking and Harmonic Generation in Metasurfaces

38

Alternating atoms with masses 𝑚 and 𝑀, joined by springs, move as either a longitudinal wave,

in which the two species can move together, or a compressional wave where different species

move oppositely. The former is called an acoustic phonon and the latter is known as an optical

phonon. As we will see, optical phonons are the only type that couple directly to incident light,

which will be extremely important for our later experiments.

Consider an atom in a lattice subject to small displacements 𝑟𝐼𝑅 due to infrared radiation.

The formalism of the harmonic oscillator tells us then that the atom must feel a restoring force

𝑭 = −𝝎𝒑𝒉𝟐 𝒓𝑰𝑹 ( 16)

for a potential that is quadratic in the position and frequency of the phonon 𝜔𝑝ℎ. Assuming the

infrared radiation drives the phonon according to the Coulomb force, then we arrive at an

equation of motion for the optical phonon amplitude under excitation by a field 𝐸:

�̈�𝑰𝑹 +𝝎𝒑𝒉𝟐 𝒓𝑰𝑹 −

𝒁

𝒎𝒆𝑬. ( 17)

Figure 3.1.1: The two types of phonon modes illustrated with a toy model of two species of masses joined by

identical springs. The springs provide a restoring force that leads to a harmonic potential, sketched on the

central atom. (a) Acoustic phonons, often drawn as longitudinal waves, do not directly couple to light and are

often seen when atoms of different species (charge) move together. (b) Optical phonons on the other hand

couple to free space radiation and are distinguished by species moving oppositely, which induces a changing

dipole.

Page 52: Symmetry Breaking and Harmonic Generation in Metasurfaces

39

Here 𝑚𝑒 is not the electron mass but the effective mass of the atom, and Z, the Born effective

charge, replaces the charge of the electron in the Coulomb force. A driven oscillator without any

channel through which to lose energy is not a particularly realistic system. The final ingredient

we will need to include recalls the type of resonance lifetime 𝛾 we explored in the context of the

bound state in the continuum. Here the lifetime will be determined by a damping constant Γ

which is inversely proportional to the lifetime. To enforce an exponential saturation on our

equations of motion we make one final addition in the form of:

�̈�𝑰𝑹 + 𝚪𝒓 𝑰𝑹 +𝝎𝒑𝒉𝟐 𝒓𝑰𝑹 −

𝒁

𝒎𝒆𝑬. ( 18)

Oscillatory electric fields with amplitude 𝐸0 therefore lead to solutions for 𝑟𝐼𝑅 with a time

and frequency dependence 𝑟𝐼𝑅 = 𝑟0𝑒−𝑖𝜔𝑡 for phonon oscillations in a damped, driven, harmonic

potential. Combining this with our equation of motion leads us to the final definition of the

phonon trajectory:

𝒓𝑰𝑹(𝒕) =𝒁𝑬𝟎𝒆

−𝒊𝝎𝒕

𝝁(𝝎𝒑𝒉𝟐 −𝝎𝟐−𝒊𝚪𝝎)

. ( 19)

Figure 3.1.2: An electric field (top) applied to a material with a phonon resonance near the electric field

frequency will oscillate and damp similar to the depiction shown on the bottom.

Page 53: Symmetry Breaking and Harmonic Generation in Metasurfaces

40

It’s clear looking at Equation (19) that the maximum amplitude of the phonon 𝑟𝐼𝑅 occurs when

the driving frequency 𝜔 approaches the phonon frequency. This is the resonance condition of a

phonon, and as we will see once we dive deeper into hexagonal Boron Nitride, the resonance

wavelength is a function of atomic masses, binding energies, as well as other factors. The

response of the phonon amplitude for an applied few-cycle pulsed electric field is illustrated in

Figure 3.1.2. The illustrated lifetime is taken to be just a few times that of the pulse duration. In

general however, the lifetime can be anywhere from a few hundred femtoseconds to many

picoseconds, depending on the system, temperature, and other experimental parameters.

Section 3.2: Lattice Anharmonicity

Short optical pulses such as the one illustrated in Figure 3.1.2 tend to be accompanied by

extreme intensities, often in the regime of terawatts per square centimeter when properly focused

down to the diffraction limit. Intensities this high (on pico- or nanosecond timescales) cannot

have their dynamics fully captured by a linear model. Once again, we are presented with a need

for anharmonicity in our model, and just as we came to powerful conclusions about electronic

harmonic generation in Chapter 1, so too will anharmonicity lead us down some novel and

interesting paths in the context of phonons.

We once again approach this problem by redefining the energy potential to include terms

that are higher-order in position. In general, there is no limit to the higher-orders that can

contribute to the potential energy, and it takes the form:

𝑈 =1

2𝜔𝑝ℎ2 𝑟𝐼𝑅

2 + ∑𝑎𝑛𝑟𝐼𝑅𝑛

𝑛=3

+ ∑ 𝑏𝑚𝑟𝐼𝑅𝑛

𝑛,𝑚=1

𝐸𝑚

(20)

Page 54: Symmetry Breaking and Harmonic Generation in Metasurfaces

41

The first summation represents all of the purely ionic contributions to the potential, and the

second summation couples the electron and phonon nonlinearities with their own set of constants

𝑏𝑚.

Let’s now focus on two different conclusions that can be drawn from the various terms of

the potential energy: (1) The generation of pure ionic high harmonics, and (2) the structural

changes induced in the lattice by the motion of the ions. First, let’s investigate the terms that are

Figure 3.2.1: (a) Top: pump electric field, unchanged from the previous section. Bottom: the response of the

ions to the electric field when the equations of motion now include higher order terms in the displacement.

This leads to anharmonicity and the generation of new frequency components. (b) Fourier Transform of the

phonon amplitudes of (a), showing new frequency components at integer multiples of the driving frequency.

Page 55: Symmetry Breaking and Harmonic Generation in Metasurfaces

42

purely phononic in nature [60]. The equation of motion considering only the potential due to the

first series is given by:

�̈�𝐼𝑅 + 2𝛾𝑟 𝐼𝑅 + 𝜔𝑝ℎ2 𝑟𝐼𝑅 + ∑ 𝑎𝑛𝑟𝐼𝑅

𝑛−1

𝑛=3

−𝑍𝐸 = 0

(21)

This series of terms represent the ionic equivalent of high harmonics (just as in Chapter 1), with

each term 𝑟𝑛 representing the 𝑛𝑡ℎ harmonic. As we will see in the following Chapter, under

resonant conditions the harmonics generated as a result of ionic motion can be many times larger

than the more standard electronic nonlinear contributions. Figure 3.2.1 generalizes the phonon

amplitudes of Figure 3.1.2 to include the higher order terms of the new equation of motion. In

Figure 3.1.2 (b), a Fourier transform of the atomic motion shows the generation of the new

frequency components resulting from the atomic anharmonicity.

Now we turn to the effect that the motion of the atoms has on the structure and symmetry

of the lattice. We can subdivide this study further into two sections: the coherent motion of the

phonons as an oscillating electric field is applied, and the rectification of the phonons or creation

of a new, temporary equilibrium position. The coherent motion of the phonons follows trivially

from the driving frequency 𝜔, observed in Equation (19) and in the phonon behaviors of both

Figures 3.1.2 ad 3.2.1.

To understand the rectification effects in the lattice positions, we turn our focus to the

second summation in Equation (20). When considering just the lower term proportional to 𝑟𝐼𝑅2 ,

the equation of motion takes the form

�̈�𝑹 + 𝟐𝜸𝒓 𝑹 +𝝎𝑹𝟐𝒓𝑹 = 𝒃 𝒓𝑰𝑹

𝟐 ( 22)

Where our attention has turned to the position of a Raman mode, specifically. The quadratic term

in the restoring force causes a non-oscillating force along the direction of 𝑟𝑅. This rectified

Page 56: Symmetry Breaking and Harmonic Generation in Metasurfaces

43

displacement, combined with the coherent oscillations, form the basis for the symmetry control

we will exert on hexagonal boron nitride in the next Chapter.

With these derivations completed, we now have both the historically relevant low- and

high- harmonic generation formalisms as well as a robust model of anharmonicity in lattice ions

based on an anharmonic potential. In theory, we understand how resonant driving of atoms can

lead to purely phononic nonlinearities and changes to the positions of lattice sites (and hence

changes to the crystal symmetries). However, in practice the material under study will determine

how different harmonics present themselves when strongly driven. Chapter 4 will now apply

these concepts to hexagonal Boron Nitride.

Page 57: Symmetry Breaking and Harmonic Generation in Metasurfaces

44

Chapter 4: Enhanced Harmonics and

Phonon-Induced Symmetry Breaking in HBN

Optical nonlinearities in solids reveal information about both the in-plane rotational and

out-of-plane inversion symmetries of a crystal. In the van der Waals material hexagonal boron

nitride (hBN) both these symmetries and the linear vibrational properties have led to the rich

physics of mid-infrared phonon-polaritons. However, the role of strong electron-phonon

nonlinearities which we began to derive in the previous Chapter requires further study. In this

Chapter, the hybrid phonon-polariton modes will be derived from first principles, and then the

specific case of hexagonal Boron Nitride will be addressed. Narrowing in on the specific

phonons and Reststrahlen bands of hBN, we will motivate the current tremendous interest in

hBN as a 2D photonics platform.

We then investigate both theoretically and experimentally the rich interplay of phonon

anharmonicity and symmetry in phonon-polariton mediated nonlinear optics. We show that large

enhancements (>30×) of third-harmonic generation occur for incident femtosecond pulses that

satisfy the previously derived resonance condition for the hBN transverse optical phonons in the

mid-infrared. In addition, we predict and observe large transient sub-picosecond duration

second-harmonic signals during resonant excitation, which in equilibrium is forbidden by

symmetry. This surprising result indicates that instantaneous crystal inversion symmetry

breaking can be optically induced and controlled via phonon interactions by both the power and

polarization of the pump laser.

Page 58: Symmetry Breaking and Harmonic Generation in Metasurfaces

45

Section 4.1: The Hyperbolic Phonon-Polaritons of Hexagonal Boron Nitride

Chapter 3 established a general formalism for nonlinear phononics that we will soon be

able to apply to hBN specifically. But before we can understand the nonlinear phononic

properties of hBN, a baseline knowledge of its linear optical and vibrational properties will help

motivate why hBN holds so much promise as a future nonlinear optical platform. Recall the

solution to the equations of motion of a phonon represented by a charged harmonic oscillator in

the previous chapter. For a bulk material we can now define the polarization in the lattice to be

equal to:

𝑷(𝒕) = 𝑵𝒁𝒁𝑬𝟎𝒆

−𝒊𝝎𝒕

𝝁(𝝎𝒑𝒉𝟐 −𝝎𝟐−𝒊𝚪𝝎)

, ( 23)

For the density of unit cells N. Substituting this expression into the definition of the dielectric

function 𝜖 = 1 + 𝑃/𝜖𝑜𝐸, we find that the dielectric function given by the presence of the phonon

resonance is

𝝐(𝝎) = 𝟏 +

𝑵𝒁𝟐

𝝐𝟎𝝁

𝝎𝒑𝒉𝟐 −𝝎𝟐−𝒊𝚪𝝎

. ( 24)

It’s convenient to define a new frequency, commonly referred to as the plasma frequency 𝜔𝑝𝑙, to

be the square root of the numerator. The dielectric function combined with Maxwell’s equations

is sufficient to fully understand the propagation of light in the medium, as well as how the

photons and phonons interact with each other. Substituting the dielectric function into 𝐷𝑖𝑣 𝐸 =

𝜖

𝑐2𝛿2𝐸

𝛿𝑡2 for a plane wave, we find the familiar dispersion relationship relating the momentum 𝑘

and frequencies 𝜔: 𝑘2 =𝜔2

𝑐2𝜖. Following the full substitutions, we find the total dispersion

relationship

Page 59: Symmetry Breaking and Harmonic Generation in Metasurfaces

46

𝝎 = √(𝝎𝑳𝑶

𝟐 𝝐∞+𝒄𝟐𝒌𝟐)± √(𝝎𝑳𝑶𝟐 𝝐∞+𝒄𝟐𝒌𝟐)

𝟐−𝟒𝝐∞𝒄𝟐𝒌𝟐𝝎𝑻𝑶

𝟐

𝟐𝝐∞ , ( 25)

where ϵ∞ is standard notation for a constant dielectric background function. Notice that the

previously named ωphonon now appears as two separate frequencies ωTO and ωLO, for transverse

optical and longitudinal optical, respectively. As depicted in Figure 4.1.1, at the point where the

photon and phonon dispersion curves meet, an anti-crossing emerges in the band structure, and

the crystal hosts new hybrid modes called phonon-polaritons [61]. The primary takeaway from

this avoided crossing is that as the momentum 𝑘 approaches 0 (the Gamma point), the lower

energy TO phonon branch behaves more likes a photon, whereas at high momentum, the

behavior switches and the LO phonon takes the role of the photon-like behavior. Furthermore,

as the momentum goes to zero, the TO phonon energy becomes infinitesimal, while the LO

Figure 4.1.1: Phonon Polariton dispersion branches. The TO phonon energy converges to 0 at small momenta,

whereas the LO branch approaches photon-like behavior (represented by the red dashed line) at large values

of k. Between the two branches is the Reststrahlen band, which plays the role of the bandgap. It is the

frequency range in which photons cannot propagate into the material.

Page 60: Symmetry Breaking and Harmonic Generation in Metasurfaces

47

phonon converges to a finite value. This is significant for our upcoming experiments. It suggests

to us that at normal incidence, the LO phonon branch is more accessible at finite frequencies.

Lastly, the two phonon branches are separated by a well-defined energy gap known as the

Reststrahlen band. Within this range, photons cannot propagate through the material and will

instead be strongly reflected.

At long last we are in a position to apply the tools we have developed to our system of

hexagonal Boron Nitride. This work will focus solely on high-quality, exfoliated flakes of hBN

such as those in figure 4.1.2. To piece together a proper understanding of hBN, we need to

describe (1) the (microscopic) equilibrium structural and symmetry properties, (2) the phononic

band structure of hBN and the particular phonons of interest, (3) the Reststrahlen bands (upper

and lower) of hBN, and briefly (4) the natural hyperbolicity of the hBN phonons.

hBN has an energetically favorable AA’ stacked lattice in equilibrium, with alternating

boron and nitrogen atoms sitting one on top of the other. Figure 4.1.3 shows the hexagonal

arrangement. Not labeled is the experimentally measured b-n bond length of 1.445 Å. hBN

Figure 4.1.2: Microscope image of hexagonal Boron Nitride flakes. The color of the flake can be used as a

means of approximating the thickness of the flake. Atomic force microscopy should be used for more precise

thickness measurements. Flakes typically range in thickness from a few tens of nanometers to a few hundred.

Exfoliated hBN flakes can have diameters of 10s to 100s of microns.

