Download as pdf or txt
Download as pdf or txt
You are on page 1of 160

Manipulation of Optical Processes in Diamond using Plasmonic Nanogap Metasurfaces

by

Andrew M. Boyce

Department of Electrical and Computer Engineering


Duke University

Date:_______________________
Approved:

___________________________
Prof. Maiken H. Mikkelsen, Supervisor

___________________________
Prof. Steven A. Cummer

___________________________
Prof. Aaron D. Franklin

___________________________
Prof. Willie J. Padilla

___________________________
Prof. Adrienne D. Stiff-Roberts

Dissertation submitted in partial fulfillment of


the requirements for the degree of Doctor of Philosophy in the Department of
Electrical and Computer Engineering in the Graduate School
of Duke University

2022

v
ABSTRACT

Large Purcell Enhancement of SiVs in Diamond Thin Films Coupled to Plasmonic

Nanogap Cavities

by

Andrew M. Boyce

Department of Electrical and Computer Engineering


Duke University

Date:_______________________
Approved:

___________________________
Prof. Maiken H. Mikkelsen, Supervisor

___________________________
Prof. Steven A. Cummer

___________________________
Prof. Aaron D. Franklin

___________________________
Prof. Willie J. Padilla

___________________________
Prof. Adrienne D. Stiff Roberts

An abstract of a dissertation submitted in partial


fulfillment of the requirements for the degree
of Doctor of Philosophy in the Department of
Electrical and Computer Engineering in the Graduate School of
Duke University

2022

v
Copyright by
Andrew M. Boyce
2022
Abstract
Solid-state quantum emitters embedded in carefully engineered nanostructures

could enable a new generation of quantum information and sensing technologies, including

networked processors for quantum computing and precise monitors of temperature and

strain at the nanoscale. The primary goal when designing these nanostructures is to utilize

the Purcell effect to improve the emission rate, directionality and brightness of quantum

emitters, as long decay times, nondirectional emission and weak fluorescence limit their

applications. One particularly promising emitter is the silicon vacancy (SiV) in diamond,

which offers excellent photostability and minimal spectral diffusion, in addition to coherent

emission at its zero-phonon line (ZPL) comprising 80% of its total fluorescence.

In this dissertation, up to 121-fold enhancement of the spontaneous emission rate

of SiVs coupled to plasmonic nanogap cavities is demonstrated. The vacancy centers are

implanted into a monolithic diamond thin film, which is then etched to nanometer-scale

thickness, an approach with a clear path towards wafer-scale fabrication. A novel approach

to creating film-coupled nanogap metasurfaces was developed to support this research and

consists of transferring EBL-fabricated nanoparticles by using a PDMS stamp. Up to seven

orders of magnitude of enhancement of nonlinear frequency conversion was also observed

in diamond thin films coupled to these metasurfaces. Furthermore, a robust mechanism for

actively tuning the nanocavity absorption resonance by integrating sub-10-nm films of the

phase-change material vanadium dioxide. This platform opens up opportunities for on-chip

quantum networks and nanoscale sensors based on nanocavity-coupled SiVs with the

iv
potential for in-situ frequency conversion to outcouple to photonic circuits and

reconfigurable properties by incorporating VO2 thin films.

v
Dedication

For my grandfather, Michael J. Francesco Sr.

vi
Contents

Abstract .............................................................................................................................. iv

List of Figures .................................................................................................................... xi

Acknowledgements ......................................................................................................... xiv

1. Introduction ................................................................................................................... 16

1.1 Perspective ......................................................................................................... 16

1.2 Background ........................................................................................................ 19

1.3 Organization........................................................................................................... 23

2. Molecular-Scale Plasmonics ........................................................................................... 25

2.1 Fundamentals of Plasmonics.................................................................................. 25

2.2 Survey of conventional plasmonic structures ........................................................ 29

2.2 Active Tuning of Metasurface Absorption Resonance .......................................... 32

2.3 Dimer Plasmonic Antennas and the Push Towards Molecular-Scale Plasmonics 35

2.4 Film-Coupled Plasmonic Nanocavities: Molecular-Scale Plasmonics for


Centimeter-Scale Devices ............................................................................................ 36

3. Vacancy Centers in Diamond and Enhancement of their Properties in Nanophotonic


Structures ........................................................................................................................... 39

3.1 Introduction and Motivation.................................................................................. 39

3.2 Fundamentals of Single Photon Sources and Hanbury-Brown-Twiss


Interferometry to Verify Single Photon Emission ....................................................... 40

3.3 The Nitrogen Vacancy Center in Diamond ........................................................... 43

3.4 The Silicon Vacancy Center in Diamond and the Search for Alternative Vacancy
Centers ......................................................................................................................... 48

vii
3.4.1 Limitations of the NV Center and Motivation to Identify New Types of Vacancy
Centers ......................................................................................................................... 48

3.4.2 Discovery of the Silicon Vacancy Center and Some Fundamental Properties ... 49

3.4.3 Evaluating the Suitability of SiVs for Quantum Information: Interlude on Hong-
Ou-Mandel Interferometry .......................................................................................... 52

3.4.4 Evaluating the Suitability of SiVs for Quantum Information: HOM


Interferometry Experiments on SiVs ........................................................................... 56

3.5 Emerging Vacancy Centers in Diamond: The SiV 0 and Additional Group IV
Vacancy Centers .......................................................................................................... 59

3.6 Purcell Enhancement of Diamond Vacancy Centers and Other Spontaneous


Emitters Embedded in Nanophotonic Structures ....................................................... 62

4. EBL Fabrication of Plasmonic Nanogap Cavities with Single- and Dual-Band


Resonances in the Near-Infrared ....................................................................................... 68

4.1 Introduction............................................................................................................ 68

4.2 Sample Design and Fabrication ............................................................................. 68

4.3 Absorption Characteristics of EBL-Fabricated Nanodisks and Nanocubes .......... 70

4.4 Dependence of Absorption Properties on Fill Fraction of the Nanodisk


Metasurface.................................................................................................................. 73

4.5 Absorption Properties Dual-Band, Polarization Sensitive Metasurfaces .............. 75

4.6 Benefits of Dual-Band Absorbers for Purcell Enhancement .................................. 77

4.7 Conclusion .............................................................................................................. 80

5. Actively Tunable Nanogap Cavity Metasurfaces Via Integrated VO2 Thin Films ........ 81

5.1 Introduction............................................................................................................ 81

5.2 Device Concept....................................................................................................... 83

5.3. Fabrication Process ................................................................................................ 85

viii
5.4 Thermal Switching Properties of Nanodisk Metasurfaces with Variable
Resonance Wavelengths in the NIR ............................................................................ 90

5.5 Simulated Metasurface Thermal Switching ........................................................... 93

5.6 In-Depth Characterization of 192 nm Nanodisk Metasurface Thermal Switching


...................................................................................................................................... 94

5.7 Conclusion .............................................................................................................. 96

6. Plasmonic Nanogap Metasurface with Integrated Silicon Vacancies in Diamond ....... 98

6.1 Introduction............................................................................................................ 98

6.2 Device Concept....................................................................................................... 99

6.3 Fabrication Process ............................................................................................... 102

6.4 Emission Lifetime Properties of Metasurfaces Containing Ensembles of SiVs ... 104

6.6 Conclusion ............................................................................................................ 107

6.7 Emission Properties of Ensembles of SiVs Coupled to Single Nanocavities ....... 109

7. Frequency Conversion in Diamond Thin Films via Nanogap-Cavity-Enhanced


Nonlinear Optics.............................................................................................................. 111

7.1 Introduction.......................................................................................................... 111

7.2 Sample Fabrication ............................................................................................... 112

7.3 Nanogap Cavity Enhanced THG and SHG ......................................................... 115

7.4 Excitation Wavelength Dependence of THG and SHG ....................................... 118

7.5 Simultaneous Generations of Multiple Nonlinear Frequencies .......................... 120

7.6 Conclusion ............................................................................................................ 123

8. Conclusion.................................................................................................................... 125

ix
Appendix A Additional Diamond Fabrication Process Details for Nanogap-Cavity-
Embedded SiVs ................................................................................................................ 129

Appendix B Additional Diamond Fabrication Process Details for Nonlinear Frequency


Conversion ....................................................................................................................... 130

Appendix C Coupled Mode Theory Analysis of Nonlinear Responses ......................... 132

Appendix D Additional Excitation Wavelength Dependence Data for SHG and THG . 133

References ........................................................................................................................ 134

Biography......................................................................................................................... 160

x
List of Figures
Figure 1.1 Photographs of the Lycurgus Cup, displaying its marvelous dichroic glass. 4 20

Figure 2.1 Schematic representation of the electromagnetic field associated with a


surface-plasmon polariton propagating along a metal–dielectric interface 12 ................... 26

Figure 2.2 Schematic representation of displacement of the electron cloud in a metallic


nanosphere resulting from incident optical fields. 13 ......................................................... 27

Figure 3.1 Experimental setup for Hanbury-Brown-Twiss interferometry. ..................... 41

Figure 3.2 Typical second-order correlation function for a single emitter, in this case a
colloidal quantum dot. 33 .................................................................................................... 42

Figure 3.3 Schematic representation of the crystallographic structure of the NV center in


diamond.40.......................................................................................................................... 44

Figure 3.4 Typical fluorescence spectrum from a combination of both NV- and NV0
centers under 532 nm laser illumination. .......................................................................... 45

Figure 3.5 Schematic of typical operation of a quantum circuit based on NV centers in


the prepare-manipulate/interact readout approach. 39 ....................................................... 47

Figure 3.6 Fundamental structure and optical properties of the SiV in diamond. 63 ......... 52

Figure 3.7 HOM interferometer utilized in Lukin and coauthors demonstration of


indistinguishable photons from separate SiVs.63............................................................... 54

Figure 3.8 Example time correlation function resulting from a HOM experiment.66 ....... 56

Figure 3.9 Hong-Ou-Mandel Interferometry with SiVs in diamond. 63 ............................. 58

Figure 3.10 Calculated energy levels of the ground state (GS) and excited state (ES) of the
negative charge state, group IV vacancy centers in diamond. 53........................................ 60

Figure 4.1: Single-Resonance, EBL Fabricated Nanogap Cavities..................................... 70

Figure 4.2: Effect of Varying the Fill Fraction on the Absorption Resonance of EBL
Nanogap Metasurfaces ...................................................................................................... 73

xi
Figure 4.3: Polarization-Dependent Absorption Properties of EBL Nanorectangle
Metasurface........................................................................................................................ 75

Figure 4.4: Simulations to Demonstrate Mode Overlap in EBL Rectangle Nanocavities . 77

Figure 5.1: Actively Tunable Nanogap Cavity Metasurface Concept ............................... 83

Figure 5.2: XPS Spectra to Characterize VO2 Film ............................................................. 86

Figure 5.3: SEM Image Revealing Pinholes in the VO2 Film ............................................. 87

Figure 5.4: Fabrication Overview and Initial Characterization of Actively Tunable


Metasurfaces with Integrated VO2 .................................................................................... 88

Figure 5.6: Thermal Switching Data from VO2 Metasurface and Control Al2O3
Metasurface........................................................................................................................ 90

Figure 5.7: Thermal Switching Spectra for Additional Sizes of Nanodisks ...................... 92

Figure 5.8: Simulated Switching Spectra ........................................................................... 93

Figure 5.9: Further Characterization of Thermal Switching For 192 nm Nanodisk


Metasurface........................................................................................................................ 94

Figure 6.1: Nanogap-Coupled Silicon Vacancy Metasurface Concept............................ 100

Figure 6.2: Fabrication Process for Nanocavity Metasurfaces with Integrated SiVs ...... 102

Figure 6.3: Cavity-Enhanced SiV Lifetime and PL Data for Metasurface Sample .......... 104

Figure 6.4: Cavity-Enhanced SiV Lifetime for Single Cavity Sample ............................. 109

Figure 7.1: Diamond Frequency Conversion Sample Overview..................................... 112

Figure 7.2: THG and SHG from Diamond Nanogap Cavity Sample .............................. 115

Figure 7.3: Excitation Wavelength Dependence of THG and SHG ................................. 118

Figure 7.4: Generation of Multiple, Simultaneous Nonlinear Wavelengths ................... 120

Figure A.1 Additional THG Response Spectra as a Function of Excitation Wavelength 133

xii
Figure A.2 Additional SHG Response Spectra as a Function of Excitation Wavelength 133

xiii
Acknowledgements
I would like to thank my advisor, Prof. Maiken Mikkelsen for her mentorship

and support during my PhD studies. I am grateful for the opportunity to have grown as

a researcher, engineer, and scientific communicator over these past six years. Joining

Prof. Mikkelsen’s group has given me tremendous opportunities to collaborate both

within our research group and with outstanding scientists at other universities and

national laboratories. I am incredibly thankful for the research environment she has

fostered in her group, which is simultaneously cooperative and welcoming, while

maintaining a strong culture of innovation and excellence.

I would also like to thank my defense committee, Professors Steven Cummer,

Aaron Franklin, Willie Padilla, and Adrienne Stiff-Roberts. I have greatly enjoyed

discussing my research with you and have found our discussions to be helpful,

stimulating, and challenging. I owe Professor Padilla double debt of gratitude both for

serving on my dissertation committee and for serving as my faculty advisor through the

Gabelli Presidential Scholars Program at Boston College during my undergraduate

studies.

I would like to thank many of my colleagues, including Dr. Andrew Traverso

and Dr. Tao Cai for their mentorship during the early part of my PhD studies. I would

like to thank my collaborators: Prof. Marko Loncar at Harvard University and his former

students Pawel Latawiec and Amirhassan Shams-Ansari, Dr. Virginia Wheeler at the

xiv
U.S. Naval Research Laboratory, and Dr. Henry Everitt of the DEVCOM Army Research

Lab. I would also to thank my current and former labmates for their collaboration and

friendship, including Jon Stewart, Wade Wilson, Qixin Shen, Nathan Wilson, Daniela

Cruz, Tamra Nebabu, Siyuan Zhang, Deniz Acil, Hengming Li, Brian Lerner, Eunso

Shin, Hiroyuki Kishida and Rachel Bangle.

Lastly, I would like to thank my family and friends for supporting me every step

of the way, especially my parents Karla and Sheldon, my late grandfather Michael, my

sister and roommate for the last 18 months Katie, and my dog Willow who has been

faithfully and quite literally by my side as I’ve written every chapter of this dissertation.

xv
1. Introduction
1.1 Perspective
It would both correct, in many ordinary circumstances, and decidedly too

simplistic to say that the diffraction limit is the smallest length scale of interest when

formulating an elementary theory of optics or designing photonic devices. Indeed, the

diffraction limit is of fundamental importance for the field of optics as it certainly limits

the length scales that can be resolved when performing ordinary light microscopy and

many conventional photonic devices like waveguides cannot function unless their size is

significantly larger than the wavelength of interest. Yet, on the other hand, the diverse

research field of nanophotonics has revealed a manifold of ways in which light can be

confined and manipulated on length scales comparable to or, in many cases, far below

the diffraction limit. Approaches for subverting the diffraction limit in this manner

include photonic crystals, metamaterial lenses and wavefront optics, and, as will be

discussed at great length in the pages to follow, plasmonic/metallic nanostructures.

It would be furthermore naïve to conclude that only artificial engineering has

devised means for manipulating light at length scales below the diffraction limit, as

nature has employed these techniques for millions of years to produce bright coloring or

camouflage in a variety of plants and animals ranging from the feathers of birds of

paradise, the skin of chameleons, butterfly wings and the spine of the sea mouse.

Indeed, evolutionary biology is capable of producing a wide enough range of optical

16
nanostructures to yield an array of visual effects from active camouflage, to narrowband

colors produced by highly-ordered nanostructures and sparkly reflectance of a

broadband nature resulting from strongly disordered, natural photonic crystals that

induce Anderson localization of light. 1

Plasmonics, or the study of metallic nanostructures capable of coupling light to

coherent oscillations of the free electrons in the metal known as plasmons, provides an

interesting case study of the surprising and complex optical phenomena that can arise

when discontinuities and structuring of materials at sub-wavelength scales are involved.

Indeed, metals are thoroughly uninteresting in a conventional optics context, merely

acting as broadband reflectors (at photon energies smaller than the energy necessary to

pump the metal’s interband transition) that are well described by elementary ray optics.

Yet, as will be expounded upon in the chapters that follow, plasmonic nanostructures

enable a host of useful and interesting photonic applications ranging from, high-

sensitivity bio sensors, ultrafast photodetection, efficient nonlinear generation,

photocatalysis and, pivotally, control of optical emitters with effects that are both

powerful and wide-ranging.2 Furthermore, plasmons have wavelengths about an order

of magnitude smaller than optical waves, opening up the possibility for packing many,

many more photonic components into a chip-scale device than would be possible with

other types of photonic technology.

17
The results presented in this dissertation all employ the use of film-coupled

plasmonic nanogap cavities. In these sorts of devices, the architecture is engineered on

the single-nanometer/molecular scale, bringing to fruition the full potential of

plasmonics to deliver devices that are scalable to potentially the same degree as

semiconductor electronic circuits and capable of squeezing light to dimensions small

enough that force a complete rethinking of what is possible given the intrinsic optical

properties of the materials involved, potentially to the benefit of many of the potential

applications for plasmonics that were listed above.3 In the context of the research

presented in this dissertation, this involves drastically altering the properties of single-

photon sources that are candidates for quantum information device components to bring

together the necessary mix of emitter properties that cannot be achieved simultaneously

with any one natural, quantum emitter. Chapter 3 gives an overview of many potential

solid-state quantum emitters and should help to make the case why nanophotonic

engineering, such as through a plasmonic nanocavity, might be necessary to push these

solid-state candidate quantum emitters over the edge to potentially achieve the status of

the material of choice for sophisticated applications in quantum information. The field of

nanophotonics and molecular-scale plasmonics, in particular, are rapidly evolving. As

research progresses in these fields, the landscape of potential applications could

transform completely, in the entire range of use cases and not just in quantum

information.

18
1.2 Background
Some may be surprised to learn that the earliest artificial use nanophotonic

structures goes back approximately two millennia to when the Romans utilized metallic

nanoparticles in the creation of stained glass. 2 A particular masterful demonstration of

this type of craftsmanship is the Lycurgus Cup, a glass cage cup that was forged in the

4th-century CE and depicts the myth of King Lycurgus, who is portrayed as he is being

dragged to the underworld by the Greek nymph Ambrosia, disguised as a vine. As

shown in Figure 1.1, the cup is a striking example of dichroic glass, as it appears red and

pink when back-illuminated (transmitted light) yet is a mix of seafoam- and yellow-

green when front-illuminated (reflected light). The dichroic effect results from colloidal

nanoparticles of gold and silver, which were dispersed throughout the glass and were

made from very finely ground silver and gold dust. 4

However, any understanding of the mechanics of plasmonic behavior would not

come until much later; the foundational mathematical descriptions of these phenomena

began to emerge around the turn of the 20th century. The development of a theory of

surface plasmon polaritons is not very clean cut and began with papers by Sommerfeld

in 1899 and Zenneck in 1907 that modeled the propagation of radio waves along the

surface of a conductor with finite conductivity. The first observation of similar

phenomena in the visible band was in 1902 by Wood who observed anomalous dips in

the reflection spectra of light after hitting a metallic grating; however, it was not until

19
1941 that Fano connected this experimental observation with the related theoretical

work of Sommerfeld and Zenneck.

Figure 1.1 Photographs of the Lycurgus Cup, displaying its marvelous dichroic glass. 4

In 1957, Ritchie observed related phenomena in the form of loss during the

diffraction of electron beams at thin metallic films; in 1968, he then linked these

observations back to the original work on loss during the reflection of visible light at a

diffraction grating. The last key to a unified description of all these loss/surface wave

phenomena was also discovered in 1968 by Kretschmann and Raether, who were able to

use visible light to excite the same surface waves that Sommerfeld had with radio

waves. A unified theory of surface plasmon polaritons was then developed. As far as the

theory of localized surface plasmons in metal nanostructures, Mie laid out a clear

20
mathematical foundation for the description of this type of plasmon in 1908 and further

development of this theory did not take so many twists and turns as did the theory of

surface plasmon polaritons. 2

Plasmonics as an entire research field of its own did not emerge until the 1990s,

with early interest being focused on plasmonic sensors. Around the turn of the

millennium, much interest arose in research topics that leveraged the ability of plasmons

to confine light in deeply subwavelength scales, such as highly integrated photonics and

optical data storage. 5 In the first decade of the 21st century, considerable research focus

began to be paid to employing plasmonic nanostructures to greatly reduced the

emission lifetime of coupled emitters via the Purcell effect. The Purcell effect, which

describes the increase in the optical density of states (and, in turn, the radiative emission

rate according to Fermi’s Golden Rule) of a radiative dipole emitter via confinement

within a subwavelength cavity, had been known since Purcell’s original paper in 1946. 6

However, advancements in nanophotonic technology, particularly in photonic crystals

and plasmonics, needed to be made for the full potential of the Purcell effect to be

realized. In the early 2000s, Purcell enhancement factors (usually shortened to Purcell

factors) of ~10-100 were achieved.

In 2009, a landmark paper by Kinkhabwala et al. demonstrated 1,340-fold

enhancement of the emission brightness of single-molecule fluorescence by coupling to

plasmonic bowtie antennas through a combination of Purcell enhancement of the

21
radiative rate, absorption rate enhancement and enhancement of the intrinsic quantum

efficiency. 7 Similarly dramatic results were achieved using film-coupled plasmonic

nanocavities to produce upwards of 1,000-fold enhancement of the radiative rate of

single dye molecules, with simulation pointing to the potential for up to 10,000-fold

enhancement of the radiative rate with optimal coupling of the dye molecule to the film-

coupled nanoantenna. 8 The results from this paper were then extended to achieve

similar enhancements of single-photon emission from single quantum dots, 9

fluorescence from monolayers of 2D materials,10 and infrared emission of PbS quantum

dots.11

These results demonstrate the unparalleled ability of plasmonic nanostructures

to alter the emission properties of coupled emitters by orders of magnitude. One of the

main results from this dissertation concerns the use of film-coupled plasmonic

nanocavities to enhance the spontaneous emission rate of silicon vacancy centers in

diamond by two orders of magnitude. Vacancy centers in diamond have their own

history, spanning three decades, as a topic of research in photonics and quantum optics,

which will be discussed in depth in Chapter 3. The results communicated in this

dissertation aim to unite the singular properties of deeply subwavelength plasmonic

nanocavities for tailoring of spontaneous emitter properties with the important research

field of vacancy centers in diamond, which are perhaps the most important type of

22
solid-state quantum emitter, as they show great promise for implementation in quantum

computing architectures and quantum key distribution schemes.

1.3 Organization
The remainder of this dissertation is organized as follows. Chapter 2 provides an

introduction to the field of plasmonics, with an emphasis on recent advances to push the

scale of plasmonic devices down to molecular sizes. Chapter 3 gives an overview of the

existing research into color centers in diamond and their potential use as components in

quantum computing or quantum communications architectures. The chapters that

follow describe research undertaken during the course of my doctoral studies beginning

with a foundational study of EBL fabrication of nanogap plasmonic cavities in Chapter

4. Next, in Chapter 5, active tuning of the plasmon resonance of nanogap cavities is

explored via integration of nanometer-scale films of VO2. In Chapter 6, ~100x Purcell

enhancement of silicon vacancies in diamond is demonstrated via coupling to plasmonic

nanogap cavities. Finally, in Chapter 7, efficient nonlinear generation in the form of

SHG, THG, SFG and FWM are demonstrated in diamond thin films via coupling to

plasmonic nanogap cavities. Chapter 8 offers some concluding thoughts and outlook.

