You are on page 1of 84

Depar tment of Applied Physics

I nteraction of light with


f unctional spatially
dispersive nanomaterials

Ville Kivijärvi

D O C TO R A L
DISSERTATIONS
Aalto University publication series
DOCTORAL DISSERTATIONS 17/2019

Interaction of light with functional


spatially dispersive nanomaterials

Ville Kivijärvi

A doctoral dissertation completed for the degree of Doctor of


Science (Technology) to be defended, with the permission of the
Aalto University School of Science, at a public examination held at
the lecture hall AS1 of the TUAS building (Maarintie 8, Espoo,
Finland) on the 22nd of March 2019 at 12.

Aalto University
School of Science
Department of Applied Physics
Optics and Photonics group
Supervising professor
Prof. Matti Kaivola, Aalto University, Finland

Thesis advisor
Dr. Andriy Shevchenko, Aalto University, Finland

Preliminary examiners
Prof. Jari Turunen, University of Eastern Finland, Finland
Prof. Sergei Popov, KTH Royal Institute of Technology, Sweden

Opponent
Prof. Carsten Rockstuhl, Karlsruhe Institute of Technology, Germany

Aalto University publication series


DOCTORAL DISSERTATIONS 17/2019

© 2019 Ville Kivijärvi

ISBN 978-952-60-8396-4 (printed)


ISBN 978-952-60-8397-1 (pdf)
ISSN 1799-4934 (printed)
ISSN 1799-4942 (pdf)
http://urn.fi/URN:ISBN:978-952-60-8397-1

Unigrafia Oy
Helsinki 2019 N
SWA ECO
C
DI

LA
NOR

BEL

Finland
Printed matter
Printed matter
1234 5678
4041-0619
Ab s t ra c t
Aalto University, P.O. Box 11000, FI-00076 Aalto www.aalto.fi

Author
Ville Kivijärvi
Name of the doctoral dissertation
Interaction of light with functional spatially dispersive nanomaterials
P u b l i s h e r School of Science
U n i t Department of Applied Physics
S e r i e s Aalto University publication series DOCTORAL DISSERTATIONS 17/2019
F i e l d o f r e s e a r c h Nano-optics
M a n u s c r i p t s u b m i t t e d 28 November 2018 D a t e o f t h e d e f e n c e 22 March 2019
P e r m i s s i o n t o p u b l i s h g r a n t e d ( d a t e ) 28 January 2019 L a n g u a g e English
Monograph Article dissertation Essay dissertation
Abstract
Progress in nanotechnology has enabled systematic development of artifi cial nanomaterials with
extraordinary optical properties, such as zeroth and negative index of refraction, perfect absorption,
and enhanced nonlinearity, anisotropy and temporal dispersion. Another property that is common
for designed nanomaterials is spatial dispersion. It makes the refractive index and wave impedance
depend on light propagation direction, which complicates the description, but also makes it
p os s ibl e t o obt ain p re viou s ly u nre achable cap abil it ie s f o r t he mat e rial s . F o r e xamp le , t he mat e rial
can be made to refl ect or absorb light differently by its different facets. In spite of a high potential
of spatially dispersive nanomaterials in science and technology, only few theoretical models and
design tools for these materials can be found in the literature.

The re sults p re se nt e d in this t he sis can p ave t he w ay to comp re he nsive charact e rization and more
efficient design of spatially dispersive nanomaterials, including metamaterials and nanostructured
optical waveguides. We develop novel theoretical methods and numerical calculation techniques
to characterize and design nanostructured materials and devices made of them. Furthermore, we
put forward a novel approach to characterize nanostructured optical media, for which
electromagnetic modes cannot be introduced due to the effect of polarization conversion by spatial
dispersion. We fi nd that a large variety of designed optical nanomaterials belong to this class of
optical media. In addition, the thesis contains an effi cient semi-analytical technique to model the
inte raction of optical be ams with sp atially dispe rsive mate rials. Making use of the tools of Fourie r
optics, the method allows treating the beam-propagation phenomenon in large-sized nanomaterial
s amp l e s . C o mp are d t o c o nve nt io nal nu me ric al ap p ro ac h e s , t he t e c hniqu e is c o mp u t at io nal l y l ig ht
and provides a deeper understanding of the light-matter interaction picture.

Using the developed methods, we design novel optical nanoscatterers and nanomaterials composed
of them. As an example, we demonstrate both theoretically and experimentally a spatially
dispersive metasurface that shows primarily quadrupole optical response when illuminated from
one side, and primarily dipole response when illuminated from the other side. Furthermore, we
design novel diffraction-compensating optical metamaterials, whose anisotropy and spatial
dispe rsion pre ve nt focuse d light be ams from spre ading up on propag ation. We also de monstrate a
diffraction-compensating metamaterial slab waveguide with low refl ection and absorption losses
and a broadband operation. The designed materials and optical components based on them may
find applications in efficient solar cells, novel laser resonators, and high-speed photonic integrated
circuits, and together with the presented theoretical tools, contribute to the progress of
nanostructured optical media towards real-life applications.
K e y w o r d s Optical nanostructures, metamaterials, wave propagation, diffraction compensation
I S B N ( p r i n t e d ) 978-952-60-8396-4 I S B N ( p d f ) 978-952-60-8397-1
I S S N ( p r i n t e d ) 1799-4934 I S S N ( p d f ) 1799-4942
L o c a t i o n o f p u b l i s h e r Helsinki L o c a t i o n o f p r i n t i n g Helsinki Y e a r 2019
P a g e s 148 u r n http://urn.fi /URN:ISBN: 978-952-60-8397-1
Tiivistelmä
Aalto-yliopisto, PL 11000, 00076 Aalto www.aalto.fi

Tekijä
Ville Kivijärvi
Väitöskirjan nimi
Valon ja spatiaalisesti dispersiivisten nanomateriaalien vuorovaikutus
J u l k a i s i j a Perustieteiden korkeakoulu
Y k s i k k ö Teknillisen fysiikan laitos
S a r j a Aalto University publication series DOCTORAL DISSERTATIONS 17/2019
T u t k i m u s a l a Nano-optiikka
K ä s i k i r j o i t u k s e n p v m 28.11.2018 V ä i t ö s p ä i v ä 22.03.2019
J u l k a i s u l u v a n m y ö n t ä m i s p ä i v ä 28.01.2019 K i e l i Englanti
Monografia Artikkeliväitöskirja Esseeväitöskirja
Tiivistelmä
Nanoteknologian kehitys on tehnyt mahdolliseksi keinotekoisten nanomateriaalien
järjestelmällisen kehittämisen. Näillä materiaaleilla voi olla erikoisia optisia ominaisuuksia, kuten
voimakas absorptio, anisotropia ja dispersio, sekä nolla- tai negatiivinen taitekerroin. Optiset
nanomateriaalit ovat usein spatiaalisesti dispersiivisiä. Tämä ilmiö saa optisen väliaineen
taitekertoimen ja impedanssin riippumaan valon etenemissuunnasta, mikä vaikeuttaa väliaineen
mallinnusta, mutta mahdollistaa joukon epätavallisia optisia ominaisuuksia. Esimerkkinä
tällaisesta väliaineesta on materiaali, jonka eri pintojen heijastavuudet poikkeavat toisistaan.
Spatiaalisesti dispersiivisillä materiaaleilla on runsaasti sovelluksia, mutta niiden teoreettiset mallit
ja suunnittelutyökalut ovat yhä suurelta osin kehittymättömiä.

Väitöskirjan tutkimus tekee tietä spatiaalisesti dispersiivisten nanomateriaalien, kuten optisten


metamateriaalien ja nanorakenteisten aaltojohteiden, laaja-alaiselle karakterisoinnille ja
suunnittelulle. Väitöstyössä kehitetään uusia teoreettisia menetelmiä ja numeerisia
mallinnustekniikoita tällaista karakterisointia ja suunnittelua varten. Kehitetyillä menetelmillä
voidaan mallintaa esimerkiksi väliaineita, joissa spatiaalinen dispersio muuttaa valon
polarisaatiotilaa ja täten estää sähkömagneettisten moodien olemassaolon. Monet optiset
nanomateriaalit kuuluvat tällaisiin väliaineisiin. Lisäksi väitöskirja sisältää tehokkaan semi-
analyyttisen menetelmän, jolla voidaan mallintaa optisten säteiden ja spatiaalisesti dispersiivisten
materiaalien vuorovaikutusta. Kehitetty menetelmä hyödyntää Fourier-optiikkaa ja sallii säteen
mallinnuksen poikkeuksellisen suurissa nanomateriaalirakenteissa. Verrattuna yleisimmin
käytettyihin numeerisiin menetelmiin, menetelmä on laskennallisesti kevyt ja antaa selkeämmän
kuvan valon ja materiaalin vuorovaikutuksesta.

Väitöstyössä on kehitetty uudenlaisia optisia nanohiukkasia ja niistä koostuvia nanomateriaaleja


käyttäen työssä esiteltyjä uusia menetelmiä. Esimerkiksi on suunniteltu spatiaalisesti
dispersiivinen metapinta, jonka optinen vaste voi olla kvadrupoli- tai dipoli-typpinen riippuen
valon kulkusuunnasta. Toinen esimerkki ovat diffraktiota kompensoivat materiaalit, joiden
anisotropia ja spatiaalinen dispersio estävät fokusoidun valon merkittävän leviämisen valon
edetessä. Tämän kaltaista materiaalia voidaan käyttää diffraktiota kompensoivassa tasomaisessa
aaltojohdossa, jolla on suuri toimintakaistanleveys, mutta matalat heijastus- ja absorptiohäviöt.
Väitöstyössä suunnitellut materiaalit ja niihin perustuvat optiset komponentit voivat löytää
sovelluksia tehokkaissa aurinkokennoissa, uudenlaisissa laser-resonaattoreissa ja nopeissa
fotonisissa integroiduissa piireissä. Kehitettyjen teoreettisten menetelmien kanssa uudet
materiaalit tuovat oman lisänsä nano-optiikan kehitykseen kohti käytännön sovelluksia.
Avainsanat Optiset nanorakenteet, metamateriaalit, optinen aaltoliike, diffraktion kompensaatio
I S B N ( p a i n e t t u ) 978-952-60-8396-4 I S B N ( p d f ) 978-952-60-8397-1
I S S N ( p a i n e t t u ) 1799-4934 I S S N ( p d f ) 1799-4942
J u l k a i s u p a i k k a Helsinki P a i n o p a i k k a Helsinki V u o s i 2019
S i v u m ä ä r ä 148 u r n http://urn.fi /URN:ISBN: 978-952-60-8397-1
Preface

Most of the research work in this thesis was carried out at the Optics and
Photonics group in the Department of Applied Physics of Aalto University.
A part of the work was done in collaboration with Prof. Klas Lindfors in the
University of Cologne in Germany. All the work was carried out between
the years 2014 and 2019.
During my doctoral studies, I learned various important skills that
are widely applicable in real life. I want to express my gratitude to my
colleagues at the Optics and Photonics group. First, I want to thank my
former course assistant Dr. Patrick Grahn for sparking my interest in nano-
optics and introducing me to the field of optical metamaterials. In addition,
I am indebted to my supervisor Prof. Matti Kaivola for accepting me as a
doctoral student and acting as an excellent supervisor. Similarly, I want to
thank my adviser Dr. Andriy Shevchenko for teaching me good research
practice and helping me with manuscript preparation. Additional thanks
go to my other colleagues Prof. Klas Lindfors, Alex Karrila, and especially
Markus Nyman, for their important contributions to the research work.
I am sincerely grateful to Prof. Carsten Rockstuhl for his willingness to
act as the opponent in my doctoral defense and making the associated trip
to Finland. Equally, I am grateful to Prof. Jari Turunen and Prof. Sergei
Popov for the careful pre-examination of my thesis.
Most of the research work was funded by personal grants from the Viljo,
Yrjö, and Kalle Väisälä Foundation of the Finnish Academy of Science and
Letters. This funding and the additional research funding by the Academy
of Finland and the Finnish Cultural Foundation are acknowledged by me.
Outside of my professional sphere, I want to express a special thanks to

i
Preface

my mother Kia Kivijärvi for the invaluable support she provided during my
studies. Special thanks also go to my grandfather Matti Kivijärvi for his
encouragement that made me undertake physics studies and eventually
led to the accomplishment of my doctoral degree. Furthermore, I want
to thank my cousin Sari Kivijärvi for spending quality time with me and
proofreading this section of my thesis. Finally, I am incredibly grateful to
my other relatives and friends both in Finland and around the world for
acting as my emotional support and enriching my life during the past five
years.

Helsinki, February 22, 2019,

Ville Kivijärvi

ii
Contents

Preface i

Contents iii

List of Publications v

Author’s contribution vii

1. Introduction 1

2. Theoretical description of optical nanomaterials 5


2.1 Metal nanoparticles . . . . . . . . . . . . . . . . . . . . . . 5
2.2 Multipole expansion . . . . . . . . . . . . . . . . . . . . . . 8
2.3 Metasurfaces . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.4 Metamaterials . . . . . . . . . . . . . . . . . . . . . . . . . 21

3. Interaction of light with optical nanomaterials 29


3.1 Plane wave decomposition . . . . . . . . . . . . . . . . . . . 29
3.2 Reflection and transmission of optical beams . . . . . . . . 33
3.3 Diffraction compensation . . . . . . . . . . . . . . . . . . . 42
3.4 Light propagation in metamaterial slab waveguides . . . 50

4. Summary and outlook 57

Bibliography 61

Publications 69

iii
Contents

iv
List of Publications

This thesis consists of an overview and the following publications which


are referred to in the text by their Roman numerals.

I V. Kivijärvi, M. Nyman, A. Karrila, P. Grahn, A. Shevchenko and M.


Kaivola. Interaction of metamaterials with optical beams. New Journal
of Physics, 17, 063019, June 2015.

II A. Shevchenko, V. Kivijärvi, P. Grahn, M. Kaivola and K. Lindfors.


Bifacial Metasurface with Quadrupole Optical Response. Physical Review
Applied, 4, 024019, August 2015.

III V. Kivijärvi, M. Nyman, A. Shevchenko and M. Kaivola. An optical


metamaterial with simultaneously suppressed optical diffraction and
surface reflection. Journal of Optics, 18, 035103, February 2016.

IV V. Kivijärvi, M. Nyman, A. Shevchenko and M. Kaivola. Optical-image


transfer through a diffraction-compensating metamaterial. Optics Ex-
press, 24, 9806–9815, April 2016.

V V. Kivijärvi, M. Nyman, A. Shevchenko and M. Kaivola. Theoretical


description and design of nanomaterial slab waveguides: application
to compensation of optical diffraction. Optics Express, 26, 9134–9147,
March 2018.

v
List of Publications

vi
Author’s contribution

Publication I: “Interaction of metamaterials with optical beams”

The author introduced the reflection and transmission matrices for spa-
tially dispersive metamaterials and developed the recursive method for
calculating these matrices for multilayered metamaterials. The method
to model the interaction of optical beams with metamaterial slabs was
developed by the author. The studied metamaterials were designed and
characterized by the author. The author played the key role in writing the
manuscript.

Publication II: “Bifacial Metasurface with Quadrupole Optical


Response”

The author built the computational model of the metasurface, computed


the spectra of the metasurface’s reflection and transmission coefficients,
and calculated the corresponding spectra of the multipole moments. In
addition, the author participated in the analysis of the experimental results
and writing the manuscript.

Publication III: “An optical metamaterial with simultaneously


suppressed optical diffraction and surface reflection”

The author designed and characterized the material and analysed its prop-
erties. In addition, the author had the key role in writing the manuscript.

vii
Author’s contribution

Publication IV: “Optical-image transfer through a


diffraction-compensating metamaterial”

The author designed and characterized the metamaterial, analysed the


obtained results, and played the key role in writing the manuscript.

Publication V: “Theoretical description and design of nanomaterial


slab waveguides: application to compensation of optical diffraction”

The author significantly contributed to the development of the theory of


spatially dispersive slab waveguides, developed the numerical calculation
methods, as well as designed and characterized the introduced diffraction-
compensating waveguide. The author played a central role in analysing
the results and writing the manuscript.

Other publications to which the author has contributed:

• A. Shevchenko, P. Grahn, V. Kivijärvi, M. Nyman and M. Kaivola. Spa-


tially dispersive functional optical metamaterials. Journal of Nanopho-
tonics 9, 093097, January 2015.

• M. Nyman, V. Kivijärvi, A. Shevchenko and M. Kaivola. Generation of


light in spatially dispersive materials. Physical Review A 95, 043802,
April 2017.