Page 61: Symmetry Breaking and Harmonic Generation in Metasurfaces

48

belongs to a class of crystals called van der Waals materials, named after the type of weak

interlayer bonding common to such other two-dimensional materials as graphene and transition-

metal dichalcogenides. Having inversion symmetry in its equilibrium shape, hBN is not expected

to have any even-order nonlinearities, just as we derived in Chapter 1. The few exceptions,

particularly at interfaces or as a result of symmetry being broken by strain, have been

experimentally observed and quantified [62–64]. These nonlinearities are present only in the

thinnest few-layer samples with an even number of hBN layers, and decrease substantially with

hBN flake thickness.

hBN, like many real materials, has a band structure that cannot be calculated by any toy

model. To pick out the phonon bands of hBN we must turn to Density Functional Theory (DFT),

capable of computing the band dispersions from first principles [65]. Figure 4.1.4 shows both

the monolayer and bulk phonon dispersions, and labels each branch according to the phonon it

represents. For the experiments in this Chapter, which occur at normal incidence only, we are

restricted to the use of those phonons that have finite energy at the Γ point, which represents 𝑘 =

Figure 4.1.3: Various atomic arrangements for the hBN hexagonal lattice in equilibrium, IR-active phonon

excitation, and Raman-active phonon excitation. The equilibrium arrangement of hBN displays inversion

symmetry due to its AA’ stacking. IR-Active phonons such as the TO(E1u) induce a temporary breaking of

this inversion symmetry.

Page 62: Symmetry Breaking and Harmonic Generation in Metasurfaces

49

0. There are only two such bunches of phonons, occurring at slightly greater than 100 meV and

200 meV. These are the out-of-plane and in-plane optical phonons, respectively. The middle

panel of Figure 4.1.3 illustrates the atomic motion corresponding to the higher energy (in-plane)

transverse optical phonon. At a resonance wavelength of 7.3 𝜇𝑚, the TO (E1u) mode can be

accessed by our tabletop laser sources. The exceptionally light constituent atoms of the lattice,

with atomic numbers 5 and 6, are the reason for the high energy, mid-infrared phonon mode.

Furthermore, the fact that the oscillating modes of these phonons (depicted in Figure 4.1.3) cause

a rapidly changing dipole moment in the material is what allows these phonons to couple to

radiation, hence the Optical label.

Figure 4.1.4: Phonon dispersion (band) diagram of monolayer (red) and bulk (black) hBN. Adapted from

Mandelli, et. al. The various phonons are labeled according to whether they are Optical or Acoustic, and

Longitudial, Transverse, or in the Z direction. For normal incidence experiments, only the portions of the

bands around 𝚪 i.e. 𝒌 = 𝟎 are accessed. At over 200 meV, the optical phonons of hBN have one of the highest

known energies, making them easily accessible to our laser system.

Page 63: Symmetry Breaking and Harmonic Generation in Metasurfaces

50

At the beginning of the Chapter, we found that the anti-crossing that occurs between the

phonon dispersion bands and the photon dispersion light-line leads to a Reststrahlen band, a

region akin to a bandgap where light is unable to propagate. Each of the two sets of optical

phonon bands computed for Figure 4.1.4 also lead to their own Reststrahlen bands of

significantly higher reflectivity. Furthermore, these phonons in hBN have one more curious

property of naturally occurring hyperbolicity. Hyperbolicity in this context is the presence of

both positive and negative principle components in the material’s permittivity tensor [66]. Both

the hyperbolicity and Reststrahlen bands can be observed in a plot of the permittivity, defined

by:

𝜺(𝝎) = 𝜺∞(𝟏 +𝝎𝑳𝑶𝟐 −𝝎𝑻𝑶

𝟐

𝝎𝑻𝑶𝟐 −𝝎𝟐−𝒊𝝎𝜸

) ( 26)

Figure 4.1.5: Real parts of the permittivity along the z direction (into the crystal) and the transverse direction

(in-plae). The previously derived phonon branches correspond to strong peaks in the real permittivity. In the

regions where one component is positive and the other negative, the dispersion is hyperbolic. These regions

correspond to the Reststrahlen bands, shaded in gray.

Page 64: Symmetry Breaking and Harmonic Generation in Metasurfaces

51

where the high energy TO and LO transverse frequencies are 1370 and 1614 𝑐𝑚−1 and the lower

energy out-of-plane phonons have TO and LO frequencies of 760 and 825 𝑐𝑚−1, respectively.

The real parts of the permittivity and the corresponding Reststrahlen bands for the two phonons

are plotted in Figure 4.1.5. In Figure 4.1.6, we also see a clear example of the linear optical

effects of the Reststrahlen region: sharply enhanced reflectivity.

The hyperbolic permittivity is crucially important for applications demanding tight

confinement [67], and is one of the driving motivations for developing hBN as a nonlinear optics

platform. This is because the hyperbolic polaritons allow for propagation of the otherwise

forbidden energies in the Reststrahlen bands. Sub-diffraction confinement [67], negative

refraction [68], and hyperlensing [69,70] are all applications made possible by the strong

confinement. In the coming sections we will develop our own application, making use of the

tight optical confinement to generate large phonon amplitudes along the lines of Chapter 3. This

will lead to enhanced and novel nonlinear optical effects.

Figure 4.1.6: Enhanced reflectivity at the upper Reststrahlen frequency range of hexagonal boron nitride.

Page 65: Symmetry Breaking and Harmonic Generation in Metasurfaces

52

Section 4.2: Frequency Dependent Third Harmonic Enhancement From Phonons

We are now in a position to properly understand the nonlinear driving of the hBN

phonons, beginning from a simulated model of the anharmonic oscillations, straight through to

the phonon-resonant experiments. We first characterize theoretically the ionic displacements in

bulk hBN under resonant excitation with 25-fs FWHM pulses by performing time dependent

density-functional theory (TDDFT) simulated atomic oscillations spanning 200-fs, or roughly 8

times the theoretical pulse duration (see Figure 4.2.1). For a modest input intensity of 7 x 1010

W/cm2, we estimate that the phonon amplitude is 1% of the equilibrium lattice constant. While

the period of the lattice oscillation is 25-fs, which is consistent with the expected phonon

frequency, the relaxation time, described back in Chapter 3 cannot be theoretically determined

due to a lack of dissipative pathways in TDDFT. The amplitudes of atomic motion are plotted as

a function of pump intensity in Figure 4.2.1 (b). The displacements predicted by TDDFT

Figure 4.2.1: (a) Simulated atomic displacements of boron and nitrogen ions in TO (E1u) excited hBN. A 25-fs

FWHM, 1 x 1012 W/cm2 pulse excites the lattice dynamics. The TDDFT simulations do not include any

damping terms through which to estimate the relaxation time. (b) Peak amplitude of atomic displacements as

a function of pump intensity, fit to I1/2 with a small linear-in-intensity correction. Displacements nearing 5%

of the equilibrium lattice constant are achievable before the onset of damage.

Page 66: Symmetry Breaking and Harmonic Generation in Metasurfaces

53

calculations are fit by I1/2 with deviations appearing at large intensities and reach nearly 5% of

the equilibrium lattice constant (2.5 Å) [71] at 10 TW/cm2. The deviations strongly indicate the

presence of anharmonicity emerging in the simulation, and as we will now see, the expected

harmonic frequencies can be observed in the computed spectrum. The time-dependent electronic

current is extracted, and from this we generate the theoretical harmonic spectra employed

throughout this work. This process is repeated for wavelengths below, at, and above 7.3 μm, and

the integrated theoretical THG yields are plotted in Figure 4.2.2 (a) as green dots.

To perform the TDDFT simulations, the time evolution of the wave functions and the

evaluation of the time-dependent electronic current were computed by propagating the Kohn-

Sham equations in real space and real time, as implemented in the open source Octopus

code [14,72], in the adiabatic LDA [73] (the findings and trends discussed in the present work

are robust with different functionals) and with semi-periodic boundary conditions [72]. All

calculations were performed using fully relativistic Hartwigsen, Goedecker, and Hutter (HGH)

pseudopotentials [74]. The real-space cell was sampled with a grid spacing of 0.4 bohr and the

Brillouin zone was sampled with a 42 × 42 × 21 k-point grid, which yielded highly converged

results. The boron nitride bond length is taken here as the experimental value of 1.445 Å. The

laser was treated in the dipole approximation using the velocity gauge (that implies that we

impose the induced vector field to be time dependent but homogeneous in space), and we used a

sin-square pulse envelope. In all of our calculations, we used a carrier-envelope phase of f =

0 [75]. The full harmonic spectrum is computed directly from the total electronic current j(r, t)

as:

𝑯𝑯𝑮(𝝎) = | 𝑭𝑻(𝜹

𝜹𝒕∫𝒅𝟑𝒓 𝒋(𝒓, 𝒕))|

𝟐, ( 27)

Page 67: Symmetry Breaking and Harmonic Generation in Metasurfaces

54

where FT denotes the Fourier transform. The atomic vibrations of phonon modes are prepared

with the following two methods: (i) the time-evolution from a distorted atomic configuration

along the phonon modes of 1% of the bulk hBN lattice. (ii) application of pump laser pulses with

the same frequencies and polarizations of phonon modes. Our calculations confirm the two

methods are equivalent in the simulations of high-harmonic generation. TDDFT simulations

were performed primarily by members of the Rubio Group, Max Planck Institute.

Turning to the experimental side of things, we show the integrated and normalized THG

amplitudes for a range of pump wavelengths from 3 μm to 9.5 μm in Figure 4.2.2 as blue dots,

which are in excellent agreement with the calculations discussed in Figure 4.2.1 and plotted as

Figure 4.2.2: (a) Normalized third-harmonic generation yields of 120-fs pulses as a function of pump

wavelengths throughout the mid-IR. THG yields are below the noise level for all wavelengths less than 6 µm or

greater than 9 µm. Within a roughly 1 µm bandwidth of the phonon-polariton resonance, a THG

enhancement of 30x is observed. The black line is a Lorentzian fit to the data (blue dots) with full-width at

half-maximum of 500 nm. The green dots were obtained by integrating the third-harmonic signal in TDDFT

simulations and show excellent agreement with experiments. (b) Normalized intensity dependence of THG

pumped on-resonance at 7.3 µm and measured at 2.43 µm. The data is a close fit to I3, indicating that the

nonlinear process is scaling perturbatively.

Page 68: Symmetry Breaking and Harmonic Generation in Metasurfaces

55

green dots in Figure 4.2.2. The third-harmonic exhibits a strong peak for pump wavelengths near

the TO phonon resonance at λ = 7.3 µm. As this wavelength is far from any electronic or

excitonic resonances, the enhancement must be phononic in nature, described by the theory of

Chapter 3. We fit the data to a Lorentzian and extract a resonance full-width at half-maximum of

500 nm. THG yields are below the noise level for all λpump less than 6 µm or greater than 9 µm,

compared to that of the resonant signal which yields at least a 30-fold increase, and thus the

phononic enhancement of the THG coefficient at the phonon-polariton wavelength is

significantly greater than the purely-electronic component in this regime. In Figure 4.2.2 (b) we

plot the measured intensity dependence of the THG signal for λpump = 7.3 µm. The fit to a cubic

function indicates that the measured nonlinearity is third-order and that the scaling is

perturbative, even at high intensities [76]. We note that a similar effect has been observed in the

phononic second-harmonic generation of LiNbO3, which also remained in the perturbative

regime at higher-than-expected intensities. Also, we propose that subwavelength structures that

support confined phonon-polaritons [67,77] can drastically enhance the phonon-induced

nonlinearity in the same way that they dramatically enhance electronic nonlinearities.

Figure 4.2.3: Experimental setup for THG experiments. Detection is performed with PbS and MCT detectors,

a lock-in amplifier, and boxcar-averaging.

Page 69: Symmetry Breaking and Harmonic Generation in Metasurfaces

56

As previously mentioned, we performed the nonlinear experiments on high-quality hBN

flakes with thicknesses of 10 to 50 nm and typical sizes of tens of microns. The flakes are

exfoliated onto a CaF2 substrate, chosen for its high transparency in both the visible and mid-

infrared and its relatively small nonlinearity. Sample fabrication was performed primarily by

members of the Hone Group, Columbia University. For our long-wave infrared pump pulses we

utilize an optical parametric amplifier (OPA, Light Conversion HE Topas Prime) pumped by an

amplified Titanium-sapphire laser system (Coherent Legend Elite) operating at a 1-kHz

repetition rate with 6 mJ of pulse energy. The OPA produces 60-fs duration signal and idler

pulses with center wavelengths in the near-IR. The parametric amplifier output is then used to

seed an additional difference frequency generation (DFG) module for all mid-infrared

measurements from λpump = 3 to 10 μm with pulse durations ranging from 70- to 120-fs. Pulse

intensities are consistently set below the hBN damage threshold, which we estimate to be 50

TW/cm2. For THG experiments the pump is focused onto an hBN flake using a 2-cm focal

Figure 4.2.4: (a) HHG spectrum for the pump laser parallel with TO (E1u) mode. (b) HHG spectrum for the

pump laser parallel with LO (E1u) mode. Here, we employ an in-plane electric field with a wavelength of 𝝀 =

800 nm and an intensity of I = 1012 W /cm2, and a pulse duration of 25-fs full width at half maximum.

Page 70: Symmetry Breaking and Harmonic Generation in Metasurfaces

57

length CaF2 lens, and the emitted THG signal is collected in a transmission geometry by an

identical lens. After the residual pump beam is rejected by a short-pass filter, the remaining THG

is measured on a PbS detector for harmonic wavelengths λTHG below 1.7 μm, and on a liquid

nitrogen-cooled MCT detector for all λTHG greater than 2 μm. Figure 4.2.3 illustrates the premise

of the THG setup.

We also performed TDDFT simulations of the wavelength dependence of a higher-order

harmonic (HHG) spectra of bulk hBN. See Figure 4.2.4 for two different pump lasers with a

wavelength of λpump = 7.3 µm (polarized parallel to the TO mode) and λpump = 6.2 µm (polarized

parallel to the LO mode). Changing the wavelength and polarization of the pump laser can lead

to the excitation of different phonon modes and lead to significant modulation of the HHG

spectra. Excitation of either the TO or LO mode leads to noticeable modifications of the high-

harmonic spectra, with the TO (E1u) enhancement being one order of magnitude greater than that

caused by LO excitation. Furthermore, more intense laser pulses can introduce larger atomic

Figure 4.2.5: HHG spectra for different pump laser intensities. Here, the polarizations of pump and probe

laser are parallel with TO (E1u) mode (pump laser with a wavelength of 𝝀 = 7300 nm). For the probe laser, we

use an in-plane driving electric field with a wavelength of 𝝀= 800 nm and an intensity of I = 1012 W/cm2, and a

pulse duration of 25-fs full width at half maximum.

Page 71: Symmetry Breaking and Harmonic Generation in Metasurfaces

58

displacements and lead to larger nonlinearity. As seen from a pump intensity of 2.5 x 1010

W/cm2, the high-harmonic yields can be increased in a wide energy regime, and the high

harmonic generation plateau is enhanced (Figure 4.2.5), which is attributed to the increased

atomic movement and enhanced nonlinearity.

The content of this Section has served primarily as confirmation of the anharmonic model

of Chapter 3. By upgrading from a ball-and-spring model to the far more sophisticated

techniques of Density Functional theory analysis, we have gained further insight. This includes

the extension of the nonlinear enhancement to much higher harmonic orders, but also has the

potential to greatly expand our understanding of the combined electron-phonon coupling effects

in the intense strongly-coupled light-matter regime. Regardless, we come away from the third

harmonic experiment understanding that when driven on resonance, the THG due to lattice

anharmonicity can be far greater than the electron processes of decades past.

Figure 4.3.1: TDDFT simulations show the emergence of even-order nonlinearity during resonant excitation of

the TO phonon mode.