I would like to make a preliminary note on contributions from my

colleagues/collaborators. Chapters 2 and 3 provide background on plasmonics and

vacancy centers in diamond, respectively, which is sourced from published and cited

literature. As is common in experimental photonics work, I had co-authors for the

23
various projects that I worked on. The work in Chapter 4 was co-authored with fellow

graduate student Qixin Shen. The results presented in Chapter 5 were co-authored with

fellow graduate student Jon Stewart. Likewise, Chapter 6 was co-authored with Nathan

Wilson and Amirhassan Shams-Ansari. Finally, Chapter 7 was co-authored with Qixin

Shen, Nathan Wilson and Amirhassan Shams-Ansari. All research in this dissertation

was advised and guided by Prof. Maiken Mikkelsen.

24
2. Molecular-Scale Plasmonics
Recently, plasmonic nanogap cavities have been employed to demonstrate

hundreds- to thousands-fold Purcell enhancement of embedded spontaneous emitters,

values that are unparalleled in other nanophotonic architectures. This and similar

architectures are pushing the sub-wavelength nature of plasmonic devices to the

extreme of single-namoeter/molecular scales. Purcell enhancement is inversely

proportional to the mode volume of the cavity to which a spontaneous emitter is

coupled. As a result, molecular-scale plasmonics is enabling unprecedented Purcell

enhancement, as well as many other uniquely tailored light-matter interactions.

2.1 Fundamentals of Plasmonics


The field of plasmonics brought a novel means of confining light at optical

frequencies to subwavelength scales and offers promise for applications ranging from

sensing to communications and photocatalysis. The field is named for surface plasmon

polaritons (SPPs), which are coherent oscillations of the electromagnetic field and charge

density at a metal-dielectric interface. SPPs occur in metals as they require the free

electron gas found in the conduction-band electrons of a metal. In a SPP, energy from

light can be coupled to subwavelength scales in the dimension perpendicular to the

metal-dielectric interface and arises due to evanescent decay on either side of the

interface because the propagation constant is greater than the wave vector in the

25
dielectric. 2 A schematic representation of a SPP excited at the metal’s surface by

incident fields propagating in the dielectric is shown in Figure 2.1.

Figure 2.1 Schematic representation of the electromagnetic field associated with a


surface-plasmon polariton propagating along a metal–dielectric interface12

Plasmons are quasi-particles, which are utilized conceptually to quantize plasma

oscillations in the metal in increments of  p , where  p is the plasma frequency and is

an intrinsic property of a metal given by:

𝑛𝑒 2
𝜔𝑝 = √
𝑚𝜀0

Where n is the free electron density, e is the fundamental charge unit, m is the effective

electron mass, and 𝜀0 is the permittivity of free space.

Plasmons in the bulk of the metal oscillate with frequency 𝜔𝑝 ; however, surface

plasmons at the interface of the metal with a dielectric oscillate at a lower frequency

given by:

𝜔𝑝
𝜔𝑠𝑝 =
√1 + 𝜀𝑑

where 𝜀𝑑 is the dielectric constant of the adjacent material.

26
Plasmons also exist in the form of localized surface plasmons (LSPs), which are

non-propagating, coherent oscillations of the conduction electrons in a metallic

nanoparticle. 2 Optical illumination provides a direct means of exciting LSPs on metallic

nanoparticles. Typical nanoparticle shapes for plasmonics include nanospheres and

nanocubes. Once the electron cloud in these particles has been displaced by an incident

field, the nonequilibrium concentration of negative/positive charge on either side will

act as a restoring force and the plasmon’s coherent oscillation will continue until

overcome by damping, primarily via Joule heating. Figure 2.2 provides a schematic

representation of LSP oscillations.

Figure 2.2 Schematic representation of displacement of the electron cloud in a metallic


nanosphere resulting from incident optical fields. 13

The LSP oscillates with a frequency that is determined by the metal the

nanoparticle is made of (via the bulk plasmon frequency of that metal), as well as the

size and shape of the nanoparticle. Resonant phenomena arise when the frequency of the

driving optical fields match the LSP frequency, which is known as the localized surface
27
plasmon resonance (LSPR) of the metal. Other factors can also influence the LSPR

wavelength, such as coupling to neighboring nanoparticles and the surrounding

dielectric environment of the particles, enabling further flexibility for tuning the LSPR of

a plasmonic system.

Traditionally, noble metals such as gold and silver have been used for

plasmonics due to their high conductivity, which limits nonetheless substantial ohmic

losses present in any plasmonic system. While appealing in terms of their conductivity,

gold and silver have other drawbacks, including high material cost, incompatibility with

silicon fabrication, inoperability in the UV due to low bulk plasmon frequencies, and, in

the case of silver, rapid degradation in performance due to oxidation in atmospheric

conditions.

A variety of alternative materials have been explored to overcome some of these

drawbacks. Copper and aluminum offer satisfactory plasmonic properties at a much

lower cost while being compatible with silicon fabrication processes. Furthermore,

aluminum has a high bulk plasmon frequency and thus can operate in the UV, opening

up applications in that band including photocatalysis. 14,15 Other plasmonic metals with

high enough bulk plasmon frequencies to allow use in the UV include rhodium, gallium

and indium. 14,16 Another application of interest is creating plasmonic devices for

extremely high-temperature applications, such as aerospace. Titanium nitride is a

28
leading candidate for these applications, as it offers a superb melting point, although it

suffers from limited conductivity and challenging fabrication. 17

2.2 Survey of conventional plasmonic structures


Light absorption is a critical phenomenon for many key applications including

solar energy, optical communication, photocatalysis and sensing. Engineered structures

known as metasurfaces have been widely utilized to realize some of these applications. A

metasurface is an array of subwavelength elements in which each element is designed to

manipulate the phase and amplitude of transmitted, reflected, and scattered light. 12 An

ideal light-absorbing structure for sensing or communications applications would have

an absorption resonance that can be easily and precisely tailored, 100% absorption on-

resonance, 0% absorption off-resonance and moderate bandwidth. On the other hand,

broad absorption over an entire band could find many applications; for example,

absorption over the entire visible range would be useful for photovoltaics. Both of these

types of metamaterial absorbers have received considerable attention in the literature.

Here we will confine our literature overview to plasmonic metasurfaces, although a

number of dielectric metasurface absorbers have been extensively studied.

An early demonstration of a plasmonic perfect absorber was by Teperik et al.,

using nanostructured metal surfaces that operate in the visible and near-infrared regimes

and are fabricated by electrochemically depositing gold on a monolayer of latex spheres

supported on a gold substrate. 14 The latex is later removed, and empty spherical voids

29
remain underneath the gold. The metasurface relies on plasmons generated around these

voids to create perfect absorption. Furthermore, they are able to tailor the absorption

resonance wavelength by varying the thickness of the deposited gold. A consequence of

their fabrication method and its reliance on depositing colloidal latex spheres is that

deterministic placement of these void plasmonic hotspots is not possible.

An example of an EBL-fabricated metasurface perfect absorber is the Atwater

group’s work on broadband absorption in the visible using crossed trapezoidal arrays in

a MIM configuration with an average measured absorption of 0.71. 17 They also simulate

metasurface designs with average absorption of 0.85, indicating the potential of this

system for super absorption over the entire visible band.

Nielsen et al. improve on the Atwater group’s figure for broadband light

absorption in the visible by demonstrating a periodic array of EBL-fabricated, differently

sized and circularly shaped gap plasmon resonators with an average absorption of ~94%

for unpolarized light from 400-750 nm.18 By finite-element simulations, they are able to

verify that the absorption comes from localized gap surface plasmon whose resonant

excitations depend only weakly on the angle of incidence.

Giessen’s group has worked extensively on developing EBL-fabricated plasmonic

perfect absorbers in a MIM configuration, which they initially demonstrated in the near

infrared. Their structure consists of Au disks with 352 nm diameter and 20 nm height

separated from an Au ground plane by a 30 nm MgF2 spacer layer.20 They were able to

30
obtain 99% absorption on resonance at 1.6 µm for normal incidence; peak absorption for

the structure remains high even out to ±80° angles of incidence. They furthermore propose

refractive index sensing as an application for the structure as simulations revealed that

the resonance of the structure blue-shifts as it is submerged in a liquid with an

increasingly larger index.

Since the initial demonstration, Giessen’s group has also demonstrated that their

structure can be fabricated using a more scalable and cost-efficient process known as

colloidal etching lithography, which leverages colloidal deposition of polystyrene

nanospheres to form the template for the Au nanodisks. 21 Argon etching serves as the

“lithography” process to transfer the pattern from the polystyrene nanospheres to the Au

nanodisks; thus, the entire fabrication method is parallel in nature and therefore scalable.

A drawback of the process is that it does not allow for deterministic placement of the

nanodisks because the template polystyrene nanospheres are deposited from solution.

Even more recently, Giessen’s group has shown that laser interference lithography

can also be employed to create their structure, this time using nanosquares instead of

nanodisks. 22 This technique relies on creating and recording a 2D interference pattern in

photoresist before completing a standard evaporation and liftoff process. Laser

interference lithography is normally used to fabricate regular arrays of nanoparticles and

thus does allow for deterministic placement of metasurface elements but is limited in how

fabricated metasurfaces can be structured. Furthermore, by making their nanosquares out

31
of palladium, they show how the structure can be applied as a hydrogen sensor. In

ambient conditions, the metasurfaces demonstrate 99% absorption, which drops to 96%

in the presence of hydrogen and corresponds to a relative signal change of nearly 400%.

The combination of a scalable fabrication process and high sensitivity makes these

metasurfaces potentially suitable for commercial plasmonic devices.

Alternative plasmonic materials have also been utilized for fabricating perfect

absorbers. Kildishev’s group demonstrated a broadband absorber with high average

absorption of 95% over the entire visible range.23 The metamaterial consists of TiN

nanostructures in a square frame shape and is fabricated by EBL. The broadband

absorption of these TiN nanostructures is found to far outperform identically patterned

Au and Ag structures.

2.3 Active Tuning of Metasurface Absorption Resonance


Recently, much interest has been devoted to actively tuning metasurface

properties. The mechanisms for tuning are numerous but are generally designed to either

shift a metasurface absorption resonance or modulate the phase and/or angle of reflected

light. Phase modulation has been explored by the Atwater group. They created a

metasurface based on conducting oxides that allows for 180° tuning of the phase and 30%

change in the reflectance by applying a 2.5 V gate bias. 24 Furthermore, they demonstrated

modulation frequencies upwards of 10 MHz and electrical switching of ±1 order diffracted

32
beams by electrically controlling subgroups of metasurface resonantors, laying the

foundation of electrically tunable, beam-steering phased array metasurfaces.

Many demonstrations have centered on controlling the metasurface absorption

resonance. Capasso’s group has shown that the resonant wavelength of a plasmonic

antennas can be actively tuned by integrating graphene under the optical antennas and

placing electrical contacts. 25 Applying a voltage results in the resonant wavelength

shifting over a range of 650 nm in the mid to far infrared, corresponding to 10% of the

resonant frequency.

Placing strain on a metasurface is another means of changing its properties. Gutruf

et al. fabricated an all-dielectric system composed of TiO2 pillars on a PMDS matrix with

resonant wavelength in the visible. 26 Stretching the metasurface causes large, 2-3

linewidth shifts in the resonant wavelength. However, mechanical tuning is not ideal for

devices and the shifted resonance for high strains demonstrates poor peak absorption.

The use of phase-change materials is another means of controlling the absorption

properties of metasurfaces. YH2 undergoes an insulator-to-metal to (IMT) transition to

YH3 when hydrogen gas is flowed past it, an effect which is leveraged by Giessen’s group

in their work on creating YH2 optical antennas in a MIM configuration. 27 Inducing the

insulator to metal transition in YH2 makes it possible to switch the metasurface absorption

resonance on or off as is shown in Figure 1. While these antennas could be very useful for

applications in hydrogen sensing, flowing hydrogen gas is not a practical switching

33
mechanism for other applications due to the large device footprint that would be required

and hydrogen’s flammability. Furthermore, the absorption resonance for the antennas is

broad and does not have high peak absorption.

Giessen’s group has also explored leveraging the amorphous-to-crystalline phase

change in germanium antimony telluride (GST) to create actively tunable metasurfaces.

In a MIM configuration, they use a GST spacer layer to separate square Al nanoantennas

with side-length of 450 nm from an Al ground plane, yielding a final structure with a

strong absorption resonance (>90% absorption) at ~2.9µm. 28 Heating the sample induces

the GST phase transition causing a redshift in the absorption of 0.7µm while maintaining

strong absorption; this shift corresponds to a high reflectance contrast of ΔR = 0.6. They

also create a sample containing arrays four different sizes of square nanoantennas

arranged in a checkerboard pattern to demonstrate multispectral imaging as a potential

application of the system.

Another phase change that can utilized for active plasmonic devices is the IMT in

VO2, which, as is shown in Fig. 2, occurs when VO2 is heated to 68°C and switches crystal

structure from monoclinic to tetragonal.29 This is accompanied by a shift in its refractive

index, which occurs with a large magnitude in the mid to far infrared, as well as to a

smaller degree in the visible. The VO2 IMT has been leveraged in a few demonstrations of

active plasmonic devices in the mid to far infrared. One of these is the Capasso group’s

work on switching the scattering resonance of Y-shaped plasmonic antennas that operate

34
in the far infrared; impressive switching magnitudes of greater than a linewidth are

demonstrated. 30

2.4 Dimer Plasmonic Antennas and the Push Towards Molecular-


Scale Plasmonics
Another quintessential plasmonic device geometry is the dimer antenna, which

consists of two plasmonic nanoparticles positioned in proximity to each other (separation

of 10s of nm) such that an antenna resonance arises that is strongly influenced by the

coupling between the two nanoparticles comprising the dimer. In this configuration, very

strong “plasmonic hot spots” – areas of large electric field confinement and, therefore,

enhancement can often be created. One prototypical dimer architecture is the bowtie

antenna, which consists of two triangular nanoparticles placed in close proximity to one

another in the shape of a bowtie. Very high electric field enhancements are possible in this

geometry since the corners of the two triangles independently form plasmonic hot spots

and this effect is further amplified by placing them very close to one another in a dimer

configuration.

Kinkhabwala et al. used this structure to show 1,340-fold enhancement of the

brightness of single-molecule fluorescence. The authors employed electron-beam

lithography (EBL) to position the two sides of the bowtie dimer with a separation of about

15 nm. At the time of publication, this work constituted the largest nanophotonic

enhancement of a spontaneous emitter’s brightness achieved in literature by a factor of

10.7 The dimer architecture that they employed allowed them to create a plasmonic cavity
35
with a uniquely small mode volume, giving rise to extremely high enhancement of

incident optical fields. Dimer antenna structures like the one described above began the

push to create plasmonic devices at the single-nanometer/molecular scale, which has

enabled more complete leveraging of the ability of plasmonics to subvert the diffraction

limit and explore exotic phenomena such as single-molecule strong coupling18, enabling

optical transitions forbidden by selection rules, 19 and creating biosensors with ultra-low

limits of detection. 20 Dimer antenna structures; however, are not without their drawbacks,

such as difficult fabrication as two nanoparticles needed to be positioned within very close

proximity to one another, something that can often only be achieved via EBL, a fabrication

technique that is expensive and not scalable. This provided researchers with a motivation

to explore other plasmonic structures that might allow for investigation of the exotic

phenomena enabled by molecular-scale gaps, while offering scalable and inexpensive

fabrication.

2.5 Film-Coupled Plasmonic Nanocavities: Molecular-Scale


Plasmonics for Centimeter-Scale Devices
A key advancement in nanoscience that enabled exploration of alternative

plasmonic geometries with ultrasmall mode volumes was the development of recipes for

the colloidal synthesis of metallic nanoparticles with a range of shapes and

compositions.21–24 Colloidal synthesis is a highly appealing fabrication technique as it

offers much greater scalability than EBL, results in high-quality nanoparticles with

36
single-crystalline facets, and is capable of producing shapes with perfect fidelity at sizes

<50 nm where EBL fabrication is challenging.

Colloidal synthesis provides a solution to the problem of scalable and reliable

nanoparticle fabrication; however, the issue remained of how to arrange nanoparticles

into a device architecture capable of very high field confinement, like that of a dimer

antenna. Here again advancements in chemistry had something to offer the plasmonics

community. Polymer chemists had developed self-limiting monolayers that could be

stacked to yield dielectric coatings with single-nanometer precision in their thickness.25,26

These advancements in chemistry formed the necessary building blocks – the only

remaining thing that was necessary was a plasmonics insight that metallic nanoparticles

could resonantly couple to metallic films resulting in mirror charges in the metallic film

upon excitation of the cavity and thus electric field profiles that are effectively analogous

to that of a dimer antenna.

Mock et al. published an early study of these film-coupled nanocavities by

examining gold nanospheres coupled to gold films with resonances that could be tuned

by adjusting the number of self-limited polymer layers as discussed above. 26 Moreau et

al. expanded upon this work by instead using colloidal silver nanocubes as the

nanoparticles in film-coupled metasurfaces; in their article, they develop a framework for

understanding the perfect absorption of film-coupled metasurfaces in a metamaterials

context. In their description, the combination of having an electric conductor in the form

of the metallic film and an effective magnetic conductor in the form of the current loop of

the film coupled nanocavities leads to two reflected optical waves that are perfectly out of

37
phase with each other, thus yielding a reflection coefficient of zero. Because the metallic

film in opaque and thus has a transmittance of zero, the combination of these two

properties results in resonant perfect absorption by the film-coupled metasurface.27

The absorption resonances in the Moreau article were only about 50% in strength.

Akselrod et al. were able to realize >99% absorption on resonance as was predicted by

the metamaterial theory developed by Moreau et al. Furthermore, they were able to

leverage the scalable nature of the self-limited polymer layers and colloidal silver

nanocubes to demonstrate perfect absorber devices over centimeter-scale areas. The film-

coupled nanocavities also maintained most of their performance as resonant perfect

absorbers for large incident angles up to 60°.

Much subsequent work has leveraged film-coupled nanocavities for a wide range

of applications, their ability to strongly absorb and confine energy from incident optical

fields having been established. The literature to date has included demonstrations of

actively tunable metasurfaces,28,29 enhanced fluorescent biosensors,20 high-speed

pyroelectric photodetectors30 and plasmonic color printing.31 Perhaps the application of

film-coupled nanocavities that has received the most attention in the literature is Purcell

enhancement of embedded emitters, including, notably the highest Purcell factor to date

observed in any nanophotonic structure. More attention will be given to this application

in the following section, which presents a broader review of Purcell enhancement of

emitters coupled to nanophotonic cavities.

38
3. Vacancy Centers in Diamond and Enhancement of
their Properties in Nanophotonic Structures
3.1 Introduction and Motivation
Single photon sources are of emerging interest in numerous applications in

which leveraging the quantum nature of light is paramount. Among these are the fields

of quantum information processing and quantum metrology. A few applications in this

domain have achieved considerable commercial success, namely quantum key

distribution (QKD) and quantum random number generation (QRNG). However, other

technologies, such as on-chip quantum networks based on solid-state single photon

sources remain much less developed. The deficiencies of naturally occurring single

photon sources for quantum computing applications include low photon count rates,

short coherence times and lack of a directional emission pattern. Coupling to

nanophotonic structures has the potential to ameliorate many of these deficiencies. 9 In

the chapter to follow, I give an introduction to the properties and metrology of single-

photon sources, as well as an in-depth survey of the development of various single-

photon sources using a particularly promising material platform: vacancy defect centers

in diamond.

39
3.2 Fundamentals of Single Photon Sources and Hanbury-
Brown-Twiss Interferometry to Verify Single Photon Emission
An idealized single-photon source emits photons one-by-one at a periodic

interval. While physical single-photon sources always deviate from these ideal

properties, they do nonetheless exist in a variety of forms including single atoms,

molecules, quantum dots, or artificial atoms, like vacancy centers in diamond. 32 In the

idealized case, a periodic optical or electrical excitation source is used to pump the

single emitter, which will in turn emit in phase with the excitation source with negligible

temporal jitter; i.e., a perfectly stable emission lifetime.

There are a few key methods of evaluating how much a practical single-photon

source deviates from the idealized case. The first and most fundamental of these is

degree of second-order coherence, which is a measure of the antibunching property of

the emitted photons, that is, how much the probability of a multiphoton emission event

has been reduced, since for actual single photon sources this probability is always

nonzero. The second-order correlation function of an emitter, 𝑔2 (𝜏), can be directly

measured by performing a Hanbury-Brown-Twiss (HBT) interferometry experiment, the

configuration of which is shown in Figure 3.1.

40
Figure 3.1 Experimental setup for Hanbury-Brown-Twiss interferometry.

Light from a candidate single-photon source is sent through a beam splitter and then to

two avalanche photodiodes (APDs) and photon arrival times are recorded using ps-

resolution timing electronics.

In this type of experiment, light from a potential single-photon source is directed

towards a 50:50 beamsplitter. At each exit port from the beamsplitter is a low-timing-

jitter avalanche photodiode (APD), the output of which is connected to a timing

electronics device with ~10 ps resolution. The timing electronics box records the arrival

times of photons at each of the detectors. To interpret the results and extract the second-

order correlation function of the emitter, the time-series data from the two detectors are

cross correlated and normalized, which yields 𝑔2 (𝜏). For a single emitter, 𝑔2 (𝜏) will

41
display a dip around 𝜏 = 0; in the idealized case, 𝑔2 (𝜏) will dip to a value of zero at 𝜏 =

0. For real emitters, a value of 𝑔2 (0) < 0.5 is considered to be within the single-photon

emitter regime, where a typical 𝑔2 (𝜏) curve for a single emitter (in this case a quantum

dot) is shown in Figure 3.2. 𝑔2 (𝜏) will usually have an exponential shape on either side

of 𝜏 = 0 and the width of the dips corresponds to the fluorescence lifetime of the emitter,

assuming that the cross correlation is performed with sufficiently small bin sizes for the

time axis.

Figure 3.2 Typical second-order correlation function for a single emitter, in this case a
colloidal quantum dot.33

A 𝒈𝟐 (𝟎) value of 0.17 confirms single photon emission. Fitting exponentials to either side

of the dip yields an approximate emitter lifetime of 7 ns.

42
Another key measure of the quality of a single-photon source is the coherence

time / coherence length. For an ideal single-photon source, the emission does not

diphase at all and emitted photons are wholly indistinguishable; quantitatively, this

corresponds to:

2𝜏𝑠
=1
𝜏𝑐

where 𝜏𝑠 is the emitter’s fluorescence decay time and 𝜏𝑐 is the coherence time of the

emitter. Two-photon interference, such as via Hong-Ou-Mandel interferometry, is the

main method for measuring the coherence photons emitted by single-photon sources.

This technique will be described in detail later in the chapter during the discussion to

follow on vacancy centers in diamond and characterization of their properties as

quantum emitters.

3.3 The Nitrogen Vacancy Center in Diamond


The field of diamond photonics can be viewed as having emerged in 1997 with

the discovery of optically detected magnetic resonance (ODMR) at room temperature

from a single diamond negatively charged nitrogen-vacancy (NV) center.34 Shortly

thereafter, photostable single photon generation was observed from NV centers. 35,36 The

discovery brought hope for realizing solid-state qubits operating in ambient conditions

due to the capabilities of the NV center’s spin to be initialized, manipulated and

monitored optically without the use of cryogenic cooling. 37,38 Following these initial

demonstrations of control of the vacancy center’s electronic spin, additional articles

43
followed showing manipulation of adjacent N and 13C nuclear spins. Progress has also

been made in scaling to many cooperative qubits via photonic coupling. Additional

applications for the NV center include room-temperature nanoscale magnetometry, bio-

magnetometry, electrometry and decoherence microscopy.39

Figure 3.3 Schematic representation of the crystallographic structure of the NV center


in diamond. 40

In terms of its crystallography, the NV center is a diamond point defect with C 3v

symmetry and consists of a nitrogen-lattice vacancy pair with a <111> orientation. 41 Its

crystallographic structure is shown in Figure 3.1. NV centers can exist as a natural

byproduct of chemical vapor deposition (CVD) synthesis of diamond42 or as a result of

irradiation damage and annealing / ion implantation and annealing43 in both bulk

diamond and diamond nanoparticles. A neutral charge state of the nitrogen vacancy

center (NV0) also exists; however, it has not been shown to demonstrate the same

appealing properties for quantum information applications as the negatively charged


44
NV center (NV-). The characteristic spectral properties of these vacancy centers include

optical zero phonon lines (ZPLs) at 637 nm (1.945 eV)44 and 575 nm (2.156 eV)45, for the

NV- and NV0 centers respectively. Both centers also possess wide phonon sidebands

that are redshifted from the ZPL and in fact comprise the vast majority of the total

fluorescence. A typical emission spectrum of combined fluorescence from both NV0 and

NV- centers is shown in Figure 3.2.