• A. Shevchenko, M. Nyman, V. Kivijärvi and M. Kaivola. Optical wave


parameters for spatially dispersive and anisotropic nanomaterials. Optics
Express 25, 8550–8562, April 2017.

viii
1. Introduction

Optical properties of ordinary materials are fundamentally limited. For


example, magnetic light-matter interactions are usually negligible, and
therefore, values of the relative magnetic permeabilities are very close to
unity [1]. These limitations exist, because the structural units of ordinary
materials, e.g., atoms, are much smaller than the optical wavelengths.
The small size of the structural units allows excitation of electric-current
dipoles in these units, but prevents efficient excitation of higher-order mul-
tipoles, such as magnetic dipoles. This restricts available optical properties
of ordinary matter, because some optical responses can only take place
in the presence of multipole excitations of higher orders than the electric
dipole. However, an optical material can be composed of nanoparticles
that are not much smaller than the wavelength. They can still appear to
light as "atoms" and constitute an effectively homogeneous optical nano-
material. Structural units of such materials can be tailored for significant
excitation of higher-order multipoles, which makes it possible to achieve
optical properties far beyond the limitations of ordinary materials. Indeed,
optical nanomaterials can exhibit such peculiar phenomena as optical
magnetism [2–4] and strong spatial dispersion [5–7]. Furthermore, two
counterpropagating plane waves can experience different reflection coeffi-
cients, if they are incident at a nanomaterial slab that exhibits prominent
current quadrupole excitations [8]. Nanomaterials that are designed for
targeted optical responses can be called optical metamaterials. Their key
applications include sub-diffractionlimited optical imaging [9–15], optical
cloaking [16–19], and perfectly coherent light sources [20–22].
Due to spatial dispersion, optical waves in metamaterials experience the

1
Introduction

material parameters, such as refractive index and impedance, that depend


on the wave propagation direction [23]. These so-called "wave parameters"
have been conventionally retrieved from the reflection and transmission
coefficients of a metamaterial slab [24–29]. However, the retrieval proce-
dures that take into account the quadrupole and higher-order multipole
excitations were introduced only recently [8, 30, 31]. Thus, the wave pa-
rameters have become the design parameters for nanomaterials, whose
optical properties are relatively freely tailorable, e.g., through adjusting
the geometry of the material’s nanostructure [32].
When designing optical nanomaterials for practical applications, one has
to consider interaction of optical beams rather than plane waves with these
materials, either analytically or numerically. This is necessary, because
all light sources, especially lasers, generate beams instead of plane waves.
So far, purely analytical methods have been unable to treat interaction
of beams with spatially dispersive metamaterials. Therefore, numerical
methods have been used instead, the most important of which include Fi-
 
nite Element Method FEM and Finite Difference Time Domain method
 
FDTD . However, such methods do not provide a clear physical interpre-
tation of the results, e.g., since they treat metamaterials as inhomogeneous
media. Often, it is necessary to consider the material as effectively homo-
geneous to understanding its interaction with light. Another drawback of
these methods is that their computational load increases fast when the
size of the modelled interaction volume is increased beyond the scale of the
wavelength [33]. Therefore, new methods are required for treating large
nanomaterial structures.
Efficient methods for treating beam-metamaterial interaction are par-
ticularly important in the design of materials with engineered spatial
dispersion, e.g., diffraction-compensating metamaterials [14, 34–38]. In
the future, diffraction-compensating metamaterials can have multiple ap-
plications, e.g., in optical imaging and information processing. To achieve
diffraction compensation, one must adjust dispersion surfaces of the de-
signed material in a specific way. Up to now, the phenomenon of diffraction
compensation has been demonstrated in photonic crystals [39–42] and
metamaterials that are composed of metal-dielectric layers [12], metal
fishnet structures [35, 36], or effectively two-dimensional metal wires

2
Introduction

[14, 34, 43]. In photonic crystals, this phenomenon is called self-collimation.


It takes place at wavelengths close to the Bragg diffraction regime, and
is, therefore, accompanied by high reflectivity of the crystal’s surface.
Same kind of high surface reflectivity is typical for the above-mentioned
diffraction-compensating metamaterials as well [43]. Furthermore, meta-
materials of this type are usually highly absorptive at optical frequencies.
Low-loss optical metamaterials can facilitate design of integrated optical
components, such as diffraction-compensating slab waveguides, that can
be used in photonic micro chips. Current research on metamaterial slab
waveguides is focused on structures that employ anisotropic, negative-
 
index or hyperbolic metamaterials see, e.g., [44–50] . Theoretical models
of such waveguides do not usually consider spatial dispersion and can-
not, therefore, comprehensively guide the design of spatially dispersive
nanostructured waveguides.
The main goal of this thesis is to develop numerical and analytical meth-
ods that can be used to design spatially dispersive optical nanomaterials for
practical applications. As an example, we develop and then demonstrate
experimentally a metasurface in which electric-current quadrupole excita-
tions become prominent. The reflection coefficient of this metasurface de-
pends strongly on the illumination side. This experimental result matches
well with the multipole theory used in the design. In addition, we have
developed a semianalytical method that can be used to treat interaction of
optical beams with almost any spatially dispersive material. Compared to
typical numerical methods, our method is computationally light and can
provide clearer understanding of the light-matter interaction effects. For
example, the method predicts that spatial dispersion can cause conversion
of light’s polarization upon propagation. Furthermore, we use the method
to design diffraction-compensating metamaterials that can be impedance
matched to the surrounding medium and exhibit very low absorption. To
be able to design and characterize spatially dispersive slab waveguides,
we develop a theoretical model, that makes it possible to evaluate the
refractive index and impedance experienced by the waveguide modes. We
use this new model to design a broadband diffraction-compensating meta-
material slab waveguide that can be used in optical chips operating at THz
modulation frequencies.

3
Introduction

The thesis is organized as follows. Chapter 2 presents the theoretical


background of the work done in the thesis and illustrates how the presented
theoretical tools can be used to design a metasurface with prominent mag-
netic dipole and electric quadrupole excitations. Chapter 3 describes new
methods for treating interaction of light with spatially dispersive struc-
tures. The chapter also presents design principles for novel diffraction-
compensating metamaterials and metamaterial slab waveguides. Also
some key examples of such materials and waveguides are presented. Chap-
ter 4 summarizes the thesis.

4
2. Theoretical description of optical
nanomaterials

2.1 Metal nanoparticles

In various branches of science, there has been a growing interest towards


metal nanoparticles. Such particles have a large diversity of shapes and
can be composed of many different materials. As an example, Fig. 2.1 shows
   
a nanosphere made of copper Fig. 2.1a , a silver nanorod Fig. 2.1b , an
   
aluminum nanodisc Fig. 2.1c , and a gold nanodimer Fig. 2.1d . These
particles have very distinctive optical scattering and absorption properties.
A large variety of metal nanoparticles, including the examples in Fig. 1, can
be fabricated by well-established nanofabrication methods, such as direct
chemical synthesis and electron-beam lithography described in [51–58].
Applications of these particles range from microscopy [59,60] and medicine
[61–64] to solar energy harvesting [65–68] and optical waveguides [69–71].
Some of the most important applications are cancer treatment [62], killing
antibiotic-resistant bacteria [64], purification of water [68], enhancement
of efficiency of solar cells [72], and artificial nanomaterials with designed
optical properties [73].
To the greatest extent, these applications stem from specific properties
of metals, such as large density of conduction electrons compared to other
materials. These "free" electrons in metals can to a good approximation
be treated as plasma in terms of the Drude model, from which optical
properties of bulk metals can be derived [74]. However, real conduction
electrons can undergo interband transitions that are not accounted by the
Drude model, but can have significant influence on the metal’s characteris-

5
Theoretical description of optical nanomaterials

 
a b
Cu Ag

 
c d

Al Au

   
Figure 2.1. Examples of metal nanoparticles: a a copper sphere, b a silver nanorod,
   
c an aluminum disc, and d a gold nanodimer.

tics. For this reason, it is recommended to use experimental values for the
optical parameters of metals. In the thesis, we use the data of [75] for the
refractive-index spectra of silver and gold.
At optical frequencies, the long-range nature of the Coulomb interaction
leads to collective behaviour of electrons, which, under proper conditions,
gives rise to wave-like excitations of the electron plasma and the cor-
responding quasiparticles that are called plasmons [76]. Longitudinal
oscillations of the bulk electron plasma, called volume plasmons, cannot be
excited by light, and are typically created by electron bombardment of the
metal. On the other hand, transverse oscillations of electron plasma can
be excited by electromagnetic waves at a metal-dielectric interface. Such
coherent field-plasma oscillations are called surface plasmon polaritons
 
SPPs . Typically, SPPs do not match the wavelengths of the exciting opti-
cal waves in free space and are excited by evanescent waves [74, 77, 78] or
using phase-matching diffraction gratings [79, 80]. The third type of plas-
 
mons that we use in the thesis are the localized surface plasmons LSPs of
metal nanoparticles. The curved surface of the particle effects the electron
density and makes it possible to excite localized surface plasmons directly
by propagating waves [81].

6
Theoretical description of optical nanomaterials

Efficient excitation of a LSP requires that the incident light frequency


essentially coincides with the resonance frequency of the plasmon. The
 
localized surface plasmon resonance LSPR differs from the Fabry-Perot
and Mie standing-wave resonances by the resonant oscillation of the elec-
tron plasma, which results in particularly strong near-fields around the
nanoparticle. Amongst all possible geometries of nanoparticles, calculation
of the LSPR frequencies is most straightforward for metal spheres that
are much smaller than the wavelength of light interacting with them. The
spheres exhibit only one major resonance, at a frequency that satisfies
the Fröhlich resonance condition, Re{εm (ω)} = −2εs (ω), where εm (ω) and
εs (ω) are the frequency dependent electric permittivities of the metal and
the surrounding dielectric, respectively [79]. More complex particles can
support many plasmon modes with different resonance frequencies. For
example, ellipsoidal metal nanoparticles support three plasmon modes
that oscillate along the three principal axes of the ellipsoid. This type
of modes also appear in cuboidal particles, such as nanorods. Resonance
frequencies of such modes depend not only on the surrounding material,
but also on the particle’s aspect ratio. Each of the modes can be excited
only by an electric field component that oscillates along the corresponding
symmetry axis of the particle, which makes the particle’s optical response
depend on the polarization of the incident light.
Efficient excitation of localized surface plasmons results in significant
scattering and absorption of the incident light by the particle. For par-
ticles that are significantly smaller than the incident light wavelength,
the scattered far-field is similar to the far-field of an electric dipole. For
larger particles, higher-order multipole excitations become important and
the far-field is given by a sum of vector spherical harmonics, which for
spherical particles was first proposed by Mie [82]. Formally, the interaction
of a nanoparticle with light can be characterized by the scattering and
absorption cross-sections, defined by

Pscat
Cscat = , (2.1)
Iin

and
Pabs
Cabs = , (2.2)
Iin
respectively, where Pscat and Pabs are the scattered and absorbed optical

7
Theoretical description of optical nanomaterials

2.4

1.6
.6

Cscat 10−13 m2

0.8
.8

0
400 700 1000 1300
λ (nm)

Figure 2.2. Numerically calculated scattering cross-section spectra for three silver
nanorods of different aspect ratios. The black arrows show the direction
of the plasmon oscillation responsible for the scattering. The nanorod di-
mensions are 40 × 40 × 90 nm3 (green line), 30 × 30 × 90 nm3 (red line), and
30 × 30 × 120 nm3 (blue line).

powers and Iin is the incident light intensity. These cross-sections are
strongly peaked at the particle’s LSPR frequencies. Figure 2.2 shows the
spectra of Cscat for three silver nanorods interacting with light polarized
along the rods. The rods have different height-to-width aspect ratios,
and therefore, different LSPR frequencies. The increase of the LSPR
wavelength with the aspect ratio allows one to shift this wavelength to
a desired value by tuning the nanoparticle’s geometry. This property
was used in the design of a diffraction-compensating metamaterial of
Publication III.

2.2 Multipole expansion

Multipole expansion is useful in the mathematical description of electro-


magnetic fields produced by localized sources [83]. As a rule, the expan-
sion is written in spherical coordinates as a series of orthogonal angle-
dependent eigenfunctions. A clear benefit of using such an expansion is

8
Theoretical description of optical nanomaterials

that its higher-order terms have smaller amplitudes, which allows one
to truncate the series and thus simplify the representation. This usually
helps one to obtain an intuitive picture of the described fields and their
sources, as well as characterize local light-matter interaction effects. For
example, the multipole expansion can be used to analytically calculate
potential energies of charge configurations [84] and scattering and absorp-
tion cross-sections of nanoparticles [32]. In addition, the eigenfunction
coefficients in the expansion determine the multipole moments and the
electric current excitations in the particle producing the field. This enables
one to design and characterize optical nanoantennas, whose scattering and
absorption properties are tailored for specific applications [32].
For a single nanoparticle scattering an electromagnetic field in a di-
electric host medium, the scattered electric and magnetic fields can be
expanded as

 ∞  l   
1 (1)
E (r, θ, ϕ) = E0 aE (l, m) ∇ × hl (kr) Xl,m (θ, ϕ)
k
l=1 m=−l

(1)
+aM (l, m) hl (kr) Xl,m (θ, ϕ) [π (2l + 1)]1/2 il , (2.3)
∞   
E0   1
l
(1)
H (r, θ, ϕ) = aM (l, m) ∇ × hl (kr) Xl,m (θ, ϕ)
η k
l=1 m=−l

(1)
+aE (l, m) hl (kr) Xl,m (θ, ϕ) [π (2l + 1)]1/2 il−1 , (2.4)

(1)
respectively, where Xl,m are the vector spherical harmonics and hl are
the spherical Hankel functions of the first kind [32]. The particle is located
at r = 0. In Eqs (2.3) and (2.4), E0 is the complex amplitude of the
incident plane wave and η and k are the wave impedance and wave number
in the host medium, respectively. The multipole coefficients aE and aM
depend on the particle geometry and the properties of the incident plane
wave. For a spherical particle, these coefficients and moments can be
calculated analytically, as proposed by Mie [82]. However, for nontrivial
particle geometries, there are no analytical solutions, and the multipole
moments must be calculated numerically from the scattered fields by using
the orthogonality of the functions Xl,m [32]. In the presence of multiple
nanoscatterers, the multipole moments of each particle cannot be found
from Eqs (2.3) and (2.4), because the fields scattered by the individual
nanoparticles cannot be extracted from the overall scattered field. However,

9
Theoretical description of optical nanomaterials

  
a b c
s
s

dipole quadrupole octupole


   
Figure 2.3. Examples of current multipoles: a an electric-current dipole, b a current
 
quadrupole, c a current octupole.

one can calculate the coefficients aE and aM from the total electric field
distribution E (r) inside each nanoparticle by using the so-called scattering
current density [85]

JS (r) = −iωε0 [εr (r) − εh ] E (r) . (2.5)

Here, ω is the angular frequency, εr (r) is the inhomogeneous dielectric


function, and εh is the dielectric constant of the host medium. In Eq. (2.5),
JS (r) is equal to 0 outside the nanoparticles. As shown in [32] and [85], aE
and aM can be written as

(−i)l−1 k 2 ηOl,m  

aE (l, m) = 1/2
ψl (kr) + ψl (kr) Plm (cos θ) r̂ · JS (r)
E0 [π (2l + 1)]
1   
+ ψl (kr) τl,m (θ) θ̂ − iπl,m (θ) ϕ̂ · JS (r) e−imϕ d3 r, (2.6)
kr
(−i)l+1 k 2 ηOl,m  
aM (l, m) = 1/2
e−imϕ jl (kr) iπl,m θ̂ + τl,m (θ) ϕ̂
E0 [π (2l + 1)]
· JS (r) d3 r, (2.7)

where jl are the spherical Bessel functions, ψl are the Riccati-Bessel func-
tions, and the prime stands for the derivative with respect to kr. The
functions Ol,m , πl,m , and τl,m are defined as in [32]:
1/2
1 2l + 1 (l − m)!
Ol,m = , (2.8)
[l (l + 1)]1/2 4π (l + m)!
d m
τl,m = P (cos θ) , (2.9)
dθ l
m m
πl,m = P (cos θ) . (2.10)
sin θ l

In Eqs. (2.6) and (2.7), the integrals are taken over the volume of each
individual nanoparticle and can be carried out numerically.

10
Theoretical description of optical nanomaterials

= +

aM (1, 0) Qyx -Qxy


Figure 2.4. Decomposition of a magnetic dipole into two current quadrupoles. The complex
amplitude of the magnetic dipole is equal to the aM (1, 0) coefficient in the
multipole expansion of the electromagnetic field. The complex amplitudes of
the current quadrupoles are denoted by Qyx and Qxy .

