Anda di halaman 1dari 150

The optical properties of metamaterial waveguide structures

by

ARVIN REZA

A thesis submitted to the Department of Physics, Engineering Physics, and Astronomy in conformity with the requirements for the degree of Master of Science

Queen's University Kingston, Ontario, Canada October,2008

Copyright c ARVIN REZA, 2008

Abstract

Metamaterials, articially engineered structures with negative average relative permittivity and permeability, provide a route to creating potential devices with exciting electromagnetic properties that cannot be obtained with natural materials. One particularly interesting metamaterial device, is a planar metamaterial waveguide structure (MWS) that has potentially exciting applications. In this thesis, the properties and potential applications of metamaterial waveguide structures are explored. In particular, we examine the properties of metamaterial waveguides when the limitations arise from fabrication techniques and physical principles are taken into account. First, we study the basic properties of dispersion curves of an idealized (without loss and dispersion) metamaterial waveguide structure. We show that there are a

rich variety of modes, such as the bound modes, the surface polariton modes, and the complex leaky modes, that are supported in MWS and have entirely dierent properties than the modes of a conventional waveguide structure. These novel modes provide more control over the electromagnetic elds and consequently lead to potential applications ranging from waveguide miniaturization to the slowing down of the light. Next, we study the eects of dispersion and loss, which are the inherent features of metamaterials, on the properties of MWS. We numerically show that the characteristic modes of the MWS are signicantly changed particularly near the slow-light-modes i

when the intrinsic loss is introduced into the system. In particular we show that the stopped-light-modes disappear even in the presence of an arbitrarily small amount of loss. Moreover, we nd several novel properties such as a splitting of complex leaky modes in a lossy MWS.

ii

Co-Authorship

Some of the results in Chapter 3 have been published in Nature [49]. created by the author unless otherwise cited.

Figures are

iii

Acknowledgments

The author would like to express his sincere gratitude hereby to his parents and his supervisors, Professor Marc Dignam and Professor Stephen Hughes, who have provided him the best of support, supervision and insight.

iv

Contents

Abstract

Co-Authorship

iii

Acknowledgments

iv

Contents

List of Tables

viii

List of Figures

ix

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

1.1 1.2 1.3 1.4 1.5

Introduction to metamaterials . . . . . . . . . . . . . . . . . . . . . . Fabrication techniques of metamaterials . . . . . . . . . . . . . . . .

1 4 7 11 14

Negative-index refraction in metamaterials . . . . . . . . . . . . . . . Metamaterial waveguiding . . . . . . . . . . . . . . . . . . . . . . . .

Thesis overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Chapter 2: Modal characteristics of MWS . . . . . . . . . . . . . . . 17

2.1 2.2 2.3 2.4

Introduction

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

17 18 24

Theoretical formulation of dispersion relation . . . . . . . . . . . . . . Orthogonality and normalization . . . . . . . . . . . . . . . . . . . .

Comparison between conventional and metamaterial modes of a slab waveguide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 30 43 45

2.5 2.6 2.7

Guided, surface polaritons, and complex surface modes

The TM modes of the lossy, dispersive metamaterial waveguide

Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Chapter 3: Slow light in MWS . . . . . . . . . . . . . . . . . . . . . . 46

3.1 3.2 3.3 3.4

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Negative energy density in dispersionless and lossless metamaterials .

46 48 52

Energy density in a dispersive and absorptive media . . . . . . . . . . Power ux, group velocity, and energy velocity in a lossy and dispersive metamaterial waveguide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

57 66 73

3.5 3.6

The role of intrinsic loss on slow-light modes Concluding remarks

. . . . . . . . . . . . . . . . . . . . . . . . . . .

Chapter 4: Excitation on the modes of MWS . . . . . . . . . . . . . 75

4.1 4.2 4.3 4.4

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Transfer matrix formalism for lossy metamaterial multilayer structures The Attenuated Total Reection technique (ATR) . . . . . . . . . . . Anomalous Goos-Hanchen shift due to surface mode excitation . . . .

75 76 82 94

vi

4.5

Concluding remarks

. . . . . . . . . . . . . . . . . . . . . . . . . . .

96

Chapter 5: Dispersion curves of MWS with complex-

. . . . . . .

98

5.1 5.2 5.3 5.4 5.5

Introduction

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

98 100 107 113 116

Comparison between

and

approaches . . . . . . . . . . . . . . . . . . . . . . . . .

The experimental conditions of the ATR method

Perturbation theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Chapter 6: Conclusions and suggestions for future work . . . . . . . 117

Bibliography

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

119

Appendix A: Orthogonality relations . . . . . . . . . . . . . . . . . . 130

vii

List of Tables

1.1

Negative-index refraction in THz metamaterials

. . . . . . . . . . . .

viii

List of Figures

1.1 1.2 1.3

Lycurgus Cup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Unusual optical eects of metamaterials . . . . . . . . . . . . . . . .

2 4 6

History of metamaterial fabrication . . . . . . . . . . . . . . . . . . .

2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9

Schematic geometry of a metamaterial waveguide structure . . . . . . Graphical solution of the TE modes . . . . . . . . . . . . . . . . . . . Dispersion curves of a dispersionless and lossless waveguide . . . . . . The

18 27 33 35 36 40 41 42 43

and

derived from Lorentz-Drude models versus frequency

. .

Dispersion curves of a dispersive LHM waveguide for TE modes

. . .

The surface polariton mode coupling with the higher order mode . . . The

T E1

and

T E5

modes coupling

. . . . . . . . . . . . . . . . . . .

The bending back of the dispersion curves near Permeability near the resonance frequency

fres

. . . . . . . . . .

. . . . . . . . . . . . . . .

2.10 Dispersion curves of the lossy dispersive metamaterial waveguide structure for the TM modes . . . . . . . . . . . . . . . . . . . . . . . . . . 44

3.1 3.2

Schematic of the double vortex-like energy ux structure

. . . . . . .

47

Power ux, energy density and energy velocity of the dispersionless waveguide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

ix

3.3

The energy velocity and the power ux for the backward propagating

T E5
3.4

and

T E1

coupling mode . . . . . . . . . . . . . . . . . . . . . . .

61

The energy velocity and the power ux for the forward propagating

T E5
3.5 3.6

and

T E1

coupling mode . . . . . . . . . . . . . . . . . . . . . . .

63 64

The energy velocity and power ux of the

T E2

mode

. . . . . . . . .

(a) The normalized energy velocity and (b) the propagation loss for the backward-propagation

T E2

mode versus frequency.

. . . . . . . . . . . . .

65 67

3.7 3.8

The schematic of the wedge-shaped metamaterial waveguide

Dispersion curves of the proposed waveguide structure in Fig. 3.7(b) with

= 0.01c

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

69

3.9

Dispersion curves of the proposed waveguide structure in Fig. 3.7(b) for TM modes with

= 0.01c .

. . . . . . . . . . . . . . . . . . . . .

70

3.10 The energy velocity and propagation loss of the proposed waveguide structure for

T M2

mode

. . . . . . . . . . . . . . . . . . . . . . . . .

71

3.11 The energy velocity of the proposed waveguide structure with and

= 0.01
73

= 0.0001

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4.1 4.2 4.3 4.4 4.5 4.6

Schematic geometry of the ATR method

. . . . . . . . . . . . . . . . . . . . . . . .

77 78 84 87 89

The schematic of a one-dimension multilayer structure.

Dispersion curves and prism coupling . . . . . . . . . . . . . . . . . . Eects of the air gap width, d, on the reectance. . . . . . . . . . . . . . . .

Reectance with an optimal air gap for dierent material loss

Comp arisen between the solutions obtained from the ART with the exact solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 91

4.7

The eect of the

of the prism on the dispersion curves . . . . . . . .

4.8 4.9

Optimum air gap and

nprism

for the

T E1

mode . . . . . . . . . . . . .

92

Schematic of the positive and negative lateral beam shift (Goos-Hanchen shift). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95 96

4.10 The lateral Goos-Hanchen (GH) shift for the reected beam

5.1

Dispersion curves of the dispersionless and lossless metamaterial slab in the complex- and in the complex- approaches . . . . . . . . . . 100

5.2

The dispersion curves of the dispersive, lossless metamaterial waveguide in the complex- and in the complex- approaches . . . . . . . . 101

5.3

The dispersion curves of the dispersive, lossy metamaterial waveguide for the

T E2

mode . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

103

5.4

Group velocity and loss propagation of the metamaterial waveguide for the

T E2

mode in the complex- approach. . . . . . . . . . . . . . . .

104

5.5

Comparison between the imaginary part of the exact obtained from Eq. 5.5

with the

106 108

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

5.6 5.7

Reectance versus the incident angle and versus frequency

Reectance versus the incident angle for the metamaterial waveguide with

= 0.1

GHz . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

111 112

5.8 5.9

The experimental conditions of the ATR method

Exact and perturbation solutions of a metamaterial waveguide structure115

xi

Chapter 1

Introduction

1.1 Introduction to metamaterials


In the last few years, the eld of metamaterials has attracted considerable attention in the scientic community due to the exciting potential applications ranging from perfect lenses [1] and cloaking devices [2, 3, 4] to sub-wavelength optical waveguides [5] and the enhancement of magnetic resonance imaging (MRI) [6]. Metamaterials are a class of composite materials articially fabricated to achieve unusual electromagnetic properties not found in nature. Their unusual properties are not determined by

their constituent materials but by man-made structures that are smaller than the electromagnetic wavelength involved. These structures that form the metamaterial act like articial atoms and can be coupled to both the electric and magnetic eld components of the electromagnetic waves, leading novel optical properties, such as a negative index of refraction. In this chapter, we provide a brief history of the

eld, a simple introduction to the fundamental properties and fabrication methods of metamaterials and an overview of the recent theoretical and experimental work in 1

CHAPTER 1.

INTRODUCTION

Figure 1.1: Lycurgus Cup (British Museum; AD fourth century) is made of ruby glass, the rst optical metamaterials. The cup has a greenish color when light reected from the glass; however, it has a reddish color when light transmitted through the glass. [source: Nature Photonics 1, 207-208 (2007).]

the eld, emphasizing the particular area of metamaterial waveguide structures. It is probably surprising to learn that metamaterials have a long history; for example, medieval ruby glass might be considered as the rst optical metamaterial [7]. Ruby glass contains nano-scale gold droplets (5-60 nm) that give an unusual

color to the glass. Actually, a resonance of the surface plasmon on the gold droplets causes the extraordinarily color of the glass. Ruby glass appears neither golden nor transparent, but depending on the size of the droplets and the direction of the light beam, looks red or green. In Fig. 1.1, Lycurgus cup, the Roman cup made of ruby glass, is shown. It has a greenish color when viewed in reected light, but it has a reddish color when light is transmitted through the glass. We can name plenty of metamaterials in the last century; for example, the work of

CHAPTER 1.