Figure 3.4 Typical fluorescence spectrum from a combination of both NV- and NV0
centers under 532 nm laser illumination.

At cryogenic temperatures, the ZPLs of both vacancy centers become quite sharp.

This fact, along with the well-defined phonon sidebands of the centers indicate that the

optical transitions take place between discrete defect levels deep within the diamond

bandgap and, thus, the continua of valence or conduction band levels are not involved.

Because of these properties, the NV center is known as a deep-level defect in diamond. 39

The NV- center possesses a zero-field magnetic resonance at 2.88 GHz, which is a result

45
of transitions between the 𝑚𝑠 = 0 and 𝑚𝑠 = ±1 spin sub-levels of the spin triplet ground

state 3A2. Under cryogenic conditions (𝑇 < 10𝐾), the excitation spectrum of the NV- ZPL

exhibits a strong dependence of its fine structure on mechanical strain 46 and applied

electric fields. 47

A key property of the NV center for applications in quantum information, is that

its ground state spin has a room-temperature single spin coherence time (T2) of ~1.8 ms,

which is the longest for an electronic spin in any solid. 48 This property facilitates

coupling the ground state spin to neighboring electronic and nuclear spins in a manner

that can be resolved and manipulated. 49,50 Implementing the NV- center as a spin qubit

typically follows a common operation scheme with prepare-manipulate and interact-

readout phases. 39 First, optical spin-polarization is used to prepare the ground state spin

by sending in an off-resonance optical pulse. Next, the spin is manipulated by applying

DC and microwave fields; coupled nuclear spins can similarly be controlled by sending

in radio-frequency fields. A delay time is allowed to pass so that the NV can interact

with other spins and undergo state evolution. Lastly, the device is excited with another

off-resonance optical pulse and the integrated, Stokes-shifted fluorescence is measured,

which serves to read-out the final state of the ground-state spin. A schematic of this form

of device operation is shown in Figure 3.3.

46
Figure 3.5 Schematic of typical operation of a quantum circuit based on NV centers in
the prepare-manipulate/interact readout approach.39

The mode of operation described above relies heavily upon the performance of

the optical spin-polarization and readout mechanism, the current understanding of

which is limited and thus hinders device performance. A large range of values (42 – 96

%) for the degree of ground state optical spin polarization are reported in literature.

More thorough characterization of the ground state spin-polarization, including its

dependence on applied fields, strain and temperature, is necessary since this property

determines the preparation fidelity of the qubit.39 Significant improvements are also

required for the readout contrast; currently, the difference in signal between the 𝑚𝑠 = 0

and ±1 is a mere fraction of a photon due to low collection efficiencies in current device

architectures. To differentiate the two spin projections, a single qubit operation must be

performed many times to thus accumulate sufficient shot-noise statistics. 39 Some

enhancement of the readout contrast has been achieved via hyperfine coupling with N51

and 13C52 nuclear spins. Hopefully, as the optical readout mechanism of the NV center is

47
more fully understood, alternative techniques for readout contrast enhancement will

emerge. 39

3.4 The Silicon Vacancy Center in Diamond and the Search for
Alternative Vacancy Centers

3.4.1 Limitations of the NV Center and Motivation to Identify New


Types of Vacancy Centers
The first decades of research into applications of the NV center in quantum

information exposed the multitude of necessary characteristics for a single-photon

source candidate. In quantum communication, high quantum efficiency, high Debye-

Waller factor (i.e. low displacement of lattice atoms due to thermal vibrations), short

emission lifetimes and minimal spectral diffusion. Quantum sensing and computing

applications add additional requirements, such as a spin state that is addressable for

initialization, manipulation and read-out and possessing a coherence of time that is a

few orders of magnitude longer than the time require for fundamental operations on the

qubit’s state. 53 The NV center falls short when considering several of these requirements;

for example, its long fluorescence lifetime of ~11 ns and low-brightness ZPL at room

temperature (4% of total emission) limit the maximum achievable photon count rates for

NV centers in quantum photonic devices (that is, without introducing coupling to e.g.

nanophotonic cavities).

Nanophotonic cavities can indeed help to improve the fluorescence lifetime and

engineered, solid immersion lenses can increase the collection efficiency. Nonetheless,

48
the motivation remains to identify and characterize additional single photon sources

with more advantageous intrinsic properties. The past few decades of research have

resulted in the discovery of a multitude of additional types of diamond color centers that

maintain the fundamental beneficial traits of diamond as a material, while offering some

advantageous variation in the properties of the particular vacancy center. The emission

wavelengths of these vacancy centers span the majority of the visible and near-infrared

spectral ranges.54 Figure 3.4 provides an overview of the emission wavelengths of

various vacancy centers.

Figure 3.6. Emission bands of various color centers in diamond54

3.4.2 Discovery of the Silicon Vacancy Center and Some


Fundamental Properties
Many of the group IV atoms on the periodic table (in particular Si, Ge, Sn and

Pb) have been found to be capable of producing vacancy centers in diamond. Among

these, the negatively charged silicon vacancy (SiV) was the first to be discovered,

initially studied in the early 1980s53 and later more thoroughly investigated in the 1990s

49
resulting in the deduction of its electronic level structure and unambiguous

identification as a silicon-related defect. 55,56 Single SiV defects were first created and

isolated using ion implantation in 2006. 57 The sharp ZPL comprising a substantial

majority of the vacancy center’s total fluorescence at room temperature, in addition to a

relatively short emission lifetime of ~1 ns was an immediate draw; however, other

properties of the SiV were initially found to be less favorable. In the initial study, low

photon count rates (a few thousand per second) were observed, and a low radiative

quantum yield was deduced (5%).

Neu et al. achieved a significant breakthrough in fabricating single SiV centers in

2011 that rekindled scholarly interest in the vacancy center. Their approach was to grow

single defects contained in nanodiamond particles via microwave-plasma-assisted

chemical vapor deposition (CVD) on an iridium substrate, which resulted in vacancy

centers with a very high brightness of 4.8E6 counts/s at saturation. 58 Fueled by this

newly discovered technique for growing nanodiamonds containing single SiVs, a flurry

of research into the vacancy center’s followed, producing improved understanding of its

electronic structure, photophysical properties, spin-coherence times, as well as new

schemes for initialization, readout, and coherent preparation of the quantum state of the

defect.53

One of the key discoveries during this new burst of research into the SiV is that

the vacancy center possesses a nearly lifetime-limited fluorescence linewidth in both

50
nanodiamond particles59,60 and a bulk crystalline lattice. 61 This is an important property

for applications in quantum information, as it facilitates the fabrication of many,

intrinsically identical emitters that could be used as qubits in a quantum computing

architecture. 53 The physical explanation for this beneficial aspect of SiV emission is that

the vacancy center possesses group D3d inversion symmetry with its silicon atom sitting

in a split-vacancy configuration aligned along the <111> crystalline axis, so termed

because the silicon atom is centered in between two empty lattice sites as depicted in the

crystalline structure shown in Figure 3.5(a). This, in turn, screens the vacancy center and

its optical transitions from local variations in the electric field, which would otherwise

have the effect of Stark shifting the emission lines. 62

51
Figure 3.6 Fundamental structure and optical properties of the SiV in diamond. 63

(a) Crystalline structure and (b) electronic structure/optical transitions of the SiV.

Doublets are found both in the ground and excited states due to spin-orbit coupling,

which partially lifts the degeneracy. (c) Example low-temperature emission spectrum

from the four possible optical transitions that comprise the SiV’s ZPL.

3.4.3 Evaluating the Suitability of SiVs for Quantum Information:


Interlude on Hong-Ou-Mandel Interferometry
Mikhail Lukin’s group was the first to demonstrate indistinguishable photons

from different SiVs in their seminal 2014 article, which served to legitimize the SiV as a

52
strong candidate for applications in quantum information. 63 The primary experimental

results reported in their study were from a Hong-Ou-Mandel (HOM) interference

experiment using two spatially separated SiVs in a single-crystalline diamond sample

cooled to 5K.

HOM interference is the gold standard for testing the indistinguishability of two

single photon sources. The technique relies on the HOM effect, which was discovered in

1987 by the three eponymous physicists at the University of Rochester and occurs when

two identical photons hit a 50:50 beam splitter, one in each input port. When the two

photons have perfect temporal overlap, they will always exit on the same side of the

beam splitter. Equivalently, there is zero probability that the photons will exit separately

from each of the exit ports, resulting in a coincidence event.64 The HOM effect can only

be explained using the particle theory of light and second quantization of quantum

many-body theory and, as such, is an experiment that provides very direct illustration of

quantum optics. Adding to its significance, the HOM effect is employed as an

underlying physical mechanism for logic gates in the linear optical approach to

quantum computing. 65

Conducting a HOM experiment involves setting up the two photon sources in

question (in this case, two separate SiVs) in the configuration depicted in Figure 3.6

53
Figure 3.7 HOM interferometer utilized in Lukin and coauthors demonstration of
indistinguishable photons from separate SiVs. 63

The experiment is allowed to progress for a certain duration of time with photons from

the two sources passing through the beam splitter and two the two detector arms, each

containing an avalanche photodiode (APD). Once the experiment has run to completion,

the time-series count data from the two detectors is correlated and the resultant time

correlation function is plotted. As a result of the HOM effect, when the time delay

between the two correlated signals is very small (a few ps or less), the correlation

function dips in value, ideally to a value of zero at zero time delay. An example

correlation function from a HOM experiment utilizing identical photon pairs generated

via spontaneous parametric down conversion is shown in Figure 3.7. In this case data

are plotted as a function of variable spatial displacement instead of the correlation

function of time-series data, but the operating principle and communicated result are

exactly the same. In this case, a spatial displacement of ~120 µm corresponded to the two

arms of the HOM interferometer having equal length and is thus equivalent to zero time

54
delay if the data had been instead plotted as the correlation of time-series data from the

two detector arms.

55
Figure 3.8 Example time correlation function resulting from a HOM experiment.66

In this case photon pairs for the experiment are generated via spontaneous parametric

down conversion. The solid grey line is the theoretical prediction and the yellow dots

correspond to experimental data. The peaks where experimental data deviates for theory

are a result of employing a rectangular filter instead of a Gaussian filter like in the

theoretical model.

3.4.4 Evaluating the Suitability of SiVs for Quantum Information:


HOM Interferometry Experiments on SiVs
The diamond sample utilized in the Lukin group’s study consisted of a low-

strain, high-pressure high-temperature <001> diamond substrate with a high-purity,

homoepitaxial layer on top, which was grown using the microwave plasma-assisted

CVD technique that had been shown to be so successful for producing high-quality SiVs

in previous reports. The SiV sites were incorporated into the top layer of the diamond
56
during the growth process by etching a piece of silicon carbide (SiC) with the growth

plasma, an approach that had been shown in a previous article to produce highly

uniform SiVs with minimal inhomogeneous broadening in their emission. 61

To perform their HOM experiment, the authors utilized dichroic mirrors to

spectrally separate the SiV fluorescence into its ZPL (~737 nm) and phonon-side-band

(~760-860 nm). The ZPL had to be then further filtered to isolate a single two-level

transition, as there are in fact four possible ZPL transitions as a result of the electronic

structure of the SiV, which is shown in Figure 3.5(b). To summarize this structure, the

ground and excited states each have the form of a fourfold-degenerate manifold in

which two degenerate orbitals are occupied by a particle with 𝑆 = 1/2. Each of the

excited state doublets has dipole transitions to the two ground states, which are labeled

A, B, C, and D in the electronic energy diagram. These four dipole transitions are

identifiable in the PL spectrum of single SiVs at single-Kelvin temperatures; an example

PL spectrum showing the resultant four distinct peaks from a single SiV identified by

the authors is shown in Figure 3.5(c).

For the HOM experiment, Lukin’s group uses solid etalons to bandpass filter

exclusively transition C, such that only a single fluorescence peak is transmitted, as is

desired for indistinguishable photon generation. Some inhomogeneous broadening is

nonetheless present among different vacancy centers on the sample, so the authors

needed to select two SiVs with a very high degree of spectral overlap. The resultant

57
second-order intensity correlation function from a HOM experiment from two suitable

selected SiVs is shown in Figure 3.8.

Figure 3.9 Hong-Ou-Mandel Interferometry with SiVs in diamond. 63

Second-order intensity correlation functions for interference of two distinct SiVs when

polarized parallel to one another (pink data) and orthogonally to one another (green

data).The solid lines correspond calculated/predicted results.

Here the authors varied the relative polarizations of light from the two SiVs to

switch between distinguishable and indistinguishable photon states. The authors also

developed a theoretical model to plot predicted correlation functions for the two

different relative polarizations, taking into account a variety of other independently

measured optical parameters for the two vacancy centers. Their model agrees perfectly

with experimental data when accounting for shot noise. Finally, the authors calculate a

58
HOM visibility (a measure of the contrast between the distinguishable and

indistinguishable states) using the two data sets of:

𝑔2⊥ (0)
𝜂= 2 = 0.72 ± 0.05
𝑔∥ (0) + 𝑔2⊥ (0)

Upon comparing this result to prior reports on NV centers, the authors conclude that the

SiVs measured have shown a superior degree of indistinguishability/optical coherence.

This fact elucidates a significant potential advantage of SiVs in terms of single photon

generation for quantum information / quantum optical applications.

3.5 Emerging Vacancy Centers in Diamond: The SiV 0 and


Additional Group IV Vacancy Centers
While the negatively charged silicon vacancy center has many beneficial

properties, it is not without drawbacks, such as a short spin coherence time even at

cryogenic temperatures and a low intrinsic quantum efficiency. 67,68 Many researchers

suspected that other group IV elements would be capable of forming vacancy centers in

diamond with the same symmetry/bi-vacancy configuration and perhaps one of these

elements might form a vacancy with largely advantageous optical properties for

quantum information and limited drawbacks. The last decade has seen the discovery of

Ge, Sb and Pb vacancy centers (ordered chronologically by discovery date).

These novel group IV vacancy centers were found to share some advantageous

properties in common with the SiV, such as a narrow linewidth and low Debye-Waller

factor. However, some key differences are expected as well, such as larger splitting in

59
the ground state due to a higher atomic mass, which would result in less spin mixing

and higher allowable qubit operation temperatures.53 Furthermore, the emission

wavelength of group IV vacancy centers blueshifts as one moves down the group IV

column of the periodic table. The calculated energy levels of each of the known,

negatively charged group IV vacancy centers is shown in Figure 3.9. Lastly, greater

control over vacancy center doping is expected, since silicon, unlike the other group IV

elements, is a common contaminant in CVD chambers used to grow diamond.

Figure 3.10 Calculated energy levels of the ground state (GS) and excited state (ES) of
the negative charge state, group IV vacancy centers in diamond. 53

The narrow linewidth, in particular, is a key property shared among all group IV

vacancy centers due to their common di-vacancy symmetry that screens against local

electric fields and has critical consequences for the quantum information potential of

these vacancy centers. In particular, this characteristic leads to much greater

indistinguishability of emitted photons from various vacancy centers across the device

and thus results in higher fidelity of quantum operations in e.g. a linear-optical quantum

60
computing scheme, such as projective Bell-state measurements. 53 The low Debye-Waller

factor of the group IV vacancy centers is also of substantial importance as it leads to the

common trait of high-probability emission into the ZPL, with much smaller phonon side

bands than the NV center. The quantum information consequences of this property are

straightforward, as only coherent light from the ZPL can be utilized in quantum

circuits. 53

This overview of key vacancy centers in diamond would be incomplete without

mentioning the neutrally charged silicon vacancy center (SiV0), which has recently been

discovered and characterized. The SiV0 possesses several advantages for quantum

information applications, as it has both a long electron-spin-coherence time and 90% of

its PL is emitted into its nearly transform-limited ZPL at 946 nm. A significant

engineering challenge for this vacancy center, however, has been reliably and

controllably producing the neutral charge state and the feasibility of controllably

producing single SiV0 defects is still an open question.

The search for novel vacancy centers and additional fundamental and device

engineering studies are still ongoing in the field of optical engineering of defect centers

in diamond. While this chapter has given an overview of the current state of the field

and the advantages and disadvantages of each vacancy as they are currently

understood, this area of research is rapidly evolving and new discoveries could

completely alter the landscape of vacancy center research, as well as the overall outlook

61
of vacancy centers in diamond as candidates for components in quantum computers or

quantum communication channels.

3.6 Purcell Enhancement of Diamond Vacancy Centers and


Other Spontaneous Emitters Embedded in Nanophotonic
Structures
Single emitters are often limited in their applications by poor intrinsic properties,

including long emission lifetimes, low quantum yield and nondirectional emission. A

powerful approach for enhancing the properties of spontaneous emitters is to utilize the

Purcell effect by embedding the emitter in a resonant micro- or nano-cavity leading to an

increase in the optical density of states and, therefore, the spontaneous emission rate.

The optical density of states and the spontaneous emission rate are linked by

Fermi’s golden rule:

𝛾𝑠𝑝 ∝ |𝒑|2 𝜌0 (1)

Purcell enhancement of the optical density of states and thus the emission rate is

quantified by the Purcell factor, 𝐹𝑝 , which is proportional to:

𝜌𝑐𝑎𝑣 𝛾 𝜆 3𝑄
𝐹𝑝 ∝ ∝ 𝛾𝑠𝑝
0 ∝ ( 𝑛) (2)
𝜌0 𝑠𝑝 𝑉

where 𝜌𝑐𝑎𝑣 is the optical density of states of the emitter in a cavity, 𝜌0 , is the

optical density of states of the emitter in free space, 𝛾𝑠𝑝 is the spontaneous emission rate

0
of the emitter in a cavity, 𝛾𝑠𝑝 is the spontaneous emitter in free space, 𝜆 is the emission

wavelength, 𝑛 is number of emitters, 𝑄 is the cavity quality factor, and 𝑉 is the cavity

62
mode volume. 32 From this equation, it is apparent that there are two possible means of

achieving large Purcell factors: creating a cavity with a high quality factor, which is a

measure of the number of oscillations that can occur in the cavity before damping takes

over, or creating one with a small mode volume. Thus, most of the approaches for

performing Purcell enhancement of embedded emitters can be grouped by whether the

goal is to optimize the cavity quality factor or the mode volume.

One recent notable approach utilizing a high-Q cavity is the work of

Ramachandrarao et al. on embedding a quantum dot in a composite photonic crystal

cavity composed of an optical nanofiber and a nanofabricated grating. The emission rate

enhancement for individual quantum dots is modest (Purcell factor of 7); however,

outcoupling the emission to a nanofiber is favorable for quantum network applications. 33

A high cavity quality factor is also leveraged in the work of Jeantet et al. on

embedding a carbon nanotube in a microcavity at low temperature. They are able to take

advantage of the cavity feeding effect on the phonon sidebands of the emission and lock

the nanotube emission at the cavity resonance wavelength.34 As a result, the single

photon character of the emission was maintained through tuning of the cavity resonance

over a 4 THz band. This points towards the feasibility of carbon-based, tunable single

photon sources at telecom wavelengths.

A successful approach using a low volume cavity is coupling a quantum dot to a

dielectric nanowire, as demonstrated by Kolchin et al. Using atomic force microscope

63
(AFM) manipulation, they are able to optimally position a single CdSe/ZnS quantum dot

to obtain a Purcell factor of 31 and couple 80% of the emission into a strongly confined

waveguide mode in a nanowire.35 Additionally, the emission dynamics of the quantum

dot are favorably modified in terms of increased brightness and strong suppression of

blinking and optical saturation. The all-dielectric nature of their system makes it

favorable for integration with CMOS components; however, positioning with an emitter

with an AFM tip is not an easily scalable approach.

The nanopatch antenna has also been utilized to demonstrate significant Purcell

enhancement of embedded emitters as a result of an ultrasmall cavity mode volume.

Rubidium dye was integrated into the NPA gap layer; the cube size and gap thickness

were chosen to match the cavity scattering resonance with dye PL. This yielded a

record-high Purcell factor of ~2000.11 The Purcell-enhanced emitter furthermore

maintains a high quantum efficiency of upwards of 50% and directional emission, which

enables an 84% collection efficiency with a 0.9 NA objective.

Single photon emission has also been dramatically enhanced with NPAs.

CdSe/ZnS quantum dots were deposited in the NPA gap, again with the antenna

resonance tailored to match with the PL from the quantum dots. The result was a 540-

fold enhancement in the quantum dot emission rate, which was limited by the

instrument response function of the single photon detectors that were used. 31

64
Furthermore, the emission brightness increases by a factor of 1900 with a high radiative

quantum yield of 50%.

Like other systems that have been reviewed, Purcell enhancement with the NPA

does not allow for deterministic coupling to individual emitters, but, rather, relies on

stochastic coupling during the metasurface self-assembly process. The present research

on NPA metasurfaces fabricated by EBL will allow for deterministic coupling, while

retaining the unique potential of the NPA for large Purcell enhancement.

Additionally, applications of colloidal quantum dots are limited because of their

susceptibility to photobleaching and degradation over time. In contrast, nitrogen-

vacancy (NV) and silicon-vacancy (SiV) centers, which are the focus of this study, have

excellent photostability and long coherence times. These color centers occur when a

nitrogen or silicon atom, respectively, substitute into the diamond’s carbon lattice and

leave an adjacent vacancy site. Single photon emission from these color centers occurs at

their zero-phonon line (ZPL) wavelength, which are 575 nm for NV 0, 637 nm for NV-

and 737 nm for SiV-.37,38

A large research community has formed around characterizing and enhancing

single photon emission from diamond. A notable recent work is the Loncar group’s

demonstration of enhanced single photon emission from single NV centers using

ordered arrays of plasmonic apertures.39 Their fabrication technique relies on defining

the pillar-array pattern using EBL and transferring it into diamond using oxygen-based

65
reactive ion etching (RIE). Silver is then evaporated on top of the patterned diamond,

yielding the diamond pillars encapsulated in a plasmonic cavity. One of the strengths of

their technique is that it allows for parallel fabrication of large numbers of embedded

NV centers. Their samples contain hundreds of diamond pillars and their procedure has

a >10% yield of single-center devices, indicating a strong potential for scalability.

Riedrich-Möller et al. present an approach to Purcell enhancement of diamond

that relies on a high cavity quality factor. They develop a unique focused-ion-beam-

based system to create photonic crystal cavities in single crystal diamond that enables

low-damage fabrication of thin, free-standing membranes. 40 Their technique yields one-

and two-dimensional photonic crystals with quality factors of up to 700 in diamond.

During post-processing etching, they tune the cavity modes into resonance with the SiV

ZPL giving a fluorescence intensity enhancement factor of 2.8. Their work is an

important step in improving precision structuring of thin diamond films, which may be

essential components of eventual quantum networks.

The larger footprint of diamond-based single photon sources presents a

challenge for their integration with NPAs, as a gap thickness of <20 nm is required to

obtain a film-coupled mode. As will be discussed, my approach for integration into

plasmonic nanogap cavities is to etch diamond slabs containing silicon vacancies (SiVs)

to ~10 nm thickness. At these miniscule sizes, one must contend with surface effects in

the color centers, which are detrimental to diamond single photon emission.

66
Nonetheless, the upside of this research is significant as combining these color centers

with NPAs could lead to fast, efficient and robust single photon sources for applications

in quantum information.