Page 72: Symmetry Breaking and Harmonic Generation in Metasurfaces

59

Section 4.3: Time- Resolved Even-Order Nonlinearity in hBN

The generation of optical harmonics in solids provides a window into the optical

susceptibility, band-structure, and underlying symmetries of crystals, each of which can

dramatically affect the nonlinear frequency-conversion process [2,16,78]. Symmetries, more so

than any other factor, dictate the allowed higher-order processes in a given nonlinear

system [79]. Following the discussions of lattice control outlined earlier in this work, hBN

presents us with a fantastic opportunity to explore specific instances of lattice motion away from

equilibrium.

Multilayer hBN has inversion (and 6-fold rotational) symmetry due to the natural 2H

stacking of its van der Waals structure [80]. As briefly described in Section 4.1, any contribution

to SHG in few- to many- layer hBN is therefore restricted only to the broken inversion symmetry

cases of interfaces and an odd number of layers and is inherently weak [64]. By conducting

ultrafast pump-probe experiments we show that the crystal inversion symmetry can instead be

Figure 4.3.2: Experimental setup for pump-probe SHG experiments. The time-delay is controlled by a

mechanical delay stage with sub-1 µm step size. The pump and probe are both focused onto the sample with a

reflective objective with 0.5 numerical aperture. Detection is performed with a silicon photomultiplier tube

and lock-in amplifier.

Page 73: Symmetry Breaking and Harmonic Generation in Metasurfaces

60

controllably broken by excitation of the IR-active TO (E1u) phonon, which leads to a finite

second-order susceptibility. For large amplitudes of the laser-driven lattice deformation, our

simulations reveal the emergence of an ultrafast, transient SHG signal from a secondary 800 nm

laser pulse, as shown in Figure 4.3.1 We distinguish this induced SHG from any nearby odd-

order processes by performing the calculations in the following two different ways: (i) with a 7.3

µm pump field, and (ii) using the time-evolution of an equivalently distorted lattice and no

resonant photons. In both cases the signal at harmonic order 2 emerges, confirming the principal

role of the broken symmetry. Furthermore, simulations confirm that in the case of the Raman-

active hBN TO (E2g) phonon which preserves inversion symmetry, even-order nonlinearity

cannot be observed.

For our SHG measurements we modify the experimental setup to a pump-probe scheme

with the addition of a 792 nm, 45-fs pulse from the same amplified Ti-Sapphire laser, shown in

Figure 4.3.2. A variable time-delay separates the 7.3 µm pump pulse that we use to excite the

phonon, from the near-IR probe which produces the SHG signal at 396 nm. The intensity of both

pulses is maintained at or below the (TW/cm2) range, which is below the hBN damage threshold.

The time delay between pulses is controlled by a mechanical delay line with sub-1-µm step size.

The polarization of the pump and probe beams are independently rotated with zero-order half-

wave plates and wire-grid polarizers before the pulses are combined on a beamsplitter. The

collinear pump and probe are focused onto an hBN flake using a reflective objective (NA = 0.5),

which ensures the same focal plane for the two beams with very different wavelengths. The SHG

signal produced by the 792 nm pulse can be collected either in the reflection geometry by the

Page 74: Symmetry Breaking and Harmonic Generation in Metasurfaces

61

reflective objective, or in transmission by a CaF2 lens. The signal is then directed through a

bandpass filter with a 10-nm bandwidth to reject the residual 792 nm and 7.3 µm light, as well as

any unwanted χ(3) signals, before detection on a fast photomultiplier tube (PMT) and lock-in

amplifier.

The experimentally measured SHG signal at 396 nm is presented in Figure 4.3.3 as a

function of the time delay between 792 nm and 7.3 µm pulses. The probe pulse from an

amplified Titanium-Sapphire laser and the pump pulse from a mid-infrared OPA and difference

frequency generation module are scanned in time by a mechanical delay line. The powers and

relative polarizations are set with filters and half wave plates (HWP), and the two beams are then

combined on a beamsplitter before being focused onto the sample by a reflective objective.

Figure 4.3.3: (a) Time-resolved SHG yield (normalized) of the 792 nm probe pulse. While the pumps are

temporally overlapped, an ultrafast second-order nonlinearity is measured. The inversion symmetry is

restored following a 200-fs time constant, or about twice the pulse duration. The appearance of wings in the

time-delay scan is a result of a non-perfectly Gaussian pulse, a result of strong atmospheric absorption. (b)

Zoom in of the region outlined by the dotted box in (a). (c) Fourier Transform of the time-series measurement.

The peak in the spectrum at 7.3 𝝁𝒎 confirms that the phonon is at least partly coherent with the field.

Page 75: Symmetry Breaking and Harmonic Generation in Metasurfaces

62

When the probe pulse precedes the pump pulse, no SHG is measured, indicating that the

interface SHG and odd layer-number contributions are below the noise floor. The time-resolved

SHG then displays a finite signal at the zero-time delay, when the probe pulse’s arrival coincides

with the strong excitation of the hBN phonon-polariton. The SHG signal relaxes back to zero

with a time constant of approximately 120 fs, which is approximately twice the pump pulse

duration. When pumped far off-resonance, no SHG is measured. The presence of fast oscillations

in the measured SHG on top of another transient but finite signal is indicative of a combination

of effects. Whereas the fast fs-timescale oscillations appear to be coherent with the excited

phonons (based on the observation that the Fourier transform of the time-resolved measurement

peaks at the pump wavelength), the constant pedestal suggests a second, rectified signal. This

can occur, for example, when the coupling of the IR-active phonon to a Raman-active mode

creates a new equilibrium position of the lattice atoms [81].

In Figure 4.3.4 (a) and (b) we show the dependence of the SHG yield on the intensity of

the probe and pump, respectively. A quadratic dependence of the SHG intensity on the probe

power is observed, as expected for a second-order nonlinear process. The linear scaling of the

Figure 4.3.4: (a) Dependence of measured SHG yield on probe power. (b) Dependence of measured SHG on

pump power. The SHG yield increases linearly with the phonon driving intensity.

Page 76: Symmetry Breaking and Harmonic Generation in Metasurfaces

63

SHG signal with respect to the mid-IR intensity in (b) implies a direct dependence of the

transient SHG process on the ionic displacements. These dependences match those reported for

SrTiO3, in which an SHG yield in low temperature experiments was found to steadily increase

over hours of total pump exposure and persist for hours after, with ps-scale modulations that also

increase in frequency with total exposure time [81]. Another similar effect with a very long

response time was observed in the naturally inversion-symmetry-broken van der Waals material

WTe2 [82]. However, in this case the effect of a shear strain and lattice displacement was to

eliminate the naturally present SHG, which is the opposite of our observed effect.

Section 4.4: Polarization Dependence of Second Harmonic Generation

In Section 4.3 we treated crystals of hBN as though they were perfectly uniform, making

no remarks on the rotational anisotropy of its crystal structure. This obviously couldn’t be further

Figure 4.4.1: Theoretical (left) and experimental (right) polarization dependent SHG. White dashed lines on

either plot represent the 60o rotational symmetry of a hexagonal lattice. The functional form of the

polarization dependence is given in the inset on the left. Here, 𝜽 and 𝝓 are the angles of the pump and probe

with respect to the ZZ axis, respectively.

Page 77: Symmetry Breaking and Harmonic Generation in Metasurfaces

64

from the truth, evidenced by our earlier discussion of the hexagonal structure of the hBN lattice.

Common nomenclature when discussing the high-symmetry directions of any hexagonal lattice

is to define two perpendicular axes, the so-called zigzag (ZZ) and armchair (AC) directions.

Looking at Figure 4.1.3, we would say that the y-direction is the ZZ axis, whereas the x-direction

must be AC. Obviously, these two axes periodically repeat. A 120o rotation of the ZZ axis will

return another ZZ axis, for example. The harmonic oscillator model used to discuss the THG of

ions on a theoretical basis did not leave room for a rotationally anisotropic restoring force. In a

real crystal of hBN, the AC and ZZ directions will feel different restoring forces, and hence will

have a substantial impact on the symmetries of measured harmonic emission. Next, we

establish the dependence of the ultrafast SHG on the orientation of the pump and probe

Figure 4.4.2: Linecuts from the full

polarization plot. Left (red linecut) shows the

typical behavior when the pump is polarized

along the AC direction. Right (blue linecut) is

when the pump is polarized along the ZZ

direction. No matter the pump polarization,

yields are largest along ZZ axes that are at

least partly excited by the pump. The yield

does not go identically to 0 along the AC axes,

perhaps because the slightly higher energy,

perpendicular phonon is weakly excited by the

tail of our pump spectrum.

Page 78: Symmetry Breaking and Harmonic Generation in Metasurfaces

65

polarizations with respect to the crystal high-symmetry axes. Figure 4.4.1 gives the total

normalized SHG yield for 360o rotation of both pulses (180o rotation is measured and the data is

then mirrored). We observe a polarization behavior unique from either the inherent 6-fold χ(2) or

isotropic χ(3) symmetries of purely electronic hBN nonlinearities [83]. Specifically, the emission

closely follows the functional form,

𝑺𝑯𝑮(𝜽,𝝓) = [𝜶 𝑪𝒐𝒔(𝟑𝜽)𝟐 + 𝜷 𝑺𝒊𝒏(𝟑𝜽)𝟐]𝑪𝒐𝒔𝟐(𝜽 − 𝝓), ( 28)

where θ and ϕ are the angles of the pump and probe relative to the zigzag (ZZ) axis of the

crystal, respectively, and α and β determine the relative strengths of the emission along the ZZ

and Armchair (AC) axes, respectively. This form was phenomenologically derived, based on the

following key observations: each of the AC and ZZ high-symmetry directions contributes a 6-

fold symmetric flower pattern to the SHG, just as one expects in a hexagonal lattice that

naturally lacks inversion symmetry such as [64]. The relative strengths of these two contributions

Figure 4.4.3: TDDFT computed high-harmonic generation spectra for pump and probe pulses co-polarized

along the ZZ (blue) and AC (red) directions. The TO(E1u) phonon present along the ZZ axes leads to the

greatest even order nonlinearity.

Page 79: Symmetry Breaking and Harmonic Generation in Metasurfaces

66

are left as fit parameter, to be compared with TDDFT theory later. Next, the SHG yield ought to

be greatest when the pump and probe are co-polarized, and could be expected to vanish when

cross-polarized. This leads to the final cos2(𝜃 − 𝜙) term at the end of the expression.

SHG yields peak only along ZZ axes that are being resonantly driven with a phonon-

polariton. This is most clearly visible in the linecuts of the probe polarization dependence for

pump fields aligned parallel to the AC and ZZ axes, given in Fig. 4.4.2. Even when the pump

excitation is aligned with an AC axis, the two adjacent ZZ oriented TO(E1u) phonons break the

inversion symmetry, and we observe SHG, whereas the ZZ axis at exactly 90o from that

excitation shows no emission. From Figure 4.4.2 we determine that the phonon-mediated SHG is

at least 3 times greater parallel to ZZ than AC directions. This is supported by time-dependent

density functional theory (TDDFT) simulations in Figure 4.4.3 which identifies even-order

nonlinearity along both symmetry axes, though much greater for the TO(E1u) phonon than the

relative π phase LO(E1u).

Section 4.5: Outlook on the Future of Nonlinear Phononics in hBN

Over the course of this Chapter we have extended the light-matter interactions confined by

the hyperbolic nature of the hBN phonon dispersion to a strongly nonlinear regime. In doing so,

we demonstrated that the large electron-phonon coupling leads to a nearly two order of magnitude

enhancement of SHG and THG. Efficient coupling of light to hBN phonon-polaritons at normal

incidence places stringent requirements on the allowed optical excitation wavelength. For the free-

space wavevector k = 0, the required photon wavelength of 7.3 µm is fixed, independent of flake

thickness [84]. With the energy-momentum conditions met, we have shown that efficient coupling

of optical energy into the hBN lattice leads to anharmonic driving of the atoms. Phonon amplitudes

can reach a few percent of the equilibrium lattice constant long before the onset of laser-induced

Page 80: Symmetry Breaking and Harmonic Generation in Metasurfaces

67

damage according to TDDFT simulations. The anharmonicity of the ionic motion leads to a novel,

enhanced nonlinearity, for which hBN is the most attractive platform in the mid-IR due to its

relatively light constituent atoms. Saturation of the THG yield below its perturbative cubic scaling

was not observed and is more likely to occur closer to the onset of sample damage. The extension

of nonlinear phonon enhancement to higher-order nonlinearities and to two-color high-harmonic

generation techniques will likely advance our understanding of electron-phonon coupling and

nonlinearly-driven lattices.

The experiments of this chapter clearly establish SHG as a sensitive probe for ultrafast

symmetry monitoring and eventually control in hBN, complementing earlier demonstrations on

broken-inversion-symmetric transition metal dichalcogenides, while reducing the relaxation time

scale by more than two orders of magnitude. It’s worth noting that the observed transient broken-

inversion symmetry and atomic displacement can also be interpreted as a strong photo-induced

strain field when the phonon mode is resonantly excited [85]. The promise of optically-

controllable strain in hBN lends itself to numerous other applications, from the tuning of

photoluminescence [86] to optical control of van der Waals heterostructures, for which hBN is a

common encapsulating material.

Page 81: Symmetry Breaking and Harmonic Generation in Metasurfaces

68

Chapter 5: Graphene Harmonic Generation and

Time- Reversal Symmetry Breaking

Graphene, a single-atom layer of carbon atoms arranged in a hexagonal lattice, has been

one of, if not the most hotly studied materials on the planet for 15 years running. First exfoliated

from its bulk form graphite by Geim and Novoselov [87] (who received the 2010 Nobel prize in

physics for their efforts), graphene has now earned its place as one of the most impressive two-

dimensional materials for its electronic [88], optical [89,90], mechanical [91], and thermal [92]

properties. To cover even a fraction of these would take (and surely has taken) many, many

dissertations. Therefore, our discussion of graphene will begin with an introduction only of its

more crucial electronic properties regarding its band structure and lack of a band gap, before

immediately moving on to its optical properties.

From there we will discuss an interesting discrepancy in the early literature on graphene

high harmonic generation, in which two seemingly identical experiments found completely

opposite optical behaviors. On the one hand, graphene appeared to behave like a solid as one

might expect [93]. And yet in the other experiment, graphene displayed all of the properties of

atomic (gaseous) harmonic generation that was characterized by the three-step model all the way

back in Chapter 1 [94]. The resolution of that discrepancy will take us down an extremely novel

path, where we will not only come to understand how both of these papers were in fact correct,

but graphene (the incredible system that it is) can be smoothly tuned from one of the two regimes

to the other.

Taking advantage of the incredible fabrication groups available to us at Columbia

university, we will then take a step beyond either of the two previous demonstrations. By

utilizing large single-crystals of exfoliated graphite, we will gain valuable insights into the

Page 82: Symmetry Breaking and Harmonic Generation in Metasurfaces

69

rotational and symmetrical behaviors of graphene in the context of elliptically pumped harmonic

generation. In a beautiful demonstration of time-reversal symmetry breaking caused by a chiral

optical field, we will be able to use third and fifth harmonic generation in graphene for two final

experiments: (i) to reveal spatial symmetries that should be entirely unmeasurable in harmonic

experiments and (ii) to probe a population imbalance of electrons in the so-called “valleys” of

the band structure. Though theoretically predicted (many times over), no experimental

measurements of valley polarization in carbon materials have been demonstrated prior to our

own experiments.