INTRODUCTION

Bose on the rotation of the plane of polarization by articial twisted structures in 1898 [8] or articial dielectrics for microwave antenna lenses by Kock in 1945 [9]. However, what we today call metamaterials was born at the beginning of this century (2000) in a lab at the University of California at San Diego [10] when D.R.Smith and his team presented evidence of both a negative electric permittivity and a negative magnetic permeability at a specic microwave range of frequencies in their experiment. The idea of metamaterials with negative permittivity and permeability was rst proposed by Victor Veselago in 1968 [12], who termed such media left-handed material (LHM) because the electric and magnetic elds and the wave vector form a lefthanded set of vectors rather than the usual right-hand set. In his study, Veselago

showed that the Poynting vector of an electromagnetic wave is anti-parallel to the wave vector in a left-handed material; thus, light propagates in the opposite direction to the energy ux. This leads to number of unusual properties in metamaterials such as negative-index refraction, the reversal of Cherenkov radiation, the reversal of the Doppler shift, and the reversal of Snell's law. These revolutionary properties give

remarkable control over electromagnetic elds with incredible possible applications such as invisibility devices. As a further consequence of negative-refraction, metamaterials produce unusual optical eects. For example, a bulb located in front of a metamaterial slab, would appear to be in front of the slab; a sh swimming in water made of metamaterial would appear to be moving above the water surface (See Fig 1.2). Negative-refraction could even lead to negative space [45]. For example, if a distance between two

objects is properly lled with a material with negative-refraction and a conventional material, the two objects would appear to be beside each other.

CHAPTER 1.

INTRODUCTION

Figure 1.2: The schematic of how a sh observed in water and in the water made of metamaterials. [source: http://www.berkeley.edu/news/media/download/]

1.2 Fabrication techniques of metamaterials


Some of the unusual and exciting properties of metamaterials were investigated a long time ago. For example, Sir Arthur Schuster showed negative phase velocity and its consequence in 1904 [13] and Dimitriy Sivukhin found that the energy ux and phase velocity are anti parallel in left-handed material in 1957 [14]. However, the

work of Veselago [12, 13, 14] is generally considered the rst comprehensive investigation of metamaterials and even it did not receive signicant attention in the scientic community until the rst modern metamaterial was fabricated in 2000 [10]. In fabricating the rst metamaterial with negative permeability and permittivity, Smith and his coworkers followed the idea of Sir John Pendry [11]. They designed a structure, composed of metallic rings and gaps, for creating negative permeability in a specic range of frequency. The proposed structure by Pendry, the so-called split ring resonator (SRR), is magnetically active though it is made of non-magnetic materials. To achieve simultaneously negative electric permittivity, he arranged metallic wires in periodic lattices with SRRs. By using a periodic arrangement of metallic wires and

CHAPTER 1.

INTRODUCTION

rings, Smith and his coworkers showed evidence of negative-index refraction in 2001 [15]. The rst metamaterials were demonstrated in microwave frequencies; however most potential applications are in the optical domain. It is a very challenging problem to extend metamaterials to optical frequency because this requires scaling the articial atoms from the millimeter size down to the nanometer size. However, the progress in nanotechnology and the thrust for optical metamaterials, made this goal possible, at least at infrared frequencies (780 nm) only ve years after the creation of the rst metamaterial [16]. Of course, some critical problems still remain unsolved. The key one is that most potential applications of metamaterials cannot be achieved if there are signicant losses in the metamaterials. [19, 20]. The high losses originate from the design required to create metamaterials and particularly are a critical problem at the infrared and optical frequencies since the metal required for fabrication departs from a perfect conductor in the optical domain. The losses of the best designs to date (see table 1.1) are too high for most of the potential applications to become practical (e.g. see [49]). Another problem in the fabrication of metamaterials is the development of truly large-area bulk 3D isotropic metamaterial structures. Although recently Valentine et al. reported to have successfully fabricated a 3D metamaterial structure at infrared frequencies [46], the proposed metamaterial structure is not a true 3D metamaterials because it only works for light coming from a narrow range of directions and with a single polarization [47]. Fig. 1.3 gives a brief history of the development of metamaterials. The early metamaterials were built from double split ring resonators and parallel wires. Changing the double SRR to a single one and reducing the size of the structure, the metamaterial frequency value reaches about 200 THz. However, by this approach, one cannot

CHAPTER 1.

INTRODUCTION

Double SRR

Single SRR

U-shape SRR

Plate pairs

Double -fishnet

Figure 1.3: History of metamaterial fabrication: From SRR to sh-net structure.

go beyond this frequency because the composition of SRR and parallel metallic wires are so sensitive at nano scale and it is very dicult to measure transmission and refraction along the direction parallel to the SRR plane; moreover, the metal of the SRR strongly departs from an ideal conductor at and above this frequency. Therefore, in 2005, several groups employed alternative designs to create metamaterials at higher frequencies [17, 18]. The new designs contained pairs of metal wires or plates which were separated by a dielectric spacer. Wire pairs can behave like SRRs and give a negative permeability,

over some frequency range; moreover, wire pairs can

simultaneously give a negative permittivity, both

However, nding the regions where

and

are negative with only wire pairs, is dicult. Therefore, a new design,

the so-called double-shnet, which contains a pair of metal shnets, was proposed in 2005 [19]. Using this structure, several groups reported negative-index metamaterials near the visible spectrum [16, 19, 20]; however, the proposed metamaterials suer from high losses that originate from the metal in the structure. To provide a

Birck Nanotechnology Center

CHAPTER 1.

Negative Refractive Index in Optics: State of the Art


1st time posted and publication Refractive index, n Wavelength Figure of Merit F=|n|/n

INTRODUCTION

Year and Research group

Structure used

2005:
Purdue
April 13 (2005) arXiv:physics/0504091 Opt. Lett. (2005) April 28 (2005) arXiv:physics/0504208 Phys. Rev. Lett. (2005)

0.3 2

1.5 m 2.0 m

0.1

Paired nanorods Nano-fishnet with round voids

UNM & Columbia

0.5

2006:
UNM & Columbia Karlsruhe & ISU Karlsruhe & ISU
J. of OSA B (2006) OL. (2006) OL (2007) OL (2006)

4 1 1 0.6 0.9 -1.1

1.8 m 1.4 m 1.4 m 780 nm 770 nm 810nm

2.0 3.0 2.5 0.5 0.7 1.3

Nano-fishnet with round voids Nano-fishnet 3-layer nanofishnet Nano-fishnet

Purdue

OL (2007)

Nano-fishnet

see review: Nature Photonics v. 1, 41 (2007)


Table 1.1: Negative-index refraction and gure of merit in THz metamaterials [96].
5

picture of the status of metamaterial at present, we summarize the recent successful observations of negative-index refraction with corresponding gure of merit by active groups in this eld in table 1.1. The gure of merit (FOM) is a suitable measure for the loss dened as the ratio of the real to the imaginary part of the refractive index. So far, the largest FOM for infrared frequencies is around 1 and for THz frequencies is around 3.

1.3 Negative-index refraction in metamaterials


We have already stated that one of the unusual properties of the metamaterials is the negative-index of refraction. Several experimental studies have been reported

conrming this property in metamaterials (e.g. [15]); moreover, several groups have

CHAPTER 1.

INTRODUCTION

mathematically shown that the index of refraction should be negative in metamaterials (e.g. [27]). In this section, for the sake of completeness, we show why the index of refraction is negative in metamaterials. We consider a plane wave propagating in a metamaterial, with electric and magnetic elds

E(r, t) = Re{E exp[i(k r t)]} H(r, t) = Re{H exp[i(k r t)]}.


From the Maxwell curl equations, we have

(1.1)

k E = i H, c
where

k H = i E, c

(1.2)

and

are the relative electric permittivity and magnetic permeability, re-

spectively and

is the wave vector. It is dened by

k2 =

2 . c2 k

(1.3)

As evidenced from Eq. (1.2), and

E, H,

and

form a left-handed set of vectors when

are simultaneously negative, and form a right-handed set of vectors when and

are positive. However, it is meaningless to dene handedness of the system when

only

or is negative because the wave vector is imaginary in this case.

Next, we will

show that the phase velocity of a wave (represented by the wave vector) is opposite to the movement of the energy ux (represented by the time-averaged Poynting vector) of the wave in a metamaterial medium. We consider an electric eld polarized in the

p direction, E = p.

Using Maxwell's

CHAPTER 1.

INTRODUCTION

equations and Poynting theorem, the time-average Poynting vector is

1 S = Re(E H ), 2 = c p (k p), 2 ck . 2

(1.4)

The result shows that the time-averaged Poynting vector is in the opposite direction of the wave vector if electric eld when

is negative.

We can also show the same result for a s-polarized

is negative. Thus, in a metamaterial with both negative

and

, the phase velocity of a wave is always opposite to the movement of the energy ux
of the wave. It is not evident from

n2 =
that having negative

k2 = 2 k0

(1.5)

and

leads to negative index of refraction.

Both negative

and positive index of refraction solutions satisfy Maxwell's equations and boundary conditions for refraction of light. However, the negative solution for the index of

refraction is selected in metamaterials to satisfy the requirement that in the refracted beam the energy ows away from the interface. The choice of negative refractive

index in metamaterials is discussed in detail in the literature (e.g. [27]). Here, we provide a simple argument to demonstrate negative index property in a passive lossy metamaterial.

CHAPTER 1.

INTRODUCTION

10

In a lossy metamaterial,

and

are complex; thus, we can write

= + i = + i ,
where

(1.6)

< 0, < 0

and

> 0, > 0

because we assume that a metamaterial

medium is absorbing. Using Eq. (1.5), the index of refraction for a passive medium with small absorption is

n = = ( ) + i( + )
+ [ + i 2 ].
The (1.7)

< 0, < 0

and

> 0, > 0

conditions lead to

+ < 0. 2

(1.8)

Since the imaginary component of refraction index is positive in a lossy medium, Eq. (1.8) demands that the negative sign of the square root in Eq. (1.7) be chosen. Therefore, the real component of index of refraction must be negative if both

and

are negative. In the reminder of this chapter, we provide a brief overview of some of the po-

tential applications of metamaterials proposed in the literature, emphasizing those applications related to metamaterial waveguide structures and the multilayer systems. It is not our aim to give a picture of all of the applications of metamaterials, but with our examples from the recent literature, we will provide a background to better understand the results of the following chapters.

CHAPTER 1.

INTRODUCTION

11

1.4 Metamaterial waveguiding


In 2000, when Smith et al. reported the rst modern metamaterial, Ruppin theoretically showed that a simple planar metamaterial waveguide can support surface polariton modes [21], modes with an evanescent prole decay from the interface. These type of modes had been observed at metal surfaces and had been used for near-eld imaging. However, these modes have richer properties in metamaterials;

for example, the metal waveguide can guide only transverse magnetic, TM, waves while the metamaterial slab can guide both polarizations of light (the evidence of TE surface polariton modes was found in 2005 by Smith [22]). Thus, the existence of surface polariton modes in metamaterials opened the door for further applications [23]. Ruppin's work, drew more attention to metamaterial waveguide structures [23, 24]. Further study showed that not only can the metamaterial waveguide support surface polariton modes but also there are new types of modes (complex surface or leaky modes) with immediate applications to leaky antennas [26]. Moreover, it was shown that the properties of the normal guided modes of the metamaterial waveguide are completely dierent than in a conventional waveguide. For example, Nader Engheta and Andrea Alu showed that there is no cut-o thickness for the rst mode of metamaterials waveguide. Thus, this feature provides a solution to the problem of energy transmission with lateral cross section below the diraction limits [5]. Moreover, Yuri Kivshar and his coworkers showed that the fundamental mode (the mode with no node in the modal prole) does not exist in a metamaterial waveguide. By using

this feature, we can introduce the concept of surface wave suppression which leads to enhance radiation eciency for microstrip antennas [26] or introduce a one dimensional periodic waveguide with a complete band gap [28]. Usually, we require 3D

CHAPTER 1.