67
4. EBL Fabrication of Plasmonic Nanogap Cavities with
Single- and Dual-Band Resonances in the Near-Infrared
4.1 Introduction
Plasmonic nanogap cavities were initially fabricated exclusively via dropcasting

or spincoating colloidal nanoparticles. 8,33,69 This approach has several advantages,

including the ability to easily and rapidly fabricate metasurfaces over centimeter-scale

areas; however, it is limited by the range of nanoparticle shapes and sizes that can be

readily synthesized. By contrast, fabrication using electron beam lithography (EBL)

would enable much greater freedom over the choice of nanoparticle size/shape. Here I

utilized EBL to fabricate nano-disks, -cubes and -rectangles with a range of sizes and fill

fractions to demonstrate near-perfect absorption on resonance with performance

comparable to that of colloidal nanogap metasurfaces. Furthermore, metasurfaces

fabricated with nanorectangles demonstrate dual-band, polarization-metasurfaces with

spectral separations of up to 500 nm.

4.2 Sample Design and Fabrication


The geometry of the sample consists of arrays of nanoparticles on a Au substrate

separated by a thin dielectric layer (7 nm of Al2O3) and is shown in Figure 1a-c. This

structure results in film-coupled cavity mode with a resonant wavelength that is

determined by the nanoparticle dimensions and thickness/effective index of the dielectric

layer. The charge distribution in the fundamental mode results in a current loop between

the gold structures and gold substrate, thus inducing a strongly enhanced magnetic field.
68
This enables the surface to be impedance matched to free space resulting in near-perfect

absorption of light. 27,69,70

To fabricate the EBL metausrfaces, 5 nm Cr followed by 75 nm of Au is first

deposited onto a silicon substrate by electron beam evaporation. Then the dielectric gap

layer is grown via followed by atomic layer deposition (ALD). EBL is then used to pattern

arrays of absorbing elements on top of the dielectric layer. The size of the elements are

characterized via scanning electron microscopy (SEM), while the thickness of the

dielectric layer is determined by the deposition rate of the ALD and verified using

ellipsometry measurements.

69
4.3 Absorption Characteristics of EBL-Fabricated Nanodisks and
Nanocubes

Figure 4.1: Single-Resonance, EBL Fabricated Nanogap Cavities

(a) Structure of the sample: patterned gold elements (height 30 nm) separated from a gold
ground plane by 7 nm of Al2O3. The sample has four distinct sections: I. cubes of different
sizes with a fixed fill fraction of ~25%; II. Disks of different sizes with a fixed fill fraction
of ~25%; III. Disks with various fill fractions; IV. Rectangles with various ratios of width
and length. Nine, twelve or sixteen elements are shown for clarity; however, the actual
size of one array is 100 µm × 100 µm and contains over ten thousand elements. (b) SEM
images of 93 nm EBL-fabricated disks. (c) SEM images of 160 nm EBL-fabricated cubes.
(scale bar is 500 nm) (d) Reflection spectra of EBL-fabricated cubes and disks of various
sizes. (e) Reflection spectra of colloidally-synthesized silver nanocubes of different sizes.
(f) Comparison of the full-width at half-maximum (FWHM) between EBL-fabricated
disks, cubes and colloidal silver nanocubes.

70
First, I study structures with a single fundamental resonance, including disks and

cubes fabricated using both EBL and colloidal synthesis, before examining dual-band

rectangular absorbers. To evaluate the performance of EBL-fabricated nanogap absorbers,

reflection spectra were measured from film-coupled gold disks and cubes using a white

light source at nearly normal incidence as shown in Figure 1d (additional spectra in SI

Figure 1). The resonance wavelength is tuned from the visible to the near-infrared by

varying the size of the absorbing elements, thus changing the volume of the cavity. With

a surface coverage density of approximately 25%, an absorption of 80% or greater can be

achieved on resonance for eight different sizes. EBL-fabricated disks are observed to have

higher peak absorption compared to EBL-fabricated cubes.

Previous literature shows that metasurfaces based on colloidal nanocubes show

near-perfect absorption with a similar nanogap structure. 27,69,70 However, to this point, it

remained unclear whether similar performance could be obtained with EBL-fabricated

nanoparticles, which are not limited by the variety of sizes and shapes that can be

synthesized. To compare the performance of the EBL-fabricated, film-coupled

nanoparticles with colloidal nanoparticles, metasurfaces of colloidal silver nanocubes

were produced using a similar 7 nm Al2O3 layer on a gold ground plane. The colloidal

nanocube metasurfaces were fabricated by first dip-coating a single polymer adhesion

layer onto the same substrate used for EBL - consisting of 7 nm Al2O3 on gold. Nanocubes

in water were then dropcast onto the sample using a 10 mm coverslip and a deposition time

of 1 hour.
71
Figure 1e shows reflection spectra of these metasurfaces for nanocubes with side

lengths from 65 to 120 nm. In these experiments, for the same resonance wavelength, EBL-

fabricated nanoparticles typically have higher peak absorption than colloidal nanocubes

which is most likely due to higher fill fractions and better size uniformity. In addition, the

full-width at half-maximum (FWHM) are extracted from the spectra for both the colloidal

nanocubes and the EBL-fabricated particles and are shown as a function of resonance

wavelength in Figure 1f. The bandwidth of EBL-fabricated and colloidal samples is

comparable at similar resonance wavelengths demonstrating that the resultant plasmonic

cavities have similar quality factors.

72
4.4 Dependence of Absorption Properties on Fill Fraction of the
Nanodisk Metasurface

Figure 4.2: Effect of Varying the Fill Fraction on the Absorption Resonance of EBL
Nanogap Metasurfaces

(a) Reflection spectra from disks with various fill fractions, which show that the FWHM
of the spectra are sensitive to the density of gold disks. The fill fraction ranges from 9% to
35%. (b) SEM images of the disk arrays in (a) in order of increasing fill fraction. (c)
Experimental contour plot of reflection spectra with variable fill fraction. For clarity,
interpolation is used between data points. (d) Simulated reflection spectra using the same
dimensions as (c) for comparison, showing good agreement with experiments.

To control the peak absorption and the width of the plasmon resonance, critical to

e.g. photovoltaics71 and photoluminescence enhancement,7,8 I fabricated arrays of 100 nm

73
nanodisks with variable pitch and, thus, different fill fractions of absorbing elements on

the surface. The reflection spectra for each of these arrays are shown in Figure 2a and the

corresponding SEM images are shown in Figure 2b. As the fill fraction is increased

incrementally from 9% to 35%, the FWHM of the plasmon resonance grows from 80 nm

to 122 nm. Additionally, the resonance wavelength redshifts with increasing fill fraction,

particularly for the highest fill fraction of 35%. The simulations shown in Figure 2d are

found to be consistent with experiments for fill fractions below 26%. However, at the

highest fill fraction, a significant redshift in the resonance wavelength is observed which

may be explained by near-field coupling between neighboring elements that, in turn,

delocalizes the plasmon mode. 72

74
4.5 Absorption Properties Dual-Band, Polarization Sensitive
Metasurfaces

Figure 4.3: Polarization-Dependent Absorption Properties of EBL Nanorectangle


Metasurface

(a) Reflection spectra from rectangle absorbers with varying white light polarization angle
from transverse electric (TE) configuration to transverse magnetic (TM) configuration.
The width of rectangular absorbers is 110 nm, while the length ranges from 150 nm to 170
nm. The inset shows the TE configuration and TM configuration when white light is
incident on the sample. (b) Reflection spectra from the rectangle absorbers shown in (a)
when the incident white light polarization is 50°. (c): Resonance wavelength of rectangular
absorbers as a function of rectangle length from 140 nm to 200 nm. The transverse mode
associated with the 110 nm width dimension is at approximately 1000 nm, while the
longitudinal mode can be tailored from 1118 nm to 1527 nm.

Next, to realize nanostructures with two resonances that can be tuned

independently of each other, I studied arrays of rectangular elements. Two resonances

arise in this structure: one each associated with the length and width dimensions of the

75
rectangle respectively and are excited to varying degrees depending on the polarization

of incident light. The transverse electric (TE) configuration occurs when the incident

polarization is along the length, while the transverse magnetic (TM) configuration occurs

when the incident polarization is along the width. As the polarization of incident light is

swept from along the length dimension to along the width, the coupling of the incident

light to the two modes varies resulting in a reduction in the amplitude of the longitudinal

mode and an increase in the amplitude of the transverse mode, which reaches a maximum

when the polarization is in the TM configuration. For the 110 nm width and 150 nm length

rectangles, shown in Figure 3a, this tuning corresponds to a gradual shift in the resonant

wavelength from 982 nm to 1172 nm. Additionally, rectangles with 160 nm and 170 nm

length were investigated to further examine tailored double resonances on a single

element; their polarization-dependent absorption properties are also shown in Figure 3a.

When the incident polarization is near 45°, there is significant absorption into both the TE

and TM modes. This can be seen in the absorption spectra for the 150, 160 and 170 nm by

110 nm rectangles with 50° incident polarization shown in Figure 3b.

76
4.6 Benefits of Dual-Band Absorbers for Purcell Enhancement

Figure 4.4: Simulations to Demonstrate Mode Overlap in EBL Rectangle Nanocavities

(a-c) Spatial distribution of electric field enhancement in the gap of a film coupled, 110 ×
160 nm rectangular nanoantenna. (a) Transverse gap plasmon mode excited by TM
polarized light at the resonant wavelength of 990 nm; (b) Longitudinal gap plasmon mode
excited by TE polarized light at the resonant wavelength of 1250 nm; (c) The product of
the electric field enhancements in (a) and (b); local maxima indicate regions where the two
gap plasmon modes are strongly spatially overlapped. (d-e) Spatial distribution of
excitation rate enhancement  ex /  ex0 , spontaneous emission rate enhancement  sp /  sp0 ,
quantum yield QY , total fluorescence enhancement EF for (d) 110 × 160 nm
nanorectangle, (e) 110 nm nanocube and (f) 160 nm nanocube.

77
To demonstrate that the transverse and longitudinal modes in rectangular

absorbing elements can be utilized simultaneously, the electric field distribution is

calculated using full-wave simulations based on the finite-element method (COMSOL

Multiphysics) at the resonance wavelengths of both the transverse and longitudinal

modes for 110 × 160 nm rectangles. For the resonance wavelengths of the both modes, the

electric field shows large, over 45-fold, enhancement along the two sides of the rectangle

that are perpendicular to the incident polarization, as seen in Figure 4a,b. In addition to

demonstrating two resonance modes with spectral separation controlled by the difference

between the rectangle’s width and length, it is further shown that the enhanced electric

fields from these two resonance modes possess good spatial overlap. To further illustrate

this spatial overlap, the spatial distribution of the field enhancement is shown in Figure

4c for 990 nm and 1250 nm light impinging on the nanorectangle, which matches the two

resonant wavelengths. This spatial distribution is calculated by multiplying the field

enhancements from Figures 4a and b. The field enhancements could be even greater by

utilizing a thinner dielectric layer than the 7 nm of Al2O3 used here.

The enhanced electric field at two wavelengths with good spatial overlap are

promising for optical processes involving multiple wavelengths. Fluorescent emission is

selected as an example to illustrate the advantage of these dual resonances; however, other

processes like harmonic generation and four wave mixing could also benefit. Specifically,

it is assumed that a dipole emitter, such as a PbS quantum dot, with an emission

78
wavelength of 1250 nm is placed in the nanogap and excited by 990 nm light with

polarization perpendicular to the nanorectangle’s width. It is shown in Figure 4d that the

total fluorescence enhancement factor EF and spontaneous emission rate enhancement

 sp /  sp0 achieve large enhancements compared with an emitter placed on glass as a

control. The simulated dependence of this enhancement on the emitter position is

obtained by sequentially placing a single dipole emitter on a discrete 41 × 41 grid in the

nanogap underneath the nanorectangle. The emission rate enhancement can be attributed

to the increased local density of states due to the presence of the longitudinal mode

centered at the emission wavelength. While the total fluorescence enhancement factor

EF is a consequence of both transverse mode and longitudinal mode, which is defined

as

 ex QY
EF = , (1)
 0 exo QY 0

where  is the collection efficiency,  ex is the excitation rate, QY is the quantum yield

and the superscript “0” refers to the emitter on glass. The transverse mode at 990 nm

causes the excitation rate enhancement distribution depicted, while the longitudinal mode

at 1250 nm enhances the photon out-coupling channel and thus results in the spontaneous

emission rate enhancement and quantum yield enhancement distribution shown. The

final EF pattern can be regarded as the result of the spatial overlap of the two modes

according to Eq. (1), similar to the product of the electric field distributions shown in

79
Figure 4c. Thus, the total fluorescence intensity and spontaneous emission rate can be

enhanced simultaneously with dual-band resonances in a single nanoparticle.

In order to compare the performance to singly resonant nanoparticles, simulations

are also performed using the same method for 110 nm and 160 nm nanocubes. In those

cases, either the total fluorescence or the spontaneous emission rate obtains large

enhancement depending on whether the resonance of the nanocube is overlapped with

the excitation or the emission wavelength. However, simultaneous enhancement cannot

be achieved without the presence of dual-band resonances.

4.7 Conclusion
The metasurfaces expounded upon above operate in the visible and near-IR and

possess resonances that are spectrally separated while spatially overlapped, which could

provide critical advantages for multiplexed sensors and detectors, as well as processes like

stimulated Raman scattering, entangled photon generation and multiphoton absorption. For

the nanocubes, particles fabricated using both colloidal nanoparticles and EBL and

displayed similar absorption performance. This shows the promise for combined bottom-

up and top-down fabrication, as ALD enables single-nanometer control of the critical

vertical gap dimension, while EBL allows for a wide variety of sizes and shapes of

nanoparticles as well as deterministic element placement. This platform also opens up

possibilities to tailor the radiation pattern, mix different sizes and shapes of nanoparticles

and create integrated waveguides.

80
5. Actively Tunable Nanogap Cavity Metasurfaces Via
Integrated VO2 Thin Films
5.1 Introduction
Integration of active materials into nanophotonic structures provides a means of

tuning optical properties in real-time and is promising for a new generation of multi-

functional, nanoscale optoelectronic devices. Active metasurfaces have been

demonstrated using a wide range of mechanisms, including electrical modulation, 28,73–77

polarization of incident light, 78 mechanical strain,79,80 hydrogen gas flow, 81,82 UV

exposure, 29,83 femtosecond laser pulses84,85 and heating. 86–91 While there have been several

demonstrations of robust tuning at mid-IR and longer wavelengths, 73,77,91 active

metasurfaces in the visible and near-IR have generally been limited by tuning

mechanisms that are not well-suited for practical devices, for reasons such as lack of

reversibility or degradation of switching over multiple cycles. 28,29,79,92 One option for

overcoming these limitations is employing the phase-change material vanadium dioxide

(VO2), which is capable of robust switching via heating above 68°C. At this temperature,

VO2 undergoes a first-order, crystalline phase transition with an concomitant change in

its refractive index, a process that is fully reversible by cooling below the transition

temperature.93 It has also been reported that this phase change can be induced by

applying moderate voltages resulting in a less than 2 ns response time 94 and, thus,

increasing the promise of VO2 for integration into practical, active metasurface devices.

81
Previously, VO2 has been used to create active metasurfaces over a large gamut

of the electromagnetic spectrum86,90,91,95–97, typically with bulk materials or sputtered

films with thicknesses of 50-150 nm98–103 or more rarely by utilizing a lattice-matched

substrate to grow of thinner films. 104 However, integrating nanometer-scale films of VO2

with a nanophotonic structure on non-lattice-matched substrates has remained an

outstanding challenge, yet offers the promise for robust, deeply subwavelength active

optoelectronic devices to enable e.g. nanoscale optical memories and holograms, as well

as dynamic control of quantum emitters. Furthermore, conventional techniques to

deposit VO2 thin films, such as pulsed laser deposition (PLD), are non-conformal, only

uniform over small areas, and are incompatible with many materials. ALD provides a

promising, alternative deposition technique to enable highly uniform films over wafer-

scale areas, conformal coating of three-dimensional or flexible structures, as well as

integration of dissimilar materials that can be deposited with abrupt interfaces. Lastly,

the resultant ultrathin layers allow heat to be removed faster, which would enable

devices with a faster thermal response.

82
Figure 5.1: Actively Tunable Nanogap Cavity Metasurface Concept

(a) Schematic of sample structure consisting of Au nanodisks (height = 30 nm) separated


from an Au ground plane by 9 nm of VO2. (b) Illustration of phase transition of VO2. Upon
heating to ~68°C, the crystal structure shifts from monoclinic to rutile leading to a
corresponding change in its refractive index. (c) Ellipsometry-measured real and
imaginary components of the refractive index of a 9 nm VO2 film on Au at room
temperature and after heating to 80°C. This data is plotted in terms of the real and
imaginary components of ε in Supporting Information Figure S1.

5.2 Device Concept


Here I demonstrate reversible tuning of near-IR absorption resonances from active

plasmonic metasurfaces via the insulator-to-metal phase transition of a 9 nm VO2 film

grown by ALD. The tunable metasurfaces are comprised of arrays of plasmonic nanodisks

on top of the VO2 and Au films, where a single nanocavity is shown in Figure 5a. The

parallel metallic interfaces in Figure 5a form a planar cavity between the two Au surfaces

separated by the VO2 film, which supports a gap-mode surface plasmon and operates

83
according to the same physics as the metasurfaces expounded upon in the previous

chapter.

The resonances are statically controlled by varying the nanoparticle diameter and

dynamically controlled via refractive index tuning of the VO2 layer. Upon heating to

~68°C, VO2 undergoes a phase transition which alters its crystalline structure from

monoclinic to rutile, as depicted in Figure 5b. In its monoclinic phase, VO2 is transparent

and insulating in the IR and possesses a bandgap of 0.7 eV. 105 When transitioned to its

tetragonal phase, VO2’s bandgap closes resulting in lossy metallic behavior i.e. a metallic

phase. Beyond a purely electronic transition, this insulator-to-metal transition possesses

an accompanying change in the real and imaginary components of VO2’s refractive index.

Figure 5c depicts the refractive index change for a 9 nm film grown on Au and measured

by ellipsometry at 24°C and 80°C for wavelengths spanning from 400-2400 nm. Spectrally,

the largest refractive index change of -0.17 occurs at 640 nm for the real component and

0.12 at 1660 nm for the imaginary component. In the monoclinic phase, the VO2 film

exhibits normal dispersion for wavelengths longer than 500 nm indicative of the

insulating/dielectric state. When in the tetragonal phase at 80°C, the refractive index

displays anomalous dispersion at wavelengths longer than 1000 nm indicative of metallic

behavior. As such, the refractive index change in Figure 5c distinctly shows the insulator-

to-metal transition induced upon heating. Although the real component of the

permittivity of the VO2 film is expected to go below zero around the plasma frequency,

84
and this has been demonstrated previously for VO2 films fabricated in a similar manner, 106

this behavior is not observed in the present samples and can likely be explained by the

ultrathin nature of the VO2 film and the underlying non-lattice-matched gold substrate.

5.3. Fabrication Process


Fabrication of the structure began with growing the VO2 on a Au film via ALD

according to a process that has been extensively characterized previously and been shown

to produce thin films capable of undergoing a phase change107–109 In detail, a 75 nm Au

film was deposited by electron-beam evaporation with a 2 Å/s deposition rate onto a 5 nm

Ti adhesion layer on a sapphire substrate. Next, 9 nm amorphous VO2 films were

deposited by atomic layer deposition (ALD) on the Au films in a Veeco Savannah 200

reactor at 150°C using tetrakis(ethylmethayl)amido vanadium (TEMAV) and ozone (O 3)

precursors. Due to the inert nature of the underlying substrate, ten sequential cycles of

ozone were used as an in-situ pretreatment to ensure the substrate surface was

supersaturated with hydroxyl groups essential for the nucleation of VO 2 films. For these

particular films, the pulse/purge times for TEMAV and O3 were 0.03/30 s and 0.075/30 s,

respectively. Additional details regarding the deposition of these films can be found in a

recent paper by Kozen et al.107 Due to poor lattice matching conditions between the VO2

and Au films, the ALD VO2 films were amorphous and did not exhibit fully reversible

behavior. Post-annealing the films formed a poly-crystalline VO2 film which exhibited full

reversibility and withstood repeated cycling; this was carried out in a custom ultra-high

85
vacuum (UHV) chamber at 400°C for 1 hr. Anneals were conducted under an ultra-high

pressure (UHP) O2 environment (3x10-5 torr) to assure the correct stoichiometry of the

films was maintained as they crystallized.

Figure 5.2: XPS Spectra to Characterize VO2 Film

(a) Survey XPS spectrum of the VO2 thin film on a gold substrate. (b) Higher resolution
XPS spectrum of the same sample resolving the various V 2p peaks

X-ray photoelectron spectroscopy (XPS) characterization of the annealed VO 2 films

was performed (shown in Figure 7) and revealed V+4 and V+5 peaks with full-widths at

half-maximum (FWHM) of around 1.6-1.7 eV, indicating that the film has a fairly high

degree of crystallinity as amorphous films typically have widths around 2-3 eV due to

slight variations in bonding strengths. While the V +4 peak is indicative of a VO2 film, the

V+5 peak is the result of some undesired contamination, likely in the form of a hydroxyl

group or V-O-C bonding in addition to the V2O5 stoichiometry. However, previous studies

have determined that the thickness of this contaminated layer is no more than 1 nm and

86
results from the subsequent fabrication steps being performed in atmospheric

conditions. 108

Figure 5.3: SEM Image Revealing Pinholes in the VO2 Film

SEM images of the VO2 film (shown in Figure 7) reveal the presence of pinholes

that likely have the effect of reducing the amount of switching that can be achieved.

However, such defects are common for depositions on a non-lattice-matched and

polycrystalline substrate. ALD growth of the VO2 film, annealing and XPS measurements

were performed by the research group of my collaborator Dr. Virginia Wheeler at the US

Naval Research Laboratory.

87
Figure 5.4: Fabrication Overview and Initial Characterization of Actively Tunable
Metasurfaces with Integrated VO2

(a) Fabrication process for transferring EBL-fabricated gold nanodisks onto the VO2 film.
(b) SEM images of gold nanodisks for a few of the fabricated sizes. (c) Simulated electric
(left) and magnetic (right) field profiles at 24°C and 80°C for the nanogap cavity structure
with 192 nm gold nanodisks. An incident plane wave with wavelength matching the
room-temperature resonance of the nanocavity at 1495 nm is utilized. (d) Room-
temperature reflectance spectra for metasurfaces consisting of eight different sizes of
nanodisks where additional spectra are shown in SI Figure S4.

The VO2 films were embedded into plasmonic nanocavities by transferring Au

nanodisks on top of them. The Au nanodisks were fabricated using EBL on a Si substrate.

Due to the weak adhesion of Au to the Si substrate, the Au nanodisks could be exfoliated

from the Si wafer after removing the native SiO2 and deposited onto the VO2 film using a

PDMS stamp as shown in Figure 6a. This technique was utilized due to the sensitivity of

VO2 to electron-beam exposures and subsequent development processes. In detail, a


88
PDMS stamp was placed on top of the nanodisk arrays on silicon and the sample was

baked at 100°C for 30 minutes. A potassium hydroxide (KOH) solution was used to etch

the thin silicon dioxide (SiO2) layer on the silicon substrate for two hours. This results in

the Au nanodisks debonding from the silicon substrate and attaching to the PDMS stamp,

which was then placed on a glass slide and stored in vacuum to dry overnight. Lastly,

with the Au nanodisks face-down, the PDMS stamp was pressed onto the VO2 film and

baked at 120°C for 4 minutes. The sample was placed in vacuum for two days and then

baked at 120°C for 4 minutes before removing the stamp

In all, thirteen metasurfaces were fabricated by varying the diameter of the

nanodisks (from 88 nm to 329 nm), where SEM images of representative metasurfaces are

shown in Figure 6b. FDTD simulations confirm that nanocavities possess a large effective

electric and magnetic response on resonance as shown in Figure 6c, which is strongly

localized in the VO2 film below the disks.