In analogy with the multipole expansion of the electrostatic potential, in


which each term comes from a configuration of point charges [84], the terms
in Eqs. (2.3) and (2.4) can be represented by orthogonal configurations of
point-like electric currents that can be called electric-current multipoles
[32]. Then, the scattering current JS (r) can be written as a sum of these
multipoles that together produce the same fields as that radiated by JS (r).
The current elements of some of these multipoles are shown schematically
in Fig. 2.3. Note that the fields produced by the current multipoles are
in general not orthogonal, although some of them are exactly the same
as those of classical multipoles, such as the electric dipole and linear
quadrupole. Also some of the classical multipoles can be decomposed into
the current multipoles, as shown in Fig. 2.4.
The current multipole expansion is particularly useful in the design
of nanoparticles with desired scattering properties, because the current
multipole excitations are easy to visualize, and as a result, also to realize
in practice. Indeed, these multipoles are simply rectangular collections of
electric dipoles. Electric dipole excitations prevail in subwavelength-size
nanoparticles, such as metal spheres or nanorods. Higher-order multipoles
can be efficiently excited in more complex particles, such as dimers. The
phases and amplitudes of the currents are adjustable by changing the
particle geometry and consequently also the multipole resonance frequency.
In principle, any current configuration that can be decomposed into the
current multipoles can be created and used to control the incident field.
The current multipoles are indexed by the arrangement of their linear
 
dipolar current elements. For example, an arbitrary current quadrupole
Qij is characterized by two indices {i, j} ∈ {x, y, z} that define a pair

11
Theoretical description of optical nanomaterials

of i-directional dipoles separated in the j-direction by an infinitesimal


 
distance s see Fig. 2.3 . The indexing of octupoles and higher-order
multipoles is done similarly [32]. Hence, the current multipoles are simple
configurations of linear dipolar electron oscillations determined by the
multipole indices. Furthermore, the amplitudes of these easily visualizable
current multipole moments are the elements of the current multipole
(l)
tensor M defined as
(l) i
M = J (r) rl−1 d3 r, (2.11)
ω (l − 1)!
where ω is the angular frequency, l is the multipole order, and J is the
complex amplitude of the current density in the particle [32]. In terms of
(l)
the elements of M , the multipole expansion of the current density J is
obtained as

   
l−1 l−a−1
∂ a ∂ b ∂ l−(a+b+1)
J (r) = v̂A (v̂, l, a, b) δ (r) , (2.12)
∂xa ∂y b ∂z l−(a+b+1)
l=1 v̂=x̂,ŷ,ẑ a=0 b=0

(−1)l (l − 1)!
A (v̂, l, a, b) = iωM (l) (v̂, a, b) , (2.13)
a!b! [l − (a + b + 1)]!
(l)
where M (l) (v̂, a, b) are the elements of M composed of v̂ oriented dipoles;
the number of dipoles in the x-direction is a + 1, in the y-direction b + 1,
and in the z-direction l − a − b [32]. The multipole coefficients aE and aM
(l) 
of a scattered field are linear combinations of the elements of M see

Eqs. (35)-(48) of [32] . Therefore, if these coefficients are known, one can
(l)
calculate the elements of M , which will reveal the amplitudes of the
current multipole excitations responsible for the scattering. Note however,
that the set of classical multipoles is incomplete, as it misses the toroidal
multipoles, while the set of current multipoles is complete. In general, the
(l)
number of elements of M is larger than the number of coefficients aE and
aM by the number of possible current excitations that do not radiate. Hence,
the expression of aE and aM in terms of current multipole moments is to
some extent arbitrary, as discussed in [32]. Typically, nanoparticles that
scatter light are designed to exhibit only a few low-order current multipole
resonances, the other ones being negligible, which makes calculation of
(l) (l)
M from aE and aM simple. Unlike M , the classical multipole tensor
[84, 85] does not explicitely contain the information on the amplitudes of
the currents in the nanoparticles. This is because the elements of this

12
Theoretical description of optical nanomaterials

Figure 2.5. An electric field distribution around a radiating current quadrupole excited
in a pair of silver discs. The black and red arrows show the directions of the
electric field and the overall scattering currents in the discs, respectively, at a
fixed instance of time.

tensor are proportional to the amplitudes of the electric charge density,


rather than the current density. However, the charge density cannot be
unambiguously related to the current density in the nanoparticle, but
rather, different current densities can result in the same charge density.
The elements of the classical multipole tensor are often difficult to associate
with some readily visualizable charge oscillations.
Two metal nanoparticles placed close to each other exhibit a pure current
quadrupole excitation if their linear scattering currents are equally strong
and oscillate out of phase. Such current quadrupoles usually have a narrow
resonance bandwidth, which is beneficial to some applications. Because
the fields radiated by the two current elements of a current quadrupole
partially cancel each other, the radiated power is low compared to that of an
electric dipole with the same amplitude of current. Therefore, molecules or
other quantum emitters that are strongly coupled to a current quadrupole
do not radiate efficiently, for which reason, their excited-state lifetimes
 
can be long see, e.g., [86] . The electric field distribution radiated by a
current quadrupole excitation in a pair of silver discs is shown in Fig. 2.5.
If this quadrupole field is superposed with an equally strong field radiated
by an electric dipole that oscillates in phase with the quadrupole, the

13
Theoretical description of optical nanomaterials


a

z
y x

xz-plane yz-plane
 
b 90o 3 c 90o 3
120o 60o 120o 60o
2 2
150o 30o 150o 30o
1 1

180o 0 180o 0
I/Iav I/Iav
210o 330o 210o 330o

240o 300o 240o 300o


270o 270o
 
Figure 2.6. The directivity of radiation scattered by the disc dimer shown in a . The
dimer consists of two 10 nm thick silver discs with 30 nm and 40 nm diameters
and a 10 nm surface-to-surface separation. The dimer is excited by a x-
polarized plane wave with a wavelength of 591 nm. The polar plots show the
scattered intensity at an arbitrary distance from the source divided by its
average over all directions. The scattered intensity is shown in the xz-plane
   
in b and in the yz-plane in c .

overall radiation can disappear in one of the directions due to destructive


interference. This explains directional radiation by current densities that
can be decomposed into electric-dipole and current-quadrupole elements of
comparable magnitudes [58]. As an example, the polar plots in Figs. 2.6b
and 2.6c show the radiation directivity diagrams of a silver disc dimer of
Fig. 2.6a in two orthogonal cross-sectional planes. The dimer is excited by
light at a wavelength of 591 nm produces a high directivity of the scattered
radiation. The dipole and quadrupole fields are seen to cancel each other
quite well at the scattering angle of 270◦ , that corresponds to the negative
z-direction. If quantum emitters are coupled to such disc dimers, their
emission can become directional [87]. This can be used, e.g., in optical
nanoantennas that, due to optical reciprocity, mostly absorb light coming
from the same directions in which they emit. Because of such directivity,
novel optical metamaterials composed of disc dimers are expected to exhibit

14
Theoretical description of optical nanomaterials

optical properties that depend on light propagation direction [8, 31, 58, 88].

2.3 Metasurfaces

Optical metasurfaces are nanopatterned surfaces, the structural units of


which are separated by subwavelength distances. Subwavelength gratings
and two-dimensional arrays of closely spaced nanoparticles are simple
examples of such metasurfaces. Although metasurface structures are
thin compared to the wavelength, light can experience drastic changes
in its phase and polarization upon its reflection or transmission at these
structures [89]. This can be used in a wide variety of applications, such as
metasurface-based holograms [90], low-aberration high-numerical-aperture
lenses [91], and ultrathin polarizers [92, 93] and waveplates [94]. Another
important application of phase-shifting metasurfaces is the flexible gen-
eration of optical beams with desired wave fronts, such as Bessel-like
beams [95] and beams that carry orbital angular momentum [96]. In fact,
most of the phase-shifting metasurfaces are simply additionally structured
diffraction gratings, whose major advantages over the conventional grat-
ings are the operation in a single diffraction order and its widely tunable
optical properties. Furthermore, some metasurfaces are designed to ex-
hibit hyperbolic dispersion that, for optical emitters coupled to the surface,
can lead to a significant enhancement of their emission, which can be used
to create more efficient light sources [97]. The same phenomenon can be
used to enhance optical absorption as well. If the metasurface is composed
of metal nanoparticles, the near-field coupling of them to surrounding
molecules can influence their plasmon resonance frequencies, which can be
used for applications in optical sensing [89, 98]. Such a plasmonic sensor
detects even minor concentrations of the molecular species of interest, and
in some cases even individual molecules [99]. Moreover, if the plasmon res-
onance frequencies of the nanoparticles coincide with the lattice resonance
frequency of the metasurface, the resonant near fields can become spec-
trally narrower and further enhanced, which increases their sensitivity to
the surroundings [100, 101]. Compared to individual nanoparticles, meta-
surfaces can exhibit much stronger interaction with light, which can also
be used to enhance optical nonlinearities [99]. In regard to enhancement

15
Theoretical description of optical nanomaterials

Au
PC403 SiO2

Figure 2.7. A gold nanodimer metasurface. Each dimer is composed of two 20 nm thick
gold discs, separated by a 70 nm distance. The diameters of the discs are
55 nm and 110 nm. The dimers are arranged in a two-dimensional lattice
on a SiO2 substrate. The lattice constant is Λ = 190 nm. The particles are
 
embedded in a polymer layer PC403 that extends 100 nm above the dimers.
The inset shows a scanning electron micrograph of a part of the fabricated
metasurface.

of optical absorption, metasurfaces are ideal candidates for applications


in thin-film solar cells, providing a longer light-matter interaction time.
Compared to the corresponding three-dimensional metamaterials, meta-
surfaces are more straightforward to fabricate and integrate on a "chip"
with other optical elements.
Metasurfaces are, as a rule, spatially dispersive, which means that the
effective multipole polarizabilities of their unit cells depend on the wave
propagation direction. This can be used to manipulate the amplitude,
phase, and polarization of the angular-spectrum components of light in-
dividually. The possibility to utilize spatial dispersion of metasurfaces in
optical beam engineering has not yet attained much research attention.
However, in the future, they will definitely find numerous applications in
such optical components as ultrathin planar lenses, advanced antireflec-
tion coatings, and apertureless spatial filters, which are difficult to realize
using traditional technologies.
In Publication II, we propose a new type of metasurface in which electric-
current quadrupole excitations can prevail the dipole excitations and lead
to strong spatial dispersion. The metasurface is a two-dimensional array
 
of gold nanodimers embedded in a layer of polymer PC403 on a SiO2

16
Theoretical description of optical nanomaterials

(a) Reflectance (b) |px | |kQxz |


0.6 10
8
0.4 6
4
0.2
2
0 0
(c) Transmittance (d) arg(ipx ) arg(kQxz )

0.8 3π
2
0.6
0.4 π

0.2 π
0 2
550 650 750 850 950 550 650 750 850 950
wavelength [nm] wavelength [nm]
Figure 2.8. Wavelength-dependent properties of the metasurface of Fig. 2.7. The power
   
reflectance and transmittance spectra are shown in a and c . The solid lines
and dots represent numerically calculated and measured results, respectively.
     
In figures a , b , and d , the red and blue colors represent the two directions
of incidence that are shown by the red and blue arrows in the insets. The
numerically calculated multipole terms ipx and kQxz contributing to the
 
reflection and transmission coefficients of the metasurface are shown in b
 
and d by the solid and dashed lines. The magnitude and phase of these
     
terms are shown in b and d , respectively. The units in b are given in
10−32 Cm.

substrate as shown in Fig. 2.7. The structure of the metasurface is simple


and relatively easy to fabricate by conventional nanofabrication methods.
To experimentally demonstrate its properties, we have fabricated the
structure using the electron beam lithography and spin coating techniques
 
see Publication II for the fabrication details . The scanning electron
micrograph of a segment of the fabricated metasurface is shown in the
inset of Fig. 2.7.
The numerically calculated and measured power reflectance spectra of
the metasurface are shown in Fig. 2.8a by solid lines and dots, respectively.
The reflectances at illumination from the larger- and smaller-disc sides of

the metasurface are depicted by the blue and red colors, respectively see

the inset . The reflectance clearly depends on the illumination direction,
which is as expected, since the dimers are asymmetric. In fact, many asym-

17
Theoretical description of optical nanomaterials

metric structures, such as stacked pairs of highly reflective and absorptive


films, exhibit side-dependent reflectivity, but usually, these structures are
thick compared to the wavelength. Note that the reflection and transmis-
sion of the metasurface has a large contribution from light scattered by
the nanodimers. The scattering characteristics of the dimers are deter-
mined by their multipole excitations. The side-dependent reflectivity of the
metasurface is in principle caused by the illumination-dependent excitable
multipole modes of the dimers that radiate primarily in the direction of
the larger disc. However, due to optical reciprocity, the forward scattering
must be the same for the two excitation directions, and therefore, the
excitation efficiencies of such modes must depend on the direction of light
propagation. For the nanodimers illuminated by an x-polarized plane
wave, the only current multipoles that can be efficiently excited are electric
dipoles, with the moments px , and quadrupoles, with the moments Qxz ,
oscillating in the planes of the discs. At normal incidence, the plane-wave
reflection and transmission coefficients for a single layer of the particles
can then be written in terms of px and Qxz as [8]

k0
ρ= (ipx − k0 nh Qxz ) , (2.14)
2ε0 nh Λ2 E0

and
k0
τ =1+ (ipx + k0 nh Qxz ) , (2.15)
2ε0 nh Λ2 E0
respectively. Here, Λ is the lattice period, E0 is the complex amplitude of
the incident plane wave, k0 is the wave number in vacuum, and nh is the
 
refractive index of the material surrounding the particles k = k0 nh . In
the experiments, these coefficients are extracted from the overall reflection
and transmission coefficients of the fabricated sample, taking into account
the reflection and transmission by the polymer-air interface. As follows
from Eq. (2.14), the reflectance of the metasurface is minimized, when the
condition ipx = k0 nh Qxz is satisfied. This means that the fields scattered
back by the dipole and quadrupole excitations cancel each other. For the
designed metasurface, this cancellation takes place when the sample is
illuminated from the smaller-disc side at a wavelength of 675 nm, around

which the measured reflectance is significantly suppressed see the red

curves in Fig. 2.8a . The spectra of the magnitudes and phases of px and
Qxz , shown in Figs. 2.8b and 2.8d, confirm that at this wavelength the

18
Theoretical description of optical nanomaterials

 16 b
a ↑ |E| /E0 ↑ |H| /H0
4

0 0
 8 
c |E| /E0 d |H| /H0

↑ 0
↑ 0
Figure 2.9. The near-field distributions of a nanodimer under plane-wave illumination.
The arrows show the direction of incidence. The electric field intensity is
   
shown in a and c for the opposite illumination directions. The magnetic
   
field intensity is shown in b and d . All the distributions are normalized to
the intensity of the incident plane wave.


condition ipx = k0 nh Qxz is indeed quite well satisfied see the red curves

dash-crossing the vertical dotted lines . These spectra are obtained by
solving Eqs. (2.14) and (2.15) for the multipole moments. These results
imply that, taking into account also the reflection and transmission of light
by the boundary between the substrate material and glass, one can design
a metasurface that acts as an ultrathin antireflection coating.
Figure 2.8c shows the power transmittance spectrum of the metasur-
face. The transmittance is seen to be relatively low in the wavelength
region where |px | and |kQxz | are high, which is associated with the high
absorption due to the strong electric current excitations in the nanodimers.
The transmittance does not depend on the illumination side, which is a
consequence of optical reciprocity. The Lorentz reciprocity theorem implies

that the extinction cross-section Cext = Cscat + Cabs see Eqs. (2.1) and

(2.2) of a nanoparticle is the same for two opposite directions of light
propagation [102, 103]. This means that also the power removed from
the incident light by the particle is the same. According to the optical

19
Theoretical description of optical nanomaterials

theorem, this power is determined from the far-field intensity behind the
nanoparticle, i.e., from the transmitted light. Hence, the transmittance of
an individual dimer, as well as the whole metasurface must be independent
of the illumination side.
The absorptance A of the metasurface satisfies the enegy conservation
requiring that R + T + A = 1, where R is the reflectance and T is the
transmittance of the metasurface. Because R depends on the illumination
side, but T does not, also A must depend on the side of illumination.
Moreover, higher absorption is connected to stronger current excitations
and the associated ohmic losses in the particles. Because stronger currents
usually create stronger near-fields, the side dependent absorption should
be accompanied by side-dependent local field enhancement. The near-field
distributions of the electric and magnetic field magnitudes around the
nanodimers are shown in Fig. 2.9. The shapes and peak values of these
distributions clearly depend on the illumination direction.
The near-field patterns around the dimers can to some extent be approx-
imated by the field patterns of two radiating electric dipoles located at
the positions of the discs. For two dipoles oscillating out of phase, the
electric field is partially canceled in the region between the dipoles, and
the magnetic field in that region is enhanced, as can be seen in Figs. 2.9a
and 2.9b. In contrast, for two dipoles oscillating in phase, the electric field
is enhanced at the center of the dimer and the magnetic field shows a
 
minimum see Figs. 2.9c and 2.9d . The former configuration of the two
dipoles corresponds to the electric-current quadrupole excitation and the
later configuration to the prevailing electric dipole excitations. It is seen in
the figure that the quadrupole moment is excited efficiently only by light
incident from the smaller-disc side. Because the quadrupole modes are not
very efficient light scatterers, the absorption and local enhancement of the
electric field by the particles are significant, as seen also in Fig. 2.9. By
optimizing the structure of the metasurface, the near-field enhancement
can possibly be made much stronger for the smaller-disc illumination side,
which can be used, e.g., for solar energy harvesting. In addition, such
spatially dispersive metasurface can have applications in photodetectors
that are sensitive to light propagation direction.

20
Theoretical description of optical nanomaterials

2.4 Metamaterials

An optical metamaterial is a designed three-dimensional material whose


structural units are subwavelength-sized. Periodic three-dimensional
arrays of closely packed nanostructured particles, such as the one shown
in Fig. 2.10a, are typical examples of optical metamaterials. Since the
materials are three-dimensional, they can be used for applications based
on wave propagation, amplification, and attenuation in the medium. A
bright example of such applications is a negative-index metamaterial
slab used as a perfect lens. The resolution of this lens is not limited by
diffraction and makes use of the contribution of the evanescent waves in
the image [9]. Other useful examples include three-dimensional slow-light
applications [104], unique applications of transformation optics [105], such
as invisibility cloaking [106], and certain superior optical imaging devices,
such as a hyperlens [107]. In addition, optical metamaterials can be used
to create smaller, deeply subwavelength resonators that are difficult to
realize using conventional materials [108].
Even though the ordinary optical materials, such as glass, are treated as
homogeneous media, they are composed of atoms and molecules, and thus,
microscopically, they are inhomogeneous. Metamaterials are not very dif-
ferent and can also be treated as effectively homogeneous media, using the
so-called effective material parameters. These parameters appear in the
macroscopic Maxwell equations describing spatially averaged electric and
magnetic fields. Many widely known effective-medium theories consider
composites, in which small particles are embedded in a transparent host
medium. In an external electric field, Eext , the particles are polarized and
acquire electric dipole moments p = αEext , where α is the polarizability of
the particle. Near-fields of the neighbouring particles do not essentially
contribute to p, since when integrated over the particle volume, they cancel
each other [109]. By using p, one can calculate the polarization P = np
of the composite, where n is the volume density of the polarized particles.
This macroscopic quantity is then equated to that of the effective medium,
i.e.,
P ≡ nαEext = (ε − εh ) E. (2.16)

Here Eext is the macroscopic field external to the dipoles, εh is the electric

21
Theoretical description of optical nanomaterials

 
a b

k1
n1 , Z 1
k2
E1
E2
n2 , Z 2

Figure 2.10. A spatially dispersive metamaterial and its equivalent effective medium.
 