INTRODUCTION

12

photonic crystals (PhC) to have complete band gap; however, Kivshar and his team theoretically showed that 1D PhC containing matamaterials have this property. A multilayer system containing metamaterials has a very important property which was discovered by Li et al. in 2003 [29]. The proposed structure supports

new types of transmission band gaps , so-called zero-n gaps, with unique properties. The novel gap is completely dierent from the conventional Bragg reection gap, as it is insensitive to the direction or polarization of the beam [30]. Another interesting problem in the 1D PhC containing matamaterials, is the investigation of the spontaneous emission of an atom embedded in the structure [31]. It is shown theoretically that the spontaneous emission of an atom can be completely suppressed (modied Purcell eect) because metamaterials provide the possibility of vanishing optical path length between two points [32]. It is not an exaggeration to state that the vortex-like energy ux structure [24] (see Fig. 3.1) is the most important property of a metamaterial waveguide structure. Almost all unusual properties of metamaterial waveguides such as eld localization [25], super-waveguiding (a theoretical design which can transmit light with high-power density [33, 34]) and trapped-rainbows (an idealized structure which could stop light over broadband range of frequency [35]) come from this feature. This phenomena

happens at the interfaces of the metamaterial waveguide because the energy in a metamaterial waveguide core ows in the opposite direction to that in the cladding. One can potentially slow down or even stop the light passing through a metamaterial waveguide structure due to the vortex-like energy structure. This concept was rst discussed by Y.Kivshar and coworkers [24] in 2003 and was investigated in detail by He et al. [38]. However, the idea lay dormant until Tsakmakidis et al. proposed

CHAPTER 1.

INTRODUCTION

13

in 2007 a theoretical wedge-shaped metamaterial waveguide structure which could stop light over a broadband range of frequencies at room temperature producing a so-called trapped-rainbow [35]. It is interesting to note that Tsakamakidis et al.

provided another approach to explain the stopped-light mode phenomenon in the metamaterial waveguide rather than vortex-like energy ux approach. They showed that by controlling the thickness of the waveguide, the optical path of the light could be set to zero because of the negative Goos-Hanchen shift in metamaterials [36, 37]. Because of the exotic properties of the theoretical structure proposed in [35], considerable interest arose in the literature, and several groups have designed dierent theoretical structures to create trapped-rainbows based on the work of Tsakmakidis
et al.; for example, the trapped-rainbows are proposed in a hybrid metamaterial-

photonic crystal structure [43], in a surface plasmon in a graded metallic structure [39], in graphene for an electronic system [44], and in an anisotropic metamaterial waveguide structure [41, 42]. We should mention here that the idea of slow light in an anisotropic metamaterial waveguide was rst demonstrated by Narimanov and Alekseyev one year before the idea of trapped-rainbows [40]. They showed not only can light be slowed down in an anisotropic metamaterial waveguide but the velocity of a particular slow-light mode is almost constant over a broadband range of frequencies. All papers mentioned up to this point, astonishingly, neglect the eect of material loss, which is an inherent feature of metamaterials (see table 1.1), in their proposed structures. As we have previously discussed in this chapter, all currently methods

to create negative-index metamaterials are based on the use of metallic structures and all real metals introduce some losses particularly at optical frequencies (see section 1.2). One might propose to use the other methods rather than to use lossy

CHAPTER 1.

INTRODUCTION

14

metallic structures to create negative-index metamaterials. One could employ a photonic crystal in its negative-index regime or introduce chirality or optically amplifying materials (gain mechanism) to remove losses in negative-index metamaterials. However, recently, Stockman showed that loss with dispersion are the inherent features of negative-index materials and it is impossible to compensate loss completely over a broadband rang of frequencies [48]. To ensure a positive energy density (see Chapter 3) and to satisfy causality, any negative-index material must have dispersion. Thus, to satisfy Kramers-Kronig relations, it should have nite loss at some frequencies.

1.5 Thesis overview


In the thesis, we study the optical properties of metamaterial waveguide structures by investigating the guided, complex leaky, and surface polariton modes of a planar metamaterial waveguide structure with dispersion and loss. We explicitly study the important role of intrinsic loss on the optical properties of the metamaterial waveguide structure (MWS). In Chapter 2, we study the dispersion curves of a planar metamaterial waveguide structure. The main contribution of this chapter is to provide an understanding

of the eect of the negative-index refraction of metamaterials on dispersion, and to present normalization and orthogonality relations. In particular, it is shown that the properties of the dispersion curves of metamaterial waveguide are quite dierent from those of a conventional waveguide. In this chapter, we also study the role of dispersion and absorption on the modes of a metamaterial waveguide structure. We show that the dispersion curves are signicantly changed in the presence of intrinsic loss. The purpose of Chapter 3 is to study the slow-light modes in a metamaterial

CHAPTER 1.

INTRODUCTION

15

waveguide structure with dispersion and loss.

The contribution of the rst half of

this chapter is to provide the correct forms of power ux, energy density, and energy velocity in a dispersive and lossy medium. As discussed above, the topic of the slowlight mode in metamaterials has received much attention in the literature after the idea of trapped-rainbows proposed in [35]. The trapped-rainbows claim is examined by performing a simple and clear numerical analysis in a realistic lossy metamaterial waveguide structure. Chapter 4 is concerned with addressing the possible methods for the excitation of the guiding modes. The contribution of the rst part of this chapter is to introduce the transfer matrix formalism in a multilayer containing metamaterials with loss. Next, we provide an understanding of the attenuated total reection method (ATR) for exciting the modes. We then discuss the limitations of the ATR method and the conditions needed to optimize the accuracy of the ATR method. The lateral GoosHanchen shifts and their relation with the guiding modes are also covered in this chapter. Chapter 5 is concerned with addressing the modal characteristics of a metamaterial waveguide structure using a complex-frequency approach with a real propagation constant. In this chapter, we make the connection between the complex frequency approach with the complex propagation constants approach. We indicate that the

dispersion curves of the metamaterial waveguide structure are dierent in these two approaches. The other topic covered in this chapter includes perturbation theory. We provide a comparison between the exact solutions with the solutions calculated by perturbation theory.

CHAPTER 1.

INTRODUCTION

16

In Chapter 6, we summarize previous chapters main points and give some suggestions for future works.

Chapter 2

Modal characteristics of metamaterial waveguide structures

2.1 Introduction
In the recent years, many investigations have been performed on metamaterial waveguide structures (e.g. [23, 24, 53, 57]) due to their unique properties, such as slow-light modes, backward propagation electromagnetic waves, and surface polariton modes. These phenomena are potentially important for devices; however, the role of intrinsic loss which is an inherent feature of negative-index metamaterials [48], has been largely neglected in previous studies. In this chapter, we derive the dispersion equations

and electromagnetic eld distributions in planar metamaterial waveguide structures. First, we consider a symmetric waveguide and then extend our results to asymmetrical waveguide structures. Although we only study planar-geometry waveguide structures, the presented results can be straightforwardly generalized to waveguides with dierent geometries [63, 64]. In section 2.3, we derive the orthogonality relations between 17

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

18

the modes by using the generalized Hermitan Hamiltonian [58, 59].

By exploiting

the orthogonality relations, we discuss the normalization of electromagnetic eld amplitude in metamaterial waveguides. We then compare the novel properties of the

proposed metamaterial structures with that of conventional waveguide structures in section 2.4. Finally, we investigate the behavior of the bulk and complex modes using three dierent models of the material in the structure: i) dispersionless and lossless, ii) dispersive and lossless, and iii) dispersive and lossy models. Our computational calculations indicate that material loss modies signicantly the dispersion curves.

2.2 Theoretical formulation of dispersion relation


(a) (b)

-d
Region(I) Region(II)

d
Region(I)

2d

Region(I) Region(II) Region(III)

Figure 2.1: Schematic geometry of (a) symmetric and (b) asymmetric metamaterial waveguide structure. An innite metamaterial slab of thickness 2d and constitutive parameters

2 < 0

and

2 < 0

is surrounded by dielectric semi-innite spaces.

Cartesian coordinate system is employed. The light propagates along the the structures are independent of

axis , and z

axis.

In this section we analyze symmetric and asymmetric guiding structures involving

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

19

metamaterials. First, we derive the electric and magnetic elds distributions of metamaterial waveguide structures by using Maxwell's equations and boundary conditions [50, 51, 52] for transverse electric (TE ) and transverse magnetic (TM ) modes. We then obtain the dispersion relations for the proposed structures. Finally, we discuss the physical and unphysical solutions of the dispersion equation. The physical so-

lutions satisfy guidance conditions. But the unphysical solutions do not. Although the form of the dispersion equations are the same in metamaterial waveguide and conventional waveguide structures, the physical solutions in these two structures are dierent. We consider an isotropic homogeneous slab waveguide in a planar geometry for two dierent structures: i) a symmetric waveguide structure with the negative-index core located in free space [Fig. 2.1(a)] and ii) an asymmetric waveguide structure [Fig. 2.1(b)]. The core, with a thickness of

2d, is sandwiched between the two semi-innite

media as shown in Fig. 2.1. For time-harmonic electromagnetic elds,

E(r, t) = Re{E(r) exp[it]} H(r, t) = Re{H(r) exp[it]},


and Maxwell's equations take the form

(2.1)

E = 0, H = 0,
where c is the speed of light in vacuum and

E = i H, c H = i E, c
and

(2.2)

are the permittivity and per-

meability, respectively. We also assume electromagnetic elds that are independent of the y-component in our proposed waveguide structure and that the propagating

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

20

modes can be represented as either transverse electric modes (TE ), or transverse magnetic modes (TM ). Thus, the y-component of the TE modes,

E y, can be expressed as

E y(x, z) = E(x)exp[i(z)], z,
and

where

is the wave vector in the propagation direction,

is the angular frequency of the eld. The time-harmonic wave equation can

be straightforwardly obtained by using Eq. (2.2). From the time-harmonic electromagnetic wave equation,

1/(r)

{1/(r)

E(r)} = 2 /c2 E(r),

(2.3)

the propagating TE modes of the electric eld are described by the following scalar equation for the proposed waveguide structure

2 1 (x) 2 2 + 2 Ey (x, z) = 2 (x)(x)Ey (x, z), z 2 x (x) x x c


The solution of Eq. (2.4) for the symmetric waveguide has the form

(2.4)

E(x, z) = y E0 exp[iz]

A exp[k1 (x d)]

(x > d)
(2.5)

B exp[k2 x] + C exp[k2 x] (d < x < d) D exp[k (x + d)] (x < d), 1


and

where

k1 =

2 1 1 2 /c2

k2 =

2 2 2 /c2 2

represent the wave number

in the cladding and the core, respectively.