Figure 5.5: Room-Temperature Reflection Spectra For Additional Sizes of Nanodisks

89
5.4 Thermal Switching Properties of Nanodisk Metasurfaces with
Variable Resonance Wavelengths in the NIR

Figure 5.5: Thermal Switching Data from VO2 Metasurface and Control Al2O3
Metasurface

(a) Thermal switching spectra for metasurfaces consisting of 88, 150, 221, 256 and 329 nm
diameter Au nanodisks. Spectra for the remaining sizes are shown in SI Figure S5 (b)
Switching spectra from a control sample with a gap consisting of 9 nm of Al 2O3 for the
same size nanodisks as in (a). (c) Comparison of the amount of switching observed
experimentally for both the control sample (9 nm Al2O3 gap) and the VO2 thermal
switching sample (9 nm VO2 gap). The grey dashed line at 0 nm of switching is intended
to guide the reader’s eye.

Due to the spectrally dependent variations in the insulator-to-metal transition,

thirteen spectrally distinct metasurfaces were used to assess the dispersion of the active

tuning. The resonant wavelengths for these absorbing metasurfaces range from ~900 nm

90
to 2100 nm, and span above and below the anomalous dispersion of the tetragonal VO 2

phase. The room-temperature reflection spectra of the metasurfaces were measured with

a white light source focused through a 0.4 NA objective as shown in Figure 6d (additional

spectra in Figure 8). All the metasurfaces demonstrate near-perfect absorption of incident

light on resonance, with typical peak absorption values of ~95%. The switching spectra

for the 88, 150, 221, 256 and 329 nm nanodisk metasurfaces are shown in Figure 7a, while

switching spectra for the remaining sizes are shown in SI Figure 9. In addition to

producing a spectral shift in the resonance, heating the metasurfaces was also observed

to modulate the peak absorbance by as much as 16%, contributing further to the dynamic

nature of this structure.

91
Figure 5.6: Thermal Switching Spectra for Additional Sizes of Nanodisks

(a,b) The solid lines correspond to the reflectance spectra at 20° C, while the dashed lines

correspond to the reflectance spectra of the same metasurfaces at 80° C.

A control sample was also fabricated with a gap comprised of a non-phase-change

material (9 nm of Al2O3). The thermal switching spectra for select sizes of nanodisks on

this sample (same sizes shown in Figure 7a) are presented in Figure 7b. An average of just

3 nm of wavelength switching was observed for the control sample. A comparison of the

92
amount of switching for each nanodisk size for both the sample with the VO2 gap and the

sample with the Al2O3 gap is shown in Figure 7c. We note that the full-width at half-

maximum (FWHM) of the resonance on the control sample is significantly less for the

control sample. This is likely due to increased surface roughness for the VO 2 film arising

from the annealing process and resulted in larger inhomogeneous broadening of the

plasmon resonance.

5.5 Simulated Metasurface Thermal Switching

Figure 5.7: Simulated Switching Spectra

To provide additional insight into the spectral dependence of the switching

behavior, the active metasurfaces were simulated with a finite-difference solver to model

the temperature-dependent reflection spectra for the various metasurfaces. The simulated

spectra for the same sizes of nanodisks as in Figure 7a are shown in Figure 10. Both the

simulations and experimental results show that spectral switching can be maximized for a

given geometry and resonant wavelength. Furthermore, simulations show good agreement
93
with experiment in terms of the magnitude of thermal switching that is predicted, although,

for some disk sizes, simulations predict a different direction for the switching than what is

observed experimentally.

5.6 In-Depth Characterization of 192 nm Nanodisk Metasurface


Thermal Switching

Figure 5.8: Further Characterization of Thermal Switching For 192 nm Nanodisk


Metasurface

(a) Thermal switching spectra over five cycles, which shows fully reversible

switching. (b) Magnitude of switching over twelve cycles of the temperature (c) Switching

94
spectra over one heating cycle taken in 5°C increments from 20°C to 85°C. (d) Extracted

resonance wavelength as a function of temperature for the spectra displayed in (c).

For further characterization of the VO2 switching, the amount of switching for the

192 nm nanodisks metasurface is analyzed over multiple cycles and smaller temperature

steps. The temperature-dependent reflectance spectra of the 192 nm nanodisk metasurface

over five cycles is shown in Figure 11a. In all, reflection spectra for this metasurface were

collected over twelve temperature cycles; the shift in peak wavelength measured during

each cycle is plotted in Figure 11b instead of the full spectra for clarity. For each cycle the

magnitude of shift was observed to be 36 ± 2 nm, demonstrating the repeatable nature of

the thermal switching.

The reflection spectra of the 192 nm nanodisk metasurface was also measured at

5°C increments to characterize the continuity of the switching in more detail, and the

results are shown in Figure 11c. The extracted resonance wavelength is shown as a

function of temperature in Figure 11d. The resonance wavelength of the 192 nm

metasurface displays little change until the temperature approaches 45°C. The resonance

then begins to blue-shift with increasing temperature, reaching a maximum rate of change

at 60°C. This implies a transition temperature of ~60°C for the fabricated VO 2 gap layer,

which is slightly less than the reported value in literature for bulk VO 2.93 Previous

investigations of ALD-grown VO2 have similarly shown a lowered transition temperature

of 61°C.108

95
Above 80°C in our experiments, the resonance wavelength stabilizes at a final

value of 1300 nm, which is blue-shifted by 40 nm compared to the resonance wavelength

at room temperature. The lowered transition temperature and gradual phase shift

occurring over a 40°C range when compared to bulk VO 2 is characteristic of the effect of

boundary conditions on the VO2 thin films. The amount of switching observed here is less

than what has been measured previously with sputtered/PLD VO 2 films or bulk material

and likely arises from a combination of the effects of interface strain on the ultrathin film

and the presence of only a small amount of the phase-change material. Further

optimization of the VO2 film deposition process on an Au substrate, such as reducing the

roughness and Au dewetting that occur during annealing as well as lessening

contamination in the form of V+5, is expected to lead to a larger tuning range.

5.7 Conclusion
In summary, I have here presented thermally tunable nanogap absorbers, which

rely upon a robust, solid-state switching mechanism and possess a range of resonant

wavelengths spanning 1200 nm in the near-IR. These metasurfaces display up to 60 nm of

switching that is fully reversible over more than 10 temperature cycles. With the current

amount of switching, we anticipate that this structure could be utilized for controllable

enhancement of processes such as fluorescence, single photon emission, and nonlinear

generation, as these effects depend strongly on the overlap of the resonance with the

wavelength of the process of interest.7,110 An additional application that should be

96
realizable with the current amount of switching is holography since the phase of the

reflection coefficient varies rapidly around the resonance wavelength. Thus, a modest

shift in the resonance wavelength should lead to a large shift in phase. With

improvements to the VO2 thin film deposition, such that the magnitude of switching is

increased, a variety of additional deeply subwavelength, all-solid-state, optoelectronic

devices could be created using this platform including optical switches and memories,

tunable photodetectors, and reconfigurable displays.

97
6. Plasmonic Nanogap Metasurface with Integrated
Silicon Vacancies in Diamond
6.1 Introduction
Solid-state quantum emitters such as quantum dots, defects in 2D materials and

color centers in diamond are crucial for the advancement of quantum information and

sensing technologies. These emitters are highly valued for their ability to produce

indistinguishable single photons on demand, 63,111,112 create interactions between single

photons113 and, furthermore, store and process quantum information. 114–116 Aside from

these applications, solid-state quantum emitters can also be used to sense

temperature,117,118 strain119 and electromagnetic fields118,120,121 at the nanoscale. Among the

various types of quantum emitters, color centers in diamond, such as nitrogen (NV -) and

silicon vacancies (SiV-), are particularly promising due to their excellent photostability

and minimal spectral diffusion. 63,122

To fully realize the potential of diamond color centers for quantum information

and sensing applications, there is a need to precisely tailor the interaction of emitters

with incident electromagnetic fields. In particular, engineering the radiative

spontaneous emission rate is key for extracting sufficient photon count rates from

ordinarily dim vacancy centers since long intrinsic lifetimes on the order of nanoseconds

act as a fundamental limit. 123 A promising solution for modifying their spontaneous

emission rate is to utilize the Purcell effect by embedding the color centers in micro- or

nanocavities. 124–127 However, previous demonstrations have been limited to 10-fold


98
enhancement or less. A recent demonstration showed 80-fold enhancement of NV-

centers contained in nanodiamond particles by integrating into plasmonic

nanocavities. 128 However, usable, coherent emission at the zero-phonon line (ZPL)

comprises just 5% of the total NV- fluorescence, 40 compared to 80% for the SiV-.63

Furthermore, atomic force microscope (AFM) probe manipulation of the nanodiamond

particles was utilized to assemble the nanocavity structure, an approach which is

difficult to scale to chip manufacturing.

6.2 Device Concept


Here I demonstrate up to 121-fold enhancement of the emission rate of SiVs

coupled to plasmonic nanocavities. The SiVs are implanted into monolithic diamond thin

films, which are reduced to nanometer-scale thickness by reactive ion etching (RIE), and

transferred onto a gold film, an approach that has a clear path to fabrication over wafer-

scale areas. Arrays of gold nanodisks are stamped onto the diamond, forming the

nanocavities. The sample architecture is shown in Figures 5a and 5b. Initial wedge in the

diamond results in a thin etched slab with a gradient of 2 nm/µm and a minimum thickness

of ~16 nm, where further fabrication details are included in Appendix A.

99
Figure 6.1: Nanogap-Coupled Silicon Vacancy Metasurface Concept

(a) Schematic of sample structure consisting of Au nanodisks (height = 30 nm) separated


from an Au ground plane by etched diamond slab. (b) Height profile of the thinnest
edge of the diamond slab measured using an imaging ellipsometer. The minimum
thickness was measured to be ~16 nm (c) Electric field profile of the structure. Simulated
via finite-difference, time-domain methods using an incident plane wave of 738 nm
light, which was selected to match the SiV ZPL emission wavelength (further details in
the Methods section). (d) Absorption spectrum of the metasurface overlaid with a PL
spectrum of embedded SiVs, showing the strong degree of overlap.

A height profile of the thinnest section of the diamond slab is shown in Figure 5c

and was measured using an imaging ellipsometer (Accurion nanofilm_ep4). In this

measurement, the gradient of the diamond film was determined by taking measurements at
100
10 wavelengths for each field of view and then stitching them together across the length of

the slab. The material model used for fitting consisted of a diamond layer, native silicon

oxide, and a silicon substrate.

The plasmonic nanocavities possess a large effective electric response on

resonance as shown in Figure 5d, which is strongly localized in the etched diamond slab

below the disks. The resonance wavelength was chosen to match the SiV ZPL wavelength

of 738 nm. The metasurface absorption spectrum and a PL spectrum of embedded SiVs

are plotted in Figure 5e, illustrating their strong overlap. Due to the Purcell effect, this

overlap should result in an enhancement in the emission rate of the embedded SiVs. 6 A

Purcell factor of 207 is predicted for an optimally positioned emitter near the outside edge

of the nanoparticle.

Finite-difference time-domain (FDTD) methods were performed to simulate the

response of the nanocavity with an integrated SiV to incident optical waves using

Lumerical. By simulating the scattering from single disks, an estimated disk size required

to achieve a nanocavity on resonance with the SiV emission was established. The diamond

thickness employed in the simulations was fixed based on imaging ellipsometry

measurements performed on the thinnest section of the diamond slab. The expected

Purcell factor for SiV coupled to the nanocavity was calculated by introducing a radiating

dipole oriented at 45° from the horizontal into the diamond-filled gap. Purcell factors for

different SiV positions were obtained by varying the position of the emitter in one

dimension within the nanocavity. The largest Purcell factor resulted from an emitter
101
positioned near the outside edge of the nanoparticle, which is consistent with previous

reports. 8,33

6.3 Fabrication Process

Figure 6.2: Fabrication Process for Nanocavity Metasurfaces with Integrated SiVs

(a) Fabrication process for creating nanogap metasurfaces with embedded SiVs. (b)
Optical microscope image of the etched diamond slab after transferring onto a gold film.
(c) SEM image of gold nanodisks after EBL fabrication on silicon.

An illustration of the sample fabrication process is shown in Figure 6a. Fabrication

of the nanogap cavity structure begins with commercial implantation of silicon ions at ~10

nm depth (Innovion, San Jose, CA) and subsequent cleaning and high-temperature, high-

vacuum annealing to form SiV- centers from the implanted ions (Figure 2 step: i). Next,

the diamond slab is inverted and transferred to a sapphire wafer (ii) for deep RIE using

102
alternating O2 and ArCl2 chemistries. Etching is terminated once a portion of the diamond

slab has been completely etched through (iii) and the slab is then debonded from the

sapphire wafer and rinsed onto a gold substrate (iv); an optical microscope image of the

sample at this point in the fabrication is shown in Figure 6b. Lastly, arrays of gold

nanodisks are fabricated by EBL and stamped onto the diamond slab (v, vi), where

additional fabrication details are provided in the Appendix A. This transfer technique was

utilized due to the delicate nature of the etched diamond slab and its relatively weak

adhesion to the gold substrate, which could be damaged during the development and

liftoff steps of EBL. A scanning electron microscope (SEM) image of the nanodisks prior

to transfer is shown in Figure 6c.

103
6.4 Emission Lifetime Properties of Metasurfaces Containing
Ensembles of SiVs

Figure 6.3: Cavity-Enhanced SiV Lifetime and PL Data for Metasurface Sample

(a) Time-resolved PL from SiVs coupled to an ensemble of gold nanodisks (red),


showing a biexponential decay with a fast component of 𝝉𝒇𝒂𝒔𝒕 = 15 ps and a slow
component of 𝝉𝒔𝒍𝒐𝒘 = 475 ps. The fast component decays nearly at the speed of the IRF
of the avalanche photodiode detector, also shown (grey). The lifetime of SiVs in an
etched diamond slab on silicon is 𝝉𝒔𝒊𝒍𝒊𝒄𝒐𝒏 = 965 ps (blue). (b,c) Histogram of the
extracted fast decay components (b) and corresponding Purcell factor (c) collected over
16 different spots on the sample. (d) PL spectra from SiVs coupled to an ensemble of
gold nanodisks (red), SiVs in an etched diamond slab on silicon (blue) and on gold
(yellow). In all cases a 705 nm CW laser was used for excitation.

Time-resolved fluorescence measurements were then performed to characterize

the SiVs embedded in nanogap cavities, as well as SiVs in a control sample consisting of

an identically prepared etched diamond slab on a silicon substrate. Optical measurements

104
were performed using a home-built microscope setup coupled into either a spectrometer

or timing SPAD. A 0.9 NA 100 × DF/BF objective was used for both excitation and

collection. Cavity resonances were identified by measuring the scattering or reflectance of

the surface under unpolarized white light illumination. In this way, areas where the

resonance overlapped with the SiV emission were identified and further characterization

was performed. The fluorescence of SiV were measured under laser excitation by a 514

nm Ar-ion laser or a 705 nm laser diode. The excitation wavelengths were rejected by a

combination of dichroic mirrors and long-pass interference filters. The filtered

fluorescence was then coupled into a Jobin Yvon spectrometer with a Symphony charge-

coupled device (CCD) camera.

Fluorescence lifetimes were similarly measured under ~150 fs pulsed excitation

with a 514 nm Ti:Sapphire laser operating at 80 MHz. The fluorescence was filtered and

then coupled into a Micro Photon Devices (MPD) SPAD. Detection events from the SPAD

were correlated with excitation pulses via a fast timing diode and a Pico-Harp 300 time-

correlated single-photon counting module with 4 ps time bins. The instrument response

function was measured by directly coupling the excitation light into the SPAD, enabling

the recovery of fast lifetimes using iterative deconvolution techniques. TCSPC data were

fit with a biexponential convolved with the measured IRF through a least squares

approach.129 In general, deconvolution approaches can allow lifetimes substantially

shorter than the FWHM of the IRF to be recovered.130–132 In this case, however, the the limit

105
imposed by the digitization of timing pulses with 4 ps time bins is relevant and therefore

restricted gains from deconvolution to recovery of no less than an 8 ps lifetime as a

consequence of the Shannon-Nyquist theorem.

The SiVs on silicon exhibit single-exponential decay with a lifetime of 𝜏𝑠𝑖𝑙𝑖𝑐𝑜𝑛 =

944 ± 93 ps, which is consistent with previously reported values for the SiV intrinsic

lifetime; 63,133,134 an example decay curve is shown in Figure 7a. The measured intrinsic

lifetime for the SiV is on the lower end of the reported range, likely because of surface

effects in the nanometer-scale diamond film that have the effect of reducing the SiV

lifetime. 62,135

SiVs embedded in the nanogap cavities, on the other hand, demonstrated

biexponential decay with a fast component of 𝜏𝑓𝑎𝑠𝑡 = 13 ± 3 ps and a slow component of

𝜏𝑠𝑙𝑜𝑤 = 580 ± 120 ps and weights of 65% and 35%, respectively. This indicates that the

majority of emitters decayed according to the fast component of the biexponential decay.

An example fluorescence decay curve is plotted in Figure 7a; also plotted is the instrument

response function (IRF) of the measurement system, which was measured at the excitation

wavelength using the femtosecond laser pulses from the excitation source and was

utilized during the deconvolution process prior to fitting the enhanced lifetime curves.

In total, time-resolved fluorescence was measured on sixteen different spots on the

sample, all in the thinnest region of the diamond slab. A biexponential fit was applied to

each of the collected fluorescence curves after deconvolution from the IRF and the

106
distribution of extracted fast components for the fits are plotted in Figure 7b. The

distribution of calculated Purcell factors is shown in Figure 7c and is computed by

dividing the average lifetime of the SiVs in the etched diamond slab on silicon by each of

the 16 measured fast components of SiVs in the nanocavities. Across all 16 measured spots

on the sample, the average Purcell factor was 76 ± 14.

I also characterized the brightness of SiVs coupled to the nanocavities. I collected

PL using a 705 nm CW excitation laser from SiVs coupled to the nanocavities, SiVs on

silicon and SiVs on gold. Example PL spectra from each of the samples are plotted in

Figure 7d. On average, fluorescence from the SiVs coupled to nanocavities was enhanced

13-fold relative to the SiVs on gold and 4-fold relative to the SiVs on silicon. I expect that

the SiV PL enhancement is lower than the emission rate enhancement because the

wavelength of the excitation laser was significantly detuned from the cavities’ resonance

wavelength and thus the excitation rate was not enhanced greatly. 33 Exciting the SiV closer

to resonance should result in larger PL enhancement; however, this was not attempted

due to lack of availability of a narrowband source in this wavelength range that could be

filtered out before collecting the fluorescence signal.

6.6 Conclusion
In this chapters, I integrated ensembles of SiVs in thin-film diamond with

plasmonic nanopatch antennas and observed up to 121-fold enhancement of their

spontaneous emission rate. This is the highest value that I am aware of for the SiV in

107
diamond. These results for ensembles of SiV indicate the high potential of this approach.

Producing ultra-short lifetime emission from single SiVs will only require that the Si ion-

implantation density be reduced, allowing for optically resolvable single emitters to be

measured. In the longer term, this platform offers the potential for creating nodes in a

quantum network with high photon count rates, as well as sensors for temperature, strain,

and electromagnetic fields at the nanoscale with high detectivity.

108
6.7 Emission Properties of Ensembles of SiVs Coupled to Single
Nanocavities

Figure 6.4: Cavity-Enhanced SiV Lifetime for Single Cavity Sample

Time-resolved PL from SiVs coupled to a single gold nanodisk (red), showing a


biexponential decay with a fast component of 𝜏𝑓𝑎𝑠𝑡 = 8 ps and a slow component of
𝜏𝑠𝑙𝑜𝑤 = 480 ps. The fast component is limited by 4 ps bin size of the TCSPC system and is
extracted via iterative deconvolution with the IRF of the system (grey). The lifetime of
SiVs in an etched diamond slab on silicon is 𝜏𝑠𝑖𝑙𝑖𝑐𝑜𝑛 = 965 ps (blue). Inset: white light
scattering spectra of the gold nanodisk coupled to SiVs. The cavity resonance is nearly
perfectly aligned with the SiV ZPL.

We also fabricated a sample consisting of gold nanodisks with a lower fill fraction

(4 µm pitch) in order to explore the effect of coupling single nanodisks to SiVs. The same

size gold nanodisks were fabricated (94 nm) and transferred onto an etched diamond slab

that was prepared in exactly the same manner as the one utilized when SiVs coupled to

ensembles of nanodisks were studied above. Unfortunately, due to poor adhesion of the
109
gold nanodisks to the diamond on this sample, only a few nanodisks were located on the

thinnest region of the etched diamond slab. One of these nanocavities was found with a

resonance wavelength matching the SiV ZPL; its white-light scattering spectrum is shown

in the inset of Figure 4.

SiVs coupled to this nanocavity were found to have more strongly enhanced

emission rates than any of the ensembles of SiVs coupled to arrays of nanodisks that were

studied, likely because the fastest SiV lifetimes were not averaged out, as in the ensemble

of nanodisks case. Time-resolved PL measurements revealed an instrument-limited

biexponential decay with a fast component of 𝜏𝑓𝑎𝑠𝑡 = 8 ps and a slow component of

𝜏𝑠𝑙𝑜𝑤 = 480 ps with weights of 84% and 16% respectively. This fast lifetime, which we

recover through the iterative deconvolution technique, is at the limit imposed by the 4 ps

time bins of the TCSPC system, in accordance with the Shannon-Nyquist theorem. 130,136

110
7. Frequency Conversion in Diamond Thin Films via
Nanogap-Cavity-Enhanced Nonlinear Optics
7.1 Introduction
The unique properties of diamond such as excellent thermal conductivity, a high

index of refraction, a large transparency window, and negligible birefringence 54,137 have

made this material a promising platform for nanophotonics both in the classical138–141 and

quantum regimes128,142–147. Additionally, with its relatively high third-order nonlinear

susceptibility (χ(3)), diamond is an appealing candidate for integrated nonlinear devices.

To date, Raman lasing, 148,149 supercontinuum generation,150 and frequency combs151 have

been demonstrated in integrated, single-crystal diamond platforms. One of the salutary

features of diamond is its ability to host defects in its crystal lattice known as color-centers,

which are promising candidates for nodes in networked quantum systems. These

optically addressable spin qubits are unexplored in terms of their interactions with

nonlinear processes such as frequency conversion to shift their emission wavelength. 152

However, this task has remained an outstanding challenge due to the weak intrinsic

response of nonlinear processes and the phase-matching requirements. Plasmonic

structures have proven to be a well suited platform to investigate and enhance nonlinear

optical processes. 110,153–156 The deeply subwavelength scale of these devices simultaneously

allows for large confinement and enhancement of electric fields, 8,157–160 as well as a

relaxation of phase matching conditions. 161–163

111
In this chapter, I utilize film-coupled, plasmonic nanogap cavities created by a

non-disruptive transfer method to enhance the nonlinear response in nanoscale diamond

films without patterning diamond itself. When the excitation wavelength overlaps with

the cavity resonance, both THG and SHG are dramatically enhanced compared to a thin

diamond slab reference. Furthermore, THG, SFG and FWM were enhanced

simultaneously by leveraging two different cavity modes, further highlighting the

versatility of this platform as a frequency converter.

7.2 Sample Fabrication

Figure 7.1: Diamond Frequency Conversion Sample Overview

112
Optical image of the sample. Area A: transferred nanoparticles on ~ 12 nm diamond,

forming nanogap cavities; area B: transferred nanoparticles on ~ 200 nm diamond,

forming decoupled nanoparticles; area C: diamond on gold. (b) AFM image of the area in

(a) delineated by the square. (c) Height profile along the red, dashed line in (b). (d)

Schematic of the sample structure: a thin diamond wedge is sandwiched in between a 75

nm evaporated gold film and EBL-fabricated gold nanoparticles with a height of 30 nm.