Section 5.1: An Introduction to Graphene

Electronic Properties

Like hBN, graphene is a material with a hexagonal lattice that belongs to the class of

easily-exfoliated van der Waals bonded crystals. Having only one species of atom (Carbon) and

a bond length of 14.2 Angstroms [95], graphene is one degree simpler in its structure than hBN.

For that reason, and because it will reveal one of the most crucial properties graphene offers, we

will go through the motions of estimating the graphene band structure. Rather than using either a

1D toy model or the far more complex Density Functional Theory approaches, for graphene we

will turn to the tight binding method for computing the band dispersions. Those readers with

some training in quantum mechanics will be familiar with the concept of the Bloch wave

functions, of which we will need two (𝜓𝐴 𝑎𝑛𝑑 𝜓𝐵), for two sublattices of graphene. We will

ultimately be interested in the total wave function

𝚿(𝐫) = 𝜶𝝍𝑨 + 𝜷𝝍𝑩 ( 29)

Page 83: Symmetry Breaking and Harmonic Generation in Metasurfaces

70

Consider N total atoms, with those in the A sublattice having positions 𝑟𝐴 and those in the B

sublattice having positions 𝑟𝐵. Then the two Bloch wave functions in terms of a sum of real

atomic orbitals 𝜙 are defined standardly as

𝝍𝑨(𝒓) =𝟏

√𝑵∑ 𝝓𝑨(𝒓 − 𝒓𝑨) 𝒆

𝒊 𝒌 𝒓𝑨𝒓𝑨 ( 30)

𝝍𝑩(𝒓) =𝟏

√𝑵∑ 𝝓𝑩(𝒓 − 𝒓𝑩) 𝒆

𝒊 𝒌 𝑩𝒓𝑩 ( 31)

Bloch’s theorem imposes a periodicity requirement on the total wavefunction:

𝚿𝐣,𝐧 (𝒓 + 𝑹𝒋) = 𝒆𝒊 𝒌 𝑹𝒋𝛙𝐀,𝐣 (𝒓) = 𝟏

√𝑵 𝒆𝒊 𝒌 𝑹∑𝝓(𝒓 − (𝒓𝒋,𝒏 − 𝑹)) 𝒆𝒊 𝒌(𝒓𝒋,𝒏−𝑹𝒋) ( 32)

Where the new subscript j indicates a Bravais lattice vector. The Bloch wavefunction must

satisfy the Schrodinger Equation to continue the derivation, where the total Hamiltonian is given

by the combinations of the two sublattices. In matrix form

Figure 5.1.1: Tight Binding Model estimation of the band energy dispersion of graphene, seen from above. The

6-fold symmetry of the lattice is reflected in its dispersion. Furthermore, some of the high-symmetry points

have been labeled, including the 𝚪 point at the Brillouin zone center and the surround degenerate K and K’

points. At the K and K’ points, the band gap between conduction and valence goes to zero, and the dispersion

is linear.

Page 84: Symmetry Breaking and Harmonic Generation in Metasurfaces

71

𝑯 = (< 𝝍𝑨(𝒓)|�̂�|𝝍𝑨

∗ (𝒓) > < 𝝍𝑨(𝒓)|�̂�|𝝍𝑩∗ (𝒓) >

< 𝝍𝑩(𝒓)|�̂�|𝝍𝑨∗ (𝒓) > < 𝝍𝑩(𝒓)|�̂�|𝝍𝑩

∗ (𝒓) >) ( 33)

and the change has been made to the bra-ket notation of Dirac for ease of viewing. We can

recognize that for a properly normalized wave function, the diagonal terms are identically 1. The

eigenvalues can be found by the solution of the characteristic equation of what remains:

det[𝐻𝑘 − 𝑆𝑘 𝐸𝑘] = 0. Here, the k subscript hints at the fact that our ultimate dispersion

equation will be found in terms of the wavevectors once again. 𝐸𝑘 are the corresponding energies

and 𝑆𝑘 are the wavefunction inner products of Equation (33). Two solutions can be found for the

energies now.

𝑬(𝒌) = ± 𝒉 √𝟏 + 𝟒𝐜𝐨𝐬 (𝟑

𝟐𝒌𝒙𝒂) 𝐜𝐨𝐬 (

√𝟑

𝟐𝒌𝒚𝒂) + 𝟒𝐜𝐨𝐬𝟐 (

𝟑

𝟐𝒌𝒙𝒂) ( 34)

Where h is the energy needed to hop to the nearest neighbor lattice site. The dispersion relation

Equation (34) is plotted top-down in Figure 5.1.1. Believe it or not, we’ve already stumbled onto

Figure 5.1.2: The three possible doping states of the graphene Dirac point. In the center, the undoped

“intrinsic” graphene has a Fermi level that is exactly at the Dirac point. If electrons are added to the system,

that Fermi level rises as in the n – doped case, and when charge is removed, the left-hand p – doped situation

occurs. The Fermi energy is defined in both doping scenarios as the energy difference from the Dirac point i.e.

the charge neutral point.

Page 85: Symmetry Breaking and Harmonic Generation in Metasurfaces

72

one of the great properties of graphene with this simple derivation. First, the hexagonal 6-fold

symmetry of the lattice is clearly reflected in the symmetries of the bands. In Figure 5.1.1, a few

of the significant high-symmetry points have been labeled, including the Γ point at the center of

the Brillouin zone, and the surrounding degenerate K and K’ points. At the K and K’ points, the

dispersion of the bands becomes completely linear, and the gap between the valence and

conduction bands completely closes. This point, know as the Dirac point, will be the source of all

of our interest in graphene from this point forward.

The reason for this extreme intrigue (for our research at least) is the way in which the

Dirac point interacts with the Fermi level of the graphene. The Fermi level (or Fermi energy) is

the energy level of the highest energy electron in the system. In an undoped graphene sample, the

Fermi level sits precisely at the Dirac point, where the conduction and valence bands meet. If

additional charge is added to the system, the Fermi level rises, a situation known as n – doping.

If the opposite were to occur, the Fermi energy would be lower, and that would be called p -

doping. All three of these are visualized in Figure 5.1.2. With that derivation completed we can

now move on to the optical properties of interest.

Optical Properties

We just showed in the previous Section that graphene lacks a band gap. However, for all

intents and purposes the presence of the easily doped Fermi level near the Dirac point gives us an

effective band gap to control. The Fermi level can be rapidly altered by applying a gate voltage

to graphene, though the exact speed and reproducibility of the doping will depend on the size and

quality of the sample, whether it is encapsulated, and other factors of the substrate and

environment. In order to see how the optical conductivity of the graphene changes as a function

of this doping level, the common theoretical treatment is to imagine graphene as a two-

Page 86: Symmetry Breaking and Harmonic Generation in Metasurfaces

73

dimensional conductive sheet. The general form of the graphene conductance [96] will not be

fully derived here. Instead, let us consider only the contributions to the optical conductivity due

to electrons that are trying to jump from the valence band up in to the conductance band. This

interband conductivity has been derived previously [97,98] and is given by

𝝈𝒊𝒏𝒕𝒆𝒓𝒃𝒂𝒏𝒅(𝝎) =𝒆𝟐

𝟒ℏ[𝟏

𝟐+

𝟏

𝝅𝐚𝐫𝐜𝐭𝐚𝐧 (

ℏ𝝎−𝟐𝑬𝑭

𝒌𝑩𝑻) −

𝒊

𝟐𝝅𝐥𝐧 (

(ℏ𝝎+𝟐𝑬𝑭𝟐)

𝟐

(ℏ𝝎−𝟐𝑬𝑭𝟐)

𝟐+(𝟐𝒌𝑩𝑻)𝟐

)] ( 35)

Where e is the electron charge, ℏ is as always, the reduced Planck constant, and 𝑘𝐵𝑇 is a

measure of the thermal energy, which at room temperature is much lower than the Fermi Energy.

The real and imaginary parts of the interband conductivity are plotted for a variety of normalized

Fermi Energies in Figure 5.1.3. As we can see, the real part of the conductivity undergoes a rapid

jump at twice the Fermi energy for all Fermi energies. There is a corresponding dip in the

imaginary part of the optical conductivity at these same distinct energies.

To understand these features of the optical conductivity, we refer back to the pictures in

Figure 5.1.2 defining the Fermi level. For optical experiments involving normal incidence

Figure 5.1.3: Real and imaginary parts of the optical conductivity of graphene due to interband electron

transitions. The energies have all been normalized to the Fermi level. There is a clear tendency for both the

real and imaginary parts of the conductivity to have a jump and an absolute minimum at twice the Fermi

energy, respectively.

Page 87: Symmetry Breaking and Harmonic Generation in Metasurfaces

74

excitation, where no dramatic change in momentum is possible for the electrons, only perfectly

vertical movements of electrons from the valence to conduction band are permitted. The

minimum energy difference that must be covered in order for the transition to occur is clearly

twice the Fermi energy. In situations where the energy of the photon is insufficient to promote

the electron up by 2𝐸𝐹, the interband transition is forbidden, and only intraband motion can be

taken into consideration. As we will soon see, this energy requirement, known as Pauli

Blocking [99], can also have profound effects on nonlinear optical processes.

Figure 5.2.1: (a) Ellipticity dependence of fifth harmonic generation in CVD grown graphene. The total

harmonic yield is shown for both monolayer and few-layer samples. The dashed line represents the expected

behavior of an atom undergoing the 3-step HHG process. The authors concluded based on this observation

that graphene had an atom-like behavior. (b) Ellipticity dependence of fifth harmonic generation in the same

CVD grown graphene (same vendor). Here the signal has been polarization resolved into components parallel

and perpendicular to the pump ellipse major axis. The sum of the two curves is the direct comparison to (a).

The authors claim a very much non-atomic behavior.

Page 88: Symmetry Breaking and Harmonic Generation in Metasurfaces

75

Section 5.2: Elliptically Polarized Excitation of Graphene

Graphene has recently emerged as an intriguing solid-state material for studies of HHG

with elliptically-polarized pump fields. The introduction of this Chapter alluded to a mystery that

has been swirling around this field ever since the publication of two fascinating works back in

2017. In one paper, graphene was shown to exhibit large enhancements in the harmonic yield for

finite pump ellipticities [93]. The ellipticity of a laser is defined as the ratio of its electric fields

perpendicular and parallel to the major axis of the ellipse. 0 ellipticity defines linearly polarized

light, and 1 corresponds to circular polarization. The magnitude, peak ellipticity, and ellipticity

range of the enhancement all vary with harmonic order. This behavior came to be associated

with the semi-metal type electronic structure of graphene. However, in an alternative experiment

under similar conditions [94], differing only by the pump wavelength, no such ellipticity

dependence from low-order harmonics was observed. Graphene in that case instead displayed

“atom-like” behavior with a monotonically decreasing yield versus pump ellipticity. Going all

the way back to Figure 1.2.3 (b), we derived this exact trend when considering atomic gases! In

order to emphasize the nature of this discrepancy, the relevant plots from either paper are

reproduced in Figure 5.2.1.

It so happens that graphene is not the only material to display large enhancements to its

harmonic yield at finite pump field ellipticities. In silicon and MgO [100,101] and ZnO [102],

the different high harmonic orders can be categorized into those displaying atomic-like behavior

and those with a far more non-trivial dependence on the polarization. The dependence which

primarily distinguishes these two regimes from each other is on the differing roles of the

interband and intraband electron dynamics, discussed at the end of Chapter 1 when making the

analogy between solid-state and atomic HHG. Relying on that same analogy, we can make a

Page 89: Symmetry Breaking and Harmonic Generation in Metasurfaces

76

guess as to the source of novel ellipticity dependence in solid-state HHG. Seeing as the biggest

difference in the two versions of the 3-step model comes in the form of intraband harmonic

contributions coming from the motion of electrons and holes, this is a natural place to look.

The TDDFT simulations necessary to explore the ellipticity dependence are beyond the

scope of this work. Though they have been performed by other groups in the past, to show all of

the results would require the copying of figures whose creation I was not involved in, and I have

done my best to avoid doing so. Instead, I will briefly summarize the findings:

(1) Lower-order harmonics- having energies up to the band gap energy of the studied

material- are far more likely to originate from interband electron-hole recombination.

Because the recombination process is not sensitive to the band dispersion, it is identical

for polarized pumps of either chirality. These harmonics therefore tend to have an

atomic-like, monotonically decreasing yield with ellipticity.

(2) Harmonics that have energies significantly larger than that of the band gap cannot be the

direct result of interband recombination effects. Instead, these emissions must be the

result of intraband anharmonic motion of the electrons and holes in the acceleration (2nd)

step of the 3-step model. The sensitivity of these frequencies to the specific shape of the

conduction band and its dispersion make the higher-order harmonics a sensitive probe of

the electron trajectory. The electron trajectory is going to depend on the ellipticity of the

pump, and hence a non-trivial dependence is found.

When thought of as a gapless material, monolayer graphene is not expected to show a

transition between intraband and interband dominated behavior with changing photon energies

Page 90: Symmetry Breaking and Harmonic Generation in Metasurfaces

77

the way a traditional semiconductor will. Graphene should, in principle, have interband

contributions allowed at every possible frequency. Hence the understanding we’ve just

developed for the HHG in solids really should be insufficient to resolve the experimental

discrepancies reported for graphene. In Section 5.1 we have shown that in fact, the tunability of

the graphene Fermi energy and the Pauli blocking behavior associated with it provides us an

effective band gap through which to try to understand the inter- and intra- band contributions, in

essence applying the observation in silicon, ZnO, and MgO to graphene.

We perform our experiments in the following way: Our samples are based on

commercially available monolayer graphene grown by chemical vapor deposition (CVD, the

same vendor as was used in all prior experiments). The CVD graphene is transferred to a

sapphire substrate with metallic source and drain contacts to measure the graphene resistance. In

order to control the Fermi energy of graphene, we use an electrolyte top gate to which a variable

Figure 5.2.2: Sample preparation for elliptically polarized excitation and harmonic generation. CVD graphene

is transferred to a transparent sapphire substrate. Elliptically polarized pump beams incident on the graphene

generate variable polarized third and fifth harmonics. Circular polarization can be achieved. Metal contacts

and an electrolyte gel allow for easy manipulation of the graphene Fermi energy.

Page 91: Symmetry Breaking and Harmonic Generation in Metasurfaces

78

gate voltage spanning +3V to -5V (limited by electrolyte degradation) can be applied. We pump

the graphene at normal incidence with 3-μm wavelength pulses at a repetition rate of 1 kHz, 80-

fs pulse duration, and 1 μJ pulse energy through a CaF2 lens. The beam is defocused such that

the intensity incident on the graphene is just below the threshold for optical damage. A zero-

order mid-infrared quarter-wave plate allows for control of the pump polarization. The generated

3rd harmonic at 1 μm and 5th harmonic at 600 nm are then collected in a transmission geometry

and analyzed on a visible- near-infrared spectrometer (Ocean Optics QE65 Pro). Neither the

sapphire substrate nor the electrolyte is observed to contribute a nonlinear signal in the

experiment. The experimental principle for the sample is shown in Figure 5.2.2. A block diagram

of the rest of the experimental setup is given in Figure 5.2.3.