The coecients A, B, C, and D are

determined by the boundary conditions. As the propagating modes in this waveguide structure may be split into even and odd modes, the electric eld of the even and odd
TE modes are

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

21

even Ey (x, z) = E0 exp[iz]

cos(k d) exp[k (x d))] (x > d) 2 1 cos(k2 x) cos(k d) exp[k (x + d)] 2 1 sin(k d) exp[k (x d)] 2 1 (d < x < d) (x < d), (x > d)

(2.6)

odd Ey (x, z, t) = E0 exp[iz]

(d < x < d) sin(k2 x) sin(k d) exp[k (x + d)] (x < d). 2 1

(2.7)

The magnetic eld, from

H,

components of the even and odd TE modes can be obtained

Ey

by using the Maxwell equation,

E = i H, c

as

even

cE0 e[iz] 1 (x, z) = x z 2 [ cos(k2 x) + ik2 sin(k2 x)] 1 [ik z + x] cos(k d) exp[k (x + d)],
1 1 2 1

1 [ik z + x] cos(k d) exp[k (x d)] (x > d) 1 1 2 1 (d < x < d) (x < d)

(2.8)

Hodd (x, z) =

cE0 e[iz]

1 [ik z + x] sin(k d) exp[k (x d)] (x > d) 1 1 2 1 1 x z (d < x < d) 2 [ sin(k2 x) + ik2 cos(k2 x)] 1 [ik z + x] sin(k d) exp[k (x + d)], (x < d) 1 2 1 1

(2.9)

Employing the boundary conditions, the tangential components of the electric, magnetic,

E, and

H,

elds are continuous at

x = d

and

x = d,

we obtain the dispersion

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

22

relations for the even and odd TE modes [50, 51, 52]

k2 d k1 = tan(k2 ) 1 2 2
(2.10)

k1 k2 d = cot(k2 ). 1 2 2
Finally, the results for the magnetic transverse wave, TM, can be found from duality by changing

and

E H .

Thus, the dispersion equations of the odd and even

TM modes are [50, 51, 52]

k1 k2 d = tan(k2 ) 1 2 2
(2.11)

k1 k2 d = cot(k2 ). 1 2 2
These dispersion equations are valid for any complex or negative value of

and

Hence, they also work for the lossy double-negative index planar geometry waveguide structure. We can easily generalize the results to the asymmetric waveguide of Fig. 2.1(b) [51]. The electromagnetic elds of the asymmetric waveguide have the form,

Ey (x, z)

E0 exp[iz]

sin( 2k d) exp[k (x 2d)] (x > 2d) 2 1 sin( + k2 x) sin() exp[k x], 3 (0 < x < 2d) (x < 0)

and

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

23

H(x, z) =

cE0 e[iz] 1 x z 2 [ sin( + k2 x) + ik2 cos( + k2 x)] 1 [ik z + x] sin() exp[k x],
3 3 3 2 2 1 1 k0

1 [ik z + x] sin( 2k d) exp[k (x 2d)] (x > 2d) 1 1 2 1 (0 < x < 2d) (x < 0)
(2.12)

where

k1 =

and

k3 = k2 =

2 2 3 3 k0

are the wave numbers in the two is the wave number in the core,

cladding regions respectively, and where from

2 2 2 k0 2

k0 = /c

represents the vacuum wave number. The angle

can be obtained

tan( k2 d) = k2 /k1 .

(2.13)

The dispersion equation for the asymmetric waveguide of the Fig. 2.1(b) can readily be put in the form

2 2d 2 2 k0 2 = arctan(

2 1

2 2 1 1 k 0 2 2 2 k0 2

) + arctan(

2 3

2 2 3 3 k0

2 2 2 k 0 2

) + m/2,
(2.14)

for the TE modes, and in the form

2 2d 2 2 k0 2 = arctan(

2 1

2 2 1 1 k 0 2 2 2 k0 2

) + arctan(

2 3

2 2 3 3 k0 2 2 2 k0 2

) + m/2,
(2.15)

for the TM modes, where

m = 0, 1, 2, ... E y (x)

represents the number of the nodes in the The roots of the above transcendental

characteristic mode prole of

eld.

equations give the propagation wave number,

which in general can be complex

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

24

due to the intrinsic loss in the structure or the light leakage below the total internal reection angle (see section 2.5.3). transverse wave number equations, e.g. The double-value of the square root in the

ki = k2 ;

2 2 i i k0 ,

leads to a branch cut

problem. It is not important which sign is chosen for the propagation constant, for the transverse wave number in the core,

or

however, the sign of the transverse

wave number in the cladding has signicant physical meaning. If the real part of it is negative, this leads to a violation of the guidance condition, which requires that the eld amplitudes vanish in the transverse direction at

x .

2.3 Orthogonality and normalization


In this section, we derive the normalization relations for electromagnetic elds in the metamaterial waveguide structure. In the previous section, we found the electromagnetic elds in the metamaterial waveguide structure by using Maxwell equations and boundary conditions. However, in Eqs. (2.6-2.9) , there is one unknown variable ,

E0 ,

which can be obtained from eld normalization. Before we normalize the elds, we need to derive the orthogonality relations in the metamaterial waveguide structures. First, we assume the permeability and permittivity are real (lossless model). We note that the negative value of

and complicate the problem as the method suggested in

[60] does not work to nd orthogonality relations in metamaterial structures. We use the Generalized Hermitian Hamiltonian technique [58, 59] to nd the orthogonality relations in a metamaterial waveguide structure. The mathematical details are given in appendix A. The orthogonality relation is

1 2

1 Ey (x)Ey (x)dx = N , , (x)

(2.16)

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

25

where N is the normalization parameter. We can interpret it as a total power ow carried by a specic mode when

= .

We have the freedom to set it to one. We

now can derive the electric eld amplitude of the proposed waveguide of Fig. 2.1(a) ,

E0 ,

by using Eqs. ( 2.6-2.9,2.16)

sym E0 =

2 c d +

2
2 1 k1
2 2 k1 +k2 2 k 2 +2 k 2 2 1 1 2

(2.17)

where d is the thickness of waveguide;

k1 and k2

are the wave number in the cladding

and the core respectively. We note that the normalized eld amplitudes of the metamaterial waveguide structure may diverge for particular sets of parameters. How-

ever, this unphysical divergence disappears when we consider a realistic metamaterial waveguide structure with material loss. In some recent work (e.g. [35]), the Authors insert the absolute value of

in Eq. (2.16) to avoid the divergence point of the nor-

malized eld amplitude; however, we lose orthogonality relations if we normalize the eld amplitude with respect to

1/|(x)|Ey (x)Ey (x)dx = , .

The normalization and orthogonality relations derived in this section are not valid for a lossy metamaterial structure with complex

and

but only apply to a lossless

metamaterial waveguide structure. However, we will use the orthogonality relations for a lossless waveguide to derive the approximate dispersion curves of the lossy metamaterial waveguide by using perturbation technique (see Chapter 5).

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

26

2.4 Comparison between conventional and metamaterial modes of a slab waveguide


In this section, we compare the properties of metamaterial and conventional waveguide structures by employing a graphical solution. We nd the properties of a metamaterial waveguide structure to be very dierent than the properties of a conventional waveguide structure. These unique properties are currently attracting considerable attention because of potential applications [53-57]. For simplicity, we only consider a symmetric waveguide structure in this section. The main results will not be changed for an asymmetric waveguide. From the section 2.2, we know the dispersion equation for the TE modes has the form:

k2 d k1 = tan(k2 m ) 1 2 2 2
2 2 k1 + k2 =

(2.18)

2 d2 (2 2 1 1 ) R2 . c2

(2.19)

We exploit the graphical method to determine the solutions of the transcendental equation. Fig. 2.2 shows the Riemann sheet for the planar waveguide structure. The vertical axis represents the real part of the wave number in the cladding,

k1 ,

and the

horizontal axis shows the real and imaginary parts of the wave number in the core,

k2 .The

intersection of the solid curves, Eq. (2.18), and the dotted curves, Eq. (2.19),

determine the solution of the guided modes. The roots located in the half top of the Riemann sheet satisfy the guidance condition since their elds exponentially decay in the cladding due to positive value of the

k1

in this region (see the inset (B) in Fig.

2.2(a)). However, the roots in the half bottom of the Riemann sheet are improper

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

46

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

27

Re[k1]
2

TE0

TE1

0 2 2 0 2

(C) Re[k2]

Img[k2] (A)
2

2 0

(B)

0 2 2 2 2 0 2 0 2

2 2
2 0

TE2
0

i/2

/2 Re[k1]
2

TE2 (B)

1 2

(A)
0 2

0 2 2 0 2

Img[k2]
2

Re[k2] (C)
0 2

0 2 2

TE0 i/2
0

TE1

/2

Figure 2.2: Graphical solution of the TE modes for (a) the dielectric waveguide with

==4

and

The intersection of the solid curves, Eq. proles of the guided modes. (2.18), and the dotted curves, Eq. (2.19), Insets show the transverse determines the solution of the guided modes. Insets show the transverse proles of the guided modes for

d = Figure 2.2: Graphicalthe metamaterial waveguide dielectric waveguide with. 2 cm , and (b) solution of the TE modes for (a) the with = = 4
==4
and

d = 2cm

, and (b) the metamaterial waveguide with

= = 4.

E y (x)

eld.

solutions because their eld amplitudes exponentially grow in the cladding (e.g. see the inset (A and B) in Fig. 2.2(a)). The right half of the Riemann sheet determines the oscillatory modes and the left half represents the evanescent or surface polariton modes. By looking at the transcendental equation, we see that the improper roots of a dielectric waveguide, which are lying on the bottom of the Riemann sheet, are proper solutions for metamaterial waveguides since the negative sign of

in Eq.

(2.18)

leads to a reection of the Riemann sheet about the horizontal axis, (compare Fig.

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

28

2.2(a) and (b)); therefore, the guided modes of the metamaterial waveguide behave as the improper modes of the conventional waveguide. In the past, the properties of the improper conventional waveguide modes, usually termed leaky modes, have been studied because of their useful applications such as leaky wave antenna, and leaky wave lters. Several new properties may be observed in the metamaterial waveguide structure [23, 24]. First of all, the metamaterial waveguide supports the surface polariton modes for both the TE and TM polarizations. By looking at the Riemann sheet, we notice that there is no solution in the second quadrant for the dielectric waveguide structure; therefore, the conventional waveguide cannot support surface polariton modes. At

the metallic interface, only the magnetic transverse, TM, modes can be supported; however, a metamaterial waveguide can support both TM and TE surface polariton modes; moreover, it supports complex modes. The complex modes are the complex solution of the dispersion equation above the cut-o frequencies. wave number, The propagation

is always complex in this type of mode, even in a lossless structure.

We can assume that the complex mode represents the wave propagation below the angle of total internal reection in the core; therefore, the wave leaks out of the waveguide and the imaginary parts of the complex modes indicate the damping factor of the leakage. Analytically, with a straightforward calculation, it can be shown that the complex modes are always improper solutions of the dispersion equation in a conventional waveguide [52]; however, a lossless metamaterial waveguide structure always supports proper complex modes [56]. By using the Eq. (2.10) and separating

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

29

the real part and imaginary part of the wave numbers, we have

r k1 =

1 y tanh(y)[1 + tan2 (x)] + x tan(x)[1 tanh2 (y)] 2 tan2 (x) + tanh2 (y)
(2.20)

i k1 =

1 x tanh(y)[1 + tan (x)] y tan(x)[1 tanh (y)] , 2 tan2 (x) + tanh2 (y)
and

where

r i k1 = k1 + ik1

k2 = x + iy.