(e) Simulated electric field enhancement distribution of nanogap cavities at 1455 nm. (f)

Measured reflection spectra from nanogap cavities (red) and decoupled nanoparticles

(blue)

The diamond slab is sandwiched in between a 75 nm gold film and gold

nanoparticles, forming plasmonic nanogap cavities. Arrays of nanoparticles were

fabricated on a silicon substrate by electron beam lithography (EBL) and then transferred

onto the diamond slab using a polydimethylsiloxane (PDMS) stamp (see Appendix B for

fabrication details). As shown in Figure 6.1a, there are three key areas on the sample: (A)

transferred nanoparticles on a ~12-nm-thick diamond film, forming nanogap cavities; (B)

transferred nanoparticles on a ~200 nm diamond film, which behave as decoupled

nanoparticles due to the increased diamond thickness (C) diamond (~12 nm thick) on a

gold substrate, serving as a reference. An additional reference consists of a bare diamond

slab on PDMS with a similar thickness and gradient as the diamond slab on gold. The

thicknesses of the different areas of the diamond slab are confirmed by atomic force

113
microscopy (AFM) as shown in Figure 6.1b. The height profile in Figure 1c demonstrates

that the thinnest section of the diamond is 12 nm thick. A schematic of the sample is shown

in Figure 1d and consists of arrays of nanoparticles (220 nm particle side length) with a

pitch of 440 nm on a wedge-shaped diamond thin film (gradient of ~2.6 nm/μm).

The simulated field distribution for a nanocavity with a 12 nm diamond gap

illustrates that the highly confined electric fields in the cavity are enhanced up to 40-fold

in comparison with the original incident field, facilitating enhanced nonlinear generation

(Figure 6.1e). A reflection spectrum is measured from 700 nm to 1600 nm to determine the

cavities’ resonance wavelengths. For the nanogap cavities, the fundamental resonance

mode is at 1455 nm and the second-order mode is at 840 nm (Figure 6.1f). The fundamental

mode is blue-shifted to 1130 nm for the decoupled nanoparticles and no clear second-

order mode is observed. We attribute the fundamental mode from the decoupled

nanoparticles to a localized surface plasmon resonance (LSPR) mode from the

nanoparticle itself. The nanogap cavity mode is not present on this area of the sample

because the ~200 nm separation between the gold nanoparticles and gold film is too large

to support gap plasmons.

114
7.3 Nanogap Cavity Enhanced THG and SHG

Figure 7.2: THG and SHG from Diamond Nanogap Cavity Sample

(a-b) THG intensity as a function of excitation power shown on a log scale. Data points

are experimental results and lines are cubic polynomial fits. (a) 1455 nm excitation. (b)

1130 nm excitation. (c-d) SHG intensity as a function of excitation power shown on a log

scale. Data points are experimental results and lines are quadratic polynomial fits. (c) 1455

nm excitation. (d) 1130 nm excitation. Nanogap cavities: transferred nanoparticles on ~12

nm diamond with an underlying gold film; decoupled nanoparticles: transferred

nanoparticles on ~200 nm diamond with an underlying gold film.

115
To leverage the field enhancement in the nanogap cavities for nonlinear

generation, a pump wavelength of 1455 nm, matching the fundamental cavity resonance,

is used to excite THG. The presence of THG is confirmed via observation of a third-order

power law during power dependence measurements on all three regions of the sample

and on the control as shown in Figure 6.2a. The THG response from the nanogap cavities

is enhanced 1.6×107-fold compared to bare diamond on PDMS. The enhancement factor is

defined as the nonlinear intensity from the plasmonic structures divided by the intensity

from bare diamond on PDMS under the same excitation conditions. Note that different

excitation powers are used to generate the nonlinear response in order to prevent damage

to the nanocavities due to heating from the localized electric field. In light of this, the

nonlinear intensities are extrapolated to the same excitation power using the confirmed

power law that was also used to derive the enhancement factor.

Next, to investigate if the large enhancement is arising from the nanogap mode or

simply the LSPR of the gold nanoparticles, THG is also measured from the decoupled

nanoparticles. Measurements are performed using both 1455 nm excitation, which

overlaps with the fundamental resonance of the nanogap mode (Figure 6.2a), and 1130

nm excitation, which overlaps with the LSPR mode of the decoupled nanoparticles (Figure

6.2b). For the decoupled nanoparticles, nearly three orders of magnitude less

enhancement is observed for both excitation wavelengths, even though a larger amount

of nonlinear medium (i.e., a thicker diamond layer) contributes to the THG response. The

116
electric field intensity is much less in the diamond for the decoupled nanoparticles,

resulting in less THG enhancement.

As seen in Figures 6.2a and 6.2b, the nonlinear response from the diamond on gold

is also enhanced compared with the diamond on PDMS. This may be explained by a

contribution from the nonlinearity of gold or surface effects, as the nonlinear response

from PDMS itself is negligible. In addition to the large enhancement, the nanogap cavities

have relatively high nonlinear conversion efficiencies considering their nanoscale

dimensions. The efficiency is defined as the power of the generated nonlinear intensity

from embedded diamond nanocavities divided by the power of the incident excitation.

The power of the nonlinear signal is derived from the photon counts and a calibrated light

source (Labsphere) positioned at the focal plane of the objective lens; photon counts are

then collected by the CCD camera-coupled spectrometer (same as used for the nonlinear

response measurements). From this, a linear relation between radiative power and

detected photon counts are obtained. The nonlinear generation power is converted from

the nonlinear response counts using this relation and the efficiency is determined as the

nonlinear generation power divided by the excitation power. With an observed damage

threshold of 5 mW of excitation power, the maximum THG conversion efficiency is

estimated to be 2.33×10-5 %.

117
7.4 Excitation Wavelength Dependence of THG and SHG

Figure 7.3: Excitation Wavelength Dependence of THG and SHG

(a) THG response spectra as a function of excitation wavelength. (b) SHG response spectra

as a function of excitation wavelength. (c) THG enhancement factor as a function of

excitation wavelength (red) compared to the reflection spectrum (blue) which shows the

plasmon resonance. The FWHM of the THG enhancement is 31 nm. (d) SHG enhancement

factor as a function of excitation wavelength (red) compared to the reflection spectrum

(blue) which shows the plasmon resonance. The FWHM of the SHG enhancement is 40

nm, while the FWHM of the reflection spectrum is 204 nm in both (c) and (d).

118
Next, the excitation wavelength was varied to further investigate the importance

of the plasmonic nanogap mode for enhanced nonlinear responses. Specifically, laser

excitation at wavelengths ranging from 1430 nm to 1480 nm with constant power are

utilized for nanogap cavities and the resulting spectra are shown in Figure 6.3a for THG

and Figure 6.3b for SHG (additional data in Appendix D. As expected, the maximum

nonlinear intensity occurs at the cavity resonance wavelength (1455 nm) for both the THG

and SHG signals. The nonlinear intensity decreases dramatically when the excitation is

detuned from 1455 nm due to the reduced electric field intensity within the diamond

layer. Therefore, the nonlinear intensity is strongly related to the enhanced electric field

intensity confined within the diamond, illustrating the observed overall nonlinear

response significantly originates from diamond. The full-widths at half-maximum

(FWHM) of the THG and SHG enhancement as a function of excitation wavelength are

noted to be much smaller than that of the cavity reflection spectrum as shown in Figure

6.3c and 6.3d, which is a consequence of the nonlinear dependence of THG and SHG

intensity on the electric field intensity.

It is observed, that THG decreases more rapidly than SHG when the excitation

wavelength is detuned from the cavity resonance as the THG depends on the third power

of the electric field in the cavity whereas SHG has a second-order power dependence.

THG is expected to have a 3rd order Lorentzian lineshape and SHG a 2nd order Lorentzian

lineshape. A coupled mode theory analysis gives a ratio between the FWHM of the THG

119
3
2 −1
and SHG enhancement of = 0.79 which agrees very well with the experimentally
2 −1

31
observed ratio of = 0.78 (further details in Appendix C).
40

7.5 Simultaneous Generations of Multiple Nonlinear Frequencies

Figure 7.4: Generation of Multiple, Simultaneous Nonlinear Wavelengths

(a) The spectra of nonlinear responses as a function of the time delay between the two

excitation pulses at 840 nm and 1455 nm. (b) Nonlinear response spectra at zero time

120
delay. The three peaks correspond to THG, SFG and FWM as indicated in the figure. (c-

d) Power dependence measurements for variable 840 nm excitation power. (c) FWM

intensity. The 1455 nm excitation powers for nanogap cavities and the diamond reference

are 180 μW and 18 mW, respectively. (d) SFG intensity. The 1455 nm excitation power for

nanogap cavities is 180 μW and SFG intensity from the diamond reference is too small to

observe. (e-f) Power dependence measurements for variable 1455 nm excitation power.

(e) FWM intensity. The 840 nm excitation powers for nanogap cavities and the diamond

reference are 400 μW and 14 mW, respectively. (f) SFG intensity. The 840 nm excitation

powers for nanogap cavities is 400 μW and the SFG intensity from diamond reference is

too small to observe. The insets in (c-f) show the energy diagram for the corresponding

frequency conversion process.

Next, we investigate nonlinear generation arising from multiple nonlinear optical

processes occurring simultaneously within a single nanocavity. Here, sum frequency

generation (SFG) and four wave mixing (FWM) are selected to demonstrate this

capability, along with the previously characterized THG process. Excitation wavelengths

of 840 nm ( 1 ) and 1455 nm (  2 ), matching the cavities’ fundamental and second order

modes, were selected. The output SFG frequency equals SFG = 1 + 2 while the

degenerate FWM frequency equals  FWM = 21 − 2 (indicated in the inset schematic in

Figure 4c-f).

A variable delay stage is employed to control the relative time delay between the

121
two excitation pulses. The SFG (532 nm) and FWM (590 nm) responses start to appear

when the time delay between the two excitation pulses is below approximately 150 fs,

which is the pulse duration of the excitation laser. The THG intensity remains constant as

it only depends on the 1455 nm excitation (Figure 4a). The largest SFG and FWM

intensities are observed at zero time delay when the two excitation pulses are perfectly

overlapped (Figure 4b). The emission peak observed at 485 nm is from THG as confirmed

above. Similar power dependence measurements are performed to demonstrate that the

other two peaks are from SFG and FWM processes, respectively. Specifically, the signal at

the longest wavelength (590 nm) scales quadratically with the 840 nm excitation power

and scales linearly with the 1455 nm excitation power. Therefore, it is confirmed that this

signal is from FWM as it obeys the expected power law and, furthermore, occurs at the

wavelength calculated from the frequency conversion relation. Similarly, the signal in the

middle at 532 nm scales linearly with both the 840 nm the 1455 nm excitation powers,

indicating that it is indeed from SFG.

Large enhancement is expected for SFG and FWM from diamond embedded in the

nanogap cavities due to the strong electric field at both excitation wavelengths.

Experimentally we find that FWM is enhanced 3.0×10 5-fold compared to bare diamond on

PDMS as extracted from Figures 4c and e. For SFG, the signal from bare diamond is too

weak to detect since only the surface of diamond contributes SFG due to the inversion

symmetry in its crystal lattice. Thus, an enhancement factor for this process cannot be

122
determined.

7.6 Conclusion
In summary, I embedded a diamond slab into plasmonic nanogap cavities formed

by a gold ground plane and EBL-fabricated nanoparticles transferred using a PDMS

stamp. By overlapping the excitation wavelength with the nanocavities’ resonance

wavelength, 1.6×107-fold enhancement is observed for both THG and SHG. This large

enhancement is accompanied by relatively high nonlinear conversion efficiencies,

approaching 2.33×10-5 % and 7.59×10-6 % for THG and SHG respectively. These efficiencies

are comparable with other reported results using plasmonic structures, 110,153,156,163–165 but

achieved with a unique material – diamond, which has not been utilized in previous

reports. Furthermore, simultaneous enhancement of multiple nonlinear processes was

demonstrated, specifically THG, SFG and FWM, enabled by the relaxed phase matching

conditions in the deeply subwavelength cavities. To further enhance nonlinear

generation, the damage threshold of the structure could be increased by utilizing

refractory materials166–168 such as TiN or gold nanoparticles coated with ultrathin ALD

layer to prevent deformation. 169 The PDMS transfer process provides a convenient and

non-disruptive method to place nanoparticles. This technique, along with the cavity’s

vertically oriented gap, makes it possible to embed diamond containing color centers, as

demonstrated in the previous chapter. These studies demonstrate a metasurface-based

diamond frequency converter that is promising for on-chip nonlinear devices and single-

123
photon frequency conversion of the emission of color-centers in diamond from visible to

telecommunication wavelengths.

124
8. Conclusion
This dissertation has presented, for the first time, the integration of silicon vacancy

centers in diamond, a prominent candidate emitter for quantum networks, into film-

coupled plasmonic nanocavities, which have demonstrated singular capabilities for

radically altering the properties of embedded spontaneous emitters to achieve desired

effects in terms of decreasing the emission lifetime, increasing the overall photon count

rate of the fluorescence and directing the emission into a single out-of-plane radiation

lobe.

To support integration of SiVs into film-coupled plasmonic nanocavities, a process

for EBL fabrication of these cavities was first developed, resulting in near-perfect

absorption from metasurfaces comprised of either nanodisks or nanocubes, as expounded

upon in Chapter 4. During this characterization, a study of the dependence of absorber

properties on the metasurface fill-fraction was also conducted, showing an increase in

performance with increasing fill-fraction, as well as the emergence of some neighboring

element coupling effects at the highest fill-fraction. Furthermore, an investigation was also

conducted on the properties of rectangular absorbers with dual-band absorption that

depends on the incident polarization. Strong absorption into both the TE and TM cavity

modes was observed when the incident polarization was appropriately tuned.

Simulations point towards the ability of nanogap cavities employing rectangular particles

to simultaneously induce large enhancement of both the absorption rate and spontaneous

125
emission rate of embedded fluorophores, leading to total fluorescence enhancement

factors of a few thousand due to synergy from these two effects. As such, this sort of

plasmonic nanocavity provides a promising pathway for creating fluorescent sources

with photon count rates approaching the THz regime.

The EBL-fabricated metasurfaces were then employed to create film-coupled

nanocavities with embedded SiVs; here a transfer process assisted by a PDMS stamp was

created to avoid damaging the diamond thin film / SiVs during direct-write EBL. The

resulting experiments, presented in Chapter 5, have shown ~100x increase in the radiative

emission rate of SiVs in diamond, which is an order of magnitude higher than previous

enhancements achieved with other nanophotonic structures. The fabrication technique

employed utilizes monolithic, etched diamond thin films and has a clearer path towards

device scaling than less controllable techniques, such as the utilization of diamond

nanoparticles where arrangement presents a significant challenge.

The results presented lay a foundation for achieving similarly dramatic

enhancements of emission from single SiVs, as this would only require adjusting the

implantation concentration. Once single vacancy centers have been integrated into

plasmonic nanocavities in this fashion, a whole host of experiments in quantum photonics

become accessible, including optical manipulation of the spin state of embedded vacancy

centers and observation of Rabi oscillations from single vacancy centers that are in the

strong-coupling regime of emitter-cavity coupling. Integration of the neutral-charge-state

126
SiV could also be potentially achieved by employing existing techniques for manipulation

of the SiV’s charge state.

Utilizing the same architecture, enhancement of several different nonlinear optical

processes in diamond, including THG, SHG, SFG and FWM, were performed, yielding

enhancement factors ranging from 105 to 107 as described in Chapter 7. The demonstrated

nonlinear enhancements offer a promising pathway for on-chip frequency conversion of

SiV emission to e.g., telecom wavelengths for larger scale photonic integration.

Alternatively, diamond is a material with superior mechanical hardness, a large

transparency window and excellent thermal conductivity and, thus, is appealing in many

ways as a photonic platform in a general sense.

The final main result presented in this dissertation is the actively tunable nanogap

metasurfaces with integrated sub-10-nm VO2 films, which was presented in Chapter 5.

Here, EBL-fabricated nanodisks were again utilized and were transferred onto the VO 2

thin film on Au with the assistance of a PDMS stamp; a lavge range of nanodisk sizes were

fabricated and, thus, active tuning of metasurfaces with resonance wavelengths spanning

1200 nm in the NIR was explored. Up to 60 nm of wavelength tuning of the resonance was

observed and the switching was found to be fully reversible over more than ten

temperature cycles. In fact, no degradation of the sample was observed over several

months of measuring it, which is critical for creating robust practical devices based on

active nanogap metasurfaces. Further improvement of the VO2 thin film deposition on

127
Au, which is a non-lattice-matched substrate and thus requires careful optimization, is

expected to both improve the amount of wavelength tuning of the resonance and improve

the switching contrast by reducing the resonance bandwidth due to reduced surface

roughness of the VO2 film. Even with the current amount of switching, applications such

as holography should be possible since the phase of the reflection coefficient varies

rapidly around the resonance wavelength. Potential for integration with the other main

results on enhanced emission from SiVs exists, which could yield a fast and efficient

quantum light source with actively tunable properties.

128
Appendix A Additional Diamond Fabrication Process
Details for Nanogap-Cavity-Embedded SiVs
A pair of electronics-grade, 1x1x0.3 mm, <100>-oriented diamond slabs were

procured from Element Six. Following a tri-acid clean, the diamonds were commercially

implanted with 7keV Si ions at a density of 1 ∗ 1012 cm-2. Following the implantation, the

samples were annealed under high vacuum for 2 hours at 450 °C, 8 hours at 900 °C, and 2

hours at 1350 °C to form the SiV and repair damage from the implantation process. An

additional tri-acid clean was performed after the annealing and the samples were

transferred to sapphire substrates for etching. The diamonds were thinned to a gradient

from ~10 nm-1 µm in a deep RIE using alternating ArCl2 and O2 chemistries. After etching,

a drop of HF was used to release the diamonds from the sapphire substrates, and they

were washed with DI water onto their respective substrates. One of these substrates had

a layer of Au deposited on the top surface which would form the ground plane for the

nanocavity. Au nanodisks were fabricated on a separate substrate with EBL and then

transferred to the top surface of the diamond slab using a PDMS stamp, completing the

diamond-integrated nanocavity.

129
Appendix B Additional Diamond Fabrication Process
Details for Nonlinear Frequency Conversion

The samples used in this work consist of a gold ground plane on a silicon substrate,

an etched diamond slab and gold nanoparticles. The 75 nm gold film was deposited by

electron-beam evaporation with a 2 Å/s deposition rate. The etched diamond slab was

fabricated on a 1 x 1 mm, 30 µm thick electronic-grade, single-crystal diamond thin film

(Element Six). It was then mounted and etched on a sapphire carrier wafer to a submicron

thickness using deep reactive-ion etching (RIE) with alternating ArCl2 and O2 chemistries.

The thickness gradient in the etched diamond slab occurs due to the mechanical polishing

process on the surface prior to etching. Once etching is complete, the slab is dewetted

from the carrier wafer using a droplet of hydrofluoric acid (HF). Once the diamond is

detached, the HF is diluted with DI water and the thin film is transferred and bonded onto

the gold film via Van der Waals forces.

The gold nanoparticles were fabricated by electron-beam lithography (EBL) and

transferred onto diamond using polydimethylsiloxane (PDMS). The nanoparticles were

first fabricated on a silicon substrate. Then, a PDMS stamp was placed on top and the

sample was baked at 100°C for 30 minutes. Potassium hydroxide (KOH) solution was

used to etch the thin silicon dioxide (SiO2) layer on the silicon substrate for 2 hours. This

results in the gold nanoparticles debonding from the silicon substrate and attaching to the

PDMS stamp, which was then placed on a glass slide and stored in vacuum to dry

130
overnight. Lastly, with the gold nanoparticles face-down, the PDMS stamp was pressed

onto the diamond slab and baked at 120°C for 4 minutes. The sample was placed in

vacuum for two days and then baked at 120°C for 4 minutes before removing the stamp.

131
Appendix C Coupled Mode Theory Analysis of Nonlinear
Responses

The dependence of the nonlinear intensity on the excitation power P1(2) and the

excitation frequency detuning 1(2) from the respective resonant frequencies are

analyzed by coupled mode theory:

2
  
P SHG
 P  2 r1 2 
2

  1 + 1 
1

  r1    r 2 
P SFG  PP
1 2  2 2  2 2
  1 + 1    2 + 2 
3 (1)
  
PTHG  P13  2 r1 2 
  1 + 1 
2
     
P FWM
 P P  2 r1 2   2 r 2 2 
2

  1 + 1    2 + 2 
1 2

where we assumed that there is a predominant mode around the resonance frequency,

1(2) , and  i ( ri ) denotes the total (radiative) decay rate of the mode at i . Therefore, the

dependence of the nonlinear intensity on the excitation frequency detuning 1(2) can

exhibit a higher-order Lorentzian response.

132
Appendix D Additional Excitation Wavelength
Dependence Data for SHG and THG

Figure A.1 Additional THG Response Spectra as a Function of Excitation Wavelength

Figure A.2 Additional SHG Response Spectra as a Function of Excitation Wavelength

133
References

(1) Singer, A.; Boucheron, L.; Dietze, S. H.; Jensen, K. E.; Vine, D.; McNulty, I.;

Dufresne, E. R.; Prum, R. O.; Mochrie, S. G. J.; Shpyrko, O. G. Domain

Morphology, Boundaries, and Topological Defects in Biophotonic Gyroid

Nanostructures of Butterfly Wing Scales. Sci. Adv. 2016, 2 (6).

(2) Maier, S. A. Plasmonics: Fundamentals and Applications; Springer US, 2007.

(3) Rivera, N.; Kaminer, I.; Zhen, B.; Joannopoulos, J. D. . .; Soljačić, M. Shrinking

Light to Allow Forbidden Transitions on the Atomic Scale. Science (80-. ). 2016, 353

(6296), 263–269.

(4) Leonhardt, U. Optical Metamaterials: Invisibility Cup. Nat. Photonics 2007, 1 (4),

207–208.

(5) Lee, B.; Lee, I. M.; Kim, S.; Oh, D. H.; Hesselink, L. Review on Subwavelength

Confinement of Light with Plasmonics. J. Mod. Opt. 2010, 57 (16), 1479–1497.

(6) Purcell, E. M. Spontaneous Emission Probabilities at Radio Frequencies. Physical

Review. 1946, pp 246–260.

(7) Kinkhabwala, A.; Yu, Z.; Fan, S.; Avlasevich, Y.; Müllen, K.; Moerner, W. E. Large

Single-Molecule Fluorescence Enhancements Produced by a Bowtie Nanoantenna.

Nat. Photonics 2009, 3 (11), 654–657.

(8) Akselrod, G. M.; Argyropoulos, C.; Hoang, T. B.; Ciracì, C.; Fang, C.; Huang, J.;

Smith, D. R.; Mikkelsen, M. H. Probing the Mechanisms of Large Purcell

134
Enhancement in Plasmonic Nanoantennas. Nat. Photonics 2014, 8 (11), 835–840.

(9) Hoang, T. B.; Akselrod, G. M.; Mikkelsen, M. H. Ultrafast Room-Temperature

Single Photon Emission from Quantum Dots Coupled to Plasmonic Nanocavities.

Nano Lett. 2016, 16 (1), 270–275.

(10) Akselrod, G. M.; Ming, T.; Argyropoulos, C.; Hoang, T. B.; Lin, Y.; Ling, X.; Smith,

D. R.; Kong, J.; Mikkelsen, M. H. Leveraging Nanocavity Harmonics for Control

of Optical Processes in 2d Semiconductors. Nano Lett. 2015, 15 (5), 3578–3584.

(11) Akselrod, G. M.; Weidman, M. C.; Li, Y.; Argyropoulos, C.; Tisdale, W. A.;

Mikkelsen, M. H. Efficient Nanosecond Photoluminescence from Infrared PbS

Quantum Dots Coupled to Plasmonic Nanoantennas. ACS Photonics 2016, 3 (10),

1741–1746.

(12) Pitarke, J. M.; Silkin, V. M.; Chulkov, E. V.; Echenique, P. M. Theory of Surface

Plasmons and Surface-Plasmon Polaritons. Reports Prog. Phys. 2007, 70 (1), 1–87.

(13) Willets, K. A.; Van Duyne, R. P. Localized Surface Plasmon Resonance

Spectroscopy and Sensing. Annu. Rev. Phys. Chem. 2007, 58, 267–297.