The three-dimensional microscopic structure of the material is shown in a .
The horizontal gray sheets represent essentially two-dimensional layers of
nanorods that continue infinitely far in the directions indicated by the black
 
dots. The equivalent effective medium is shown in b . The two plane waves
 
in b experience different refractive indices and impedances, ni and Zi , even
 
though their electric fields Ei point in the same direction i ∈ {1, 2} .

permittivity of the host medium, ε is the effective permittivity of the


composite, and E is the spatial average of the overall electric field. This
equation can be solved for ε by noticing that E is directly proportional to
Eext , which can be shown by using the Lorentz local field correction [109].
The solution, known as the Lorentz formula, is
nα/εh
ε = εh 1 + . (2.17)
1 − (1/3) nα/εh
Alternatively, one can write α as a function of ε and obtain the well known
Clausius-Mossotti relation. In Eq. (2.17), the unknown parameter α is
often approximated by the quasistatic polarizability of a sphere, which
yields the Maxwell-Garnett mixing formula
1 + 2f (εp − εh ) / (εp + 2εh )
ε = εh . (2.18)
1 − f (εp − εh ) / (εp + 2εh )
Here εp is the electric permittivity of the particles and f is their volume
fraction in the composite. Equation (2.18) is usually valid even if the
particles are not spherical, provided that the quasistatic approximation is
not violated. The theory can thus be generalized to anisotropic media [109].
The equation can be further generalized to composites of many particle
types, in which case

ε − εh  N
εn − εh
= fn , (2.19)
ε + 2εh εn + 2εh
n=1

22
Theoretical description of optical nanomaterials

where εn is the permittivity of the n:th type of the particles and fn is


its volume fraction. However, Eq. (2.19) is limited to relatively disperse
nanocomposites, because it breaks down, if the combined volume of the
particles approaches the volume of the host medium. In such cases, one
can solve for ε from a more suitable Bruggeman mixing formula that does
not treat the host medium separately:


N
εn − ε
fn = 0, (2.20)
εn + 2ε
n=1

where

N
fn = 1. (2.21)
n=1

In fact, all the aforementioned mixing formulas have been generalized to


anisotropic and magnetic effective materials [109–112]. However, none
of these formulas is applicable to materials in which the structural units
exhibit quadrupole or higher-order multipole excitations. In addition, ho-
mogenization of a periodic material is only meaningful if all higher-order
Bloch modes, except the fundamental, can be neglected. Due to these
constraints, the above effective-medium theories only work for composite
materials with the unit-cell size much smaller than the wavelength. In
some relatively novel effective-medium theories, the effective material
parameters are derived by analyzing spatial averages of numerically cal-
culated field distributions [113–115]. Even though such theories are less
constrained than the Maxwell-Garnett and Bruggeman theories, they are
still limited to materials whose unit cells are clearly smaller than the
wavelength.
Spatial dispersion is a common property of optical metamaterials, since
the structural units of these materials are typically large enough to support
higher-order multipole modes. Excitation of such modes appears in the
Fourier-transformed macroscopic Maxwell equations

k × E = k0 n (k) Z (k) H, (2.22)


k0 n (k)
k×H=− E, (2.23)
Z (k)
k · E = 0, (2.24)

k · H = 0, (2.25)

23
Theoretical description of optical nanomaterials

as the k-dependence of the refractive index n (k) and the impedance Z (k).
Here, the wave parameters n (k) and Z (k) are associated with a plane wave
characterized by an electric field E, a magnetic field H, and a wave vector k
in a homogenized isotropic spatially dispersive medium. In general, these
parameters also depend on the polarization of the wave. Alternatively, one
could use the k-dependent permittivity and permeability, ε (k) and μ (k), as
the wave parameters. Even though the most of the realistic metamaterials
are anisotropic, such materials can often be treated as effectively isotropic,
which means that Eqs. (2.22)-(2.25) can still be used. For example, when
calculating the reflection and transmission coefficients of an anisotropic
slab, one can always find an equivalent isotropic material that yields the
same coefficients as the anisotropic one for any slab thickness. Optical
phenomena tightly linked to optical anisotropy in spatially dispersive
materials can be treated by using the theory of [116].
For en electrically anisotropic material with no spatial dispersion, the
k-dependence of n (k) and Z (k) is entirely associated with the rotation of
the E vector when the wave propagation direction is changed. By contrast,
in the presence of spatial dispersion, these parameters depend on the
 
direction of k even if the electric-field direction is fixed see Fig. 2.10b .
In a spatially dispersive isotropic medium the electric displacement field
is given by D (k) = ε (k) E (k). According to the convolution theorem,
multiplication in the Fourier space corresponds to a convolution in the real
space, i.e.,
   
D (r) = ε r − r E r  d 3 r  , (2.26)

as also noted in [1]. This equation illustrates that in a spatially dispersive


material the polarization density at r is determined by the electric field
values at multiple points r , which means that the optical response is
nonlocal. Such nonlocality arises in realistic metamaterials due to the
near- and far-field coupling of different parts of the material, that leads
to mutual dependence of the polarization densities in these parts. Similar
coupling would also occur in conventional optical materials if the overall
contribution of these fields to the polarization density would not cancel
[109, 117]. The canceling effect is usually due to the dipolar character
of the fields and, thus, does not occur under the excitation of higher-
order multipoles. If a realistic metamaterial is replaced by an equivalent

24
Theoretical description of optical nanomaterials

homogeneous material, different parts of the material must still be coupled,


which is only possible if the equivalent material is spatially dispersive.
Other possible sources of spatial dispersion that are not considered in this
thesis include diffusion of charge carriers and various quantum phenomena
[118].
Characterization of an optical metamaterial by wave parameters is not
particularly useful, if the material does not support polarization eigen-
modes that would experience these parameters. Such modes preserve
their polarization state upon propagation. In Publication I, we show that
in many materials, for almost any propagation direction, no plane wave
can preserve its polarization, which is due to the combined influence of
spatial dispersion and anisotropy on the wave polarization. Therefore, one
cannot really introduce polarization eigenmodes and the associated wave
parameters for such materials. This polarization conversion by spatial
dispersion is a new phenomenon and differs from, e.g., spatial-dispersion-
induced birefringence [119], since for birefringent materials, polarization
eigenmodes can be introduced.
For a metamaterial that can be characterized by well-defined n and Z,
the wave parameters can be retrieved from the reflection and transmission
coefficients, r and t, of a slab of this material [24, 116]. These coefficients
can be calculated, e.g., numerically for a slab illuminated by a plane wave.
Under such illumination, the fields scattered by the nanoconstituents
of the material sum up to reflected and transmitted plane waves. The
light scattered into the metamaterial forms two fundamental Bloch modes
that propagate back and forth between the slab boundaries. Close to the
boundaries, also other modes can appear, but their amplitude decays fast
away from the interfaces. To be able to retrieve the wave parameters from
r and t, one must relate these coefficients to n and Z, which can be done by
writing r and t as functions of the reflection and transmission coefficients of
the slab boundaries. These interface coefficients must, then, be derived as
functions of the wave parameters. This is done, as a rule, by applying the
continuity conditions for the tangential electric and magnetic fields. One
might think that Fresnel coefficients cannot be derived for metamaterial
interfaces, since the tangential field at the interface includes contributions
from the decaying modes. However, as shown in [120], the decaying modes

25
Theoretical description of optical nanomaterials

  
a b c
1

|E| (a.u.)
z z z
x x x 0
 
Figure 2.11. Near-field coupling between the nanorod constituents rectangles of a three-
dimensional metamaterial. The figures show electric field norm in arbitrary
units when a plane wave is incident on the material. The coupling is weak
     
in a , stronger in b , and negligible in c .

can very often be neglected in such derivations, if the metamaterial layer


is thick enough, which means that the Fresnel coefficients can still be used.
Such slabs can be treated as Fabry-Perot resonators with r and t related to
the wave parameters by well known equations. In general, the wave that
propagates forward in the slab can experience different wave parameters
than the wave that propagates backward. Therefore, r and t depend on
four wave parameters and must be calculated for two incidence angles, θ
and π − θ, in order to retrieve these parameters [8,31,116]. The calculation
yields four coefficients, that form a system of four equations, from which
the wave parameters are solved.
Some metamaterials can be considered as stacks of multiple identical
layers composed of nanoparticles. Such layers can be coupled by the
near-fields of their constituents. Examples of weak, strong, and negligible
interlayer coupling are shown in Figs. 2.11a, 2.11b, and 2.11c for a material
composed of nanorods. The layers are assumed to be perpendicular to the
light propagation direction chosen to be along the z-axis. The coupling can
be reduced by increasing the separation between the rods in the z-direction
or decreasing it in the x-direction. If the coupling is weak, the layers act
effectively as infinitely thin sheets that reflect and transmit plane waves
propagating between them. In such cases, the reflection and transmission

26
Theoretical description of optical nanomaterials

coefficients, r and t, of a multilayered metamaterial slab are determined


by the single-layer coefficients ρ and τ [30]. Therefore, also the wave
parameters of the slab that are tightly related to r and t are determined
by the single-layer properties. Indeed, one can calculate n and Z from ρ
and τ , e.g., by using the equations presented in [30]. The coefficients ρ and
τ can be calculated numerically, which usually requires less computational
resources than calculating the coefficients of the entire slab.
In Publication I, we consider polarization-converting materials that do
not support any polarization eigenmodes, but can still be treated as stacks
of weakly coupled layers. For a single layer of such a material, we propose
using dyadic reflection and transmission coefficients
⎡ ⎤
ρss ρsp
ρ=⎣ ⎦ (2.27)
ρps ρpp

and ⎡ ⎤
τss τsp
τ=⎣ ⎦. (2.28)
τps τpp
Assume that a plane wave incident on the layer is described by a Jones
vector Ji = [Eis Eip ]T , where Eis and Eip are the complex amplitudes of
the s- and p-polarized components of the wave. The Jones vectors of
the reflected and transmitted fields are given by Jr = ρJi and Jt = τJi ,
respectively. The elements of the matrices ρ and τ can be obtained by
calculating Jr and Jt first for Ji = [E0 0]T and then for Ji = [0 E0 ]T . In
general, ρ and τ depend on the illumination side. From now on, these
matrices for the incidence angle θ are denoted by ρ+ and τ+ and those for
the angle π − θ by ρ− and τ− . We have proven that the wave parameters
can be introduced for the considered materials only if the matrices τ±
 
and ρ∓ ρ± share the same eigenvectors see Publication I . If this is not
the case, the matrices ρ± and τ± itself can be used as characterization
parameters.
For metamaterial slabs of many weakly coupled layers, the reflection
and transmission matrices, r and t, are determined by the single-layer
matrices analytically for any interlayer separation Λz . In Publication I, we
show that each individual layer of such a slab can be treated as a four-port
network, whose inputs and outputs are the waves approaching and leaving
 
the layer see Fig. 2.12 . The Jones vectors of these waves are denoted by

27
Theoretical description of optical nanomaterials

J+ single layer J−


of metamaterial
J− T1 J+

Figure 2.12. A four-port network equivalent to a single layer of metamaterial. The Jones
vectors J+ and J+ describe the plane waves that approach the layer from
both of its sides. The vectors J− and J− describe the waves leaving the layer.
All four vectors are related to each other by the transfer matrix T1 .

J+ and J− , respectively, on the entrance side of the layer, and by J+ and
 
J− on the exit side see Fig. 2.12 . These four Jones vectors are related to
each other by the equation
⎡ ⎤ ⎡ −1 −1
⎤⎡ ⎤
J+ f+ −f + g− 
⎥ ⎣ J− ⎦
⎣ ⎦=⎢
⎣ ⎦ (2.29)
−1 −1
J− g+ f + f − − g+ f + g− J+

where g± = ρ± eikz Λz , f ± = τ± eikz Λz , and kz is the longitudinal wave vector


component. The matrix in Eq. (2.29) is called the transfer matrix T1 of the
N
layer. The corresponding matrix for a slab of N layers is T1 and can be
written as ⎡ ⎤ ⎡ ⎤
−1 −1
T11 T12 t+ −t+ r−
⎣ ⎦=⎣ ⎦, (2.30)
−1 −1
T21 T22 r+ t + t − − r+ t + r−

where r± and t± are the illumination-side-dependent reflection and trans-


mission matrices of the slab. This matrix connects the Jones vectors of
the incident waves and the waves reflected and transmitted by the entire
multilayered metamaterial slab. We propose the following set of equations
from which r± and t± can be solved:
−1
t+ = T11 , (2.31)

r+ = T21 t+ , (2.32)

r− = −t+ T12 , (2.33)


−1
t− = T22 + r+ t+ r− . (2.34)

Here the blocks Tij only depend on ρ± and τ± and can, thus, be easily
solved.

28
3. Interaction of light with optical
nanomaterials

3.1 Plane wave decomposition

Optical experiments deal with a large variety of electromagnetic fields,


but only part of them closely resemble plane waves. Very often, the fields
cannot be accurately approximated by individual plane waves even locally.
Examples of such optical fields include focused laser beams and fields
forming optical images, e.g., in optical microscopes. In general, one cannot
predict the properties of an optical beam propagating in a nanostructured
medium based on the properties of an individual plane wave. A powerful
method to make such predictions is based on considering beams as coher-
 
ent superpositions of infinitely many polarized plane waves see Fig. 3.1
obtained by Fourier transforming the field. The interaction of each such
Fourier-component with the medium is then treated separately. The repre-
sentation is known as the plane-wave decomposition [79]. The calculated

Fourier components satisfy the Maxwell equations for plane waves see

Eqs. (2.22)-(2.25) . In isotropic media, that still can be spatially dispersive
the electric field E, magnetic field H, and wave vector k of each component
can be considered to be perpendicular to each other. In addition, the wave
vector k satisfies the equation kx2 + ky2 + kz2 = k 2 , where ki are the Cartesian
components of k and k is the wave number in the medium. In anisotropic
media, equations k · E = 0 and k · H = 0 do not in general hold, but the
components of k are still interdependent.
To calculate the complex amplitudes of the plane-wave components, it is
useful to introduce a plane z = z0 transverse to the beam axis and consider

29
Interaction of light with optical nanomaterials

k1 k2 k3
H3
E (x, y, z)
H (x, y, z)
= E1 + H2 E2 + +...
H1 E3
z
x y
 
Figure 3.1. The plane wave decomposition of a Gaussian beam. The beam left has an
electric field distribution E (x, y, z) and a magnetic field distribution H (x, y, z).
The beam is represented as a sum of infinitely many plane waves characterized
by wave vectors ki , electric fields Ei , and magnetic fields Hi . The figures show
the yz cross-sections of the field distributions.

the cross-section of the beam on that plane. The cross-sectional fields of


the plane-wave components oscillate in the plane z = z0 with wavelengths
determined by their transverse wave-vector components kT = x̂kx + ŷky
 
see Fig. 3.2 . The overall field distribution on the plane is a superposition
of such plane-wave oscillations. The amplitudes of these oscillations are
obtained by calculating the two-dimensional Fourier transform of the
spatial distribution of the overall electric field vector E (x, y, z0 ) at z = z0
as

1
Ê (kx , ky ; z0 ) = 2 E (x, y, z0 ) e−i(kx x+ky y) dx dy. (3.1)

−∞

The distribution of amplitudes Ê (kx , ky ; z0 ) is known as the angular spec-


trum of the beam [79, 121]. If the amplitudes Ê (kx , ky ; z0 ) are known,
the corresponding amplitudes at an arbitrary z = z0 + Δz are obtained

as Ê (kx , ky ; z) = Ê (kx , ky ; z0 ) eikz Δz , where kz = k 2 − kx2 − ky2 . Just as
E (x, y, z0 ) is the inverse Fourier transform of Ê (kx , ky ; z0 ), the electric field
distribution at z is the inverse Fourier transform of Ê (kx , ky ; z), i.e.,

E (x, y, z) = Ê (kx , ky ; z0 ) ei(kx x+ky y+kz Δz) dkx dky . (3.2)


−∞

Here, the electric field is expressed as a sum of plane waves with ampli-
tudes Ê (kx , ky ; z0 ). Hence, if E (x, y, z0 ) is known, the field distribution
E (x, y, z) at any z > z0 can be calculated [122]. The magnetic field distri-
bution H (x, y, z) can then be calculated from Faraday’s law.
Optical beams are always generated and spatially modified by light-
emitting, scattering, and absorbing objects, such as atoms, molecules,

30
Interaction of light with optical nanomaterials

ΛT
k
E H

y HT
x
ET k T
Figure 3.2. Projection of a plane wave on the xy-plane. The wave is characterized by its
wave vector k, electric field E, and magnetic field H. The projections of these
vectors on the xy-plane are kT , ET , and HT , respectively. Vectors E and H are
not necessarily parallel to the wave fronts. The xy cross-section of the wave
oscillates in the xy-plane with a wavelength ΛT determined by kT = 2π/ΛT .

and their microscopic and macroscopic ensembles. If these objects are


subwavelength-sized and located close to the plane z = z0 , the fields
on this plane can have subwavelength variations due to the near-fields
of these objects. Such fields necessarily have plane-wave components
with transverse wave numbers larger in value than k. These waves have
imaginary kz , and decay away from the plane z = z0 even if the medium is
transparent. Hence, the subwavelength details in the field distribution are
lost upon propagation. This is the fundamental reason for the diffraction
limit in optics [79]. Equations (3.1) and (3.2) can be used to treat the near
fields of a beam as well. The contribution of the evanescent waves is taken
into account by not limiting the integration limits in Eq. (3.2) to a circle of
radius k.
In Eq. (3.2), the longitudinal wave vectors kz of different plane-wave
components usually differ from each other. Therefore, the distribution
of the norm |E (x, y, z) | is a function of z. As an example, a laser beam
diverges away from its focal point exactly due to this effect. If the values
of kz were the same, propagation would result in the same phase factor
for all the plane-wave components and the norm of Eq. (3.2) would be the
same at z = z0 as at z0 . In fact, far from z0 , the amplitude distribution
of any electromagnetic field will be proportional to the angular spectrum
formed by the propagating components [79].
It is remarkable that, in Eqs. (3.1) and (3.2), light is treated as a vector
field, because in many optical applications light polarization plays an

31
Interaction of light with optical nanomaterials

important role. For example, if light is focused with a high-numerical-


aperture lens, its polarization can influence significantly the location,
shape, size, and peak intensity of the focal spot [79, 84, 123–127]. The
minimum achievable focal-spot size of a radially polarized beam can go
even below the diffraction limit [125]. Equations (3.1) and (3.2) do not only
form a fundamental basis for light propagation phenomenon [79, 128–131],
but can also be used as a design tool for optical components and systems
[132]. Furthermore, the aberrations and the resolution of systems are
well known to be influenced by the polarization of light [133–135]. Such
effects can be studied using Eqs. (3.1) and (3.2). On the other hand, the
information about the intensity distribution in the focal plane obtained
with these equations can help in preventing optical damage of a sample
exposed to a focused beam [136]. Equations (3.1) and (3.2) can also
be used to design optical trapping systems with polarization-dependent
functionalities [137–139] and diffractive optical elements that provide full
control over the polarization of the diffracted light [140, 141].
An alternative way to calculate the vectorial field distribution of a beam
interacting with complex structures is to use a numerical method, such as
 
the finite-element method FEM . A considerable problem associated with
such methods is that they become increasingly computationally heavy and
expensive when the size of the modelled geometry increases far above the
wavelength. Other numerical methods that can treat large objects, e.g.,
based on geometrical optics, are not always applicable. Equations (3.1)
and (3.2) can be used to efficiently calculate electromagnetic field distri-
butions that occupy large regions of space, e.g., by making use of the fast
 
Fourier transform FFT . Very often these equations yield the calculated
distributions faster than any conventional numerical method. Also the
Stratton-Chu integral [84, 142], the Richards-Wolf integral [128, 143], and
several other diffraction integrals [144, 145] can be used to model the evo-
lution of vector beams upon propagation. However, these integrals are not
applicable, if the beam considerably interacts with an anisotropic and/or
spatially dispersive material. In contrast, Eqs. (3.1) and (3.2) can be used
even if the effects of optical anisotropy and spatial dispersion are strong as
long as the field distributions in the material can be Fourier transformed.