For now, let us assume that

is a real

number (lossless structure). The term (2.19); hence, the sign of

x tan(x) must be always positive to satisfy Eq.


has

r k1

only depends on the sign of the permeability. If

a positive value (the dielectric waveguide case), the real part of the wave number in the cladding is always negative which leads to an unphysical solution; and if

<0

(the metamaterial waveguide case), the real part of the wave number in the cladding is always positive and all complex mode solutions are on the top Riemann sheet. Second, it is possible for two modes to have the same number of nodes since one of them propagates in the same direction of the group velocity, (forward propagating mode), and the other propagate in the opposite direction of the group velocity, (backward propagating mode); hence, both forward and backward modes are supported by the metamaterial slab. Third, the rst bound mode,

T E0 , does not exist in

metamaterial waveguide structures. This property gives an opportunity to introduce one dimensional periodic waveguides with a complete band gap [28], introduce the concept of surface wave suppression which leads to enhance radiation eciency for microstrip antennas [26], and introduce the slow-light modes under the single mode operation [54]. However, in the realistic metamaterial waveguide, which should be

dispersive and lossy, and also by considering the role of the complex modes, we cannot see this property anymore.

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

30

Finally, not only does the metamaterial waveguide support evanescent and oscillatory eld distributions, but also supports linear eld distribution. The point (A) on the Fig. 2.2(b), where the root is located on the top vertical axes of the Riemann sheet,

k2 = 0,

indicates this type of mode. This unique mode is a transitional mode

between the rst magnetic or electric transverse,

T E1

or

T M1 , oscillatory and evanes-

cent modes. The electric eld distributions for the TE mode, considering boundary condition in the Eq. (2.5), is

2 k1 d 1 exp[k1 (x d)] (x > d) Ey (x, z) = E0 exp[iz] 2 k1 x (d < x < d) 1 2 k1 d exp[k1 (x + d)] (x < d), 1
which leads to constant longitudinal magnetic eld for these modes.

(2.21)

The constant

longitudinal magnetic eld gives more degrees of freedom for potential applications in a metamaterial waveguide. [61]

2.5 Guided, surface polaritons, and complex surface modes


2.5.1 Lossless and dispersionless metamaterial slab

In this section, we numerically study the behavior of dispersion curves for three different material models of metamaterial waveguide structures: i) lossless and dispersionless ii) lossless and dispersive iii) lossy and dispersive models. We know the rst

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

31

two models are unrealistic since these models violate the causality principle and positive energy density [48]. However, we investigate these two models for two reasons. First, almost all previous studies in the literature employ one of these two models. Thus, we can compare our results with others and also compare the properties of the dispersion curves between metamaterial and conventional waveguide with the same material models. Second, by comparing results from these models with the realistic model, lossy and dispersive model, we can highlight the eects of loss on the dispersion curves. Our analysis indicates that loss signicantly modies the dispersion curves and some astonishing properties in unrealistic metamaterial waveguide, such as a stopped-light mode, are destroyed or disappear in the realistic material model [49]. We begin by investigating the properties of the dispersion curves for the lossless and disprsionless metamaterial waveguide. We consider the symmetric metamaterial waveguide structure, illustrated in Fig. 2.1(a), with a thickness of free space, Fig.

d = 2 cm
and

in the

1 = 1, 1 = 1,

with the metamaterial parameters

2 = 4,

2 = 4.

2.3(b) indicates the dispersion curves for the three lowest transverse electric,

TE, modes for the proposed structure. To compare the properties of the dispersion

curves for the metamaterial waveguide with a conventional waveguide, we plot in Fig. 2.3(a) the dispersion curves of a conventional waveguide with

2 = 4,

and

2 = 4.

The rest of the parameters are chosen to be the same as the proposed metamaterial waveguide. Although the dispersion curves of the metamaterial and conventional

waveguide are the solutions of the same transcendental equations (2.10,2.11), Fig. 2.3 clearly indicates that the dispersion curve characteristics are dramatically dierent for the metamaterial and dielectric waveguide structures . As mentioned before, this

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

32

dierence comes from the fact that the physical dispersion equation roots for the metamaterial waveguide are the unphysical roots for the conventional waveguide. From our calculations, we can nd several signicant dierences between the conventional and metamaterial waveguide structures. First, the metamaterial waveguide structure supports new types of modes: so-called surface polariton modes and complex leaky modes. Hence, as shown in Fig. 2.3(b), a rich variety of modes (bound, surface polariton, and complex modes) exist in the metamaterial waveguide structure. The bound modes are restricted between the vacuum light-line and metamaterial lightline. We plot

T E1 , T E2 ,

and

T E3

modes in Fig. 2.3(b). The

T E1

mode behaves like

a bound mode whenever its propagation constant,

is between the vacuum light-

line and metamaterial light-line. As the propagation constant of

T E1

goes below the

metamaterial light-line, it behaves like a surface polariton mode. The

T E2 ,

and

T E3

modes behave like a bound mode when the propagation constant of the dispersion curve is purely real and behave like a complex leaky mode when the propagation constant is complex. Although the proposed structure is lossless, the propagation

constant has a complex value for complex modes. In the complex modes, the energy is not conned in the core and leaks out through the cladding. In Fig. 2.3(b), the dispersion curve of a complex

T E2

mode with a real propagation constant starts from

the cut-o frequency where the bound mode curve is at, e.g. at f = 2.8 GHz, and goes to lower frequencies, and the dispersion curve of a complex mode with a imaginary propagation constant, shown by the thin curve in the Fig. monotonically by decreasing the frequencies. 2.3(b), increases

As the complex modes do not carry

energy in the lossless metamaterial waveguide structure [52], they do not play an important role in the structure. Second, we have domains of frequencies where there are

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

33

no bound modes; for instance, there is no bound mode in the range of 2 and 3 GHz in Fig. 2.3(b). Third, in contrast with the dielectric waveguide modes, the modes in the metamaterial waveguide are not restricted to lie between the vacuum light-line and the metamaterial light-line. For example, the complex modes go above the vacuum light-line and the

T E1

mode goes below the metamaterial light-line. Finally, we

note that the dispersionless and lossless metamaterial is a non-realistic model since it violates causality principle; for example, in our dispersion curve model, one can nd the points where the group velocity exceeds the speed of light in vacuum and even goes to innity! We discuss this issue in the next chapter in detail, where we consider
CHAPTER 2. MODAL CHARACTERISTICS OF MWS

47

a more realistic model by including dispersion and loss.

(a)Light line
Frequency (GHz) TE
2

(b)Light line
5

TE3

Frequency (GHz)

TE1
2

Dielectric light line

TE2

Metamaterial light line

TE0

TE1

0 0

0.5

1.5

(1/cm)

2.5

3.5

0 0

0.5

1.5

(1/cm)

2.5

3.5

Figure 2.3: The dispersion curves for the dielectric and metamaterial frequency independent waveguide structure with the thickness of with

=4

and

= 4.

(b) The metamaterial

thick and thin curves correspond to the real and imaginary components of

d = 2 cm. (a) slab with = 4

The dielectric slab and

= 4. .

The

Figure 2.3: The dispersion curves for the dielectric and metamaterial frequency inde-

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

34

2.5.2

Lossless, dispersive metamaterial slab


2.1(a) in a free space with a

We next consider the symmetric waveguide of Fig. dispersive, lossless metamaterial core.

Before we investigate the properties of the

dispersion curves for this model, we describe the complex frequency-dependent electric permittivity and magnetic permeability by using the standard Drude and Lorentz Models [62]. These are given respectively by

() = 1

2 p , 2 + ie

(2.22)

() = 1

F 2 , 2 2 0 + im

(2.23)

where F is the fractional area of the unit cell occupied by the interior of the split-ring,

is the plasma frequency,

is the magnetic resonance frequency, and

and

are respectively the electric and magnetic broadening constants associated with loss. We plot the permeability and permittivity in Fig. rameters: 2.4 with the following pawhich are

F = 0.56, p /2 = 10 GHz, 0 /2 = 4 GHz, = 0.1 GHz

chosen to be in agreement with recent experiments [62]. Fig. 2.4 indicates that in the range of frequencies between 4 and 6 GHz, the proposed model behaves as a doublenegative-material (DNG) with both negative permeability and permittivity. Below 4 GHz frequency and above 6 GHz Frequency, it behaves as a single-negative-material (ENG) or plasmonic media. Although the model provides a negative-index refraction condition between 4 and 6 GHz, only the range of frequencies between 4 and 5.2 GHz satises the guiding condition waveguide structure with

|ncore | > ncladding

[50, 51, 52] for the proposed

ncladding = 1.

Hence, the bound modes are restricted to lie

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

35

between 4 and 5.2 GHz.

5 4 Re[] Im[] Re[] Im[]

Permeability & Permittivity

3 2 1 0 1 2 3 4 ENG 5 4

= 0.1 =0

fsm

fsp

DNG |n| > 1


5 6 7

ENG
8 9

f(GHz)
Figure 2.4: The permeability and permittivity of the metamaterial slab as a funcp = 10 GHz, 0 = tion of frequency using Lorentz-Drude models with F = 0.56, 2 2 4 GHz, and = 0.1 GHz. The solid curves show the real part of and , and the dashed curves show their imaginary part. Im[] and Im[] are scaled up by a factor of 20. If the metamaterial slab is in the free space,0 polariton frequency, and 7.03 GHz, respectively.

= 0 = 1, the magnetic surface fsm , and the plasmon surface polariton frequency, fsp , are at 4.71

Employing the Lorentz-Drude model for

and

2 ,

we present the calculated ve

lowest-order TE modes of the proposed metamaterial waveguide structure without loss in Fig. 2.5(a). When we compare the dispersion curves of Fig. 2.3(b) and Fig. 2.5(a), we note that the dispersion curves qualitatively rotate 180 degree when we include dispersion in our model. The dispersion curves have a cuto frequency that is an upper bound. However, the dispersion curves of the dispersionless waveguide have a cuto frequency that is a lower bound. Also, in the dispersive case higher-order

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

36

modes appear at lower frequencies rather than higher frequencies. Fig. 2.5(a) shows that for higher-order modes, the frequency depends only weakly on the propagation constant,

and

f () fres = 4

GHz . However, no mode goes below the resonance


48

frequency.

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

5.2

TE0 TE1

(a)

frequency (GHz)

4.7

fms TE2 TE3 TE4


1 2 3 4 5 6 Metamaterial Light Line

4.4

4 0

fres

Propagation constant (1/cm)


5.2

TE0

(b)

frequency (GHz)

TE1
4.7

fms TE2 TE3 TE4


Metamaterial Light Line

4.4

4 0 1 2 3 4 5 6

fres

Propagation constant (1/cm)


Figure 2.5: The dispersion curves for the rst four TE modes.(a) the lossless and

Figure 2.5: The(b) the lossy dispersive metamaterial waveguide structure with the parameters given lossless and dispersion curves for the rst four TE modes for (a) the
in Fig.4 for 2 and 2 . The 2-cm-thickness metamaterial slab is in (b) the lossy dispersive metamaterial waveguide structure withthe freeparameters given the space

in Fig. 2.4 for shown in Fig.

as shown in Fig.1 (a). The thick and thin curves represent the real and imaginary

2 of the 2 The 2-cm-thickness metamaterial slab is in free part and propagation constant, ; and the dashed curves indicate the light-line and
metamaterial light-line. 2.1(a). The thick

space as

and thin curves represent the real and imaginary and the dashed curves indicate the light-line and

part of the propagation constant metamaterial light-line.