The third and fifth harmonic signals generated from the graphene monolayer are plotted

in Figure 5.2.4 for three different gate voltages, each of course corresponding to a different

Figure 5.2.3: Basic experimental setup for generation of graphene harmonics. Femtosecond pulses are

generated by an amplified Ti-sapphire laser. The 795 nm wavelength pulses then pump an OPA and

difference frequency generation system to produce mid-infrared pulses. A quarter wave plate sets the pulse

ellipticity. An applied gate voltage sets the Fermi energy of graphene.

Page 92: Symmetry Breaking and Harmonic Generation in Metasurfaces

79

Fermi energy. We observe a clear trend towards enhancement at finite ellipticity as the gate

voltage is swept through the full range. For positive gate voltages, a linearly polarized input

produces the highest yield. At the largest negative gate voltage (highest Fermi energy), the 5th

harmonic yield is 22% greater for an ellipticity of 0.35 than for a linearly polarized input.

The shifted yield curve can be attributed to a change in the relative contributions of

harmonics emitted parallel and perpendicular to the major axis of the input ellipse as described

in [93]. A very nice consequence of this also happens to be that for a fixed input ellipticity, the

output ellipticity varies with gating. This result confirms our theory relying on the forbidden

inter-band excitations graphene possesses due to the Pauli-blocking of photons with energy less

than twice its Fermi energy [103]. For a lower Fermi level, the pump photon energies are

sufficiently large compared to twice the Fermi energy, making inter-band excitations possible.

For this case, the “atom-like” behavior due to high pump photon energy in [94] is dominant. As

the Fermi energy is increased to a point where Pauli-blocking of the excitations takes place, the

Figure 5.2.4: Third and fifth harmonics of a 3 𝝁𝒎 pump as a function of the ellipticity of the pump. Colors of

the curves correspond to three different gate voltages ranging from +3V to -5V. In all cases the graphene is n –

doped, and has the largest Fermi energy at the most negative voltage. Every curve is normalized to its own

maximum. In all cases, the peak yield shifts towards finite ellipticities with increasing Fermi energy.

Page 93: Symmetry Breaking and Harmonic Generation in Metasurfaces

80

enhancement at finite ellipticities observed due to low pump photon energy becomes more

apparent and the “non-atomic” case takes over.

This experimental result resolves the discrepancy between previously reported

measurements and establishes two unique regimes for harmonic generation in graphene. One is

an “atom-like” regime with a monotonic ellipticity dependence for which ℏ𝜔𝑝𝑢𝑚𝑝 > 2𝐸𝐹. The

other is a regime that graphene can be smoothly tuned into requiring ℏ𝜔𝑝𝑢𝑚𝑝 < 2𝐸𝐹 and

displaying rich intraband electron dynamics. Having a system that can cover all variations of this

behavior may find use as a powerful tool both for generating specific polarized harmonics at

will, and for understanding the interplay between inter- and intra- band processes.

The emergence of the finite ellipticity peak in the low photon energy regime reveals a

tunable, perpendicularly polarized harmonic component and we’ll benefit from being able to see

the separation of the two orthogonal components. Figure 5.2.5 resolves the parallel and

perpendicular polarization components of the THG for two extreme cases of graphene gating.

Figure 5.2.5: Polarization resolved THG of graphene for two extreme doping situations. Whereas the parallel

component of THG (co-polarized with the pump ellipse major axis) shows little variation with gating, the

perpendicular component can be greatly enhanced by raising the Fermi energy. At the point where the two

components cross, circularly polarized harmonics are obtained.

Page 94: Symmetry Breaking and Harmonic Generation in Metasurfaces

81

The parallel component always appears to follow an atom-like monotonic trend, whereas the

perpendicular component can be greatly enhanced by raising the Fermi energy. Recent

theoretical progress [104] has drawn the same conclusions, based off of TDDFT simulations of

the various inter- and intra- band harmonic contributions. This simulation work performed by the

Angel Rubio group clearly shows that for a fixed input ellipticity, the perpendicular harmonic

component steadily increases with Fermi energy up to some critical value. The parallel

component shows a relatively minor change.

Finally, looking at Figure 5.2.5, and specifically at the crossing of the two perpendicular

components, we see that circularly polarized harmonics can be obtained from finite elliptical

pumping. This can be particularly useful, as we now have access to a tool for rapidly controlling

the output polarization of harmonics via electrostatic gating of the nonlinear medium.

Figure 5.3.1: (a) Representative example of CVD grown graphene, containing many randomly oriented ~1𝝁𝒎

sized grains of graphene. (b) Single crystal exfoliated graphene flake measuring 10s of microns x 10s of microns.

The exfoliated sample has one single orientation of atoms.

Page 95: Symmetry Breaking and Harmonic Generation in Metasurfaces

82

Section 5.3: A Brief Review of the Role of Point-Groups of Symmetry in Harmonic

Generation from Solids

All of the experiments that have been conducted on graphene harmonic generation thus

far (including our own) have taken place on samples grown by chemical vapor deposition. Figure

5.3.1 compares the arrangement of the CVD graphene lattice (composed of small, randomly

oriented zones) grown in [105] to that of a single crystal exfoliated sample prepared by

collaborator to this work Sang Hoon Chae of the Hone lab. The difference is stark: CVD grown

graphene consists of very many randomly oriented “grains” of graphene measuring only about

1 𝜇𝑚 in size. By comparison, exfoliated graphene (and graphite) is a single crystal, all oriented

in the same direction, with average sizes on the order of 10s of microns. Consider the typical

HHG experiment in which a laser spot size is easily on the order of 10s to 100s of 𝜇𝑚. It is clear

that when using CVD graphene samples many different orientations of graphene will be

contained within a single spot area. Any information that may be contained within the orientation

of the graphene will be averaged over the whole area, and much information will be lost.

The loss of orientation information is felt particularly hard in the context of harmonic

generation, where the crystal point-group symmetry plays the deciding role in the emission

symmetries of all harmonic orders [2]. Building on the very earliest description in this

dissertation, Section 1.1, we now briefly describe the role of symmetry in low-order even and

odd harmonic generation.

Going a step beyond a simple inversion-symmetry argument, we finally circle back to

give the nonlinear susceptibilities 𝜒(𝑛) of Chapter 1 the full tensor treatment they deserve. In

order to prevent these next few pages from getting completely out of hand, we will make use of

Page 96: Symmetry Breaking and Harmonic Generation in Metasurfaces

83

the more concise contracted notation for the susceptibilities. In this notation, the second order

susceptibility tensor is a 3 x 6 matrix:

𝒅𝒊𝒍 = [

𝒅𝟏𝟏 𝒅𝟏𝟐 𝒅𝟏𝟑𝒅𝟐𝟏 𝒅𝟐𝟐 𝒅𝟑𝟑𝒅𝟑𝟏 𝒅𝟑𝟐 𝒅𝟑𝟑

𝒅𝟏𝟒 𝒅𝟏𝟓 𝒅𝟏𝟔𝒅𝟐𝟒 𝒅𝟐𝟓 𝒅𝟐𝟔𝒅𝟑𝟒 𝒅𝟑𝟓 𝒅𝟑𝟔

] ( 36)

and anything higher than second order is too large and too high in dimension for me to want to

write out in matrix form. We can immediately begin to make arguments based off of various

symmetries and conditions to reduce the number of unique and independent terms in 𝑑𝑖𝑙. First,

the permutation symmetry known as Kleinman’s symmetry [106] reduces the matrix

representation to

𝒅𝒊𝒍 = [

𝒅𝟏𝟏 𝒅𝟏𝟐 𝒅𝟏𝟑𝒅𝟏𝟔 𝒅𝟐𝟐 𝒅𝟑𝟑𝒅𝟏𝟓 𝒅𝟐𝟒 𝒅𝟑𝟑

𝒅𝟏𝟒 𝒅𝟏𝟓 𝒅𝟏𝟔𝒅𝟐𝟒 𝒅𝟏𝟒 𝒅𝟏𝟐𝒅𝟐𝟑 𝒅𝟏𝟑 𝒅𝟏𝟒

]. ( 37)

From here, it is necessary to consider the specific spatial symmetries of the crystal in order to

further reduce the elements of 𝑑𝑖𝑙 from the 10 independent elements under Kleinman’s symmetry

to as few as 3 independent elements in some cases. We must therefore specify a point symmetry

Figure 5.3.2: Rotational symmetry of second harmonic generation copolarized with pump, for crystals of

symmetry point group D3h. There is no clear comparison for the higher symmetry D6h point group representing

graphene, as the centrosymmetry of the lattice forbids SHG.

Page 97: Symmetry Breaking and Harmonic Generation in Metasurfaces

84

group before we can take this process any further, and once doing so, we may calculate the

polarization according to

[

𝑷𝒙

𝑷𝒚

𝑷𝒛

] = 𝟑𝝐𝟎 [

𝒅𝟏𝟏 𝒅𝟏𝟐 𝒅𝟏𝟑𝒅𝟏𝟔 𝒅𝟐𝟐 𝒅𝟑𝟑𝒅𝟏𝟓 𝒅𝟐𝟒 𝒅𝟑𝟑

𝒅𝟏𝟒 𝒅𝟏𝟓 𝒅𝟏𝟔𝒅𝟐𝟒 𝒅𝟏𝟒 𝒅𝟏𝟐𝒅𝟐𝟑 𝒅𝟏𝟑 𝒅𝟏𝟒

]

[

𝑬𝒙(𝝎𝟏)𝑬𝒙(𝝎𝟐)𝑬𝒚(𝝎𝟏)𝑬𝒚(𝝎𝟐)

𝑬𝒛(𝝎𝟏)𝑬𝒛(𝝎𝟐)

𝑬𝒚(𝝎𝟏)𝑬𝒛(𝝎𝟐) + 𝑬𝒛(𝝎𝟏)𝑬𝒚(𝝎𝟐)

𝑬𝒙(𝝎𝟏)𝑬𝒛(𝝎𝟐) + 𝑬𝒛(𝝎𝟏)𝑬𝒙(𝝎𝟐)

𝑬𝒙(𝝎𝟏)𝑬𝒚(𝝎𝟐) + 𝑬𝒚(𝝎𝟏)𝑬𝒙(𝝎𝟐)]

( 38)

For this to have any real meaning at all, let’s look at a specific example common to 2D materials

such as monolayer hBN and various 2D transition metal dichalcogenides. These materials often

have point group symmetry D3h, for which every element in 𝑑𝑖𝑙 vanishes with the exception of

yyy, yxx, xxy, and xyx, which are all of the same magnitude (different signs though)! From

there, the calculation of the polarization and greatly simplified, and using equations of an

arbitrarily polarized electric field, with no z-component and x and y components given by

𝐸 cos(𝜃) and 𝐸𝑠𝑖𝑛(𝜃), we find that the rotational symmetry of the second harmonic radiation

(parallel to the pump polarization) emitted from a material of this spatial symmetry group goes

Figure 5.3.3: Rotational symmetry of third harmonic generation copolarized with pump, for crystals of

symmetry point group D6h. There is no symmetry information present in the perfectly isotropic signal.

Page 98: Symmetry Breaking and Harmonic Generation in Metasurfaces

85

as cos2(3𝜃), depicted in Figure 5.3.2. We will be seeing many such patterns over the course of

the rest of this Chapter, so we give the 6-fold rotational structure of Figure 5.3.2 the nickname of

a “flower” pattern.

The higher symmetry D6h group that materials like graphene belong to cannot be

compared to the cos2(3𝜃) shape we’ve shown, since the inversion symmetry of graphene’s

lattice forbids the second order nonlinearity. Therefore, we must go up one order of nonlinearity

to 𝜒(3). The third-order nonlinear susceptibility has 81 terms. To write them all out just to study

a single material would not add much to our understanding beyond the derivation we’ve

performed for the second order. Instead, let’s just jump ahead to the plot of the co-polarized

THG from a centrosymmetric hexagonal crystal such as bulk hBN or graphene. Figure 5.3.3

shows that in fact, the THG in such a system is perfectly isotropic, revealing no information

about the rotational symmetries of the underlying structure.

It is apparent from the above discussion that for lower symmetry crystals, polarization-

resolving harmonics can lead to a wealth of information about the lattice. As it so happens,

orientation-resolved SHG has become one of the most common tools for determining the lattice

orientation of thin samples. In 2D transition metal dichalcogenides (TMDs), polarized SHG is

greatest along the AC crystallographic axis, and due to the 𝐷3ℎ point group symmetry of most

TMDs, it is periodically repeating. SHG is then perfectly applied for measuring orientation with

exceptional resolution [107]. Graphene, a centro-symmetric material in both its monolayer and

few-layer forms, has its SHG and other even order nonlinearities forbidden. Furthermore, as a

member of the 𝐷6ℎ point symmetry group, graphene’s low odd-order (third and fifth) harmonics

are perfectly isotropic as seen in Figure 5.3.3. Knowing this, the results of the next and final

Page 99: Symmetry Breaking and Harmonic Generation in Metasurfaces

86

experimental section will be very surprising, as we demonstrate the emergence of the 6-fold

rotational symmetry in these lower-order odd nonlinearities in graphene.

Section 5.4: Time- Reversal Symmetry Breaking and Valley Polarization Effects

Experimental Setup

In order to go from the non-orientation dependent measurements of Section 5.2 to a new

setup with better control and higher signal to noise ratio, we make some significant changes to

the experimental design. Our new setup utilizes a much higher 1 MHz repetition rate optical

parametric amplifier (OPA) pumped by a Ytterbium fiber laser (KM Labs/ Thor Labs Y-Fi OPA)

centered at 3 µm wavelength, and with pulse durations of 60 fs. A combination of a half wave

plate (HWP) and quarter wave plate (QWP) set the ellipticity of the pump beam while

maintaining a constant ellipse orientation. Single crystal graphene flakes (including monolayer,

up toa10s of nm thickness) are exfoliated onto a 500 µm thick sapphire substrate, chosen for its

high transparency at the pump wavelength and relatively low nonlinearity. The sample is

Figure 5.4.1: Orientation dependent 5th harmonic generation in few-layer graphite sample at three different

pump ellipticities. Red points and curves show the harmonic component parallel to the pump ellipse major axis.

Blue is measured by an analyzer in the perpendicular direction. Each plot is normalized to its own parallel

component.

Page 100: Symmetry Breaking and Harmonic Generation in Metasurfaces

87

mounted on a rotating holder so that its orientation with respect to the incident laser can be easily

varied. The mid-IR pump beam is focused onto the graphite flakes with a 10 cm focal-length

CaF2 lens, and the generated third and fifth harmonics are collected in a transmission geometry

by a 5 cm UV-fused silica lens. A visible-wavelength polarization analyzer selects for parallel or

perpendicularly oriented harmonics before the signals are measured on a thermo-electrically

cooled ultraviolet-visible spectrometer (Ocean Optics) with a 350 to 1100 nm detection range.

Peak pump intensities are maintained below the regime of optical damage.

Measurements

The surprising results of the first ever orientation dependent 5th harmonic generation

experiment in multilayer graphene are presented in Figure 5.4.1 for different pump ellipticities.

This few-layer graphite sample, when pumped with perfectly linearly polarized light, has an

isotropic odd-order harmonic process predetermined by its 𝐷6ℎ symmetry group. In addition, no

cross-polarized signal can be measure for 0 ellipticity. However, as the ellipticity of the pump is

increased, a modulation of the uniform harmonic pattern emerges. What’s more, that modulation

Figure 5.4.2: Modulation depth of emergent “flower” symmetry in graphite 5HG vs pump ellipticity. Linear

best-fit also shown.