32
Interaction of light with optical nanomaterials

3.2 Reflection and transmission of optical beams

Consider an optical beam that is partially reflected and partially transmit-


ted by a slab of a spatially dispersive material. One can ask, whether, in
such a situation, the reflected and transmitted beams can be decomposed
into reflected and transmitted plane waves. In the thesis, we answer this
question by introducing such a decomposition. The decomposition is used
to calculate the electric field distributions of the reflected and transmitted
beams. As far as we know, the method has not been introduced previously.
It is not only suitable for studying the interaction of light with spatially
dispersive materials, but it also provides better understanding of the inter-
actions and can reveal new optical phenomena. For example, in Publication
I, we use the method to demonstrate polarization conversion by spatial
dispersion.
In the presence of spatial dispersion, fully numerical methods can be used
to study reflection and transmission of optical beams by nanostructures.
However, due to the limited computational resources, such methods can
only treat the fields in relatively small regions of space, e.g. containing
several hundreds of unit cells. Moreover, since spatial dispersion appears
in many metamaterial structures, such as split-ring resonators [7], metal
fishnet-like structures [146], and alternating metal-dielectric layers [147],
it is important to develop new semianalytical methods for modelling the
interaction of optical beams with spatially dispersive metamaterials. In
the thesis, such a method is introduced and used to demonstrate some
applications of spatial dispersion. For example, in Publication I, we develop
an apertureless spatial filter that makes use of spatial dispersion, and
in Publications III-V we make spatial dispersion compensate for optical
diffraction.
The key features of the method are as follows. The incident beam is
 
decomposed into polarized plane waves see Sec. 3.1 , each of which is
then reflected and transmitted by the material. It is practical to select the
two orthogonal polarizations for each plane wave in an appropriate way.
Often, they are chosen such that neither of them is affected by reflection
and transmission. In this case, one does not need to consider polarization
conversion between the two components, which simplifies calculations.

33
Interaction of light with optical nanomaterials

 
a p b S3

P3
P
χ
E ψ
s P1 P2
k
S2

S1

Figure 3.3. Two ways to represent a polarization state of a plane wave. The polarization
 
ellipse of the wave is shown in a . It lyes on a plane perpendicular to the
wave vector k. The rotating electric field vector E can be expanded into its s-
 
and p-polarized components. The polarization state in a is represented by a
 
point P on the surface of the Poincaré sphere shown in b .

For example, s- and p-polarized waves are not converted to each other
when reflected and transmitted by an ordinary dielectric slab. This can
also be the case when the slab is made of a spatially dispersive material.
Therefore, we choose to use the s- and p-polarized waves as a basis for
the plane-wave decomposition of the incident beam. However, unlike
ordinary dielectric materials, some spatially dispersive materials do not
support polarization modes for all incidence angles. This means that, no
matter what polarization state is selected, it will necessarily change upon
reflection and transmission by the material. In such cases, the energy
transfer between the chosen orthogonal polarizations is unavoidable and
must be taken into account in the calculations.
Any polarization state of a plane wave can be represented as a two-
element Jones vector containing the complex amplitudes, Es and Ep , of the
 
s- and p-polarized components of the wave see Sec. 2.4 . It is not always
easy to see what is the polarization state from the form of one complex-
valued Jones vector. An alternative way to represent the polarization is to
use the so-called polarization ellipse that is the trajectory of the rotating
electric-field vector of the wave. The shape and the orientation of the
ellipse are characterized by two angles, ψ and χ, as shown in Fig. 3.3a.
One more way to represent the polarization state is to use the Poincaré
sphere. The angles ψ and χ determine a single point on the surface of the

34
Interaction of light with optical nanomaterials

 
Poincaré sphere see Fig. 3.3b . The coordinates P1 , P2 , and P3 of any point
P on the sphere can be calculated from the corresponding Jones vector
JP = [Es , Ep ]T as follows

P0 = |Es |2 + |Ep |2 = P12 + P22 + P32 , (3.3)

P1 = |Es |2 − |Ep |2 = P0 cos (2χ) cos (2ψ) , (3.4)


 
P2 = 2Re Es Ep∗ = P0 cos (2χ) sin (2ψ) , (3.5)
 
P3 = 2Im Es Ep∗ = P0 sin (2χ) , (3.6)

∗ denotes the complex conjugate of E


where the Es,p s,p [74]. The quantity

P0 is the radius of the sphere equal to the intensity of the wave. The
coordinates P1 , P2 , and P3 are known as the Stokes parameters. Later on in
this section, we use these parameters to visualize some of the polarization
states treated in the thesis.
When calculating the electric field distribution of a beam using the plane
wave decomposition, one must sum the electric field vectors of the beam
plane-wave components. To be able to do this at any point in space, we
represent the fields in terms of Cartesian vector components connected to
a fixed coordinate system. Then, the Jones vectors J = [Es , Ep ]T yield the
Cartesian vector components in question through the transformation
⎡ ⎤ ⎡ ⎤
Ex cos ϕ cos θ sin ϕ ⎡ ⎤
⎢ ⎥ ⎢ ⎥ Es
⎢ Ey ⎥ = CJ = ⎢ − sin ϕ cos θ cos ϕ ⎥ ⎣ ⎦ (3.7)
⎣ ⎦ ⎣ ⎦
Ep
Ez 0 − sin θ

written for a plane wave with a wave vector k = k(x̂ sin θ sin ϕ+ŷ sin θ cos ϕ+
ẑ cos θ). Here, the angles θ and ϕ are the polar and azimuthal angles,
respectively, showing the direction of k. These angles are, in general,
complex-valued, because the considered plane-wave component can be
attenuating upon propagation or evanescent. Because each plane-wave
component of the beam is associated with a Jones vector, the entire angular
 
spectrum Ê of the beam see Sec. 3.1 is associated with the spectrum of
the Jones vectors Ĵ. By using Eq. (3.7), one can obtain Ê = CĴ for each
direction of k. Therefore, the electric field distribution of the beam can be
written as

E (x, y, z) = C (kx , ky ) Ĵ (kx , ky ) ei(kx x+ky y+kz z) dkx dky . (3.8)


−∞

35
Interaction of light with optical nanomaterials

 
a b
D D
k3 k3
tC,3 Ê3 rC,3 Ê3
k2 k2
k3 y tC,2 Ê2 rC,2 Ê2 y
Ê3 z z
k2 x k1 k1 x
Ê2 tC,1 Ê1 rC,1 Ê1
k1
Ê1

Figure 3.4. Interaction of plane-wave components of an optical beam with a slab of a


spatially dispersive material. The thickness of the slab is D. Three incident
 
and transmitted plane-wave components are shown in a and the reflected
 
components are shown in b . The wave vectors k1 , k2 , and k3 of the trans-
mitted waves are equal to those of the incident waves, but differ from those of
the reflected waves.

Note that the Jones-vector spectrum can also be obtained from the angular
−1
spectrum Ê as Ĵ = C Ê.
In Publication I, by using the plane-wave decomposition, we formulate
the electric field distributions of an optical beam reflected and transmitted
by a slab of a spatially dispersive material. We start by calculating the
−1
Jones-vector spectrum Ĵi = Ci Êi of the incident beam, where Ci is the
transformation matrix of Eq. (3.7). Then, we calculate the Jones vector
spectra of the reflected and transmitted beams by utilizing the reflection
 
and transmission matrices r and t see Sec. 2.4 . These spectra can be
written as Ĵr = rĴi and Ĵt = tĴi , respectively. The corresponding angular
spectra are given by Êr = Cr Ĵr and Êt = Ct Ĵt . Note that, in general, the
three matrices Ci , Cr , and Ct differ from each other as long as the directions
of the wave vectors ki , kr , and kt are different. One can also write the
calculated angular spectra as Êr = rC Êi and Êt = tC Êi , where we introduce
−1 −1
the matrices rC = Cr rCi and tC = Ct tCi . The Cartesian components
of the electric field vectors of the reflected and transmitted plane waves
can be obtained by using these matrices. The waves are illustrated in
Figs. 3.4a and 3.4b. The reflected and transmitted plane-wave fields can
then be summed, which yields the full electric field distributions of the

36
Interaction of light with optical nanomaterials

reflected and transmitted beams


Er (x, y, z) = rC (kx , ky ) Êi (kx , ky ) ei(kx x+ky y+kz z) dkx dky (3.9)
−∞

and

Et (x, y, z) = tC (kx , ky ) Êi (kx , ky ) ei[kx x+ky y+kz (z−D)] dkx dky , (3.10)
−∞

respectively. Here, z = 0 at the entrance surface of the slab and D denotes


the thickness of the slab.
Some spatially dispersive materials do support polarization modes and
can therefore be characterized by the wave parameters. Consider a plane-
wave excitation of a polarization mode in a slab of such a material. The
reflection and transmission of the polarized incident plane wave in this
case can be treated by using the matrices r and t. The Jones vector Ji of
the wave must satisfy rJi = rJi and tJi = tJi . Therefore, one can use the
scalar-form reflection and transmission coefficients r and t for the slab,
and retrieve the wave parameters from them. Reversely, these coefficients
can be calculated from the wave parameters, if the materials surrounding
the slab are known. These calculations are carried out analytically for any
surrounding media, which makes the analysis of the materials fast. If a
beam incident onto the slab can be represented as a superposition of the
polarization modes of the slab, the matrices rC and tC can be calculated
analytically from the wave parameters. Therefore, the fields Er and Et can
be calculated nearly instantaneously. This is a major benefit of using the
wave parameters.
Spatially dispersive materials with a fixed refractive index can be made
impedance-matched to the surrounding medium in order to reduce reflec-
tion losses of light at the material boundaries. However, if the impedance
depends on the propagation angle, the reflection can be fully eliminated
only at a single incidence angle. The only exception is the Brewster’s angle
at which the reflection coefficient can be zero. An incident beam with a
small divergence angle is as well only weakly reflected by an interface
between two approximately impedance-matched media. The same is valid
for a slab of an impedance-matched spatially dispersive material. Hence,
the possibility to evaluate the wave parameters allow us to evaluate and
reduce the losses of light interacting with the material.

37
Interaction of light with optical nanomaterials

Equation (3.10) can be easily applied to calculate the transmitted field


distribution Et at the exit surface of a designed metamaterial slab. This
can be done for a set of different thicknesses of the slab in order to study
possible changes that the transmitted field can show when the thickness
of the slab is changed. The results of these calculations can be used for
optimizing the geometry of optical elements made of the material. For
example, one can find a minimum thickness for a metamaterial slab that
absorbs essentially all the incident power. Equation (3.10) can also be
used to find an approximate field distribution of a beam propagating inside
a metamaterial slab. The internal field distribution at any distance d
from the entrance surface of the slab is approximately the same as the
distribution of the field transmitted by a slab of thickness d. This is often
a good approximation because, the distribution of the transmitted field
at the exit surface of the slab does not significantly differ from the inter-
nal field distribution just behind the surface, as long as the transmission
coefficient is nearly the same for all plane-wave components of the beam.
Hence, if multiple reflections inside the slab can be ignored, the entire field
distribution in the slab can be approximated by the field distributions at
the output surfaces of many slabs of various thickness. Standing optical
waves inside the slab can be ignored, if the slab is impedance matched to
the surrounding media or has a low surface reflectivity and nonnegligible
absorption. Rigorously, the internal field can be calculated as a superpo-
sition of polarized plane waves that propagate back and forth inside the
slab, e.g., using Eq. (3.2). However, such calculations require the slab to
support polarization modes.
In Publication I, we demonstrate that Eqs. (3.9) and (3.10) are applicable
even to such metamaterial slabs that do not support polarization modes.
We consider a metamaterial, in which spatial dispersion leads to inevitable
conversion of light polarization nearly for all propagation directions in the
material. The structure of the material is shown in Fig. 3.5a. The material
is composed of silver discs, each of which is 20 nm thick and has a 30 nm
radius. The discs form a cuboidal lattice in glass. They are twisted by
an angle of 45◦ with respect to the lattice. The lattice constants in the
x-, y-, and z-directions are Λx = 120 nm, Λy = 120 nm, and Λz = 180 nm,
respectively. The material can be viewed as a stack of identical layers

38
Interaction of light with optical nanomaterials

 
a b S3 /S0

P4
P2
P3 S2 /S0
z P1

y
x S1 /S0
  
10 c d e IM

5
y (μm)

I (a.u.)
0

-5
IM = 1 IM ≈ 0.012 IM ≈ 0.006
-10 0
-10 -5 0 -10 -5 0 5 10 -10 -5 0
z (μm) z (μm) z (μm)
Figure 3.5. A polarization converting metamaterial. The structure of the material is
 
shown in a . The polarization states that are not changed if the plane of
incidence is the yz-plane and the incidence angle is 45◦ , either by reflection
 
or transmission are shown in b as points P1 and P3 , respectively, on the
Poincaré sphere. The point P2 represents the polarization state of the reflected
light when the incident light has the polarization state P1 . The point P4 rep-
resents the polarization state of the transmitted light when the incident light
has the polarization state P3 . The black dots are the projections of the points
P1 -P4 onto the horizontal symmetry plane of the sphere. A Gaussian beam
 
that is incident on a slab of the material is shown in c . The figure shows the
longitudinal cross-section of the intensity distribution of the beam. Similar
   
distributions are shown in d and e for the components of the reflected and
transmitted beams that are polarized orthogonally to the incident beam. The
maximum intensity IM is calculated separately for each of the beams.

parallel to the xy-plane and having a period Λz . The distance Λz was


chosen to be relatively large, such that no significant evanescent-wave
coupling can occur between the layers. Therefore, the optical properties
of the material are entirely determined by the properties of a single layer
 
see Sec. 2.4 . Each such layer is symmetric with respect to its middle
plane that is normal to the z-axis. Hence, the reflection and transmission

39
Interaction of light with optical nanomaterials

matrices, ρ and τ, of the layer do not depend on the illumination side.


Light that propagates along a symmetry plane of the material and is
polarized either parallel or perpendicular to this plane preserves its polar-
ization state upon propagation. In Publication I, we show that polarization
modes that do not propagate in such directions do not exist. Consider, for
example, a plane wave that is incident on a single layer of the material

at an angle of 45◦ with the plane of incidence being the yz-plane see

Fig. 3.5a . If the wave was a polarization mode of the configuration, its
polarization state would be unchanged upon both reflection and transmis-
sion. In this case, the Jones vector associated with the wave would be a
 
common eigenvector of the matrices ρ and τ see Sec. 2.4 . However, such
polarization states do not exist for the considered direction of incidence.
Figure 3.5b shows various polarization states as points on the surface of
the Poincaré sphere. The blue point P1 represents the polarization state
that, in the considered situation, is preserved upon transmission. The
red point P2 , on the other hand, stands for the polarization state of the
reflected light. The difference between these two points illustrates the fact
that the material does not support polarization modes for the considered
direction of incidence. It is also possible to find the polarization state pre-
served upon reflection. It is shown by the green point P3 . The violet point
P4 , however, represents a different polarization state of the transmitted
light.
Figures 3.5c-e illustrate some polarization-conversion properties of inter-
action of a Gaussian beam with a slab of the metamaterial. The beam is
focused on the surface of the slab. It has a vacuum wavelength of 633 nm.
The divergence angle of the beam is 15◦ , which corresponds to a beam
waist diameter of approximately 500 nm. The intensity distribution of
the beam is shown in Fig. 3.5c. On the entrance surface of the slab, the
transverse electric field of the beam is entirely polarized along the rota-
tional symmetry axes of the discs. If the material was anisotropic, but not
spatially dispersive, one would expect the reflected and transmitted beams
to have the same polarization as the incident one. However, the orthogonal
polarization is observed in both transmitted and reflected beams, as shown
in Figs. 3.5d and 3.5e, respectively. This means that the polarization of
light is indeed converted upon interaction with the slab. The illustrated

40
Interaction of light with optical nanomaterials

 20 b  10
a c

10

y (μm)

I (a.u.)
0

z -10

y -20 0
x -10 0 10 -10 0
z (μm) z (μm)

Figure 3.6. A metamaterial used as an apertureless spatial filter. The structure of the
 
material is shown in a . A slab of the material is illuminated by a beam that
is focused onto its surface. The yz cross-section of the intensity distribution of
 
the beam is shown in b . The corresponding cross-section for the transmitted
 
beam is shown as well. The cross-section of the reflected beam is shown in c .

intensity profiles have been verified by using direct numerical methods.