As in the non-dispersive structure, this structure does not support the nodeless

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

37

bound mode. But, the nodeless surface polariton and complex modes,

T E0 , are always fn = 5.2 GHz,

observed. In Fig. 2.5, the nodeless surface polariton mode starts from where

ncore (fn ) = 1, and goes to innity at the magnetic surface polariton frequency,
4.714 GHz, where

fms =

(fms ) + 0 = 0. T E0
mode. As the propa-

The

T E1

mode appears at frequencies lower than the

gation constant,

, increases, the T E1

mode changes from a bound mode to a surface

polariton mode. There is a transition point on the metamaterial light-line between the

T E1

bound mode and surface polariton mode.

At this point, the electric eld

distribution is linear or the magnetic eld is constant in the core of the waveguide structure. Thus, the

T E1

mode takes on dierent diverse types of modes, such as

bound mode, complex mode, surface polariton mode, at dierent Although the shape of waveguide's core, the

().

T E0

is almost insensitive to changes in the thickness of the

T E1

curve changes dramatically when the thickness is changed.

Nevertheless, it is important to note that in principle there is no limit on the reduction of the slab thickness [66]: all the modes are preserved by changing the thickness of the slab. This feature, which is contrary to what we have in a conventional waveguide, leads to possibility to design ultra-thin slab waveguides. When we decrease

the slab thickness, the dispersion curves shift to lower frequencies and concentrate near resonance frequency region. However, in the dispersionless waveguide structure, modes shift to higher frequencies with decreasing slab thickness . In summary, our simulations indicate that many of the exciting properties of dispersionless and lossless metamaterial waveguide structure are preserved for the dispersive metamaterial waveguide structure although the dispersion changes the dispersion curves of the bound modes.

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

38

2.5.3

Lossy and dispersive metamaterial slab

We now study the dispersion curves behavior of realistic metamaterial waveguide structures including both dispersion and loss. We nd several new properties in a

metamaterial waveguide structure when loss is included, such as a splitting of the complex modes into two curves, coupling the surface polariton modes with higher order modes, and continuing dispersion curves into frequencies below the resonance frequency . We study these properties in detail in the following paragraphs. We now consider the proposed structure in the previous subsection but with a loss parameter of

= m = e = 0.1 GHz.

Fig. 2.5(b) shows the dispersion curves The thick curve corresponds

for the rst ve TE modes in the presence of loss. to the real parts of the propagation constant, the imaginary part of

and the thin curve corresponds to

First, comparing with Fig. 2.5(a), the introduction of loss

splits the complex modes into the two modes: a complex backward propagation mode and a complex forward propagation mode. We cannot ignore the importance of the complex modes in a lossy waveguide structure since the complex modes carry energy. The average energy ux in the guidance direction is always zero for the complex modes of a lossless waveguide structure [52]. The imaginary part of imaginary part of

for the backward mode has a negative value while the We note that the In

for the forward mode has a positive value.

absolute values of the imaginary parts of the modes is plotted in our graphs.

a lossless waveguide structure, the real parts of the forward and backward complex modes overlap with each other and the imaginary part of them have the same value with dierent sign. But, the loss breaks this symmetry as shown in Fig. 2.5(b). The forward and backward complex curves of the particular mode have approximately the

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

39

same behavior except near the magnetic surface polariton frequency region. Second, the surface polariton modes,

T E0 and T E1 do not touch at fms as ; , at that frequency. T E0


mode with the backward complex

instead they join to higher order modes with nite wave number, For example, Fig. 2.6 shows the coupling of

T E4

mode.

In Fig.

2.6, the dark blue curve shows the

T E0

mode of the lossless

waveguide structure. It goes to innity at

fms = 4.71 GHz;

however, it bends back at

fms = 4.71 T E4

GHz when we insert loss in the structure and couples with the complex

modes. The thick and thin light blue curves represent the real and imaginary

parts of the complex

T E4

mode for the lossless structure, and the thick and thin red

curves show the real and imaginary parts of the

T E0 &T E4 coupling mode for the lossy

structure respectively in Fig. 2.6. We note that below 4.75 GHz and above 4.68 GHz, the inuence of material loss on mode structure is highly non-perturbative; however, the red curve approaches the blue curves above 4.75 GHz and below 4.68 GHz. To have a better insight into this phenomenon, we plot the normalized electric eld at several points on the

T E5

and

T E1

curves in Fig. 2.7. We normalize the electric eld

to its amplitude in the core. The rst point at 4.05 GHz is on the bound

T E5

mode.

As predicted, the electric eld has ve nodes inside the waveguide core at this point. The second point is chosen on the complex

T E5

mode at 4.63 GHz. The electric eld

has ve dips instead of nodes at this point. When frequency increases, the amplitude of these dips gradually decrease. Finally, four of them disappear at

fms = 4.71

GHz.

At this frequency, the electric eld is sharply conned at the waveguide surfaces as the normalized electric eld intensity approaches the electric eld at 4.85 GHz on the

2 E 2 /E0 = 104 !

Finally, we plot

T E1

curves.

The electric eld, as expected,

has only one node at this frequency. We conclude that the backward complex

T E5

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

40

mode seamlessly changes to the

T E1 surface polariton mode near the magnetic surface T E5

polariton frequency region. However, there is no specic boundary between the and

T E1

modes.

4.78

TE4 frequency (GHz)


4.74

TESP 0

fms

4.7 TESP & TEB 0 4 coupling

4.66 0 2 4

(1/cm)

10

Figure 2.6: The surface polariton mode coupling with the higher order mode. The blue curves shows the

T E0

and the complex

and the thick red curve indicates the coupled propagating

T E4 modes in the lossless structure T E0 &T E4 mode in presence of loss.

The dashed lines represent light-line and metamaterial light-line. Note: The forward

T E4

and

T E0

are not shown in the gure.

Third, the dispersion curves go below the resonance frequency when we introduce material loss in the waveguide structure. The modes of the lossless waveguide touch at

fres = 4

GHz as

and no TE mode can be observed below resonance

frequency. The red curve shows the

T E4

mode of the lossless waveguide in Fig. 2.8.

However, the TE modes of the lossy waveguide bend near resonance frequency and go below it until

Re[] 0.

The thick blue curve in Fig.

2.8 show this behavior

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

41

5.2 5

1 0.5 0 2 2 1 0 2 0 2 x 10
4

4.8 f(GHz)
2

4.6
1

4.4 4.2

0 2
0.5 0 2

TEB
1

4 0

TEB
Figure 2.7: The eld intensities,

3 (1/cm)

T E1 and T E5 2 2 Ey (x)/E0 T E4

modes coupling. Insets show the normalized electric

of the real part of

mode. We note that they behave as backward modes below

resonance frequency although their slope is positive. This result comes from the fact that the group velocity in the lossy waveguide is not equal with energy velocity. We study this issue in detail in Chapter 3. We can explain the penetration of the modes below the resonance frequency by looking at Fig. 2.9. It shows magnetic permeability near resonance frequency. The blue and red curves represent permeability without and with material loss respectively. For the lossless Lorentz model, permeability becomes innite,

( ) at the

resonance frequency. Since in the presence of material loss, the permeability has a

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

42

4.05

4.03

frequency (GHz)

TE4
4.01

frs
3.99

TEB 4
3.97

3.95 0

(1/cm)

10

15

20

Figure 2.8: The blue curve indicates that the TE mode extends below the resonance frequency in the presence of loss. The red curve represents the backward imaginary parts of the

T E4

mode

of the lossless waveguide and the thick and thin blue curves represents the real and

T E4

mode for the lossy waveguide structure.

nite value at the resonance frequency (see Fig. 2.9), the TE modes can penetrate into the single negative-index media, where the Fig. 2.8 for

has the positive value, as shown in

T E4

mode, but the real part of the wave number rapidly goes to zero.

Since the damping factor enters in the denominator of the characteristic LorentzDrude equation, both imaginary and real part of

and

are changed; therefore, the

resonance frequency also is shifted to higher values.

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

43

150 100 50 0 Im[]

permeability

= 0.1
50 Re[] 100 150 3.97 3.98 3.99 4 4.01 4.02 4.03

=0

frequency (GHz)
Figure 2.9: 2 2 The permeability near the resonance frequency with

The solid and dashed curves represents the real and imaginary parts of the permeability with

0.56f /(f 16 + if ). /2 = 0.1,

The blue curve represents the permeability with respectively.

(f ) = 1 = 0 .

2.6 The

TM modes of the lossy, dispersive metamaThe results and

terial waveguide
In this section, we calculate the rst ve TM modes of the lossy, dispersive metamaterial waveguide structure with the parameters used in section 2.5. for The TM modes can be obtained from duality by changing

E H .

Fig. 2.10(a) indicates that the dispersion curves of the TM modes behave like the
TE modes.

There is only one signicant dierence between them: the TM modes

are insensitive to the magnetic surface plasmon frequency,

fms = 4.71

GHz, where

(fms ) + 0 = 0.

As predicted from duality, the TM modes are sensitive to the

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

44

surface plasmon frequency,

fsp = 7.07

GHz, where

(fsp ) + 0 = 0.

Since the surface

plasmon frequency is in the single-negative-index region and we are only interested in


CHAPTER 2. MODAL CHARACTERISTICS OF MWS 52

double-negative-index region, we do not plot surface plasmon region in Fig. 2.10(a). If we extend Fig. 2.10(a) to higher frequencies, we see that the TM surface polariton modes,

T M0

and

T M1 ,

touch at

fsp = 7.07 fsp

GHz as

and the complex TM

modes show the same odd behavior at


CHAPTER 2.

as the TE modes at

fms .
53

MODAL CHARACTERISTICS OF MWS

5.2

TM0

(a)

5.2

TE0

(b)

frequency (GHz)

TM1
4.7

frequency (GHz)

TE1
4.7

fms TM2

fms TE2 TE3 TE4


Metamaterial Light Line

4.4

4.4

TM3 TM4
4 0 1 2 3 4

Metamaterial Light Line

fres
5 6

4 0 1 2 3 4 5 6

fres

Propagation constant (1/cm)

Propagation constant (1/cm)

Figure 2.10: (a) The dispersion curves of the lossy dispersive metamaterial waveguide structure for the TM modes with the same parameters as in the Fig. 2.5. The curves of the TM modes in contrast with the TE modes (b) are insensitive to the magnetic surface polariton frequency,

fms .

There are two types of surface modes in the metamaterial waveguide structure. One is called a surface plasmon mode. It is a result of the coupling of photons and electrons. Its frequency is determined by

(fsp )+0 = 0, where 0 is the permittivity of

Figure the cladding. The other is called2.10: (a) The dispersion curves ofplasmon mode. It is a result of the a magnetic surface the lossy dispersive metamaterial waveguide the TM modes in contrast with the TE modes (b) are insensitive to the magnetic

Figure 2.10: (a)structure for the TM modes with the same parameters in the Fig. 5. The curves of The dispersion curves of the lossy dispersive metamaterial waveguide structure for the TM modes with the same parameters in the Fig. 5. The curves of the TM modes in contrast with the TE modes (b) are insensitive to the magnetic

surface polariton Its frequency is determined by coupling of photon and electron spins. frequency, fms .