Page 101: Symmetry Breaking and Harmonic Generation in Metasurfaces

88

quickly converges to the recognizable 6-fold “flower” symmetry pattern of lower-symmetry

materials of different harmonic orders. The modulation depth as a function of the pump

ellipticity is plotted in Figure 5.4.2, and appears to follow a linear dependence. One further key

observation of Figure 5.4.1 beyond the emergent symmetry is the slight rotation of the flower

Figure 5.4.3: Orientation dependent 5th harmonic generation in few-layer graphite sample at three different

pump ellipticities. The sample used here is thicker than that in Figure 5.4.1. and a greater modulation depth is

observed for the same ellipticities.

Figure 5.4.4: Orientation dependent 5th harmonic generation in few-layer graphite sample at three different

pump powers and a fixed ellipticity of 0.4. The relative size of the blue perpendicular harmonic component

slightly increases.

Page 102: Symmetry Breaking and Harmonic Generation in Metasurfaces

89

with increasing modulation. This rotation will be a key supporting element when we come to

understand the origin of this rotational periodicity.

To round out the experimental observations before jumping into the theoretical

discussions, a similar ellipticity and orientation sweep on a thicker graphite flake is presented in

Figure 5.4.3, and the pump power dependence is investigated in Figure 5.4.4. The thickness has a

dramatic effect on the modulation depth at a given ellipticity, indicating a strong dependence on

the bandgap, which opens in graphite when compared with gapless graphene. The power

dependence presents itself primarily in the perpendicular component of the 5th harmonic.

Theoretical Explanation of Emergent Flower Symmetry

To properly understand the emergence of the flower symmetry pattern in thin graphite

samples under elliptical excitation, we must briefly describe an emerging field of 2D materials

and optics: that of valleytronics [108–113]. Recall the calculation of the graphene dispersion in

Figure 5.1.1 and Equation (34). At the time, some of the high-symmetry points of the Brillouin

zone were given names, in particular the K and K’ points. In an equilibrium system with time-

reversal symmetry, the K and K’ points are degenerate and equivalent. A look below the surface

reveals, however, that the two classes of points display opposite Berry curvatures [114]. A

proper derivation of the Berry curvature involves taking the curl of the Berry vector potential:

𝑨(𝒌) = 𝒊 < 𝝍|𝛁𝐤|𝝍 > ( 39)

Where 𝜓 recalls the Bloch functions that went into deriving the dispersion relations. Jumping

ahead past the derivation (which is beyond our scope), the Berry curvature which is defined in

momentum space plays the same role that a magnetic field does in real space. Hence the

opposite Berry curvatures of the K and K’ points can be thought of as opposite sign magnetic

fields in momentum space at the two types of points. There is an immediate and powerful

Page 103: Symmetry Breaking and Harmonic Generation in Metasurfaces

90

consequence of the analogy to magnetic fields: despite being degenerate under normal

conditions, the K and K’ “valleys” (local minima regions in the momentum space) are time-

reversal opposites of each other. This provides a neat way to lift the degeneracy by breaking

time-reversal symmetry!

It may seem as though we have strayed quite far away from our initial goal of

understanding the emergent flower symmetry of graphene 5HG. On the contrary, we are now

armed with a powerful new tool. The 6-fold rotational symmetry of graphene and graphite can be

decomposed into two sublattices in momentum space, with each of the K and K’ lattices

contributing a 3-fold rotationally symmetric trigonal lattice. In total, graphene will still belong to

the D6h symmetry group, and all contributions from the two valleys must sum to the equivalent

properties of D6h, but we have a new direction of attack to work from.

Figure 5.5.1 depicts our proposed theory. Each of the K and K’ sub-lattices in momentum

space display a 3-fold rotational symmetry. The odd-order harmonics have a 6-fold rotational

flower pattern for the 3-fold symmetry groups as seen for example in the experimental work

Figure 5.5.1: Decomposition of the graphene D6h 5HG rotational symmetry into the sum of two lower-dimension

“flower” patterns from the K and K’ valleys. Each of the K and K’ valleys has trigonal symmetry, and they are

related to each other by a 90 degree rotation.

Page 104: Symmetry Breaking and Harmonic Generation in Metasurfaces

91

done on sapphire and silicon [115]. Due to the natural 90-degree rotational relationship between

the two valleys, their independent contributions are oriented with a 𝜋 phase between them. Just

as the sum of sin2 and cos2 is 1, so too do the contributions of the two sublattice return the

isotropic, unmodulated harmonic emission of Figure 5.3.3.

This is all well and good, but unless we are capable of changing one of these two valley

contributions with respect to the other (in essence lifting the degeneracy) we will not be able to

explain the experimental observations. As previously stated, the well-established method for

lifting the K and K’ degeneracy is by breaking time reversal symmetry. It has been shown [116]

that chiral photons are sufficient to break time reversal symmetry, selectively driving electrons

into either of the two valleys. The now- famous valley hall effect is a consequence of circularly

polarized pumping for example [116]. One valley can be selectively pumped with right-hand

chiral excitation, while the opposite valley requires left-handedness. By driving electrons

selectively into one type of valley, a population imbalance is generated and the uneven

distribution of charge leads to a new polarization, called the valley polarization (VP) that

depends on that population difference.

Figure 5.5.2: The addition of a nonzero valley polarization to the total 𝝌𝟐 leads to a 6-fold symmetric flower that

is slightly rotated with respect to the intrinsic second order nonlinearity. The direction of the tilt is determined

by the phase of the valley polarization term, which is experimentally set by the handedness of the chiral photon

used to excite either the K or K’ valley.

Page 105: Symmetry Breaking and Harmonic Generation in Metasurfaces

92

Many reports have been published on the role of the valley polarization in allowing

second harmonic generation from graphene [117–119] and TMDs [120,121], though there have

not been any corresponding experiments performed in graphene or few-layer graphite. In the

context of SHG, VP contributes a new susceptibility term 𝜒𝑉𝑃 in addition to 𝜒𝑖𝑛𝑡, the intrinsic

susceptibility. This generalizes the matrix equation in Section 5.3 according to

[𝑷𝒙

𝑷𝒚] ∝ [

𝝌𝑽𝑷 −𝝌𝑽𝑷 −𝝌𝒊𝒏𝒕−𝝌𝒊𝒏𝒕 𝝌𝒊𝒏𝒕 −𝝌𝑽𝑷

] [

𝑬𝒙𝟐

𝑬𝒚𝟐

𝟐𝑬𝒙𝑬𝒚

]. ( 40)

The change that this new term induces on the symmetry of the 6-fold flower pattern for SHG that

was derived above for TMDs and other D3h materials is shown in Figure 5.5.2. As we can see,

the new contribution to the nonlinearity induced by 𝜒𝑉𝑃 is 𝜋 out of phase with the intrinsic

susceptibility. The sum of these two components squared manifests as a small rotation of the

overall measured 6-fold pattern, where the size of the rotation is determined by the size of the VP

contribution, and the direction of the rotation depends on the valley that’s been selectively

excited.

Figure 5.5.3: When each of the K and K’ 5HG contributions are altered with the addition of a VP, they rotate

in opposite directions. The grayed-out shapes show the total and component-wise contributions to the nonlinear

signal before valley polarization is induced. The black, red, and blue curves include VP terms. The 6-fold flower

begins to appear as the opposite rotations modulate the total signal.

Page 106: Symmetry Breaking and Harmonic Generation in Metasurfaces

93

We now address the question of how to control the degree of valley polarization. [120]

shows that for SHG in MoSe2 (a transition metal dichalcogenide), the degree of pump ellipticity

is directly proportional to the measured rotation of the 6-fold flower, and thus ellipticity is a

natural knob by which to control the degree of valley polarization. Hopefully now all of the

puzzle pieces are beginning to fit into place, and the final steps of our phenomenological

derivation are becoming clear. We will apply the concept of the rotation of the 6-fold flower to

higher-order nonlinearities, and observe how the 5th harmonic evolves.

To each of the spatially mirrored K and K’ harmonic contributions, we add a new term,

their respective valley polarizations, which we will assume for simplicity’s sake are equal in

magnitude and opposite in sign to each other. To write out the 6th rank tensor required to derive

the 5HG polarization components would be a crazy undertaking, so we will instead plot out the

components again in Figure 5.5.3. The K and K’ flowers undergo a slight rotation as a result of

VP, just as in the SHG case experimentally demonstrated in TMDs. The two flowers (which are

time-reversal and spatial mirrors of each other) rotate in opposite directions, bringing their petals

Figure 5.5.4: Total co-polarized 5HG for three different values of the valley population, which is experimentally

controlled by the laser ellipticity. The rotation of the flower is also recovered as a consequence of unequal

contributions from the two valleys.

Page 107: Symmetry Breaking and Harmonic Generation in Metasurfaces

94

closer together. The result is a modulation of the total, symmetric D6h harmonic emission,

causing it to reproduce the experimental observations of Figure 5.4.1.

The final experimental observations that remain to be accounted for are the dependence

on ellipticity as well as the slight rotation that seems to occur with increasing modulation depth.

Figure 5.5.3 is produced by the total function:

5𝐻𝐺(𝜃) = 𝛼 (cos(3𝜃) +𝜒𝑉𝑃

𝜒𝑖𝑛𝑡sin(3𝜃))

2

+ 𝛽 (sin(3𝜃) +𝜒𝑉𝑃

𝜒𝑖𝑛𝑡cos(3𝜃))

2

( 41)

where 𝛼 and 𝛽 determine the relative contributions from the two valleys and the VP contribution

is some small fraction of the total, determined by the ratio of the susceptibilities. Figure 5.5.4

plots three different values of that ratio as well as the situation 𝛼 > 𝛽. The unequal contributions

follow from the fact that the population of one valley is greater than the other, hence its flower

has a larger yield. The ratio 𝜒𝑉𝑃/𝜒𝑖𝑛𝑡 can be controlled with the laser ellipticity, with a nearly

linear dependence. The Figure confirms the experimentally observed modulation depth as a

function of ellipticity. Furthermore, the inequality of 𝛼 and 𝛽 reproduces the gradual turning of

the flower. Valley polarization effects are therefore shown to explain all of the fascinating

observations.

There are tremendous implications associated with the developments of this last section.

Although emergent SHG has already been proposed as a means of measuring valley polarization

in graphene, no experimental demonstration has been done. This is likely due to the fact that

graphene lacks any inherent SHG, and the contribution from valley effects may be too small to

easily measure. We’ve greatly extended this theory to include higher, odd-order nonlinearities

including fifth harmonic generation, and confirmed that VP makes a significant, easily

measurable change to the harmonic emission. This change is exerted on the rotational profile,

creating modulations as well as rotations of a few degrees.

Page 108: Symmetry Breaking and Harmonic Generation in Metasurfaces

95

Future efforts on this experiment can go in a number of interesting new directions. On the

one hand, the exact degree of valley polarization needs to be extracted from one of our harmonic

plots. In addition, twisted graphene and graphite samples, TMD homostructures, or

heterostructures containing a combination of materials are likely to have an inherent valley

polarization [122]. Using 5HG and linearly polarized light, the presence of modulations in the

rotational profile may provide a method of extracting the inherent VP with a straightforward all-

optical method.

Page 109: Symmetry Breaking and Harmonic Generation in Metasurfaces

96

Conclusions

The role of symmetry in nonlinear optics can present itself in any number of ways,

forbidding certain processes while dramatically enhancing others. Control over various

symmetries, then, becomes one of the most useful tools in engineering the next-generation

platforms of optical experiments. Low- and high- order harmonic generation are ideal candidates

for exploring the impact of symmetries on nonlinear processes. The strong scaling of high

harmonics with power greatly emphasizes resonant enhancements and tight confinement due to

symmetry breaking. Spatial symmetries and time-reversal become linked to each other in the

context of harmonic generation, and forbidden processes are deeply linked to material symmetry

groups.

In the first main experimental Chapter of this thesis, spatial symmetries were leveraged to

manufacture resonances call quasi-bound states in the continuum. Perfectly periodic structures

are capable of hosting modes with odd symmetry properties in a continuum of even symmetry

states. Being unable to couple energy in or out of that continuum, the odd symmetry state is

bound and inaccessible. We showed in Chapter 2 that by making a periodic perturbation to the

periodic unit cell, we are able to introduce coupling between the even symmetry incident field

and the mode. The size of the perturbation directly determined the quality factor (life time) of the

new resonance, while the materials and duty cycle selected the wavelength. In our particular

system which consisted of silicon-on-insulator gratings, we took further steps towards resonance

control. Understanding that the periodic perturbation of the grating effectively folds the Brillouin

zone in half, we used that knowledge to select for a mode located in the grating gaps. This

overcame two of the great shortcomings of recent semiconductor-based and plasmonics-based

nonlinear enhancement metasurfaces: damage and signal absorption. We proved the

Page 110: Symmetry Breaking and Harmonic Generation in Metasurfaces

97

effectiveness of the field confinement within the empty volume by performing third and fifth

harmonic generation on argon gas. The observed wavelength, polarization, and pressure

dependences confirmed all of the design principles of the dimerized high-contrast gratings.

From there, we tackled inversion symmetries in the two-dimensional material hexagonal

boron nitride. Being an inversion-symmetric crystal in equilibrium, hBN is forbidden to show

even-order nonlinearities in most cases. We developed a simple but robust theory of anharmonic

motion of lattice ions based on a perturbed harmonic oscillator model from which we were able

to generate new frequency components from ionic displacements. From there, a more

sophisticated theory based on time-dependent density function theory showed the exact response

of an hBN lattice to incident femtosecond pulses. These pulses, when tuned to the hyperbolic

phonon-polariton frequency of hBN, had two incredible impacts. First, the odd-order harmonics

generated by the driven atoms had yields far exceeding the more standard electronic

contributions. Second, the properties of an IR-active phonon, necessitating the creation or change

of a dipole moment in the material, allowed us to transiently break inversion symmetry. That

symmetry breaking was perturbed by a secondary laser pulse from which we were able to time-

resolve ultra-fast second order nonlinearities. There are many implications of this sub-

picosecond phase change in hBN, the applications of which may greatly impact van der Waals

heterostructures.

The final symmetry explored in this dissertation was time-reversal symmetry. We began

with a motivation to resolve an experimental discrepancy in the literature for elliptically driven

harmonic generation in graphene. By demonstrating that graphene could be put in an atomic or

non-atomic state by controlling the ratio between the pump energy and the Fermi level, we

definitively resolved that open question. In the end, it’s now becoming well known that the novel

Page 111: Symmetry Breaking and Harmonic Generation in Metasurfaces

98

ellipticity dependence originates from intra-band harmonic generation, and the degree to which

intra-band harmonics contribute can be tuned with Pauli blocking. Following those experiments,

we were presented with an opportunity to extend the field towards orientation sensitive

measurements due to our pristine single-crystal samples. What we found was surprising:

isotropic nonlinearities (with respect to polarization) display periodic modulations when time-

reversal is broken by chiral light. These modulations point us towards two extremely useful

future experiments. On one hand, we are presented with a simple, all-optical means of measuring

graphene’s lattice orientation in a way analogous to second harmonic spectroscopy used in

lower-symmetry 2D materials. Second, we have built out a theory understanding that the

emergence of the modulations in the orientation profile is a direct result of valley polarization.