Because the considered metamaterial has a relatively simple structure, we
conclude that such polarization conversion is expected to commonly occur
in optical metamaterials.
Since Eqs. (3.9) and (3.10) can be applied to nanomaterials with higher-
order multipoles efficiently excitable in their structural units, we success-
fully apply them to one of our designed metamaterials. A slab made of the
material in question acts as a new kind of spatial filter. The material is
composed of silver-disc dimers, as shown in Fig. 3.6a. Similar dimers have
been used as structural units of a metasurface in Sec. 2.3. Each dimer
consists of two 10 nm thick discs that are separated by a gap of 40 nm.
The radii of the discs are 20 nm and 10 nm. The dimers are embedded
in glass forming a cubic lattice with a lattice constant Λ = 175 nm. They
support a prominent current quadrupole excitation at the wavelength of
525 nm. If a plane wave is incident on the slab, this excitation can strongly
 
influence the reflection coefficient see Secs. 2.2 and 2.3 . In particular,
the reflection coefficient can depend significantly on the incidence angle.
The dependence can be tuned by tuning the quadrupole polarizabilities of
the dimers, through adjusting their geometry. As an example, we design a
metamaterial that efficiently reflects only nearly normally incident plane
waves. We show that a slab of such a material can be used as an aperture-

41
Interaction of light with optical nanomaterials

less reflective spatial filter. In Publication I, we consider a 5 μm thick slab


of the material illuminated by a linearly polarized optical beam focused
onto its surface with a convergence angle of 60◦ . The wavelength of the
beam is 525 nm. Figure 3.6b shows the longitudinal intensity profiles of
the incident and transmitted beams and Fig. 3.6c shows the profile of the
reflected beam. The reflected beam is seen to diverge much slower than the
incident and transmitted beams. This means that the angular spectrum of
the reflected beam is narrower than that of the incident beam. Therefore,
the slab can be considered as a spatial filter. Note that also the transmitted
beam is spatially filtered by the slab as its components with low spatial
frequencies are removed by reflection. The slab differs fundamentally from
a conventional spatial filter that consists of two lenses and a pinhole placed
between them. The beam, the lens, and the pinhole must be accurately
aligned in such a system to obtain appropriate filtering. In contrast, the
metamaterial filter presented here is completely independent of both the
lateral and longitudinal positions of the focal spot.

3.3 Diffraction compensation

Diffraction makes optical beams diverge and optical images blur upon
propagation in free space. It can be considered compensated, if these ef-
fects are significantly reduced. Diffraction compensation can find a variety
of applications, e.g., in solar panels [41, 148], antireflection coatings [149],
integrated optics [41], and laser cavities [150]. Some photonic crystals
can provide diffraction-free propagation of light along certain directions in
the crystal [39–41, 151–156]. This phenomenon, known as self-collimation,
occurs at wavelengths close to the crystal’s Bragg-reflection regime and
is, therefore, accompanied by a high surface reflectivity of the crystal. In
certain applications of self-collimating photonic crystals, the reflection loss
in question is suppressed either by using an optical coupler [157–161]
or by adjusting the thickness of the crystal [43]. These complications,
however, could be removed, if one had succeeded in reducing the sur-
face reflectivity of the material. Some metamaterials can also suppress
diffraction of light for particular direction of propagation in the mate-
rial [14, 34–36, 162–164], e.g., along the chains of their structural units

42
Interaction of light with optical nanomaterials

coupled to each other by surface plasmons. This type of optical energy


transfer is often accompanied by relatively high absorption losses in the
 
structures see, e.g., [163] . Moreover, typically optical modes supported by
such diffraction-compensating metamaterials exhibit the impedances that
differ significantly from that of air or glass, which can lead to difficulties
in coupling of light into the material. This is the case, e.g., for meta-

materials that operate in the so-called canalization regime see, e.g., [5]

and [165] . Because the absorption and coupling losses are undesirable,
the demonstrated diffraction-compensating metamaterials typically oper-
 
ate at relatively low THz frequencies. An important question is whether
low-loss diffraction-compensating metamaterials can be designed for opti-
cal frequencies. In the thesis, several optical metamaterials that exhibit
low absorption and reflection losses are introduced.
Compensation for optical diffraction by materials that support polariza-
tion modes can be studied with the help of the plane-wave decomposition
method described in Secs. 3.1 and 3.2. The use of this method is justified,
if the longitudinal evanescent-field coupling between the metamolecular
layers of the material is weak, because then optical fields are accurately
represented by the fundamental Bloch modes of the material via the de-
composition of Eqs. (3.1) and (3.2) [30, 166]. The method can be applied
if the wave parameters in the material are known. To calculate the wave
parameters, one can apply the theory of Sec. 2.4. For each plane wave
incident onto the surface of the material, one can write the complex Snell’s
law
ns sin θi = n (θ) sin θ. (3.11)

Here, θi is the incidence angle, θ is the angle of refraction, n (θ) is the


complex-valued refractive index of the material, and ns is the refractive
index of the surrounding medium. If θ represents a polar angle in spher-
ical coordinates, the wave parameters are, in general, functions of the
azimuthal angle ϕ as well, and Eq. (3.11) holds for each ϕ separately.
When the wave parameters are known, the electric field distribution of a
beam in the material can be written as

E (x, y, z) = Ê (kx , ky ; 0) ei(kx x+ky y+kz z) dkx dky , (3.12)


−∞

where kz = k0 n (θ, ϕ) cos θ and k0 is the wave number in the free space see

43
Interaction of light with optical nanomaterials


Sec. 3.1 . In a diffraction-compensating medium, E (x, y, z) = E (x, y, 0) eikz z ,
which means that the factor eikz z can be taken out of the integral, and the
beam preserves its transverse intensity profile upon propagation. This is
possible, if the refractive index in the material is

n0
n (θ, ϕ) = , (3.13)
cos θ

where n0 is constant. The refractive index n (θ, ϕ) can be made independent


of ϕ by designing the material such that its structural units are radially
 
symmetric and have their symmetry axes parallel to the z-direction θ = 0 .
If the surface representing nr (θ, ϕ) = Re {n (θ, ϕ)} is flat and Im{n (θ, ϕ)}
is negligible, Eq. (3.13) is satisfied. The Poynting vector must in this
case point everywhere in the same direction [74]. In a material, whose
refractive-index surface has multiple flat parts, the energy flow direction
is predominantly perpendicular to those parts. Hence, Eq. (3.13) can be
used as a design tool for diffraction-compensating metamaterials.
To realize a metamaterial whose refractive index approximately satisfies
Eq. (3.13), one must achieve a situation, in which plane waves propagating
at larger angles θ experience higher real parts of the refractive index.
In fact, electric dipole resonances in the structural units of a nanomate-
rial can lead to an increase of the real part of n on the long-wavelength
side of the resonance, when the strength of the resonance is increased.
Therefore, diffraction-compensating metamaterials can be realized as ar-
rays of nanoparticles, whose dipole resonances are more efficiently excited
by plane waves with larger angles θ. The material can for example be
composed of metal nanoparticles with only the longitudinal plasmon reso-
nance to be excitable. Because waves with larger angles θ have stronger
z-directional electric field components, they can more efficiently excite the
resonances, which leads to higher values of nr experienced by the waves.
However, if the current multipole excitations of the nanoparticles were
purely of electric-dipole character, the refractive-index surface of the mate-
rial would have no truly flat part. In such a case, the material would be
an ordinary anisotropic medium, and its refractive index surface would be
an ellipsoid [74]. Therefore, the structural units of the material must be
large enough and properly shaped to support excitations of higher-order
multipoles and make the refractive-index surface flat.

44
Interaction of light with optical nanomaterials

  
a c e

x x x
y y y
z z z
  
b λ0 = 525 nm d λ0 = 695 nm f λ0 = 913 nm
90◦ 90◦ 90◦
120◦ 60◦ 120◦ 60◦ 120◦ 60◦
2
150◦ 30◦ 150◦ 30◦ 150◦ 30◦
1
nr 2 3 nr 2 3
180◦ 0 180◦ 1 0 180◦ 1 0
nr

210◦ 330◦ 210◦ 330◦ 210◦ 330◦


240◦ 300◦ 240◦ 300◦ 240◦ 300◦
270◦ 270◦ 270◦
 
Figure 3.7. Examples of diffraction-compensating metamaterials. The material in a is
composed of tilted disc dimers that form a cubic lattice in glass with a lattice
constant Λ = 120 nm. Each dimer is composed of two 10 nm thick silver discs
with radii of 25 nm and 40 nm separated by a gap of 20 nm. The material in
 
c is composed of 10 nm thick silver discs with a radius of 40 nm. The discs
form a cuboidal lattice in glass with the lattice constants Λx = Λy = 100 nm
 
and Λz = 160 nm. The material in e consists of silver nanorods and has
the lattice constants Λx = Λy = 120 nm and Λz = 180 nm. Each nanorod is
130 nm long and has a square cross-section with a side length of 30 nm. The
   
refractive-index functions nr (θ) of these materials are shown in b , d , and

f , respectively, at the operation wavelengths λ0 of the materials. The empty
white sectors represent the angles at which no plane wave refracted into the
material from glass can propagate.

We have designed several diffraction-compensating metamaterials that


exhibit negligible imaginary part of the refractive index. Three examples
of these materials are shown in Fig. 3.7 together with the real parts of
their refractive-index functions nr (θ, ϕ). Figure 3.7a shows a material com-
posed of silver disc dimers that form a cubic lattice in glass. The dimers
are twisted by an angle of 45◦ with respect to the lattice. Both large and
small discs support plasmon resonances perpendicular to the rotational
symmetry axes of the dimers. At the wavelength of 525 nm, the plasmon
excitations in the xy-plane result in a diffraction-compensation effect in
the material. Note that the material supports polarization modes only in

45
Interaction of light with optical nanomaterials

the xy-plane. Figure 3.7b shows a polar plot of the nr (θ) in this plane. The
plot is shown for TM-polarized plane waves. The refractive index contour
is nearly flat at θ ≈ 45◦ and θ ≈ 225◦ . Therefore, optical beams with propa-
gation angles close to these two angles do not experience significant diffrac-
tion upon propagation, which was shown by direct numerical calculations.
As shown in [167] the material can be impedance-matched to vacuum,
which is a significant benefit over many other diffraction-compensating
materials. In principle, similar metamaterials can be designed to have
two wavelength bands of diffraction compensation, corresponding to sepa-
rate plasmon-resonance wavelengths of the large and small discs, which
potentially can result in a wider band of diffraction compensation.
A simpler metamaterial consisting of individual silver discs arranged in a
cuboidal lattice in glass is shown in Fig. 3.7c. The discs support a plasmon
resonance with the charge oscillation perpendicular to the y-direction. At
a wavelength of 695 nm, this leads to the diffraction-compensation effect.
The wave parameters, however, can be defined only for plane waves that
propagate in the symmetry planes of the material. Figure 3.7d shows
a polar plot of nr (θ) of the material in the yz-plane. The plot is shown
for TM-polarized waves at a wavelength of 695 nm. The refractive index
contour is flat around θ = 0 and θ = 180◦ . Therefore, TM-polarized beams
that are wide in the x-direction experience diffraction compensation in the
yz-plane while propagating along the z-axis in the material. Furthermore,
TM-polarized plane waves propagating along the z-axis do not significantly
interact with the nanoparticles, since they cannot efficiently excite any
plasmon resonance in them. Therefore, for such waves, the material
appears as glass and is, thus, impedance-matched to glass. It is clear that
the material can be impedance-matched to any medium whose refractive
index is equal to that of the material surrounding the discs.
Yet another metamaterial, with a cuboidal lattice of silver nanorods in
glass, is shown in Fig. 3.7e. The material exhibits diffraction compensation
at a wavelength of 913 nm, at which the nanorods exhibit a non-negligible
longitudinal plasmon resonance. Unlike the two previously presented
metamaterials, the nanorod material supports polarization modes for
all propagation directions and can be described by a three-dimensional
refractive-index surface. In the yz-plane, the cross-section of this surface

46
Interaction of light with optical nanomaterials


10 a Im

y (μm)

I (a.u.)
0

glass metamaterial
-10 0
-20 -10 0 10 20 30 40

20 b Im

I (a.u.)
y (μm)

glass metamaterial glass


-20 0
-20 0 20 40 60 80

20 c Im
y (μm)

I (a.u.)
0

glass metamaterial glass


-20 0
-40 -20 0 20 40 60 80 100
z (μm)

Figure 3.8. Propagation of optical beams in diffraction-compensating metamaterials. A


cross-section of a three-dimensional beam propagating in the silver nanorod
 
material is shown in a . A two-dimensional beam obliquely incident at a slab
 
of the silver disc material from glass is shown in b . Interference pattern
 
of two such beams propagating through the same slab is shown in c . The
beams are focused behind the entrance surface of the slab. The interfaces
between the material and the surrounding glass are indicated by the vertical
white lines.

calculated for TM-polarized plane waves at a wavelength of 913 nm is a


polar plot shown in Fig 3.7f. The refractive-index contour is flat symmetri-
cally around θ = 0 and θ = 180◦ . Hence, three-dimensional TM-polarized
optical beams should preserve their field profiles upon propagation along
the z-axis in the material. Just as the silver disc metamaterial, the nanorod

material is impedance-matched to glass that is the host medium of the

rods for beams propagating along z.
The nanorod material resembles some of the other diffraction-compensat-
 
ing metamaterials by its structure see, e.g., [162, 163, 165] . However,
it differs from them substantially, because the energy transfer in the z-
direction is not provided by evanescent-field coupling between the particles.
Instead, the energy is transferred by plane waves propagating between the

47
Interaction of light with optical nanomaterials

nanoparticle layers. Compared to near-fields, the propagating plane waves


are not confined to metal nanostructures, which results in significantly
reduced absorption losses. Because the near-field coupling between the
nanorod layers is weak, the optical properties of the nanorod material are
determined by those of a single layer of the material and do not depend
 
on the material thickness see Sec. 2.4 . As far as we know, this is not the
case for the other known diffraction-compensating metamaterials.
To demonstrate the diffraction-compensating properties of the designed
materials, we use the methods of Sec. 3.2. Figure 3.8a shows a longitudi-
nal intensity profile of a radially polarized optical beam focused onto an
interface between glass and the nanorod metamaterial described above.
The divergence angle of the beam in glass is 16◦ corresponding to a beam
radius of about 1 μm. After entering the metamaterial, the beam is seen to
propagate without spreading or attenuation over a remarkable distance of
ca. 50 μm. Furthermore, because the metamaterial is impedance-matched
to glass, also the reflection loss at the interface is low, which can be seen
from the nearly equal intensities on both sides of the interface.
Figure 3.8b shows a longitudinal intensity profile of a two-dimensional
Gaussian beam focused onto the surface of a slab of the silver disc metama-
terial. The slab is 70 μm thick and surrounded by glass on both sides. The
divergence angle of the beam in glass is 20◦ . The beam is polarized in the
yz-plane and incident onto the surface at an angle of 10◦ . Inside the slab,
the beam propagates along the normal to the slab almost without spread-
ing. The incidence angle does not influence much the beam propagation
direction, which is exactly as predicted. Similar unidirectional propagation
of light takes place also in the nanorod metamaterial.
The diffraction-compensation effect prevents arbitrary profiles E (x, y, 0)
from spreading upon propagation. To demonstrate this, we consider an
interference pattern produced by two Gaussian beams at the entrance of a
 
70 μm thick slab of the silver disc metamaterial see Fig. 3.8c . This pattern
is observed to propagate across the slab nearly without distortions. The
photons, therefore, stop crossing the fringes in the material. Absorption
in the material makes the fringes with lower intensity look disappearing
faster upon propagation.
Also two-dimensional intensity distributions can propagate without dis-

48
Interaction of light with optical nanomaterials

z = 0 Pm z = 200 Pm z = 200 Pm
  
a b c

z = 0 Pm z = 33 Pm z = 50 Pm
  
d e f

Figure 3.9. Transfer of optical images through the silver nanorod metamaterial. The
image of the Leonardo da Vinci’s Mona Lisa to be transferred through the
 
material is shown in a . The image is formed by unpolarized light at the
wavelength of 913 nm and its smallest features have a size of ca. 750 nm. Fig-
   
ures b and c show the image after propagation over 200 μm distance in the
metamaterial and glass, respectively. Another transferred image consisting
 
of three lines is shown in d . The lines, from the leftmost to the rightmost,
 
are formed by TE-, TM-, and circularly polarized light. The images in e and

f show the consequence of light propagation over 33 μm and 50 μm in the
material, respectively. From [168].

tortion through an appropriately designed diffraction-compensating meta-


material. In Publication V, we demonstrate this by using the nanorod
metamaterial of Fig. 3.7c. As an example, we consider propagation of an

image of the Leonardo da Vinci’s Mona Lisa through the material see

Fig. 3.9 . Interestingly, the image quality remains very good even after
propagation through a 200 μm thick slab of the material. Figure 3.9a shows
the original image formed by unpolarized light and Fig. 3.9b shows the
transmitted image. For comparison, Fig. 3.9c illustrates the image after
propagation over the same distance in glass.
The nanorod material can compensate for optical diffraction of TM-
polarized light only. Therefore, TE-polarized light patterns should vanish
upon propagation being washed out by diffraction. Indeed, the image in
Fig. 3.9b is predominantly formed by TM-polarized waves. We have found
that nearly any optical image can be constructed solely with TM-polarized
waves. To consider the sensitivity of the diffraction compensation to light

49
Interaction of light with optical nanomaterials

polarization, we considered propagation of three line patterns shown in


Fig. 3.9d in the material. The first two lines are TE- and TM-polarized,
and the third one is circularly polarized. Figures 3.9e and 3.9f show the
pattern after propagation over 33 μm and 50 μm in the material, respec-
tively. The TE-polarized line becomes wide and eventually disappears
as it propagates, while the TM-polarized one remains sharp. The third,
circularly polarized line also stays sharp, but its brightness is lower by
a factor of two. This observation is explained by the fact that circularly
polarized light is a superposition of TE- and TM-polarized fields with equal
amplitudes. Because unpolarized light waves can be considered as an inco-
herent superposition of the TE- and TM-polarized waves, also unpolarized
light patterns must preserve their shapes, as described in the example of
the image of Mona Lisa. The material can therefore be considered as a new
type of optical polarizer, removing TE-polarized plane-wave components
from field distribution.