(fms )+0 = 0,

where

surface polariton frequency, fms . is the permeability of the cladding. If we were to set the optical parameters

of the metamaterial waveguide as

fsp = fms ,

the dispersion curves of the TM modes

would be indistinguishable from those of the TE modes.

CHAPTER 2.

MODAL CHARACTERISTICS OF MWS

45

2.7 Concluding remarks


In this chapter, we have investigated the dispersion curves of a metamaterial waveguide structure. By making a connection to the recent literature, we have shown

that the properties of the metamaterial waveguide structure are completely dierent from those of a conventional waveguide system. It was shown that the metamaterial waveguide structure supports a rich variety of modes with unusual properties. These properties provide more control over electromagnetic elds and open a door to exciting potential applications, such as miniaturizing devices and systems with slow-light modes. In the last section of this chapter, we have examined these properties by considering the practical limitations. We have explicitly studied the roll of dispersion and loss, which are inherent features in metamaterials, on the dispersion curves. We have found that the properties of modes can change greatly when loss comes into play. In the following chapter, we will study the eect of intrinsic loss on the one of the most important potential properties of the metamaterial waveguide structure, slow-light modes, in more detail.

Chapter 3

Slow light in metamaterial waveguide structures

3.1 Introduction
Metamaterials can change the electromagnetic properties of many dierent physical structures. Even a slab waveguide containing a metamaterial shows exotic features that may have important potential applications, such as a vortex-like energy ux structure[24], and stopped-light modes [35]. In 2003, Y.Kivshar and coworkers showed that a double vortex-like energy ux pattern forms at the two interfaces of a waveguide with the core made of metamaterial and claddings made of positive-indexof-refraction materials (see Fig 3.1). The vortex-like energy structure forms because the propagating energy in a metamaterial waveguide core ows in the opposite direction to that in the cladding [24]. This property plays a key role in controlling light propagation in a metamaterial waveguide [25]. The vortex-like energy ux structure property leads to a slowing down or even a 46

CHAPTER 3.

SLOW LIGHT IN MWS

47

Figure 3.1: The schematic of the double vortex-like energy ux structure in a metamaterial waveguide (based on Fig.4 in Ref. [24]).

stopping of the light traveling through the metamaterial waveguide structure. Such an anomalous feature has aroused strong interest recently due to its potential applications in data storage, optical buers, and optical sensing [5-10]. Moreover, the vortex-like surface wave property may lead to an extremely high power density in a metamaterial waveguide, which may nd potential applications in medical treatments, and industry [33, 34]. In this chapter, we investigate the fascinating features of slow-light and highpower density in a realistic metamaterial waveguide, by including both dispersion and loss. model. First, we consider a dispersionless and lossless metamaterial waveguide

We calculate the energy density, energy velocity, and the group velocity Our

for a particular mode for this model over a broadband range of frequencies.

numerical calculations show that the group velocity can exceed the velocity of light in vacuum, and the total energy density can take negative values in some frequency ranges, which of course contradicts the principle of causality and the second law of

CHAPTER 3.

SLOW LIGHT IN MWS

48

thermodynamics [97, 98]. Thus, we repeat our calculations for a lossy and dispersive metamaterial waveguide. It is well-known that the group velocity is not equal to the energy velocity, where the latter one is the correct measure of energy transport for a lossy medium. To nd the energy velocity, we need to derive the correct form of the energy density and the total power ux. In section 3.3, we derive the appropriate energy density formulation in a lossy metamaterial medium for the Lorentz medium model and the Drude-Lorentz medium model. We derive the closed form of the

total power ux and the energy velocity for a symmetric metamaterial waveguide in section 3.4. Then, we numerically investigate the energy velocity and the power density for various types of modes and structures. Finally, we study the role of

intrinsic loss on the slow-light modes of the metamaterial waveguide structures. It has been proposed recently [35], that it could be possible to slow down or even stop light in a metamaterial waveguide structure. Because of the exotic properties of the theoretical structure proposed in [35], there is considerable interest in the literature (e.g. [39, 41, 44]). However, the intrinsic loss, which is the inherent feature of any realistic metamaterial, is completely neglected in the proposed structure. At the end of this chapter, we critically re-examine the structure proposed in [35] in the presence of material loss.

3.2 Negative energy density in dispersionless and lossless metamaterials


In this section, we investigate the energy density, power ux, and energy velocity in a dispersionless and lossless metamaterial waveguide. We consider a metamaterial

CHAPTER 3.

SLOW LIGHT IN MWS

49

slab with

1 = 4

and

1 = 4

and a thickness of 2 cm, with an air cladding.

The dispersion curves of this structure have been studied in Chapter 2; from Fig. 2.3, we know that the proposed structure supports both complex modes and normal modes. In this section, we only investigate the group velocity of the particular normal mode,

T E2 ,

because the velocity of the complex modes is always zero in a lossless

medium.

We plot the dispersion curve for the

T E2

mode in Fig.

3.2(a) over the

range of frequencies for which the backward and forward propagation modes coexist. In Chapter 2, we showed that the dispersion curve of a metamaterial waveguide, in contrast to a dielectric one, is not monotonic at all frequencies and the transcendental equation may have two proper solutions at some frequencies. These two possible

solutions have two dierent power distributions. In one solution, (the red curve in Fig. 3.2) the averaged power ux in the core is greater than that in the cladding

(forward propagating mode) while in the other solution, the averaged power ux in the core is less that that in the cladding (backward propagating mode) . We dene the normalized total power ux per unit length (in

y axis)

as

total Sz

1 = Re 2

dx (E H )z .

(3.1)

We can calculate the total power ux for the proposed structure by substituting and

from Eqs.

(2.6-2.9) into Eq.

(3.1).

The results for the total power ux

normalized to the square of the electric eld amplitude,

2 E0

dened previously in

Chapter 2, are plotted in Fig. 3.2(c). As shown in Fig. 3.2(c), the total power ux can be positive (forward propagating mode), negative (backward propagating mode) or even zero when the averaged power ux in the core is completely canceled by that in the cladding (here, at 2.85 GHz). Thus, light can be stopped at that critical frequency

CHAPTER 3.

SLOW LIGHT IN MWS

50

point as shown in Fig. 3.2(d). We plot the slope of the proposed mode (group velocity) in Fig. 3.2(d) to highlight the critical point with zero group velocity. Fig. 3.2(d)

demonstrates that the group velocity behaves anomalously versus frequency in the proposed structure. Although we know that the energy velocity should be equal to the group velocity for the proposed lossless structure, we calculate the energy velocity of the proposed mode to nd an answer for the unusual behavior of the velocity. Our results conrm that the slope of the dispersion curves of the proposed mode, which indicates the group velocity velocity, density.

vg =

d , is completely in agreement with the energy d

vE = S / w

. To nd the energy velocity, we need to calculate the energy

The total energy density for real and frequency independent

and

is

obtained from [72]

w =

1 dx ( |E|2 + |H|2 ). 4

(3.2)

We calculate the total energy density over the range of the frequencies in Fig. 3.2(b) for the proposed structure. Astonishingly, Fig. 3.2(b) indicates that

takes neg-

ative values at certain frequencies or goes to zero at a critical frequency; however, classical physics ticular point,

restricts the energy density to have a positive value. At this parthe energy velocity diverges and changes sign suddenly. Fig. at 3.45 GHz, which obviously contradicts the causal-

w = 0,

3.2(d) indicates that

vE

ity principle. To address this unphysical result we next investigate, more carefully, the electromagnetic energy density in metamaterials. It is clear from Eq. (3.2) that energy density is always negative in a dispersionless metamaterial structure with both negative

and

This energy density formulation When the medium is

is only applicable for a lossless and dispersionless medium.

1 Unlike classical physics, quantum physics does not restrict the energy density to have a positive
value [75].

CHAPTER 3.

SLOW LIGHT IN MWS

51

(a)
f (GHz)

<w> = 0
f (GHz)

(b)
3.5

3.5 3 2.5 0 4

<w> = 0
3 2.5 20 4

<Sz> = 0
1
(1/cm)

<w>/|E0|2

20

40

(c)
f (GHz) f (GHz)

(d)
3.5 3 2.5 10

3.5 3 2.5 10

vE

<Sz> = 0
0 10 20
<Sz total>/|E0|2

0
vE/c

10

Figure 3.2: (a)The dispersion curve of the dispersionless metamaterial waveguide for the

T E2

mode with

this graph: a point with zero total power ux, which leads to zero energy velocity, and a point with zero total energy density, which leads to innite energy velocity. The red and blue curves indicate the mode with negative and positive total power 2 ux, respectively. (b) Total energy density (Normalized with respect to |E0 |) versus frequency. (c) The average total energy ux versus frequency. (d) The normalized energy velocity versus frequency. It shows that

1 = 4

and

1 = 4.

We highlight two important points on

vE

where

w = 0.

dispersive, the average energy density for time harmonic waves assumes the form [67]

w=

1 4

[()] [()] |E|2 + |H|2 .


(3.3) reduces to Eq. (3.2) when

(3.3)

It is immediately evident that Eq. assumed dispersionless.

and

are

It can be shown that the new-dened energy density, in a

lossless medium, is always positive even if

()

and

()

take on negative values as

in metamaterial. Hence, dispersion is an inherent feature of metamaterials since it is

CHAPTER 3.

SLOW LIGHT IN MWS

52

required in order to secure positive energy density. However, the energy density expression with Eq. (3.3) is inappropriate in a lossy metamaterial. Ziolkowski numerically showed that the above expression yields negative energy density value in a high-loss region near the resonance frequency [73]. After he pointed out this issue, several approaches have been presented to nd an appropriate expression for the energy density in a dispersive and lossy metamaterial [68, 69, 70]. In the next section, we investigate this important energy density problem in detail.

3.3 Energy density in a dispersive and absorptive media


Formulating the energy density in a dispersive and absorptive metamaterial for a number of dierent models have been addressed several times [68, 69, 70]. We make contact with previous work on this issue to present the correct form of the energy densities for a Lorentz and Drude-Lorentz medium model. Although there is no general formulation for the electromagnetic energy density in such material, the electromagnetic energy density can be derived for each specic model by employing directly Poynting's theorem. Poynting's theorem, which is a statement of conservation of energy in the electromagnetic eld, is also valid in a lossy, dispersive medium. Poynting's theorem can be easily obtained from Maxwell's equations [67]. For a source-free medium, it takes the form

S+

we wm + = 0, t t

(3.4)

CHAPTER 3.

SLOW LIGHT IN MWS

53

where

is the Poynting vector of the electromagnetic wave, and

we

and

wm

are the

electric and magnetic energy densities, which can be generally dened as [72]

D we = E, t t

wm B = H . t t

(3.5)

Now, we derive energy densities in the Lorentz medium model. This simple classical model consists of a collection of non-interacting electrons of displacement coupled to the external electric eld according to the equation of motion

r,

m(

2r r 2 + e + eo r) = qE, 2 t t eo
is the resonance frequency,

(3.6)

where

is the damping factor,

and

are the

eective mass and charge of the electrons. We can rewrite Eq. (3.6) in terms of the electric polarization vector,

P = qr ,

as follows

2P P 2 2 + e + eo P = 0 ep E, 2 t t
where

(3.7)

ep q 2 /m0

is the plasma frequency and indicates the strength of the in-

teraction between the electron oscillators and the electric eld. By using Eq. (3.5) and the relationship between the electric displacement, namely

D,

and the electric eld,

E,

D = 0 E + P ,

we have

we D E P = E = 0 E+ E . t t t t

(3.8)

CHAPTER 3.