(14) McMahon, J. M.; Schatz, G. C.; Gray, S. K. Plasmonics in the Ultraviolet with the

Poor Metals Al, Ga, In, Sn, Tl, Pb, and Bi. Phys. Chem. Chem. Phys. 2013, 15 (15),

5415–5423.

(15) Knight, M. W.; King, N. S.; Liu, L.; Everitt, H. O.; Nordlander, P.; Halas, N. J.

Aluminum for Plasmonics. ACS Nano 2014, 8 (1), 834–840.

135
(16) Watson, A. M.; Zhang, X.; Alcaraz De La Osa, R.; Sanz, J. M.; González, F.;

Moreno, F.; Finkelstein, G.; Liu, J.; Everitt, H. O. Rhodium Nanoparticles for

Ultraviolet Plasmonics. Nano Lett. 2015, 15 (2), 1095–1100.

(17) Li, W.; Guler, U.; Kinsey, N.; Naik, G. V.; Boltasseva, A.; Guan, J.; Shalaev, V. M.;

Kildishev, A. V. Refractory Plasmonics with Titanium Nitride: Broadband. Adv.

Mater. 2014, 26 (47), 7959–7965.

(18) Chikkaraddy, R.; De Nijs, B.; Benz, F.; Barrow, S. J.; Scherman, O. A.; Rosta, E.;

Demetriadou, A.; Fox, P.; Hess, O.; Baumberg, J. J. Single-Molecule Strong

Coupling at Room Temperature in Plasmonic Nanocavities. Nature 2016, 535

(7610), 127–130.

(19) Rivera, N.; Kaminer, I.; Zhen, B.; Joannopoulos, J. D. . .; Soljačić, M. Transitions on

the Atomic Scale. Science (80-. ). 2016, 353 (6296), 263–269.

(20) Cruz, D. F.; Fontes, C. M.; Semeniak, D.; Huang, J.; Hucknall, A.; Chilkoti, A.;

Mikkelsen, M. H. Ultrabright Fluorescence Readout of an Inkjet-Printed

Immunoassay Using Plasmonic Nanogap Cavities. Nano Lett. 2020, 20 (6), 4330–

4336.

(21) Sun, Y.; Xia, Y. Shape-Controlled Synthesis of Gold and Silver Nanoparticles.

Science (80-. ). 2002, 298 (5601), 2176–2179.

(22) Long, N. N.; Vu, L. Van; Kiem, C. D.; Doanh, S. C.; Nguyet, C. T.; Hang, P. T.;

Thien, N. D.; Quynh, L. M. Synthesis and Optical Properties of Colloidal Gold

136
Nanoparticles. J. Phys. Conf. Ser. 2009, 187.

(23) Sun, Y.; Chen, M.; Zhou, S.; Hu, J.; Wu, L. Controllable Synthesis and Surface

Wettability of Flower-Shaped Silver Nanocube-Organosilica Hybrid Colloidal

Nanoparticles. ACS Nano 2015, 9 (12), 12513–12520.

(24) Mahmoud, M. A.; Tabor, C. E.; El-Sayed, M. A.; Ding, Y.; Zhong, L. W. A New

Catalytically Active Colloidal Platinum Nanocatalyst: The Multiarmed Nanostar

Single Crystal. J. Am. Chem. Soc. 2008, 130 (14), 4590–4591.

(25) Decher, G. Fuzzy Nanoassemblies: Toward Layered Polymeric Multicomposites.

Science (80-. ). 1997, 277 (5330), 1232–1237.

(26) Mock, J. J.; Hill, R. T.; Degiron, A.; Zauscher, S.; Chilkoti, A.; Smith, D. R.

Distance-Dependent Plasmon Resonant Coupling between a Gold Nanoparticle

and Gold Film. Nano Lett. 2008, 8 (8), 2245–2252.

(27) Moreau, A.; Ciracì, C.; Mock, J. J.; Smith, D. R.; Hill, R. T.; Chilkoti, A.; Wang, Q.;

Wiley, B. J. Controlled-Reflectance Surfaces with Film-Coupled Colloidal

Nanoantennas. Nature 2012, 492 (7427), 86–89.

(28) Hoang, T. B.; Mikkelsen, M. H. Broad Electrical Tuning of Plasmonic

Nanoantennas at Visible Frequencies. Appl. Phys. Lett. 2016, 108 (18).

(29) Wilson, W. M.; Stewart, J. W.; Mikkelsen, M. H. Surpassing Single Line Width

Active Tuning with Photochromic Molecules Coupled to Plasmonic

Nanoantennas. Nano Lett. 2018, 18 (2), 853–858.

137
(30) Stewart, J. W.; Vella, J. H.; Li, W.; Fan, S.; Mikkelsen, M. H. Ultrafast Pyroelectric

Photodetection with On-Chip Spectral Filters. Nat. Mater. 2020, 19 (2), 158–162.

(31) Stewart, J. W.; Akselrod, G. M.; Smith, D. R.; Mikkelsen, M. H. Toward

Multispectral Imaging with Colloidal Metasurface Pixels. Adv. Mater. 2017, 29 (6).

(32) Chunnilall, C. J.; Degiovanni, I. Pietro; Kück, S.; Müller, I.; Sinclair, A. G.

Metrology of Single-Photon Sources and Detectors: A Review. Opt. Eng. 2014, 53

(8), 081910.

(33) Hoang, T. B.; Akselrod, G. M.; Mikkelsen, M. H. Ultrafast Room-Temperature

Single Photon Emission from Quantum Dots Coupled to Plasmonic Nanocavities.

Nano Lett. 2016, 16 (1), 270–275.

(34) Gruber, A.; Dräbenstedt, A.; Tietz, C.; Fleury, L.; Wrachtrup, J.; Von

Borczyskowski, C. Scanning Confocal Optical Microscopy and Magnetic

Resonance on Single Defect Centers. Science (80-. ). 1997, 276 (5321), 2012–2014.

(35) Wrachtrup, J.; Nizovtzev, A. Low-Temperature Microscopy and Spectroscopy on

Single Defect Centers in Diamond. Phys. Rev. B - Condens. Matter Mater. Phys. 1999,

60 (16), 11503–11508.

(36) Kurtsiefer, C.; Mayer, S.; Zarda, P.; Weinfurter, H. Stable Solid-State Source of

Single Photons. Phys. Rev. Lett. 2000, 85 (2), 290–293.

(37) Jelezko, F.; Gaebel, T.; Popa, I.; Domhan, M.; Gruber, A.; Wrachtrup, J.

Observation of Coherent Oscillation of a Single Nuclear Spin and Realization of a

138
Two-Qubit Conditional Quantum Gate. Phys. Rev. Lett. 2004, 93 (13), 1–4.

(38) Jelezko, F.; Gaebel, T.; Popa, I.; Gruber, A.; Wrachtrup, J. Observation of Coherent

Oscillations in a Single Electron Spin. Phys. Rev. Lett. 2004, 92 (7), 1–4.

(39) Doherty, M. W.; Manson, N. B.; Delaney, P.; Jelezko, F.; Wrachtrup, J.; Hollenberg,

L. C. L. The Nitrogen-Vacancy Colour Centre in Diamond. Physics Reports. 2013.

(40) Rondin, L.; Dantelle, G.; Slablab, A.; Grosshans, F.; Treussart, F.; Bergonzo, P.;

Perruchas, S.; Gacoin, T.; Chaigneau, M.; Chang, H. C.; Jacques, V.; Roch, J. F.

Surface-Induced Charge State Conversion of Nitrogen-Vacancy Defects in

Nanodiamonds. Phys. Rev. B - Condens. Matter Mater. Phys. 2010, 82 (11), 1–5.

(41) Optical Studies of the 1.945 EV Vibronic Band in Diamond. Proc. R. Soc. London. A.

Math. Phys. Sci. 1976, 348 (1653), 285–298.

(42) Edmonds, A. M.; D’Haenens-Johansson, U. F. S.; Cruddace, R. J.; Newton, M. E.;

Fu, K. M. C.; Santori, C.; Beausoleil, R. G.; Twitchen, D. J.; Markham, M. L.

Production of Oriented Nitrogen-Vacancy Color Centers in Synthetic Diamond.

Phys. Rev. B - Condens. Matter Mater. Phys. 2012, 86 (3), 1–7.

(43) Meijer, J.; Burchard, B.; Domhan, M.; Wittmann, C.; Gaebel, T.; Popa, I.; Jelezko,

F.; Wrachtrup, J. Generation of Single Color Centers by Focused Nitrogen

Implantation. Appl. Phys. Lett. 2005, 87 (26), 1–3.

(44) Loubser, J. H. N.; Van Wyk, J. A. Electron Spin Resonance in the Study of

Diamond. Reports Prog. Phys. 1978, 41 (8), 1201–1248.

139
(45) Davies, G. Dynamic Jahn-Teller Distortions at Trigonal Optical Centres in

Diamond. J. Phys. C Solid State Phys. 1979, 12 (13), 2551–2566.

(46) Batalov, A.; Jacques, V.; Kaiser, F.; Siyushev, P.; Neumann, P.; Rogers, L. J.;

McMurtrie, R. L.; Manson, N. B.; Jelezko, F.; Wrachtrup, J. Low Temperature

Studies of the Excited-State Structure of Negatively Charged Nitrogen-Vacancy

Color Centers in Diamond. Phys. Rev. Lett. 2009, 102 (19), 1–4.

(47) Tamarat, P.; Manson, N. B.; Harrison, J. P.; McMurtrie, R. L.; Nizovtsev, A.;

Santori, C.; Beausoleil, R. G.; Neumann, P.; Gaebel, T.; Jelezko, F.; Hemmer, P.;

Wrachtrup, J. Spin-Flip and Spin-Conserving Optical Transitions of the Nitrogen-

Vacancy Centre in Diamond. New J. Phys. 2008, 10.

(48) Balasubramanian, G.; Neumann, P.; Twitchen, D.; Markham, M.; Kolesov, R.;

Mizuochi, N.; Isoya, J.; Achard, J.; Beck, J.; Tissler, J.; Jacques, V.; Hemmer, P. R.;

Jelezko, F.; Wrachtrup, J. Ultralong Spin Coherence Time in Isotopically

Engineered Diamond. Nat. Mater. 2009, 8 (5), 383–387.

(49) Hanson, R.; Mendoza, F. M.; Epstein, R. J.; Awschalom, D. D. Polarization and

Readout of Coupled Single Spins in Diamond. Phys. Rev. Lett. 2006, 97 (8), 1–4.

(50) Gurudev Dutt, M. V. 基于单个电子和核自旋量子位的钻石量子寄存器 Quantum

Register Based On. Science 2007, 316 (June), 1312–1316.

(51) Steiner, M.; Neumann, P.; Beck, J.; Jelezko, F.; Wrachtrup, J. Universal

Enhancement of the Optical Readout Fidelity of Single Electron Spins at Nitrogen-

140
Vacancy Centers in Diamond. Phys. Rev. B - Condens. Matter Mater. Phys. 2010, 81

(3), 1–6.

(52) Jiang, L.; Hodges, J. S.; Maze, J. R.; Maurer, P.; Taylor, J. M.; Cory, D. G.; Hemmer,

P. R.; Walsworth, R. L.; Yacoby, A.; Zibrov, A. S.; Lukin, M. D. Repetitive Readout

of a Single Electronic Spin via Quantum Logic with Nuclear Spin Ancillae. Science

(80-. ). 2009, 326 (5950), 267–272.

(53) Bradac, C.; Gao, W.; Forneris, J.; Trusheim, M. E.; Aharonovich, I. Quantum

Nanophotonics with Group IV Defects in Diamond. Nat. Commun. 2019, 10 (1), 1–

13.

(54) Aharonovich, I.; Greentree, A. D.; Prawer, S. Diamond Photonics. Nat. Photonics

2011, 5 (7), 397–405.

(55) Clark, C. D.; Kanda, H.; Kiflawi, I.; Sittas, G. Silicon Defects. Phys. B 1995, 51 (23),

16681–68.

(56) Goss, J. P.; Jones, R.; Breuer, S. J.; Briddon, P. R.; Öberg, S. The Twelve-Line 1.682

EV Luminescence Center in Diamond and the Vacancy-Silicon Complex. Phys.

Rev. Lett. 1996, 77 (14), 3041–3044.

(57) Wang, C.; Kurtsiefer, C.; Weinfurter, H.; Burchard, B. Single Photon Emission

from SiV Centres in Diamond Produced by Ion Implantation. J. Phys. B At. Mol.

Opt. Phys. 2006, 39 (1), 37–41.

(58) Neu, E.; Steinmetz, D.; Riedrich-Möller, J.; Gsell, S.; Fischer, M.; Schreck, M.;

141
Becher, C. Single Photon Emission from Silicon-Vacancy Colour Centres in

Chemical Vapour Deposition Nano-Diamonds on Iridium. New J. Phys. 2011, 13.

(59) Li, K.; Zhou, Y.; Rasmita, A.; Aharonovich, I.; Gao, W. B. Nonblinking Emitters

with Nearly Lifetime-Limited Linewidths in CVD Nanodiamonds. Phys. Rev.

Appl. 2016, 6 (2), 1–7.

(60) Jantzen, U.; Kurz, A. B.; Rudnicki, D. S.; Schäfermeier, C.; Jahnke, K. D.;

Andersen, U. L.; Davydov, V. A.; Agafonov, V. N.; Kubanek, A.; Rogers, L. J.;

Jelezko, F. Nanodiamonds Carrying Silicon-Vacancy Quantum Emitters with

Almost Lifetime-Limited Linewidths. New J. Phys. 2016, 18 (7).

(61) Rogers, L. J.; Jahnke, K. D.; Teraji, T.; Marseglia, L.; Müller, C.; Naydenov, B.;

Schauffert, H.; Kranz, C.; Isoya, J.; McGuinness, L. P.; Jelezko, F. Multiple

Intrinsically Identical Single-Photon Emitters in the Solid State. Nat. Commun.

2014, 5, 1–2.

(62) Evans, R. E.; Sipahigil, A.; Sukachev, D. D.; Zibrov, A. S.; Lukin, M. D. Narrow-

Linewidth Homogeneous Optical Emitters in Diamond Nanostructures via Silicon

Ion Implantation. Phys. Rev. Appl. 2016, 5 (4), 1–8.

(63) Sipahigil, A.; Jahnke, K. D.; Rogers, L. J.; Teraji, T.; Isoya, J.; Zibrov, A. S.; Jelezko,

F.; Lukin, M. D. Indistinguishable Photons from Separated Silicon-Vacancy

Centers in Diamond. Phys. Rev. Lett. 2014, 113 (11), 1–5.

(64) C.K. Hong, Z. Y. Ou, L. M. Measurement of Subpicosecond Time Intervals

142
between Two Photons by Interference. Phys. Rev. Lett. 1987, 59 (18), 2044–2046.

(65) Knill, E.; Laflamme, R.; Milburn, G. J. A Scheme for Efficient Quantum

Computation with Linear Optics. Nature 2001, 409 (6816), 46–52.

(66) Buckley, S.; Rivoire, K.; Vučković, J. Engineered Quantum Dot Single-Photon

Sources. Reports Prog. Phys. 2012, 75 (12), 126503.

(67) Rogers, L. J.; Jahnke, K. D.; Metsch, M. H.; Sipahigil, A.; Binder, J. M.; Teraji, T.;

Sumiya, H.; Isoya, J.; Lukin, M. D.; Hemmer, P.; Jelezko, F. All-Optical

Initialization, Readout, and Coherent Preparation of Single Silicon-Vacancy Spins

in Diamond. Phys. Rev. Lett. 2014, 113 (26), 1–5.

(68) Pingault, B.; Becker, J. N.; Schulte, C. H. H.; Arend, C.; Hepp, C.; Godde, T.;

Tartakovskii, A. I.; Markham, M.; Becher, C.; Atatüre, M. All-Optical Formation of

Coherent Dark States of Silicon-Vacancy Spins in Diamond. Phys. Rev. Lett. 2014,

113 (26).

(69) Akselrod, G. M.; Huang, J.; Hoang, T. B.; Bowen, P. T.; Su, L.; Smith, D. R.;

Mikkelsen, M. H. Large-Area Metasurface Perfect Absorbers from Visible to Near-

Infrared. Adv. Mater. 2015, 27 (48), 8028–8034.

(70) Bowen, P. T.; Smith, D. R. Coupled-Mode Theory for Film-Coupled Plasmonic

Nanocubes. Phys. Rev. B - Condens. Matter Mater. Phys. 2014, 90 (19), 1–12.

(71) Huang, J.; Liu, C.; Zhu, Y.; Masala, S.; Alarousu, E.; Han, Y.; Fratalocchi, A.

Harnessing Structural Darkness in the Visible and Infrared Wavelengths for a

143
New Source of Light. Nat. Nanotechnol. 2016, 11 (1), 60–66.

(72) Rozin, M. J.; Rosen, D. A.; Dill, T. J.; Tao, A. R. Colloidal Metasurfaces Displaying

Near-Ideal and Tunable Light Absorbance in the Infrared. Nat. Commun. 2015, 6,

7325.

(73) Huang, Y. W.; Lee, H. W. H.; Sokhoyan, R.; Pala, R. A.; Thyagarajan, K.; Han, S.;

Tsai, D. P.; Atwater, H. A. Gate-Tunable Conducting Oxide Metasurfaces. Nano

Lett. 2016, 16 (9), 5319–5325.

(74) Peng, J.; Jeong, H. H.; Lin, Q.; Cormier, S.; Liang, H. L.; De Volder, M. F. L.;

Vignolini, S.; Baumberg, J. J. Scalable Electrochromic Nanopixels Using

Plasmonics. Sci. Adv. 2019, 5 (5), 1–9.

(75) König, T. A. F.; Ledin, P. A.; Kerszulis, J.; Mahmoud, M. A.; El-Sayed, M. A.;

Reynolds, J. R.; Tsukruk, V. V. Electrically Tunable Plasmonic Behavior of

Nanocube-Polymer Nanomaterials Induced by a Redox-Active Electrochromic

Polymer. ACS Nano 2014, 8 (6), 6182–6192.

(76) Xiong, K.; Tordera, D.; Emilsson, G.; Olsson, O.; Linderhed, U.; Jonsson, M. P.;

Dahlin, A. B. Switchable Plasmonic Metasurfaces with High Chromaticity

Containing Only Abundant Metals. Nano Lett. 2017, 17 (11), 7033–7039.

(77) Yao, Y.; Kats, M. A.; Genevet, P.; Yu, N.; Song, Y.; Kong, J.; Capasso, F. Broad

Electrical Tuning of Graphene-Loaded Plasmonic Antennas. Nano Lett. 2013, 13

(3), 1257–1264.

144
(78) Franklin, D.; Frank, R.; Wu, S. T.; Chanda, D. Actively Addressed Single Pixel

Full-Colour Plasmonic Display. Nat. Commun. 2017, 8 (May), 1–10.

(79) Gutruf, P.; Zou, C.; Withayachumnankul, W.; Bhaskaran, M.; Sriram, S.; Fumeaux,

C. Mechanically Tunable Dielectric Resonator Metasurfaces at Visible

Frequencies. ACS Nano 2016, 10 (1), 133–141.

(80) Huang, F.; Baumberg, J. J. Actively Tuned Plasmons on Elastomerically Driven

Au Nanoparticle Dimers. Nano Lett. 2010, 10 (5), 1787–1792.

(81) Duan, X.; Kamin, S.; Liu, N. Supporting Information Dynamic Plasmonic Colour

Display. Nat. Publ. Gr. 2017, 8, 1–4.

(82) Strohfeldt, N.; Tittl, A.; Schäferling, M.; Neubrech, F.; Kreibig, U.; Griessen, R.;

Giessen, H. Yttrium Hydride Nanoantennas for Active Plasmonics. Nano Lett.

2014, 14 (3), 1140–1147.

(83) Shao, L.; Zhuo, X.; Wang, J. Advanced Plasmonic Materials for Dynamic Color

Display. Adv. Mater. 2018, 30 (16), 1–18.

(84) Michel, A. K. U.; Zalden, P.; Chigrin, D. N.; Wuttig, M.; Lindenberg, A. M.;

Taubner, T. Reversible Optical Switching of Infrared Antenna Resonances with

Ultrathin Phase-Change Layers Using Femtosecond Laser Pulses. ACS Photonics

2014, 1 (9), 833–839.

(85) Yang, Y.; Kelley, K.; Sachet, E.; Campione, S.; Luk, T. S.; Maria, J. P.; Sinclair, M.

B.; Brener, I. Femtosecond Optical Polarization Switching Using a Cadmium

145
Oxide-Based Perfect Absorber. Nat. Photonics 2017, 11 (6), 390–395.

(86) Kats, M. A.; Blanchard, R.; Genevet, P.; Yang, Z.; Qazilbash, M. M.; Basov, D. N.;

Ramanathan, S.; Capasso, F. Thermal Tuning of Mid-Infrared Plasmonic Antenna

Arrays Using a Phase Change Material. Opt. Lett. 2013, 38 (3), 368.

(87) Huang, J.; Traverso, A. J.; Yang, G.; Mikkelsen, M. H. Real-Time Tunable Strong

Coupling: From Individual Nanocavities to Metasurfaces. ACS Photonics 2019, 6

(4), 838–843.

(88) Sautter, J.; Staude, I.; Decker, M.; Rusak, E.; Neshev, D. N.; Brener, I.; Kivshar, Y.

S. Active Tuning of All-Dielectric Metasurfaces. ACS Nano 2015, 9 (4), 4308–4315.

(89) Yin, X.; Steinle, T.; Huang, L.; Taubner, T.; Wuttig, M.; Zentgraf, T.; Giessen, H.

Beam Switching and Bifocal Zoom Lensing Using Active Plasmonic Metasurfaces.

Light Sci. Appl. 2017, 6 (7), e17016–e17016.

(90) Rensberg, J.; Zhang, S.; Zhou, Y.; McLeod, A. S.; Schwarz, C.; Goldflam, M.; Liu,

M.; Kerbusch, J.; Nawrodt, R.; Ramanathan, S.; Basov, D. N.; Capasso, F.;

Ronning, C.; Kats, M. A. Active Optical Metasurfaces Based on Defect-Engineered

Phase-Transition Materials. Nano Lett. 2016, 16 (2), 1050–1055.

(91) Driscoll, T.; Palit, S.; Qazilbash, M. M.; Brehm, M.; Keilmann, F.; Chae, B. G.; Yun,

S. J.; Kim, H. T.; Cho, S. Y.; Jokerst, N. M.; Smith, D. R.; Basov, D. N. Dynamic

Tuning of an Infrared Hybrid-Metamaterial Resonance Using Vanadium Dioxide.

Appl. Phys. Lett. 2008, 93 (2).

146
(92) Wang, G.; Chen, X.; Liu, S.; Wong, C.; Chu, S. Mechanical Chameleon through

Dynamic Real-Time Plasmonic Tuning. ACS Nano 2016, 10 (2), 1788–1794.

(93) Chain, E. E. Optical Properties of Vanadium Dioxide and Vanadium Pentoxide

Thin Films. Appl. Opt. 1991, 30 (19), 2782–2787.

(94) Zhou, Y.; Chen, X.; Ko, C.; Yang, Z.; Mouli, C.; Ramanathan, S. Voltage-Triggered

Ultrafast Phase Transition in Vanadium Dioxide Switches. IEEE Electron Device

Lett. 2013, 34 (2), 220–222.

(95) Kocer, H.; Butun, S.; Banar, B.; Wang, K.; Tongay, S.; Wu, J.; Aydin, K. Thermal

Tuning of Infrared Resonant Absorbers Based on Hybrid Gold-VO2

Nanostructures. Appl. Phys. Lett. 2015, 106 (16).

(96) Song, Z.; Wang, K.; Li, J.; Liu, Q. H. Broadband Tunable Terahertz Absorber

Based on Vanadium Dioxide Metamaterials. Opt. Express 2018, 26 (6), 7148.

(97) Wang, S.; Kang, L.; Werner, D. H. Hybrid Resonators and Highly Tunable

Terahertz Metamaterials Enabled by Vanadium Dioxide (VO2). Sci. Rep. 2017, 7

(1), 1–8.