3.4 Light propagation in metamaterial slab waveguides

In this section, we focus on metamaterial slab waveguides, whose core is


composed of a metamaterial. Such waveguides can have unusual light-
guiding properties derived from the peculiar properties of metamaterials,
such as ultralow, ultrahigh, negative, or highly anisotropic refractive index.
Extraordinary optical phenomena demonstrated in metamaterial wave-
guides include, e.g., slow light propagation [47, 48] and sub-wavelength
confinement of optical power [49, 50, 169]. In addition, potentially, spatially
dispersive metamaterial slab waveguides can be used to create photonic cir-
cuits based on light self-collimation and self-guidance, similarly to photonic
crystals [41, 170].
Optical energy in conventional waveguides propagates along channels
delimited by material boundaries. In contrast, in diffraction-compensating
metamaterials, such channels are formed by light itself. Hence, the possibil-
ity to create slab waveguides that exhibit the diffraction-compensation ef-
fect can facilitate the design and fabrication of photonic chips utilizing this
effect. Unlike most of conventional waveguides, diffraction-compensating
slab waveguides can transfer multiple signals along the same self-formed

50
Interaction of light with optical nanomaterials


a nclad cladding

y
ncore MM x z ncore core

b (y) a (y) b (y)


nsubs substrate


b
dr
x

y z
βMM
βx
ϕ βMM,z
β βx dt
βz W

Figure 3.10. A metamaterial slab waveguide between two ordinary dielectric slab wave-
 
guides. Its core is made of a metamaterial MM . An effective electric-field
 
profile a (y) of a mode of the waveguide is shown in the side view a . The
profile is illustrated by a black solid line. Also an electric-field profile b (y) of
a mode of the dielectric waveguides is shown. The metamaterial waveguide
acts as a Fabry-Perot resonator, in which light propagates back and forth, as
 
shown in the top view b . The green dashed lines show the guided-mode
propagation. The reflection and transmission coefficients of the metamaterial
waveguide are calculated from the overall field distributions at the distances
dr and dt from the waveguide.

channel without significant crosstalk. Therefore, the information transfer


capacity of such waveguides can be made high. In fact, spatially dispersive
metamaterial slab waveguides have not been previously considered in the
literature. In Publication V, we study the properties of such waveguides
and show that divergence-free self-guiding of light can be obtained in them.
The modes of metamaterial slab waveguides are similar to the modes of
conventional slab waveguides. They are orthogonal sets of self-consistent
solutions of Maxwell’s equations that can propagate along the waveguide
without changing their geometric properties. The energy of such a mode
is confined mostly in the waveguide core and its density decays smoothly
away from the core. The amplitudes of the field electric and magnetic

51
Interaction of light with optical nanomaterials


components of each mode vary only along the waveguide normal the

y-direction in Fig. 3.10 . Therefore, this variation can be described by

a function b (y) whose values are restricted to the interval [−1, 1] see

Fig. 3.10a . A mode can be classified as a TE or TM mode depending on
which of its electric and magnetic fields is perpendicular to the yz-plane.
At any fixed y, the modes have constant intensities and their wave fronts
are flat. Thus, the modes can be seen as plane waves that are modulated
in the y-direction.
Optical properties of metamaterial slab waveguides can be character-
ized by the mode parameters, i.e., the effective refractive index nm and
impedance Zm experienced by the waveguide modes. The refractive in-
dex, also known as the mode index, characterizes the wave vector of the
mode β = x̂βx + ẑβz . The mode index is determined by the wave number

β = βx2 + βz2 as nm = β/k0 , where k0 is the wave number at the mode
wavelength in free space. In general, nm differs from the waveguide core
index. The impedance of the mode can be calculated by dividing the com-
plex amplitude of its transverse electric field by that of the transverse
magnetic field. However, the impedance defined in this way is generally a
function of y. Therefore, for a mode, the impedance must be defined as the
ratio between the field amplitudes averaged over the y-axis. This mode
impedance depends on the spatial distribution of energy of the mode and
is different for different modes.
The key difference between an ordinary dielectric and a metamaterial
slab waveguide is the presence of nanoscatterers in the metamaterial
waveguides. The interaction of light with these scatterers leads to excita-
tion of electromagnetic multipoles in them, which in turn influences the
mode propagation and absorption. These characteristics can be controlled
by a proper design of the scatterers. It should also be noted that the scat-
tered light can be divided into two parts, one trapped in the waveguide and
the other escaping from it. The latter contributes to the mode attenuation
along with the absorption by the scatterers.
Let us for simplicity consider a single-mode metamaterial slab waveguide.
If the mode propagating in such a waveguide does not strongly interact
with the metamaterial nanoscatterers, it is similar to a mode of the same
waveguide, but without the scatterers. Therefore, the mode of such a single-

52
Interaction of light with optical nanomaterials

mode waveguide must have a single intensity maximum in the waveguide


core and evanescent tails in the claddings. The situation can change if the
interaction with the scatterers considerably increases. In general, unlike in
a single-mode dielectric waveguide, the mode parameters in a single-mode
metamaterial waveguide depend on the mode propagation direction. In
addition, these parameters are complex-valued. In spite of this, the modes
of a metamaterial slab waveguide can be quite well matched to that of an
ordinary dielectric waveguide.
Let us now consider a segment of a metamaterial slab waveguide that
connects two identical dielectric slab waveguides as shown in Fig. 3.10a.
Optical modes propagating from one dielectric waveguide to the other one
are scattered by the metamaterial. Light scattered back and forward can,
respectively, be considered as light reflected and transmitted by the meta-
material waveguide. Inside the metamaterial waveguide, light is reflected
by the waveguide facets, and therefore, the waveguide acts analogously to
 
a Fabry-Perot etalon see Fig. 3.10b .
The fields reflected and transmitted by the metamaterial waveguide can
be calculated numerically, e.g., using the COMSOL Multiphysics software.
The incident mode can be generated by a current distribution in the first
dielectric waveguide at some distance from the metamaterial waveguide.
The mode reflection and transmission coefficients, r and t, can be defined in
analogy with the corresponding Fabry-Perot coefficients dealing with plane
waves. Defined as such, the coefficients r and t are the complex-amplitude
ratios between the incident, reflected, and transmitted modes. In terms of
the numerically calculated total electric-field distribution ET (x, y, z), we
write
∞
−∞ b (y) v̂ · ET (0, y, −W/2 − dr ) dy −iβz dr
r= ∞ e − v̂ · û e−2iβz dr (3.14)
Ei −∞ b2 (y) dy

and ∞
−∞ b (y) û · ET (0, y, W/2 + dt ) dy −iβz dt
t= ∞ e . (3.15)
Ei 2
−∞ b (y) dy
Here the fields are averaged over y. The coefficients are evaluated at dis-
tances dr and dt from the metamaterial waveguide, as shown in Fig. 3.10b.
The width of the waveguide is W . The wave vector component of the inci-
dent mode along the z-axis is βz . The unit vectors û and v̂ point along the
electric-field components of the incident and reflected modes, respectively.

53
Interaction of light with optical nanomaterials

  
a b c
90° Imax
120° 60°
3
150° 2 30°
1
nr
180° 0

ZnO
210° 330°
SiO2 Ag 300°
240° 270° 0

Figure 3.11. Diffraction compensation in a metamaterial slab waveguide. The structure


   
of the waveguide is shown in a . The polar plot b shows the real part of
the mode index, nr , of the waveguide as a function of the mode propagation
angle. A cross-sectional intensity distribution of a beam propagating along
 
the waveguide is shown in c . The dark rectangles represent the nanorods.

The y-dependence of these components is given by the mode profile function


b (y). The complex amplitude of the incident mode is Ei .
The question is whether the mode parameters of a metamaterial slab
waveguide can be retrieved from its reflection and transmission coeffi-
 
cients Eqs. (3.14) and (3.15), respectively . In the thesis, we show that

this retrieval is indeed possible, if the mode profiles b (y) and a (y) see

Fig. 3.10a are approximately equal. Otherwise, the mode mismatch leads
to an additional loss that is seen as an increased imaginary part of the
mode refractive index in the metamaterial waveguide and a corresponding
error in the retrieved value of the impedance. The core index ncore of the
dielectric waveguide should therefore be adjusted in the numerical calcula-
tions to minimize the difference between a (y) and b (y). The accuracy of
the retrieval is maximized if the mode indices of the dielectric and metama-
terial waveguides are equal to each other. To make this happen, we use an
iterative method. The index ncore is first simply guessed, e.g., based on the
knowledge of the metamaterial’s host medium or the refractive index of the
bulk metamaterial. With the guessed value of ncore , we retrieve the mode
parameters, from which we obtain the effective index of the metamaterial
core using the dispersion relation of a slab waveguide [74]. Then, the ob-
tained index can be used as a more accurate guess for ncore in the retrieval
to evaluate more accurate values of the mode parameters. By repeating
this procedure, the difference between the successively evaluated mode

54
Interaction of light with optical nanomaterials

parameters becomes negligible along with the error in their calculation.


In the thesis, a new type of a metamaterial slab waveguide that sup-
presses the effects of optical diffraction is introduced. In the design of
this waveguide, we utilize both the theory of the previous section and the
aforementioned characterization methods. The waveguide is composed of a
thin metamaterial layer sandwiched between significantly thicker layers of
   
glass SiO2 and air see Fig. 3.11a . The metamaterial layer is composed
 
of a single layer of silver nanorods in zinc oxide ZnO . The nanorods
form a rectangular array with periods Λx = 120 nm and Λz = 180 nm
in the x- and z-directions, respectively. The dimensions of the rods are
130 nm in the z-direction and 30 nm in the x- and y-directions. The ZnO
layer has a thickness of 120 nm. The material is designed for TE-polarized
modes. The modes resemble in their profiles the fundamental TE mode of
an ordinary dielectric slab waveguide, but the mode parameters depend on
 
the mode propagation angle ϕ see Fig. 3.10b . The polar plot in Fig. 3.11b
 
shows the real part of the mode index nr as a function of this angle at a
wavelength of 1310 nm. The imaginary part of the mode index is very low
 
c.a. 5.1 × 10−4 at ϕ = 0 and 1.4 × 10−2 at ϕ = 30◦ and not shown in the
figure. The contour of nr is flat over a significant angular range around the
directions ϕ = 0 and ϕ = 180◦ . Therefore, guided light beams propagating
in these directions will experience suppressed diffraction. Physically, the
diffraction compensation effect results from the multipole excitations in
the nanorods in the same way as the diffraction compensation in a bulk
metamaterial considered in the previous section. Figure 3.11c shows a
numerically calculated cross-sectional intensity distribution of a beam that
propagates in the waveguide. The beam 1/e2 diameter is only 300 nm.
Clearly, the beam divergence is negligible. To our surprise we have found
that the effect takes place over a 90-nm wide wavelength range around the
wavelength of 1310 nm. Therefore, the waveguide is promising for applica-
tions in photonic integrated circuits, in which a high-rate modulation of
optical signals, on the THz scale, is to be used.

55
Interaction of light with optical nanomaterials

56
4. Summary and outlook

The results of the thesis can be used to better understand, characterize,


and design spatially dispersive nanostructures. Characterization methods
for metasurfaces, metamaterial slab waveguides, and bulk metamaterials
have been considered. In particular, we have applied the electromagnetic
multipole theory to design and characterize a bifacial metasurface with
dominating quadrupole optical response. Such strong higher-order mul-
tipole excitations lead to significant spatial dispersion. As a result, the
reflectance, absorption, and local-field enhancement exhibited by the struc-
ture depend strongly on light propagation direction. We have also shown
that many spatially dispersive materials can be accurately characterized
by the wave parameters, such as the refractive index and impedance evalu-
ated for each propagation direction and polarization of the wave separately.
The validity of the method has been verified by direct numerical calcula-
tions.
In analogy with the wave parameters, we have introduced the mode
parameters for spatially dispersive metamaterial slab waveguides. These
parameters are the refractive index and impedance experienced by the
waveguide modes. They both can be retrieved from the reflection and
transmission coefficients of a segment of the metamaterial slab waveguide
for any relevant propagation direction of the mode. Hence, these parame-
ters pave a way to comprehensive characterization and design of spatially
dispersive metamaterial slab waveguides. We have found that the electric
and magnetic field profiles of the same mode can differ from each other,
when magnetic excitations in the structural units are pronounced. In such
cases, the impedance cannot be defined unambiguously. Because of this,

57
Summary and outlook

the two mode parameters do not fully characterize such strongly spatially
dispersive waveguides. The theory should therefore be further generalized
to such waveguides. We also think our findings can be used to retrieve
effective mode parameters of other types of nanostructured waveguides.
We have also noticed that, due to both optical anisotropy and spatial
dispersion, most optical metamaterials do not support polarization modes,
which for bulk metamaterials are plane waves with preserved polarization.
Instead, any possible polarization state of any mode in the material con-
stantly changes upon propagation. Obviously, such materials cannot be
 
characterized by the mode or wave parameters. In the presence of such
unavoidable polarization conversion, however, a slab of the metamaterial
can still be described by the slab reflection and transmission matrices that
take into account the energy transfer between the othogonal polarization
states of the incident wave. Almost any metamaterial slab can be charac-
terized by these matrices. Furthermore, if the near-field coupling between
the adjacent metamolecular layers of the material is weak, the reflection
and transmission matrices of any multilayered metamaterial slab can be
calculated from the single-layer matrices.
We also succeeded to model the interaction of optical beams with spa-
tially dispersive metamaterials, using a plane-wave decomposition method
described in Publication I. In the developed method, the incident beam is
decomposed into plane waves and the field distributions of the reflected and
transmitted beams are calculated in terms of their reflected and transmit-
ted plane-wave components. Also propagation of a beam inside the material
can be studied using this method. The method provides fast and physi-
cally insightful alternative to direct numerical simulations that become
impractical when studying the interaction of light with large metamaterial
structures. In addition, the method can be used to design spatially disper-
sive metamaterial components, such as high-numerical-aperture lenses
and advanced diffractive optical elements. As an example, we have de-
signed an apertureless spatial filter whose operation is independent of the
focal-spot position of the filtered beam. The filter is a metamaterial slab
that efficiently reflects only the normally incident plane-wave components
of the beam, transmitting and absorbing the other components. The filter
can be used to enhance the spatial coherence of light or to generate hollow

58
Summary and outlook

beams. In the future, this and other similar optical components can be
further developed and optimized also experimentally.
We have applied our theoretical tools to design diffraction-compensating
optical metamaterials, in which optical beams do not diverge upon propa-
gation. Compared to similar photonic-crystal-band structures, our struc-
tures are designed to exhibit very high optical coupling efficiency due
to negligible relection loss on the material surfaces. We also designed a
diffraction-compensating metamaterial slab waveguide with a single layer
of silver nanorods in the core. The bandwidth of diffraction compensation
in the waveguide is 90 nm implying that optical signals propagating in the
waveguide without spreading can be modulated at a rate of several THz.
The designed optical metamaterials can find applications in high-resolu-
tion imaging systems, future solar cells, low-loss photonic integrated cir-
cuits, and novel laser cavities. In addition, some of the demonstrated
unusual optical properties of our metamaterials can lead to completely
new types of applications. For example, one of our diffraction-compensating
metamaterials was shown to act as an unusual polarizer that transmits any
beam-like field or an optical image without spreading, making it composed
of TM-polarized waves only. In the future, new diffraction-compensating
metamaterials can be designed to provide divergence-free propagation also
for sub-diffraction-limited beams. For all metamaterials proposed in this
work, the most important development phase is to be their experimental
demonstration. The existing nanofabrication techniques, however, do not
always meet the precision requirements dictated by the proposed nano-
structure designs. We hope that this situation will change already in the
near future and many of the designed nanostructures will be released to
real-life applications.

59
Summary and outlook

60
Bibliography

[1] Landau L D and Lifshitz E M 1984 Electrodynamics of Continuous Media


(Oxford: Pergamon Press)

[2] Alu A and Engheta N 2008 Phys. Rev. B 78 085112

[3] Tserkezis C, Papanikolaou N and Gantzounis G Stefanou N 2008 Phys. Rev.


B 78 165114

[4] Ginn J and Brener I 2012 Phys. Rev. Lett. 108 097402

[5] Belov P, Marqués R, Maslovski S, Nefedov I, Silveirinha M, Simovski C and


Tretyakov S 2003 Phys. Rev. B 67 113103

[6] Cabuz A, Felbacq D and Cassagne D 2008 Phys. Rev. A 77 013807

[7] Gompf B, Krausz B, Frank B and Dressel M 2012 Phys. Rev. B 86 075462

[8] Grahn P, Shevchenko A and Kaivola M 2013 Opt. Express 21 23471

[9] Pendry J B 2000 Phys. Rev. Lett. 85 3966

[10] Fang N, Lee H, Sun C and Zhang X 2005 Science 308 534

[11] Wood B and Pendry J 2006 Phys. Rev. B 74 115116

[12] Belov P and Hao Y 2006 Phys. Rev. B 73 113110

[13] Zhao Y, Belov P and Hao Y 2009 J.Opt 11 075101

[14] Casse B D F, Lu W T, Huang Y J, Gultepe E, Menon L and Sridhar S 2010


Appl. Phys. Lett. 96 023114

[15] Pastuszczak A and Kotynski R 2011 J. Appl. Phys. 109 084302

[16] Pendry J B, Schurig D and Smith D R 2006 Science 312 1780

[17] Alu A and Engheta N 2008 J. Opt 10 093002

[18] Ni X, Wong Z J, Mrejen M, Wang Y and Zhang X 2015 Science 349 1310

[19] Monti A, Alu A, Toscano A and Bilotti F 2015 J. Appl. Phys. 117 123103

61
Bibliography

[20] Engheta N 2013 Science 340 286

[21] Moitra P, Yang Y, Anderson Z, Kravchenko I I, Briggs D P and Valentine J


2013 Nature Photon. 7 791

[22] Yanget J J, Francescato Y, Maier S A, Mao F and Huang M 2014 Opt.