Valley polarization measurements have been proposed in graphene SHG experiments, but have

not yet been successfully demonstrated. Using much larger nonlinearities such as the intrinsic

third and fifth, we have established a superior all-optical means of probing valley populations in

few-layer graphene samples.

Three seemingly unrelated material systems have been explored in this work, each

demonstrating the ways in which harmonic generation is inseparable from symmetry. In some

cases, it was symmetries that were applied as a way of enhancing the harmonics, while other

times the harmonics were applied as a means of understanding deep physical concepts in the

materials. I hope this work has proved to you that not only is harmonic generation one of the

longstanding pillars of nonlinear optics, but that it has much, much more to offer with each new

cutting edge material or platform to be developed.

Page 112: Symmetry Breaking and Harmonic Generation in Metasurfaces

99

References

1. P. A. Franken, A. E. Hill, C. W. Peters, and G. Weinreich, "Generation of Optical

Harmonics," Phys. Rev. Lett. 7, 118–119 (1961).

2. R. Boyd, Nonlinear Optics, 3rd ed. (Elsevier, 2008).

3. A. McPherson, G. Gibson, H. Jara, U. Johann, T. S. Luk, I. A. McIntyre, K. Boyer, and C. K.

Rhodes, "Studies of multiphoton production of vacuum-ultraviolet radiation in the rare

gases," J. Opt. Soc. Am. B, JOSAB 4, 595–601 (1987).

4. A. Gorlach, O. Neufeld, N. Rivera, O. Cohen, and I. Kaminer, "The quantum-optical nature

of high harmonic generation," Nat Commun 11, 4598 (2020).

5. F. Krausz and M. Ivanov, "Attosecond physics," Rev. Mod. Phys. 81, 163–234 (2009).

6. P. B. Corkum, "Plasma perspective on strong field multiphoton ionization," Phys. Rev. Lett.

71, 1994–1997 (1993).

7. M. Lewenstein, Ph. Balcou, M. Yu. Ivanov, A. L’Huillier, and P. B. Corkum, "Theory of

high-harmonic generation by low-frequency laser fields," Phys. Rev. A 49, 2117–2132

(1994).

8. X. Liu, X. Zhu, X. Zhang, D. Wang, P. Lan, and P. Lu, "Wavelength scaling of the cutoff

energy in the solid high harmonic generation," Opt. Express, OE 25, 29216–29224 (2017).

9. M. V. Frolov, N. L. Manakov, W.-H. Xiong, L.-Y. Peng, J. Burgdörfer, and A. F. Starace,

"Scaling laws for high-order-harmonic generation with midinfrared laser pulses," Phys. Rev.

A 92, 023409 (2015).

10. A. Blättermann, C.-T. Chiang, and W. Widdra, "Atomic line emission and high-order

harmonic generation in argon driven by 4-MHz sub-$\ensuremath{\mu}$J laser pulses,"

Phys. Rev. A 89, 043404 (2014).

11. A. Flettner, J. König, M. B. Mason, T. Pfeifer, U. Weichmann, and G. Gerber, "Atomic and

molecular high-harmonic generation: A comparison of ellipticity dependence based on the

three-step model," Journal of Modern Optics 50, 529–537 (2003).

12. S. Ghimire and D. A. Reis, "High-harmonic generation from solids," Nature Physics 15, 10–

16 (2019).

13. C. R. McDonald, G. Vampa, G. Orlando, P. B. Corkum, and T. Brabec, "Theory of high-

harmonic generation in solids," J. Phys.: Conf. Ser. 594, 012021 (2015).

14. N. Tancogne-Dejean, M. J. T. Oliveira, X. Andrade, H. Appel, C. H. Borca, G. Le Breton, F.

Buchholz, A. Castro, S. Corni, A. A. Correa, U. De Giovannini, A. Delgado, F. G. Eich, J.

Flick, G. Gil, A. Gomez, N. Helbig, H. Hübener, R. Jestädt, J. Jornet-Somoza, A. H. Larsen,

I. V. Lebedeva, M. Lüders, M. A. L. Marques, S. T. Ohlmann, S. Pipolo, M. Rampp, C. A.

Rozzi, D. A. Strubbe, S. A. Sato, C. Schäfer, I. Theophilou, A. Welden, and A. Rubio,

"Octopus, a computational framework for exploring light-driven phenomena and quantum

dynamics in extended and finite systems," J. Chem. Phys. 152, 124119 (2020).

15. N. Klemke, O. D. Mücke, A. Rubio, F. X. Kärtner, and N. Tancogne-Dejean, "Role of

intraband dynamics in the generation of circularly polarized high harmonics from solids,"

Phys. Rev. B 102, 104308 (2020).

16. A. A. Lanin, E. A. Stepanov, A. B. Fedotov, and A. M. Zheltikov, "Mapping the electron

band structure by intraband high-harmonic generation in solids," Optica 4, 516 (2017).

17. J. L. Krause, K. J. Schafer, and K. C. Kulander, "High-order harmonic generation from

atoms and ions in the high intensity regime," Phys. Rev. Lett. 68, 3535–3538 (1992).

Page 113: Symmetry Breaking and Harmonic Generation in Metasurfaces

100

18. S. Ghimire, A. D. DiChiara, E. Sistrunk, P. Agostini, L. F. DiMauro, and D. A. Reis,

"Observation of high-order harmonic generation in a bulk crystal," Nature Phys 7, 138–141

(2011).

19. H. Liu, C. Guo, G. Vampa, J. L. Zhang, T. Sarmiento, M. Xiao, P. H. Bucksbaum, J.

Vučković, S. Fan, and D. A. Reis, "Enhanced high-harmonic generation from an all-

dielectric metasurface," Nature Phys 14, 1006–1010 (2018).

20. G. Vampa, B. G. Ghamsari, S. Siadat Mousavi, T. J. Hammond, A. Olivieri, E. Lisicka-

Skrek, A. Y. Naumov, D. M. Villeneuve, A. Staudte, P. Berini, and P. B. Corkum, "Plasmon-

enhanced high-harmonic generation from silicon," Nature Phys 13, 659–662 (2017).

21. S. Kim, J. Jin, Y.-J. Kim, I.-Y. Park, Y. Kim, and S.-W. Kim, "High-harmonic generation by

resonant plasmon field enhancement," Nature 453, 757–760 (2008).

22. S. Han, H. Kim, Y. W. Kim, Y.-J. Kim, S. Kim, I.-Y. Park, and S.-W. Kim, "High-harmonic

generation by field enhanced femtosecond pulses in metal-sapphire nanostructure," Nat

Commun 7, 13105 (2016).

23. M. Sivis, M. Taucer, G. Vampa, K. Johnston, A. Staudte, A. Yu. Naumov, D. M. Villeneuve,

C. Ropers, and P. B. Corkum, "Tailored semiconductors for high-harmonic optoelectronics,"

Science 357, 303–306 (2017).

24. I.-Y. Park, S. Kim, J. Choi, D.-H. Lee, Y.-J. Kim, M. F. Kling, M. I. Stockman, and S.-W.

Kim, "Plasmonic generation of ultrashort extreme-ultraviolet light pulses," Nature Photon 5,

677–681 (2011).

25. N. Pfullmann, C. Waltermann, M. Kovačev, V. Knittel, R. Bratschitsch, D. Akemeier, A.

Hütten, A. Leitenstorfer, and U. Morgner, "Nano-antenna-assisted harmonic generation,"

Appl. Phys. B 113, 75–79 (2013).

26. S. Ghimire, A. D. DiChiara, E. Sistrunk, G. Ndabashimiye, U. B. Szafruga, A. Mohammad,

P. Agostini, L. F. DiMauro, and D. A. Reis, "Generation and propagation of high-order

harmonics in crystals," Phys. Rev. A 85, 043836 (2012).

27. C. W. Hsu, B. Zhen, J. Lee, S.-L. Chua, S. G. Johnson, J. D. Joannopoulos, and M. Soljačić,

"Observation of trapped light within the radiation continuum," Nature 499, 188–191 (2013).

28. C. W. Hsu, B. Zhen, A. D. Stone, J. D. Joannopoulos, and M. Soljačić, "Bound states in the

continuum," Nat Rev Mater 1, 16048 (2016).

29. D. C. Marinica, A. G. Borisov, and S. V. Shabanov, "Bound States in the Continuum in

Photonics," Phys. Rev. Lett. 100, 183902 (2008).

30. Y. Plotnik, O. Peleg, F. Dreisow, M. Heinrich, S. Nolte, A. Szameit, and M. Segev,

"Experimental Observation of Optical Bound States in the Continuum," Phys. Rev. Lett. 107,

183901 (2011).

31. M. Zhang and X. Zhang, "Ultrasensitive optical absorption in graphene based on bound

states in the continuum," Sci Rep 5, 8266 (2015).

32. L. Carletti, S. S. Kruk, A. A. Bogdanov, C. De Angelis, and Y. Kivshar, "High-harmonic

generation at the nanoscale boosted by bound states in the continuum," Phys. Rev. Research

1, 023016 (2019).

33. S. Romano, G. Zito, S. Managò, G. Calafiore, E. Penzo, S. Cabrini, A. C. De Luca, and V.

Mocella, "Surface-Enhanced Raman and Fluorescence Spectroscopy with an All-Dielectric

Metasurface," J. Phys. Chem. C 122, 19738–19745 (2018).

34. J. von Neumann and E. P. Wigner, "Über merkwürdige diskrete Eigenwerte," in The

Collected Works of Eugene Paul Wigner, A. S. Wightman, ed. (Springer Berlin Heidelberg,

1993), pp. 291–293.

Page 114: Symmetry Breaking and Harmonic Generation in Metasurfaces

101

35. W. Suh, M. F. Yanik, O. Solgaard, and S. Fan, "Displacement-sensitive photonic crystal

structures based on guided resonance in photonic crystal slabs," Appl. Phys. Lett. 82, 1999–

2001 (2003).

36. V. Liu, M. Povinelli, and S. Fan, "Resonance-enhanced optical forces between coupled

photonic crystal slabs," Opt. Express, OE 17, 21897–21909 (2009).

37. M. Robnik, "A simple separable Hamiltonian having bound states in the continuum," J. Phys.

A: Math. Gen. 19, 3845–3848 (1986).

38. J. U. Nöckel, "Resonances in quantum-dot transport," Phys. Rev. B 46, 15348–15356 (1992).

39. R. Parker, "Resonance effects in wake shedding from parallel plates: Some experimental

observations," Journal of Sound and Vibration 4, 62–72 (1966).

40. F. Ursell, "Trapping modes in the theory of surface waves," Mathematical Proceedings of the

Cambridge Philosophical Society 47, 347–358 (1951).

41. L. S. Cederbaum, R. S. Friedman, V. M. Ryaboy, and N. Moiseyev, "Conical Intersections

and Bound Molecular States Embedded in the Continuum," Phys. Rev. Lett. 90, 013001

(2003).

42. B. Zeng, A. Majumdar, and F. Wang, "Tunable dark modes in one-dimensional “diatomic”

dielectric gratings," Opt. Express, OE 23, 12478–12487 (2015).

43. S. Lan, S. P. Rodrigues, M. Taghinejad, and W. Cai, "Dark plasmonic modes in diatomic

gratings for plasmoelectronics," Laser & Photonics Reviews 11, 1600312 (2017).

44. H. S. Nguyen, F. Dubois, T. Deschamps, S. Cueff, A. Pardon, J.-L. Leclercq, C. Seassal, X.

Letartre, and P. Viktorovitch, "Symmetry Breaking in Photonic Crystals: On-Demand

Dispersion from Flatband to Dirac Cones," Phys. Rev. Lett. 120, 066102 (2018).

45. K. Koshelev, S. Lepeshov, M. Liu, A. Bogdanov, and Y. Kivshar, "Asymmetric

Metasurfaces with High- Q Resonances Governed by Bound States in the Continuum," Phys.

Rev. Lett. 121, 193903 (2018).

46. A. C. Overvig, S. Shrestha, and N. Yu, "Dimerized high contrast gratings," Nanophotonics 7,

1157–1168 (2018).

47. C. J. Chang-Hasnain, "High-contrast gratings as a new platform for integrated

optoelectronics," Semicond. Sci. Technol. 26, 014043 (2010).

48. V. Karagodsky, F. G. Sedgwick, and C. J. Chang-Hasnain, "Theoretical analysis of

subwavelength high contrast grating reflectors," Opt. Express, OE 18, 16973–16988 (2010).

49. A. C. Overvig, S. Shrestha, S. C. Malek, M. Lu, A. Stein, C. Zheng, and N. Yu, "Dielectric

metasurfaces for complete and independent control of the optical amplitude and phase,"

Light Sci Appl 8, 92 (2019).

50. A. C. Overvig, S. C. Malek, M. J. Carter, S. Shrestha, and N. Yu, "Selection rules for

quasibound states in the continuum," Phys. Rev. B 102, 035434 (2020).

51. S. S. Wang and R. Magnusson, "Theory and applications of guided-mode resonance filters,"

Appl. Opt., AO 32, 2606–2613 (1993).

52. D. Rosenblatt, A. Sharon, and A. A. Friesem, "Resonant grating waveguide structures,"

IEEE Journal of Quantum Electronics 33, 2038–2059 (1997).

53. C. W. Neff, T. Yamashita, and C. J. Summers, "Observation of Brillouin zone folding in

photonic crystal slab waveguides possessing a superlattice pattern," Appl. Phys. Lett. 90,

021102 (2007).

54. M. S. Bigelow, N. N. Lepeshkin, and R. W. Boyd, "Observation of Ultraslow Light

Propagation in a Ruby Crystal at Room Temperature," Phys. Rev. Lett. 90, 113903 (2003).

Page 115: Symmetry Breaking and Harmonic Generation in Metasurfaces

102

55. J. Ni, J. Yao, B. Zeng, W. Chu, G. Li, H. Zhang, C. Jing, S. L. Chin, Y. Cheng, and Z. Xu,

"Comparative investigation of third- and fifth-harmonic generation in atomic and molecular

gases driven by midinfrared ultrafast laser pulses," Phys. Rev. A 84, 063846 (2011).

56. S.-Y. Hong, J. I. Dadap, N. Petrone, P.-C. Yeh, J. Hone, and R. M. Osgood, "Optical Third-

Harmonic Generation in Graphene," Phys. Rev. X 3, 021014 (2013).

57. Y. Song, R. Tian, J. Yang, R. Yin, J. Zhao, and X. Gan, "Second Harmonic Generation in

Atomically Thin MoTe2," Advanced Optical Materials 6, 1701334 (2018).

58. J. W. You, S. R. Bongu, Q. Bao, and N. C. Panoiu, "Nonlinear optical properties and

applications of 2D materials: theoretical and experimental aspects," Nanophotonics 8, 63–97

(2019).

59. J. Chen, K. Wang, H. Long, X. Han, H. Hu, W. Liu, B. Wang, and P. Lu, "Tungsten

Disulfide–Gold Nanohole Hybrid Metasurfaces for Nonlinear Metalenses in the Visible

Region," Nano Lett. 18, 1344–1350 (2018).

60. A. von Hoegen, R. Mankowsky, M. Fechner, M. Först, and A. Cavalleri, "Probing the

interatomic potential of solids with strong-field nonlinear phononics," Nature 555, 79–82

(2018).