(98) Xu, G.; Chen, Y.; Tazawa, M.; Jin, P. Surface Plasmon Resonance of Silver

Nanoparticles on Vanadium Dioxide. J. Phys. Chem. B 2006, 110 (5), 2051–2056.

(99) Shu, F. Z.; Yu, F. F.; Peng, R. W.; Zhu, Y. Y.; Xiong, B.; Fan, R. H.; Wang, Z. H.;

Liu, Y.; Wang, M. Dynamic Plasmonic Color Generation Based on Phase

Transition of Vanadium Dioxide. Adv. Opt. Mater. 2018, 6 (7), 1–10.

147
(100) Kakiuchida, H.; Jin, P.; Nakao, S.; Tazawa, M. Optical Properties of Vanadium

Dioxide Film during Semiconductive-Metallic Phase Transition. Japanese J. Appl.

Physics, Part 2 Lett. 2007, 46 (4–7).

(101) Earl, S. K.; James, T. D.; Davis, T. J.; McCallum, J. C.; Marvel, R. E.; Haglund, R. F.;

Roberts, A. Tunable Optical Antennas Enabled by the Phase Transition in

Vanadium Dioxide. Opt. Express 2013, 21 (22), 27503.

(102) Dicken, M. J.; Aydin, K.; Pryce, I. M.; Sweatlock, L. A.; Boyd, E. M.; Walavalkar, S.;

Ma, J.; Atwater, H. A. Frequency Tunable Near-Infrared Metamaterials Based on

VO_2 Phase Transition. Opt. Express 2009, 17 (20), 18330.

(103) Muskens, O. L.; Bergamini, L.; Wang, Y.; Gaskell, J. M.; Zabala, N.; De Groot, C.

H.; Sheel, D. W.; Aizpurua, J. Antenna-Assisted Picosecond Control of Nanoscale

Phase Transition in Vanadium Dioxide. Light Sci. Appl. 2016, 5 (10), e16173-9.

(104) Peter, A. P.; Martens, K.; Rampelberg, G.; Toeller, M.; Ablett, J. M.; Meersschaut,

J.; Cuypers, D.; Franquet, A.; Detavernier, C.; Rueff, J. P.; Schaekers, M.; Van

Elshocht, S.; Jurczak, M.; Adelmann, C.; Radu, I. P. Metal-Insulator Transition in

ALD VO 2 Ultrathin Films and Nanoparticles: Morphological Control. Adv. Funct.

Mater. 2015, 25 (5), 679–686.

(105) Koethe, T. C.; Hu, Z.; Haverkort, M. W.; Schüßler-Langeheine, C.; Venturini, F.;

Brookes, N. B.; Tjernberg, O.; Reichelt, W.; Hsieh, H. H.; Lin, H. J.; Chen, C. T.;

Tjeng, L. H. Transfer of Spectral Weight and Symmetry across the Metal-Insulator

148
Transition in VO2. Phys. Rev. Lett. 2006, 97 (11), 1–4.

(106) Morsy, A. M.; Barako, M. T.; Jankovic, V.; Wheeler, V. D.; Knight, M. W.;

Papadakis, G. T.; Sweatlock, L. A.; Hon, P. W. C.; Povinelli, M. L. Experimental

Demonstration of Dynamic Thermal Regulation Using Vanadium Dioxide Thin

Films. Sci. Rep. 2020, 10 (1), 1–10.

(107) Kozen, A. C.; Joress, H.; Currie, M.; Anderson, V. R.; Eddy, C. R.; Wheeler, V. D.

Structural Characterization of Atomic Layer Deposited Vanadium Dioxide. J.

Phys. Chem. C 2017, 121 (35), 19341–19347.

(108) Currie, M.; Mastro, M. A.; Wheeler, V. D. Characterizing the Tunable Refractive

Index of Vanadium Dioxide. Opt. Mater. Express 2017, 7 (5), 1697.

(109) Currie, M.; Mastro, M. A.; Wheeler, V. D. Atomic Layer Deposition of Vanadium

Dioxide and a Temperature-Dependent Optical Model. J. Vis. Exp. 2018, 2018

(135), 1–9.

(110) Shen, Q.; Hoang, T. B.; Yang, G.; Wheeler, V. D.; Mikkelsen, M. H. Probing the

Origin of Highly-Efficient Third-Harmonic Generation in Plasmonic Nanogaps.

Opt. Express 2018, 26 (16), 20718.

(111) Somaschi, N.; Giesz, V.; De Santis, L.; Loredo, J. C.; Almeida, M. P.; Hornecker, G.;

Portalupi, S. L.; Grange, T.; Antón, C.; Demory, J.; Gómez, C.; Sagnes, I.;

Lanzillotti-Kimura, N. D.; Lemaítre, A.; Auffeves, A.; White, A. G.; Lanco, L.;

Senellart, P. Near-Optimal Single-Photon Sources in the Solid State. Nat. Photonics

149
2016, 10 (5), 340–345.

(112) Huber, D.; Reindl, M.; Huo, Y.; Huang, H.; Wildmann, J. S.; Schmidt, O. G.;

Rastelli, A.; Trotta, R. Highly Indistinguishable and Strongly Entangled Photons

from Symmetric GaAs Quantum Dots. Nat. Commun. 2017, 8 (May), 1–7.

(113) Sun, S.; Kim, H.; Luo, Z.; Solomon, G. S.; Waks, E. A Single-Photon Switch and

Transistor Enabled by a Solid-State Quantum Memory. J. Phys. B At. Mol. Opt.

Phys. 2018, 361 (6397), 57–60.

(114) Sukachev, D. D.; Sipahigil, A.; Nguyen, C. T.; Bhaskar, M. K.; Evans, R. E.; Jelezko,

F.; Lukin, M. D. Silicon-Vacancy Spin Qubit in Diamond: A Quantum Memory

Exceeding 10 Ms with Single-Shot State Readout. Phys. Rev. Lett. 2017, 119 (22), 1–

6.

(115) Serrano, D.; Karlsson, J.; Fossati, A.; Ferrier, A.; Goldner, P. All-Optical Control of

Long-Lived Nuclear Spins in Rare-Earth Doped Nanoparticles. Nat. Commun.

2018, 9 (1), 1–7.

(116) Carter, S. G.; Sweeney, T. M.; Kim, M.; Kim, C. S.; Solenov, D.; Economou, S. E.;

Reinecke, T. L.; Yang, L.; Bracker, A. S.; Gammon, D. Quantum Control of a Spin

Qubit Coupled to a Photonic Crystal Cavity. Nat. Photonics 2013, 7 (4), 329–334.

(117) Neumann, P.; Jakobi, I.; Dolde, F.; Burk, C.; Reuter, R.; Waldherr, G.; Honert, J.;

Wolf, T.; Brunner, A.; Shim, J. H.; Suter, D.; Sumiya, H.; Isoya, J.; Wrachtrup, J.

High-Precision Nanoscale Temperature Sensing Using Single Defects in Diamond.

150
Nano Lett. 2013, 13 (6), 2738–2742.

(118) Clevenson, H.; Trusheim, M. E.; Teale, C.; Schröder, T.; Braje, D.; Englund, D.

Broadband Magnetometry and Temperature Sensing with a Light-Trapping

Diamond Waveguide. Nat. Phys. 2015, 11 (5), 393–397.

(119) Trusheim, M. E.; Englund, D. Wide-Field Strain Imaging with Preferentially

Aligned Nitrogen-Vacancy Centers in Polycrystalline Diamond. New J. Phys. 2016,

18 (12).

(120) Acosta, V. M.; Jensen, K.; Santori, C.; Budker, D.; Beausoleil, R. G.

Electromagnetically Induced Transparency in a Diamond Spin Ensemble Enables

All-Optical Electromagnetic Field Sensing. Phys. Rev. Lett. 2013, 110 (21), 1–6.

(121) Wolfowicz, G.; Whiteley, S. J.; Awschalom, D. D. Electrometry by Optical Charge

Conversion of Deep Defects in 4H-SiC. Proc. Natl. Acad. Sci. U. S. A. 2018, 115 (31),

7879–7883.

(122) Doherty, M. W.; Manson, N. B.; Delaney, P.; Jelezko, F.; Wrachtrup, J.; Hollenberg,

L. C. L. The Nitrogen-Vacancy Colour Centre in Diamond. Phys. Rep. 2013, 528 (1),

1–45.

(123) Pelton, M. Modified Spontaneous Emission in Nanophotonic Structures. Nat.

Photonics 2015, 9 (7), 427–435.

(124) Hausmann, B. J. M.; Shields, B. J.; Quan, Q.; Chu, Y.; De Leon, N. P.; Evans, R.;

Burek, M. J.; Zibrov, A. S.; Markham, M.; Twitchen, D. J.; Park, H.; Lukin, M. D.;

151
Loncǎr, M. Coupling of NV Centers to Photonic Crystal Nanobeams in Diamond.

Nano Lett. 2013, 13 (12), 5791–5796.

(125) Riedrich-Möller, J.; Arend, C.; Pauly, C.; Mücklich, F.; Fischer, M.; Gsell, S.;

Schreck, M.; Becher, C. Deterministic Coupling of a Single Silicon-Vacancy Color

Center to a Photonic Crystal Cavity in Diamond. Nano Lett. 2014, 14 (9), 5281–

5287.

(126) Zhang, J. L.; Sun, S.; Burek, M. J.; Dory, C.; Tzeng, Y. K.; Fischer, K. A.; Kelaita, Y.;

Lagoudakis, K. G.; Radulaski, M.; Shen, Z. X.; Melosh, N. A.; Chu, S.; Lončar, M.;

Vučković, J. Strongly Cavity-Enhanced Spontaneous Emission from Silicon-

Vacancy Centers in Diamond. Nano Lett. 2018, 18 (2), 1360–1365.

(127) Benedikter, J.; Kaupp, H.; Hümmer, T.; Liang, Y.; Bommer, A.; Becher, C.;

Krueger, A.; Smith, J. M.; Hänsch, T. W.; Hunger, D. Cavity-Enhanced Single-

Photon Source Based on the Silicon-Vacancy Center in Diamond. Phys. Rev. Appl.

2017, 7 (2), 1–12.

(128) Bogdanov, S. I.; Shalaginov, M. Y.; Lagutchev, A. S.; Chiang, C. C.; Shah, D.;

Baburin, A. S.; Ryzhikov, I. A.; Rodionov, I. A.; Kildishev, A. V.; Boltasseva, A.;

Shalaev, V. M. Ultrabright Room-Temperature Sub-Nanosecond Emission from

Single Nitrogen-Vacancy Centers Coupled to Nanopatch Antennas. Nano Lett.

2018, 18 (8), 4837–4844.

(129) Enderlein, J.; Erdmann, R. Fast Fitting of Multi-Exponential Decay Curves. Opt.

152
Commun. 1997, 134 (1–6), 371–378.

(130) Becker, W. Advanced Time-Correlated Single Photon Counting Techniques; 2005; Vol.

81.

(131) Wahl, M. Time-Correlated Single Photon Counting. Time-Correlated Single Photon

Counting. PicoQuant GmbH: Berlin 2017, p 5.

(132) Becker, W.; Bergmann Becker, A. Timing Stability of TCSPC Experiments.

(133) Jahnke, K. D.; Sipahigil, A.; Binder, J. M.; Doherty, M. W.; Metsch, M.; Rogers, L.

J.; Manson, N. B.; Lukin, M. D.; Jelezko, F. Electron-Phonon Processes of the

Silicon-Vacancy Centre in Diamond. New J. Phys. 2015, 17.

(134) Leifgen, M.; Schröder, T.; Gädeke, F.; Riemann, R.; Métillon, V.; Neu, E.; Hepp, C.;

Arend, C.; Becher, C.; Lauritsen, K.; Benson, O. Evaluation of Nitrogen- and

Silicon-Vacancy Defect Centres as Single Photon Sources in Quantum Key

Distribution. New J. Phys. 2014, 16.

(135) Evans, R. E.; Bhaskar, M. K.; Sukachev, D. D.; Nguyen, C. T.; Sipahigil, A.; Burek,

M. J.; Machielse, B.; Zhang, G. H.; Zibrov, A. S.; Bielejec, E.; Park, H.; Lončar, M.;

Lukin, M. D. Photon-Mediated Interactions between Quantum Emitters in a

Diamond Nanocavity. Science (80-. ). 2018, 362 (6415), 662–665.

(136) Wahl, M. The Principle of Time-Correlated Single Photon Counting.

(137) Kaminskii, A. A.; Ralchenko, V. G.; Konov, V. I. CVD-Diamond - A Novel χ(3)-

Nonlinear Active Crystalline Material for SRS Generation in Very Wide Spectral

153
Range. Laser Phys. Lett. 2006, 3 (4), 171–177.

(138) Burek, M. J.; Chu, Y.; Liddy, M. S. Z.; Patel, P.; Rochman, J.; Meesala, S.; Hong, W.;

Quan, Q.; Lukin, M. D.; Loncar, M. High Quality-Factor Optical Nanocavities in

Bulk Single-Crystal Diamond. Nat. Commun. 2014, 5 (May), 1–7.

(139) Atikian, H. A.; Latawiec, P.; Burek, M. J.; Sohn, Y. I.; Meesala, S.; Gravel, N.;

Kouki, A. B.; Lončar, M. Freestanding Nanostructures via Reactive Ion Beam

Angled Etching. APL Photonics 2017, 2 (5).

(140) Gao, F.; Van Erps, J.; Huang, Z.; Thienpont, H.; Beausoleil, R. G.; Vermeulen, N.

Directional Coupler Based on Single-Crystal Diamond Waveguides. IEEE J. Sel.

Top. Quantum Electron. 2018, 24 (6).

(141) Khanaliloo, B.; Mitchell, M.; Hryciw, A. C.; Barclay, P. E. High-Q/V Monolithic

Diamond Microdisks Fabricated with Quasi-Isotropic Etching. Nano Lett. 2015, 15

(8), 5131–5136.

(142) Bray, K.; Fedyanin, D. Y.; Khramtsov, I. A.; Bilokur, M. O.; Regan, B.; Toth, M.;

Aharonovich, I. Electrical Excitation and Charge-State Conversion of Silicon

Vacancy Color Centers in Single-Crystal Diamond Membranes. Appl. Phys. Lett.

2020, 116 (10).

(143) Li, L.; Schröder, T.; Chen, E. H.; Walsh, M.; Bayn, I.; Goldstein, J.; Gaathon, O.;

Trusheim, M. E.; Lu, M.; Mower, J.; Cotlet, M.; Markham, M. L.; Twitchen, D. J.;

Englund, D. Coherent Spin Control of a Nanocavity-Enhanced Qubit in Diamond.

154
Nat. Commun. 2015, 6.

(144) Sohn, Y. I.; Meesala, S.; Pingault, B.; Atikian, H. A.; Holzgrafe, J.; Gündoǧan, M.;

Stavrakas, C.; Stanley, M. J.; Sipahigil, A.; Choi, J.; Zhang, M.; Pacheco, J. L.;

Abraham, J.; Bielejec, E.; Lukin, M. D.; Atatüre, M.; Lončar, M. Controlling the

Coherence of a Diamond Spin Qubit through Its Strain Environment. Nat.

Commun. 2018, 9 (1), 1–6.

(145) Siampour, H.; Wang, O.; Zenin, V. A.; Boroviks, S.; Siyushev, P.; Yang, Y.;

Davydov, V. A.; Kulikova, L. F.; Agafonov, V. N.; Kubanek, A.; Mortensen, N. A.;

Jelezko, F.; Bozhevolnyi, S. I. Ultrabright Single-Photon Emission from

Germanium-Vacancy Zero-Phonon Lines: Deterministic Emitter-Waveguide

Interfacing at Plasmonic Hot Spots. Nanophotonics 2020, 9 (4), 953–962.

(146) Evans, R. E.; Bhaskar, M. K.; Sukachev, D. D.; Nguyen, C. T.; Sipahigil, A.; Burek,

M. J.; Machielse, B.; Zhang, G. H.; Zibrov, A. S.; Bielejec, E.; Park, H.; Lončar, M.;

Lukin, M. D. Photon-Mediated Interactions between Quantum Emitters in a

Diamond Nanocavity. Science (80-. ). 2018, 362 (6415), 662–665.

(147) Schröder, T.; Mouradian, S. L.; Zheng, J.; Trusheim, M. E.; Walsh, M.; Chen, E. H.;

Li, L.; Bayn, I.; Englund, D. Quantum Nanophotonics in Diamond [Invited]. J.

Opt. Soc. Am. B 2016, 33 (4), B65.

(148) Latawiec, P.; Venkataraman, V.; Burek, M. J.; Hausmann, B. J. M.; Bulu, I.; Lončar,

M. On-Chip Diamond Raman Laser. Optica 2015, 2 (11), 924.

155
(149) Latawiec, P.; Venkataraman, V.; Shams-Ansari, A.; Markham, M.; Loncar, M. An

Integrated Diamond Raman Laser Pumped in the Near-Visible. Opt. Lett. 2017, 43

(2), 318–321.

(150) Latawiec, P.; Shams-Ansari, A.; Okawachi, Y.; Venkataraman, V.; Yu, M.; Atikian,

H.; Harris, G. L.; Picque, N.; Gaeta, A. L.; Loncar, M. Supercontinuum Generation

in Angle-Etched Diamond Waveguides. 2018 Conf. Lasers Electro-Optics, CLEO

2018 - Proc. 2018, 44 (16), 4056–4059.

(151) Hausmann, B. J. M.; Bulu, I.; Venkataraman, V.; Deotare, P.; Loncar, M. Diamond

Nonlinear Photonics. Nat. Photonics 2014, 8 (5), 369–374.

(152) Lin, Z.; Johnson, S. G.; Rodriguez, A. W.; Loncar, M. Design of Diamond

Microcavities for Single Photon Frequency Down-Conversion. Opt. Express 2015,

23 (19), 25279.

(153) Aouani, H.; Rahmani, M.; Navarro-Cía, M.; Maier, S. A. Third-Harmonic-

Upconversion Enhancement from a Single Semiconductor Nanoparticle Coupled

to a Plasmonic Antenna. Nat. Nanotechnol. 2014, 9 (4), 290–294.

(154) Genevet, P.; Tetienne, J. P.; Gatzogiannis, E.; Blanchard, R.; Kats, M. A.; Scully, M.

O.; Capasso, F. Large Enhancement of Nonlinear Optical Phenomena by

Plasmonic Nanocavity Gratings. Nano Lett. 2010, 10 (12), 4880–4883.

(155) Zhang, Y.; Wen, F.; Zhen, Y. R.; Nordlander, P.; Halas, N. J. Coherent Fano

Resonances in a Plasmonic Nanocluster Enhance Optical Four-Wave Mixing. Proc.

156
Natl. Acad. Sci. U. S. A. 2013, 110 (23), 9215–9219.

(156) Celebrano, M.; Wu, X.; Baselli, M.; Großmann, S.; Biagioni, P.; Locatelli, A.; De

Angelis, C.; Cerullo, G.; Osellame, R.; Hecht, B.; Duò, L.; Ciccacci, F.; Finazzi, M.

Mode Matching in Multiresonant Plasmonic Nanoantennas for Enhanced Second

Harmonic Generation. Nat. Nanotechnol. 2015, 10 (5), 412–417.

(157) Lassiter, J. B.; McGuire, F.; Mock, J. J.; Ciracì, C.; Hill, R. T.; Wiley, B. J.; Chilkoti,

A.; Smith, D. R. Plasmonic Waveguide Modes of Film-Coupled Metallic

Nanocubes. Nano Lett. 2013, 13 (12), 5866–5872.

(158) Shen, Q.; Boyce, A. M.; Yang, G.; Mikkelsen, M. H. Polarization-Controlled

Nanogap Cavity with Dual-Band and Spatially Overlapped Resonances. ACS

Photonics 2019, 6 (8), 1916–1921.

(159) Baumberg, J. J.; Aizpurua, J.; Mikkelsen, M. H.; Smith, D. R. Extreme

Nanophotonics from Ultrathin Metallic Gaps. Nat. Mater. 2019, 18 (7), 668–678.

(160) Rose, A.; Hoang, T. B.; McGuire, F.; Mock, J. J.; Ciracì, C.; Smith, D. R.; Mikkelsen,

M. H. Control of Radiative Processes Using Tunable Plasmonic Nanopatch

Antennas. Nano Lett. 2014, 14 (8), 4797–4802.

(161) Sartorello, G.; Olivier, N.; Zhang, J.; Yue, W.; Gosztola, D. J.; Wiederrecht, G. P.;

Wurtz, G.; Zayats, A. V. Ultrafast Optical Modulation of Second- and Third-

Harmonic Generation from Cut-Disk-Based Metasurfaces. ACS Photonics 2016, 3

(8), 1517–1522.

157
(162) Liu, S.; Vabishchevich, P. P.; Vaskin, A.; Reno, J. L.; Keeler, G. A.; Sinclair, M. B.;

Staude, I.; Brener, I. An All-Dielectric Metasurface as a Broadband Optical

Frequency Mixer. Nat. Commun. 2018, 9 (1), 1–6.

(163) Shen, Q.; Jin, W.; Yang, G.; Rodriguez, A. W.; Mikkelsen, M. H. Active Control of

Multiple, Simultaneous Nonlinear Optical Processes in Plasmonic Nanogap

Cavities. ACS Photonics 2020, 7 (4), 901–907.

(164) Park, S.; Hahn, J. W.; Lee, J. Y. Doubly Resonant Metallic Nanostructure for High

Conversion Efficiency of Second Harmonic Generation. Opt. Express 2012, 20 (5),

4856.

(165) Zhang, Y.; Grady, N. K.; Ayala-Orozco, C.; Halas, N. J. Three-Dimensional

Nanostructures as Highly Efficient Generators of Second Harmonic Light. Nano

Lett. 2011, 11 (12), 5519–5523.

(166) Albrecht, G.; Ubl, M.; Kaiser, S.; Giessen, H.; Hentschel, M. Comprehensive Study

of Plasmonic Materials in the Visible and Near-Infrared: Linear, Refractory, and

Nonlinear Optical Properties. ACS Photonics 2018, 5 (3), 1058–1067.

(167) Guler, U.; Boltasseva, A.; Shalaev, V. M. Refractory Plasmonics. Science (80-. ).

2014, 344 (6181), 263–264.

(168) Wells, M. P.; Bower, R.; Kilmurray, R.; Zou, B.; Mihai, A. P.; Gobalakrichenane, G.;

Alford, N. M.; Oulton, R. F. M.; Cohen, L. F.; Maier, S. A.; Zayats, A. V.; Petrov, P.

K. Temperature Stability of Thin Film Refractory Plasmonic Materials. Opt.

158
Express 2018, 26 (12), 15726.

(169) Albrecht, G.; Kaiser, S.; Giessen, H.; Hentschel, M. Refractory Plasmonics without

Refractory Materials. Nano Lett. 2017, 17 (10), 6402–6408.

159
Biography
Andrew M. Boyce is a PhD candidate at Duke University. In May of 2016, he

received his bachelor's degree in Physics with a minor in mathematics (magna cum

laude) at Boston College, where he was also a Gabelli Presidential Scholar. He entered

the graduate program in Electrical and Computer Engineering at Duke University in

September, 2016. In June, 2017, he was selected as a Fitzpatrick Foundation Scholar,

which fully supported two years of his graduate research. In April, 2019, he passed his

Preliminary Exam and became a PhD candidate. Shortly thereafter, in December, 2019,

he received his Masters Degree in Electrical and Computer Engineering from Duke

University. He is expected to receive his PhD in the same department in September,

2022. During his graduate career, Andrew has published several academic articles,

including “Polarization-Controlled Nanogap Cavity with Dual-Band and Spatially

Overlapped Resonances,” “Actively Tunable Metasurfaces via Plasmonic Nanogap

Cavities with Sub-10-nm VO2 Films,” and “A Metasurface-Based Diamond Frequency

Converter using Plasmonic Nanogap Resonators.”

160

You might also like