Express 22 9107

[23] Agranovich V M and Ginzburg V 2013 Crystal optics with spatial dispersion,
and excitons (Berlin: Springer Science and Business Media)

[24] Smith D R, Schultz S, Markos P and Soukoulis C M 2012 Phys. Rev. B 65


195104

[25] Smith D R, Vier D C, Koschny T and Soukoulis C M 2005 Phys. Rev. E 71


036617

[26] Menzel C, Rockstuhl C, Paul T, Lederer F and Pertsch T 2008 Phys. Rev. B
77 195328

[27] Kwon D, Werner D H, Kildishev A V and Shalaev V M 2008 Opt. Express


16 11822

[28] Andryieuski A, Malureanu R and Lavrinenko A V 2010 Opt. Express 15


15498

[29] Yang J, Sauvan C, Paul T, Rockstuhl C, Lederer F and Lalanne P 2010 Appl.
Phys. Lett. 97 061102

[30] Grahn P, Shevchenko A and Kaivola M 2013 New J. Phys. 15 113044

[31] Shevchenko A, Grahn P and Kaivola M 2014 J. Nanophoton. 8 083074

[32] Grahn P, Shevchenko A and Kaivola M 2012 New J. Phys. 14 093033

[33] Lavrinenko A V, Legsgaard J, Gregersen N, Schmidt F and Sondergaard T


2015 Numerical Methods in Photonics (London: Taylor and Francis Group)

[34] Belov P, Zhao Y, Tse S, Ikonen P, Silveirinha M, Simovski C, Tretyakov S,


Hao Y and Parini C 2008 Phys. Rev. B 77 193108

[35] Paul T, Rockstuhl C, Menzel C and Lederer F 2009 Phys. Rev. B 79 115430

[36] Paul T, Menzel C, Rockstuhl C and Lederer F 2010 Adv. Mater. 22 2354

[37] Kivijärvi V, Nyman M, Shevchenko A and Kaivola M 2016 J.Opt 18 035103

[38] Kivijärvi V, Nyman M, Shevchenko A and Kaivola M 2016 Opt. Express 24


9806

[39] Kosaka H, Kawashima T, Tomita A, Notomi M, Tamamura T, Sato T and


Kawakami S 1999 Appl. Phys. Lett. 74 1212

[40] Lu Z, Shi S, Murakowski J, Schneider G, Schuetz C and Prather D 2006


Phys. Rev. Lett. 96 173902

62
Bibliography

[41] Prather D, Shi S, Murakowski J, Schneider G, Sharkawy A, Chen C, Miao


B and Martin R 2007 J. Phys. D: Appl. Phys. 40 2635

[42] Park W 2010 Laser Phys. Lett. 7 93

[43] Belov P and Simovski C 2005 Phys. Rev. B 71 193105

[44] Shadrivov I V, Sukhorukov A A and Kivshar Y S 2003 Phys. Rev. E 67


057602

[45] He J and Sailing H 2006 IEEE Microw. Wirel. Compon. Lett. 16 96

[46] Cheng Q and Tie J C 2006 J. Opt. Soc. Am. A 23 1989

[47] Huang Y, Lu W T and Sridhar S 2008 Phys. Rev. A 77 063836

[48] Jiang T, Zhao J and Feng Y 2009 Opt. Express 17 170

[49] He Y, He S, Gao J and Yang X 2012 J. Opt. Soc. Am. B. 29 2559

[50] Babicheva V E, Shalaginov M Y, Ishii S, Boltasseva A and Kildishev A V


2015 Opt. Express 23 31109

[51] Prasad P 2004 Nanophotonics (New Jersey: John Wiley & Sons)

[52] Franssila S 2010 Introduction to Microfabrication (Chichester: John Wiley


& Sons)

[53] Hu J, Chen Q, Xie Z, Han G, Wang R, Ren B, Zhang Y, Yang Z and Tian Z
2004 Adv. Funct. Mater. 14 183

[54] Wang I, Chen P and Liu M 2006 Nanotechnology 17 6000

[55] Yu Y, Chen C and Wang C 2008 J. Phys. Chem. 112 4176

[56] Wang Y and Asefa T 2010 Langmuir 26 7469

[57] Huang T, Cao W, Elsayed-Ali H and Xu X 2012 Nanoscale 4 380

[58] Shevchenko A, Kivijärvi V, Grahn P, Kaivola M and Lindfors K 2015 Phys.


Rev. Appl. 4 024019

[59] Kalkbrenner T, Ramstein M, Mlynek J and Sandoghdar V 2001 J. Microsc.


202 72

[60] Stranahan S and Willets K 2010 Nano Lett. 10 3777

[61] Huang X, El-Sayed I, Qian W and El-Sayed M 2006 J. Microsc. 128 2115

[62] Jain P, El-Sayed I and El-Sayed M 2007 Nano Today 2 18

[63] Rai M, Yadav A and Gade A 2009 Biotechnol. Adv. 27 76

[64] Ramyadevi J, Jeyasubramanian K, Marikani A, Rajakumar G and Rahu-


man A 2012 Mater. Lett. 71 114

[65] Lee J and Peumans P 2010 Opt. Express 18 10078

63
Bibliography

[66] Atwater H and Polman A 2010 Nat. Mater. 9 205

[67] Linic S, Christopher P and Ingram D 2011 Nat. Mater. 10 911

[68] Neumann O, Urban A, Day J, Lal S, Nordlander P and Halas N 2013 ACS
Nano 7 42

[69] Maier S, Kik P and Atwater H 2003 Phys. Rev. B 67 205402

[70] Krenn J 2003 Nat. Mater. 2 210

[71] Lal S, Link S and Halas N 2007 Nat. Mater. 1 641

[72] Derkacs D, Lim S H, Matheu P, Mar W and Yu E T 2006 Appl. Phys. Lett.
89 093103

[73] Soukoulis C and Wegener M 2011 Nat. Mater. 5 523

[74] Saleh B E A and Teich M C 2007 Fundamentals of Photonics 2nd ed (New


Jersey: Wiley)

[75] Johnson P and Christy R 1972 Phys. Rev. B. 6 4370

[76] Bohm D and Pines D 1953 Phys. Rev. 92 609

[77] Otto A 1968 Z. Phys. 216 398

[78] Kretchmann E and Raether H 1968 Z. Naturforsch. 23 2135

[79] Novotny L and Hecht B 2012 Principles of Nano-Optics 2nd ed (Cambridge:


Cambridge University Press)

[80] Steele J M, Liu Z, Wang Y and Zhang X 2006 Opt. Express 14 5664

[81] Maier S 2007 Plasmonics: Fundamentals and Applications (New York:


Springer Science + Business Media)

[82] Hergert W and Wriedt T 2012 The Mie Theory: Basics and Applications
(Berlin: Springer Science + Business Media)

[83] Edmonds A R 1996 Angular Momentum in Quantum Mechanics (Princeton,


New Jersey: Princeton University Press)

[84] Stratton J 1941 Electromagnetic theory (New York: McGraw-Hill)

[85] Jackson J D 1999 Classical Electrodynamics (United States of America:


John Wiley & Sons)

[86] Decker M, Staude I, Shishkin I I, Samusev K B, Parkinson P, Sreenivasan


V K A, Minovich A, Miroshnichenko A E, Zvyagin A, Jagadish C, Neshev
D N and Kivshar Y S 2013 Nat. Commun. 4 2949

[87] Nyman M, Kivijärvi V, Shevchenko A and Kaivola M 2017 Phys. Rev. A 95


043802

64
Bibliography

[88] Kivijärvi V, Nyman M, Karrila A, Grahn P, Shevchenko A and Kaivola M


2015 New J. Phys 17 063019

[89] Kildishev A V, Boltasseva A and Shalaev V 2013 Science 339 1232009

[90] Walther B, Helgert C, Rockstuhl C and Pertsch T 2011 Appl. Phys. Lett. 98
191101

[91] Aieta F, Genevet P, Kats M A, Yu N, Blanchard R, Gaburro Z and Capasso


F 2012 Nano Lett. 12 4932

[92] Zhao Y, Belkin M A and Alu A 2012 Nat. Commun. 3 870

[93] Kang M, Feng T, Wang H and Li J 2012 Opt. Express 20 15882

[94] Yu N, Aieta F, Genevet P, Kats M A, Gaburro Z and Capasso F 2012 Nano


Lett. 12 6328

[95] Li X, Pu M, Zhao Z, Ma X, Jin J, Wang Y, Gao P and Luo X 2016 Sci. Rep. 6
20524

[96] Litchinitser N M 2012 Science 337 1054

[97] Smalley J S T, Vallini F, Montoya S A, Ferrari L, Shahin S, Riley C T, Kante


B, Fullerton E E, Liu Z and Fainman Y 2016 Nat. Commun. 8 13793

[98] Chau Y C, Syu J, Chao C C, Chiang H and Lim C M 2016 J. Phys. D: Appl.
Phys. 50 045105

[99] Gwo S, Wang C, Chen H, Lin M, Sun L, Li X, Chen W, Chang Y and Ahn H
2016 ACS Photonics 3 1371

[100] Zou S and Schatz G C 2005 Chem. Phys. Lett. 403 62

[101] Nikitin A G, Kabashin A V and Dallaporta H 2012 Opt. Express 20 27941

[102] de Hoop A T 1960 Appl. Sci. Res., Section B 8 135

[103] Sounas D L and Alu A 2014 Opt. Lett. 39 4053

[104] Tsakmakidis K, Boardman A D and Hess O 2007 Nature 450 397

[105] Chen H, Chan C T and Sheng P 2010 Nat. Mater. 9 387

[106] Ergin T, Stenger N, Brenner P, Pendry J B and Wegener M 2010 Science


328 337

[107] Liu Z, Lee H, Xiong Y, Sun C and Zhang X 2007 Science 315 1686

[108] Yang X, Yao J, Rho J, Yin X and Zhang X 2012 Nat. Photonics 6 450

[109] Markel V A 2016 J. Opt. Soc. Am. A 33 1244

[110] Doyle W T 1989 Phys. Rev. B 39 9852

[111] Ruppin R 2000 Opt. Commun. 182 273

65
Bibliography

[112] Bohren C F 2009 J. Nanophotonics 3 039501

[113] Smith D R and Pendry J B 2006 J. Opt. Soc. Am. B 23 391

[114] Simovski C R and Tretyakov S A 2007 Phys. Rev. B 75 195111

[115] Andryieuski A, Ha S, Sukhorukov A A, Kivshar Y S and Lavrinenko A V


2012 Phys. Rev. B 86 035127

[116] Shevchenko A, Nyman M, Kivijärvi V and Kaivola M 2017 Opt. Express 25


8550

[117] Capolino F 2009 Theory and Phenomena of Metamaterials (Boca Raton:


CRC Press)

[118] Mortensen N A, Raza S, Wubs M, Sondergaard T and Bozhevolnyi S I 2014


Nat. Commun. 5 3809

[119] Chebykin A V, Gorlach M A and Belov P A 2015 Phys. Rev. B 92 045127

[120] Acher O, Adenot A L and Duverger F 2000 Phys. Rev. B 62 13748

[121] Matsushima K, Schimmel H and Wyrowski F 2003 J. Opt. Soc. Am. A 20


1755

[122] Goodman J W 1996 Introduction to Fourier optics 2nd ed (Singapore:


McGraw-Hill)

[123] Greene P L and Hall D G 1999 Opt. Express 4 411

[124] Pu J and Lü B 2001 J. Opt. Soc. Am. A 18 2760

[125] Dorn R, Quabis S and Leuchs G 2003 Phys. Rev. Lett. 91 233901

[126] Ciocci L, Echarri R M and Simon J M 2010 J. Opt. 12 015408

[127] Li Q, Ledoux-Rak I and Lai N D 2015 Advanced Device Materials 1 4

[128] Wolf E 1959 Proc. R. Soc. Lond. A 253 349

[129] Carney P S and Schotland J C 2002 J. Opt. A: Pure Appl. Opt. 4 S140

[130] Abouraddy A F and Toussaint K C J 2006 Phys. Rev. Lett. 96 153901

[131] Lazar J, Bondar A, Timr S and Firestein S J 2011 Nat. Methods 8 684

[132] Liu T, Tan J, Liu J and Wang H 2013 Opt. Express 21 15090

[133] Oldenbourg R 1996 Nature 381 811

[134] Biss D P and Brown T G 2004 Opt. Express 12 384

[135] Serrels K A, Ramsay E, Warburton R J and Reid D T 2008 Nat. Photonics 2


311

[136] Ashkin A, Dziedzic J M and Yamane T 1989 Nature 330 769

66
Bibliography

[137] Ashkin A and Gordon J P 1983 Opt. Lett. 8 511

[138] Ashkin A, Dziedzic J M, Bjorkholm J E and Chu S 1986 Opt. Lett. 11 288

[139] Ganic D, Gan X and Gu M 2005 Opt. Express 13 1260

[140] Jabbour T G and Kuebler S M 2008 Opt. Express 16 7203

[141] Kenny F, Lara D, Rodríguez-Herrera O G and Dainty C 2012 Opt. Express


20 14015

[142] Hsu W and Barakat R 1994 J. Opt. Soc. Am. A 11 623

[143] Richards B and Wolf E 1959 Proc. R. Soc. Lond. A 253 358

[144] McCutchen C W 1964 J. Opt. Soc. Am. 54 240

[145] Marathay A S and McCalmont J F 2001 J. Opt. Soc. Am. A 18 2585

[146] Menzel C, Paul T, Rockstuhl C, Pertsch T, Tretyakov S and Lederer F 2010


Phys. Rev. B 81 035320

[147] Sun L, Li Z, Luk T S, Yang X and Gao J 2015 Phys. Rev. B 91 195147

[148] Chuang Y and Suleski T 2010 J. Opt. 12 035102

[149] Liu H 2011 Metamaterials with hyperbolic dispersion relations and their
near-field and radiated-wave applications Ph.D. thesis Purdue University
West Lafayette, Indiana

[150] Zhao D, Zhou C, Gong Q and Jiang X 2008 J. Phys. D: Appl. Phys. 41 115108

[151] Witzens J, Lončar M and Scherer A 2002 IEEE J. Sel. Topics Quantum
Electron. 8 1246

[152] Iliew R, Etrich C and Lederer F 2005 Opt. Express 13 7076

[153] Shin J and Fan S 2005 Opt. Lett. 30 2397

[154] Chuang Y and Suleski T 2011 J. Opt. 13 035103

[155] Arlandis J, Centero E, Polles R, Moreau A and Campos J 2012 Phys. Rev.
Lett. 108 037401

[156] Rumpf R, Pazos J, Garcia C, Ochoa L and Wicker R 2013 Prog. Electromagn.
Res. 139 1

[157] Baba T and Ohsaki D 2001 Jpn. J. Appl. Phys. 40 5920

[158] Miao B, Chen C, Shi S and Prather D W 2005 IEEE Photonics Technol. Lett.
17 61

[159] Momeni B and Adibi A 2005 Appl. Phys. Lett. 87 171104

[160] Lee S, Choi J, Kim J and Park H Y 2008 Opt. Express 16 4270

[161] Park J, Lee S, Park H and Lee M 2010 Opt. Express 18 13083

67
Bibliography

[162] Ono A, Kato J and Kawata S 2005 Phys. Rev. Lett. 95 267407

[163] Kawata S, Ono A and Verma P 2008 Nat. Photonics 2 438

[164] Yao J, Liu Z, Liu Y, Wang Y, Sun C, Bartal G, Stacy A and Zhang X 2008
Science 321 930

[165] Belov P, Hao Y and Sudhakaran S 2006 Phys. Rev. B 73 033108

[166] Lombardet B, Dunbar L A, Ferrini R and Houdré 2005 J. Opt. Soc. Am. B
22 1179

[167] Shevchenko A, Grahn P, Kivijärvi V, Nyman M and Kaivola M 2015 J.


Nanophoton. 9 093097

[168] Kivijärvi V, Nyman M, Shevchenko A and Kaivola M 2016 Propagation of


optical fields through a three-dimensional diffraction-compensating meta-
material Advanced Electromagnetic Materials in Microwaves and Optics
(METAMATERIALS), 2016 10th International Congress on (IEEE) pp 184–
186

[169] Silveirinha M and Engheta N 2006 Phys. Rev. Lett. 97 157403

[170] Prather D W, Shi S, Pustai D M, Chen C, Venkataraman S, Sharkawy A,


Schneider G J and Murakowski J 2004 Opt. Lett. 29 50

68
Aalto-DD 17/2019

ISBN 978-952-60-8396-4 (printed) BUSINESS +


9HSTFMG*aidjge+

ISBN 978-952-60-8397-1 (pdf) ECON OMY


ISSN 1799-4934 (printed)
ISSN 1799-4942 (pdf) AR T +
DESIGN +
Aalto Universit y ARC HITECTURE
Sc hool of Scienc e
Depar tment of Applied Physics SCIEN CE +
www.aalto.fi TECHNOLOGY
CROSSOVER
D O C TO R A L
DISSERTATIONS

You might also like