61. S. Foteinopoulou, G. C. R. Devarapu, G. S. Subramania, S. Krishna, and D. Wasserman,

"Phonon-polaritonics: enabling powerful capabilities for infrared photonics," Nanophotonics

8, 2129–2175 (2019).

62. Y. R. Shen, "Optical Second Harmonic Generation at Interfaces," Annual Review of

Physical Chemistry 40, 327–350 (1989).

63. K. Yao, N. R. Finney, J. Zhang, S. L. Moore, L. Xian, N. Tancogne-Dejean, F. Liu, J.

Ardelean, X. Xu, D. Halbertal, K. Watanabe, T. Taniguchi, H. Ochoa, A. Asenjo-Garcia, X.

Zhu, D. N. Basov, A. Rubio, C. R. Dean, J. Hone, and P. J. Schuck, "Enhanced tunable

second harmonic generation from twistable interfaces and vertical superlattices in boron

nitride homostructures," Sci. Adv. 7, eabe8691 (2021).

64. Y. Li, Y. Rao, K. F. Mak, Y. You, S. Wang, C. R. Dean, and T. F. Heinz, "Probing

Symmetry Properties of Few-Layer MoS2 and h-BN by Optical Second-Harmonic

Generation," Nano Lett. 13, 3329–3333 (2013).

65. D. Mandelli, W. Ouyang, M. Urbakh, and O. Hod, "The Princess and the Nanoscale Pea:

Long-Range Penetration of Surface Distortions into Layered Materials Stacks," ACS Nano

13, 7603–7609 (2019).

66. Z. Jacob, "Hyperbolic phonon–polaritons," Nature Mater 13, 1081–1083 (2014).

67. J. D. Caldwell, A. V. Kretinin, Y. Chen, V. Giannini, M. M. Fogler, Y. Francescato, C. T.

Ellis, J. G. Tischler, C. R. Woods, A. J. Giles, M. Hong, K. Watanabe, T. Taniguchi, S. A.

Maier, and K. S. Novoselov, "Sub-diffractional volume-confined polaritons in the natural

hyperbolic material hexagonal boron nitride," Nat Commun 5, 5221 (2014).

68. D. R. Smith, P. Kolinko, and D. Schurig, "Negative refraction in indefinite media," J. Opt.

Soc. Am. B, JOSAB 21, 1032–1043 (2004).

69. Z. Jacob, L. V. Alekseyev, and E. Narimanov, "Optical Hyperlens: Far-field imaging beyond

the diffraction limit," Opt. Express, OE 14, 8247–8256 (2006).

70. A. Salandrino and N. Engheta, "Far-field subdiffraction optical microscopy using

metamaterial crystals: Theory and simulations," Phys. Rev. B 74, 075103 (2006).

71. W. Paszkowicz, J. B. Pelka, M. Knapp, T. Szyszko, and S. Podsiadlo, "Lattice parameters

and anisotropic thermal expansion of hexagonal boron nitride in the 10–297.5 K temperature

range," Appl Phys A 75, 431–435 (2002).

Page 116: Symmetry Breaking and Harmonic Generation in Metasurfaces

103

72. X. Andrade, D. Strubbe, U. D. Giovannini, A. H. Larsen, M. J. T. Oliveira, J. Alberdi-

Rodriguez, A. Varas, I. Theophilou, N. Helbig, M. J. Verstraete, L. Stella, F. Nogueira, A.

Aspuru-Guzik, A. Castro, M. A. L. Marques, and A. Rubio, "Real-space grids and the

Octopus code as tools for the development of new simulation approaches for electronic

systems," Phys. Chem. Chem. Phys. 17, 31371–31396 (2015).

73. G. Onida, L. Reining, and A. Rubio, "Electronic excitations: density-functional versus many-

body Green’s-function approaches," Rev. Mod. Phys. 74, 601–659 (2002).

74. C. Hartwigsen, S. Goedecker, and J. Hutter, "Relativistic separable dual-space Gaussian

pseudopotentials from H to Rn," Phys. Rev. B 58, 3641–3662 (1998).

75. N. Tancogne-Dejean and A. Rubio, "Atomic-like high-harmonic generation from two-

dimensional materials," Science Advances 4, eaao5207 (2018).

76. A. Spott, A. Becker, and A. Jaroń-Becker, "Transition from perturbative to nonperturbative

interaction in low-order-harmonic generation," Phys. Rev. A 91, 023402 (2015).

77. M. Tamagnone, A. Ambrosio, K. Chaudhary, L. A. Jauregui, P. Kim, W. L. Wilson, and F.

Capasso, "Ultra-confined mid-infrared resonant phonon polaritons in van der Waals

nanostructures," Science Advances 4, eaat7189 (2018).

78. N. Klemke, N. Tancogne-Dejean, G. M. Rossi, Y. Yang, F. Scheiba, R. E. Mainz, G. Di

Sciacca, A. Rubio, F. X. Kärtner, and O. D. Mücke, "Polarization-state-resolved high-

harmonic spectroscopy of solids," Nature Communications 10, 1319 (2019).

79. O. Neufeld, D. Podolsky, and O. Cohen, "Floquet group theory and its application to

selection rules in harmonic generation," Nat Commun 10, 405 (2019).

80. S. M. Gilbert, T. Pham, M. Dogan, S. Oh, B. Shevitski, G. Schumm, S. Liu, P. Ercius, S.

Aloni, M. L. Cohen, and A. Zettl, "Alternative stacking sequences in hexagonal boron

nitride," 2D Mater. 6, 021006 (2019).

81. T. F. Nova, A. S. Disa, M. Fechner, and A. Cavalleri, "Metastable ferroelectricity in optically

strained SrTiO 3," Science 364, 1075–1079 (2019).

82. E. J. Sie, C. M. Nyby, C. D. Pemmaraju, S. J. Park, X. Shen, J. Yang, M. C. Hoffmann, B. K.

Ofori-Okai, R. Li, A. H. Reid, S. Weathersby, E. Mannebach, N. Finney, D. Rhodes, D.

Chenet, A. Antony, L. Balicas, J. Hone, T. P. Devereaux, T. F. Heinz, X. Wang, and A. M.

Lindenberg, "An ultrafast symmetry switch in a Weyl semimetal," Nature 565, 61–66

(2019).

83. A. A. Popkova, I. M. Antropov, J. E. Fröch, S. Kim, I. Aharonovich, V. O. Bessonov, A. S.

Solntsev, and A. A. Fedyanin, "Optical Third-Harmonic Generation in Hexagonal Boron

Nitride Thin Films," ACS Photonics 8, 824–831 (2021).

84. S. Dai, Z. Fei, Q. Ma, A. S. Rodin, M. Wagner, A. S. McLeod, M. K. Liu, W. Gannett, W.

Regan, K. Watanabe, T. Taniguchi, M. Thiemens, G. Dominguez, A. H. C. Neto, A. Zettl, F.

Keilmann, P. Jarillo-Herrero, M. M. Fogler, and D. N. Basov, "Tunable Phonon Polaritons in

Atomically Thin van der Waals Crystals of Boron Nitride," Science 343, 1125–1129 (2014).

85. S. H. Rhim, Y. S. Kim, and A. J. Freeman, "Strain-induced giant second-harmonic

generation in monolayered 2H-MoX2 (X = S, Se, Te)," Appl. Phys. Lett. 107, 241908

(2015).

86. N. Mendelson, M. Doherty, M. Toth, I. Aharonovich, and T. T. Tran, "Strain-Induced

Modification of the Optical Characteristics of Quantum Emitters in Hexagonal Boron

Nitride," Advanced Materials 32, 1908316 (2020).

Page 117: Symmetry Breaking and Harmonic Generation in Metasurfaces

104

87. K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V.

Grigorieva, and A. A. Firsov, "Electric Field Effect in Atomically Thin Carbon Films,"

Science 306, 666–669 (2004).

88. A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, and A. K. Geim, "The

electronic properties of graphene," Rev. Mod. Phys. 81, 109–162 (2009).

89. V. Kumar, "Linear and Nonlinear Optical Properties of Graphene: A Review," Journal of

Elec Materi 50, 3773–3799 (2021).

90. "Graphene-based active slow surface plasmon polaritons | Scientific Reports,"

https://www.nature.com/articles/srep08443.

91. D. G. Papageorgiou, I. A. Kinloch, and R. J. Young, "Mechanical properties of graphene and

graphene-based nanocomposites," Progress in Materials Science 90, 75–127 (2017).

92. A. A. Balandin, "Thermal properties of graphene and nanostructured carbon materials,"

Nature Mater 10, 569–581 (2011).

93. N. Yoshikawa, T. Tamaya, and K. Tanaka, "High-harmonic generation in graphene enhanced

by elliptically polarized light excitation," Science 356, 736–738 (2017).

94. M. Taucer, T. J. Hammond, P. B. Corkum, G. Vampa, C. Couture, N. Thiré, B. E. Schmidt,

F. Légaré, H. Selvi, N. Unsuree, B. Hamilton, T. J. Echtermeyer, and M. A. Denecke,

"Nonperturbative harmonic generation in graphene from intense midinfrared pulsed light,"

Phys. Rev. B 96, 195420 (2017).

95. D. R. Cooper, B. D’Anjou, N. Ghattamaneni, B. Harack, M. Hilke, A. Horth, N. Majlis, M.

Massicotte, L. Vandsburger, E. Whiteway, and V. Yu, "Experimental review of graphene,"

arXiv:1110.6557 [cond-mat] (2011).

96. L. A. Falkovsky and A. A. Varlamov, "Space-time dispersion of graphene conductivity,"

Eur. Phys. J. B 56, 281–284 (2007).

97. L. A. Falkovsky and S. S. Pershoguba, "Optical far-infrared properties of a graphene

monolayer and multilayer," Phys. Rev. B 76, 153410 (2007).

98. L. A. Falkovsky, "Optical properties of graphene," J. Phys.: Conf. Ser. 129, 012004 (2008).

99. K. F. Mak, L. Ju, F. Wang, and T. F. Heinz, "Optical spectroscopy of graphene: From the far

infrared to the ultraviolet," Solid State Communications 152, 1341–1349 (2012).

100. Y. S. You, D. A. Reis, and S. Ghimire, "Anisotropic high-harmonic generation in bulk

crystals," Nature Phys 13, 345–349 (2017).

101. N. Tancogne-Dejean, O. D. Mücke, F. X. Kärtner, and A. Rubio, "Ellipticity dependence

of high-harmonic generation in solids originating from coupled intraband and interband

dynamics," Nat Commun 8, 745 (2017).

102. X. Zhang, J. Li, Z. Zhou, S. Yue, H. Du, L. Fu, and H.-G. Luo, "Ellipticity dependence

transition induced by dynamical Bloch oscillations," Phys. Rev. B 99, 014304 (2019).

103. 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. 101, 196405

(2008).

104. S. A. Sato, H. Hirori, Y. Sanari, Y. Kanemitsu, and A. Rubio, "High-order harmonic

generation in graphene: Nonlinear coupling of intraband and interband transitions," Phys.

Rev. B 103, L041408 (2021).

105. J. Ryu, Y. Kim, D. Won, N. Kim, J. S. Park, E.-K. Lee, D. Cho, S.-P. Cho, S. J. Kim, G.

H. Ryu, H.-A.-S. Shin, Z. Lee, B. H. Hong, and S. Cho, "Fast Synthesis of High-

Performance Graphene Films by Hydrogen-Free Rapid Thermal Chemical Vapor

Deposition," ACS Nano 8, 950–956 (2014).

Page 118: Symmetry Breaking and Harmonic Generation in Metasurfaces

105

106. D. A. Kleinman, "Nonlinear Dielectric Polarization in Optical Media," Phys. Rev. 126,

1977–1979 (1962).

107. S. Psilodimitrakopoulos, L. Mouchliadis, I. Paradisanos, A. Lemonis, G. Kioseoglou, and

E. Stratakis, "Ultrahigh-resolution nonlinear optical imaging of the armchair orientation in

2D transition metal dichalcogenides," Light Sci Appl 7, 18005–18005 (2018).

108. J. Isberg, M. Gabrysch, J. Hammersberg, S. Majdi, K. K. Kovi, and D. J. Twitchen,

"Generation, transport and detection of valley-polarized electrons in diamond," Nature Mater

12, 760–764 (2013).

109. A. Rycerz, J. Tworzydło, and C. W. J. Beenakker, "Valley filter and valley valve in

graphene," Nature Phys 3, 172–175 (2007).

110. J. R. Schaibley, H. Yu, G. Clark, P. Rivera, J. S. Ross, K. L. Seyler, W. Yao, and X. Xu,

"Valleytronics in 2D materials," Nat Rev Mater 1, 1–15 (2016).

111. Y. P. Shkolnikov, E. P. De Poortere, E. Tutuc, and M. Shayegan, "Valley Splitting of

AlAs Two-Dimensional Electrons in a Perpendicular Magnetic Field," Phys. Rev. Lett. 89,

226805 (2002).

112. D. Xiao, W. Yao, and Q. Niu, "Valley-Contrasting Physics in Graphene: Magnetic

Moment and Topological Transport," Phys. Rev. Lett. 99, 236809 (2007).

113. W. Yao, D. Xiao, and Q. Niu, "Valley-dependent optoelectronics from inversion

symmetry breaking," Phys. Rev. B 77, 235406 (2008).

114. M. V. Berry, "Quantal phase factors accompanying adiabatic changes," Proceedings of

the Royal Society of London. A. Mathematical and Physical Sciences 392, 45–57 (1984).

115. G. Petrocelli, E. Pichini, F. Scudieri, and S. Martellucci, "Anisotropic effects in the third-

harmonic-generation process in cubic crystals," J. Opt. Soc. Am. B, JOSAB 10, 918–923

(1993).

116. K. F. Mak, K. L. McGill, J. Park, and P. L. McEuen, "The valley Hall effect in MoS2

transistors," Science 344, 1489–1492 (2014).

117. M. S. Mrudul, Á. Jiménez-Galán, M. Ivanov, and G. Dixit, "Light-induced valleytronics

in pristine graphene," Optica, OPTICA 8, 422–427 (2021).

118. H. K. Kelardeh, M. Eidi, T. Oka, and J. M. Rost, "Photoinduced Nonperturbative Valley

Polarization in Graphene," arXiv:2012.14025 [cond-mat, physics:physics] (2021).

119. L. E. Golub and S. A. Tarasenko, "Valley polarization induced second harmonic

generation in graphene," Phys. Rev. B 90, 201402 (2014).

120. Y. W. Ho, H. G. Rosa, I. Verzhbitskiy, M. J. L. F. Rodrigues, T. Taniguchi, K.

Watanabe, G. Eda, V. M. Pereira, and J. C. Viana-Gomes, "Measuring Valley Polarization in

Two-Dimensional Materials with Second-Harmonic Spectroscopy," ACS Photonics 7, 925–

931 (2020).

121. F. Hipolito and V. M. Pereira, "Second harmonic spectroscopy to optically detect valley

polarization in 2D materials," 2D Mater. 4, 021027 (2017).

122. J. Liu, Z. Ma, J. Gao, and X. Dai, "Quantum Valley Hall Effect, Orbital Magnetism, and

Anomalous Hall Effect in Twisted Multilayer Graphene Systems," Phys. Rev. X 9, 031021

(2019).