SLOW LIGHT IN MWS

54

Substituting

from Eq. (3.7) into Eq. (3.8), we obtain

1 1 we = 0 [E]2 + 2 t t 2 20 ep
Hence, the electric energy density is

P t

2 eo

[P ]

e P + 2 0 ep t

(3.9)

1 1 we = 0 [E]2 + 2 2 20 ep

P t

+
2 eo

[P ]

P e dt 2 0 ep t

,
(3.10)

we + we .
We can separate the electric energy density in Eq. (3.10) into two parts. The second part,

we ,

represents the heat energy due to loss; as this is not energy in the eld, we

neglect this part from now. The rst part represents the stored electric eld energy density, which has the form

1 1 we = 0 [E]2 + 2 2 20 ep

P t

2 2 + eo [P ]2 .
(3.11)

We can simplify Eq. (3.11) by adopting a time harmonic electromagnetic eld. Hence, the time-averaged stored electric energy density that follows directly from Eq. (3.11), is

1 1 2 we = 0 |E|2 + 2 + eo |P |2 . 2 4 40 ep
We then express the time-harmonic polarization,

(3.12)

P,

in terms of the electric eld by

using Eq. (3.7). Finally, the electric energy density in terms of electric eld for the

CHAPTER 3.

SLOW LIGHT IN MWS

55

Lorentz medium model can be written as

we =

2 2 ( 2 + eo ) ep 0 1+ |E|2 . 2 2 )2 + 2 2 4 (eo e

(3.13)

From Eq. (3.13), it is obvious that the electric energy density always takes a positive value in the lossy, dispersive Lorentz model even though the permittivity may be negative. Similarly, we can write the equation of motion of the magnetic polarization. It has the form

M 2M 2 2 + m + mo M = F mp H, 2 t t
where

(3.14)

is the damping factor,

mo

is the magnetic resonance frequency,

mp

is

the plasma frequency,

indicates the strength of the interaction between the dipole

and the external magnetic eld, and

is the magnetization vector.

After some

algebraic manipulations, we can derive the time average magnetic energy density for the Lorentz medium model with

2 F mp . () = 1 + m = 1 + 2 mo 2 im
The stored magnetic energy density is

(3.15)

2 2 F ( 2 + mo ) mp 0 wm = 1+ |H|2 . 2 2 )2 + 2 2 4 (mo m
From Eq.

(3.16)

(3.16), it is clear that the magnetic energy density remains positive at

all frequencies. Hence, the total energy density is always positive in a metamaterial structure that obeys the Lorentz medium model. Next, we derive the energy density

CHAPTER 3.

SLOW LIGHT IN MWS

56

for the Drude-Lorentz medium model with

() = 1 +

F 2 , 2 mo 2 im
2 ep

(3.17)

() = 1

2 + ie

.
This means

We note that there is no general formulation for the energy density.

that the problem of nding the energy density should be solved separately for each model. However, we can always dene the energy density properly by using the

Poynting's theorem. The energy density must be positive even in a double-negativeindex material according to the second principle of thermodynamics. It has been

shown [69] that the time-average electric and magnetic energy densities for the DrudeLorentz medium model dened by Eq. (3.17) are

2 ep 0 we = 1+ 2 |E|2 , 2 4 + e
(3.18)

wm =

2 2 F 2 [mo (3mo 2 ) + 2 2 ] 0 m |H|2 . 1+ 2 2 2 )2 + 2 2 4 mo (mo m

Again, the total energy density remains positive at all frequencies, even near resonance frequencies or high loss regions. If the losses are negligible,

e = m = 0,

Eqs.

(3.13,3.16,3.18) reduce to the well-known energy density formulation, namely Eq. (3.3) in a dispersive medium for a narrow-band electromagnetic waves [67].

CHAPTER 3.

SLOW LIGHT IN MWS

57

3.4 Power ux, group velocity, and energy velocity in a lossy and dispersive metamaterial waveguide
In this section, we rst derive the closed-form of the total power ux in a symmetric metamaterial waveguide structure. We then show that the total power ux is always zero in a lossless metamaterial waveguide for the complex modes. Second, employing the total energy ux and energy density from the previous section, we derive a correct measure of transport, the energy velocity, in a lossy metamaterial waveguide. Finally, we analyze the behavior of the metamaterial waveguide structure, previously studied in Chapter 2, by calculating the power ux and the energy velocity of the proposed structure for some particular modes . By denition, the total power ux of the waveguide structure is the spatial average of the Poynting vector over the guide's cross section, which has the following normalized (per length in

y axis)

form [72]

total = 1 Re Sz 2

dx (E H )z .

(3.19)

The total power ux of a waveguide structure can be expressed as the sum of the power ux in the core and in the cladding,

total core cladding Sz = Sz + Sz .

For a

symmetric waveguide with the thickness of 2d, the total power ux can be expressed as

1 total Sz = 2 Re

+d
d

dx (E H )z + 2

+
+d

dx (E H )z ,

(3.20)

where the rst term is the power ux in the core and the second term is that in the cladding. Substituting Eqs. (2.6-2.9) into Eq. (3.20), the closed-form of the total

CHAPTER 3.

SLOW LIGHT IN MWS

58

power ux is

total Sz

2 r i c |E0 | e2 z sinh (2k1 d) sin (2k1 d) = Re + r i 4 1 k1 1 k1


(3.21)

2 c |E0 | e 2

2 i z

Re

i r cosh (2k1 d) + cos (2k1 d) , r 2 k2

where

r i r i k1 = k1 + ik1 , and k2 = k2 + ik2

are the complex wave numbers in the core and

cladding, respectively. The rst two terms in Eq. (3.21) corresponds to the energy ux in the core and the last two terms corresponds to the energy ux in the cladding. Since the sign of the permeability in the core and in the cladding are dierent, the energy ux in the core is in the opposite direction of that in the cladding. Thus,

the total energy ux can become negative or even zero in a metamaterial waveguide. It can be analytically shown that the total energy ux is always zero,

total = 0, Sz

for the complex modes of the lossless metamaterial waveguide. After some algebraic manipulations and separating the real and imaginary parts of wave numbers in the transcendental equation, Eq. (2.10), for complex mode, we obtain

r k2 =

r r i i 1 sinh (2k1 d) /k1 + sin (2k1 d) /k1 . r i 2 cosh (2k1 d) + cos (2k1 d)

(3.22)

Substituting Eq. (3.22) into Eq. (3.21) gives

total Sz = 0.

Thus, the complex modes

in a lossless metamaterial waveguide never carry energy, although the normal modes do. However, we note that

total Sz = 0 does not remain valid in a lossy metamaterial


Therefore, in a lossy model, the

waveguide because of the complex permeability.

complex modes carry energy as we show later in our numerical calculations (see e.g. Fig. 3.4 or Fig. 3.5).

CHAPTER 3.

SLOW LIGHT IN MWS

59

Finally, in this subsection, we nd the closed-form of the energy velocity for the proposed structure. Dening the energy velocity,

vE , as the magnitude of the average

of the Poynting vector divided by the average energy density [70], we have

+ dx Sz Sz . = + vE w dx (we + wm )
Using Eqs. (3.18,3.21), and introducing

(3.23)

P1 Re

r i sinh (2k1 d) sin (2k1 d) + , r i 2k1 2k1

P2 Re

r i cosh (2k1 d) + cos (2k1 d) , r k2


(3.24)

0 1+ 2 , 4 + 2 e

2 p

2 2 F 2 [o (3o 2 ) + 2 2 ] 0 m 1+ , 2 2 4 o (o 2 )2 + 2 2 m

the closed-form of the energy velocity for a symmetric waveguide structure, is

vE = w + w

c 2 ( k1 )c 1
2

P1 P2 + 1 2 P1 + 2 + 2 ( + k2 )c 2
2

. P2

(3.25)

Since the energy velocity measures the transport of the electromagnetic energy, it must be always smaller than the speed of light in vacuum. In the following, we show that

CHAPTER 3.

SLOW LIGHT IN MWS

60

this condition is preserved in the proposed metamaterial waveguide at all frequencies even near the resonance frequency or high-loss regions. Though, it can be negative for the backward propagation modes. It is important to note that the energy velocity is dierent from the group velocity in general. Since the energy velocity the group velocity,

vE

is equal to

vg d/d ,

in a lossless medium [67, 70], the group velocity can

be considered as a correct measure of transport in a lossless medium. However, the group velocity can be a misleading concept in lossy media. For example, from Fig. 2.5(b), we can conclude that the group velocity goes to innity near the resonance frequency or the magnetic surface polariton frequency. Thus, if the group velocity

shows the energy transport, the causality principle would certainly be violated in a lossy metamaterial waveguide. In some recent publications (e.g. [42, 76]), the group velocity is incorrectly used as the measure of transport in lossy metamaterial media; however, it is well-established in the community that the energy velocity, correct measure of transport in lossy media [67].

vE ,

is the

3.4.1

Numerical analysis

We consider a 2-cm-thick metamaterial slab with

() = 1

0.56 2 2 16 + i0.016

&

() = 1

100 , 2 + i0.016

(3.26)

with a free-space cladding as shown in Fig. 2.1(a). In Chapter 2, we have previously studied the modal characteristics of the proposed structure for the TE and the TM modes. Now, we investigate the energy velocity,

vE , and the total power ux behavior

of this proposed structure . In Chapter 2, we have also calculated the dispersion curves of this structure for the rst ve TE modes (see Fig. 2.5(b)). Here, we study the

CHAPTER 3.

SLOW LIGHT IN MWS

61

energy velocity behavior of the

T E1 , T E2 ,

and

T E5

modes. The other modes have

similar energy velocity behavior.

5.5

5.5

(a)
5
f (GHz)

(b)
f (GHz)

4.5 4 3.5 0 5.5 5

TEB 1

5 4.5 4 15 20 3.5 0 10 5.5


1 2 3

TEB 5
(1/cm)

10

10

10

10

vE/c

(c)
5
f (GHz)

(d)
5
f (GHz)

4.5 4 3.5
4 2 0 2

4.5 4 3.5
4 2 0 2

10

10
<Sz

10
>/|E0|
2

10

10

10

10

10

total

<Score>/|E0|2 z

Figure 3.3: (a) The blue curve indicates the real part of propagation constant for the backward propagating

T E5

and

T E1

coupling mode. A metamaterial slab, with

a thickness of 2 cm, is in free-space and we use the standard Lorentz-Drude model p = 10 GHz, 0 = 4 GHz, = 0.1 GHz to calculate its permittivity with F = 0.56, 2 2 and permeability. (b) the energy velocity versus frequency. (c) The normalized total energy ux and (d) The normalized energy ux in the core versus frequency.

In Fig. 3.3, we investigated the total power ux and the energy velocity of the backward-propagating

T E1

and

T E5

modes. In the previous chapter, we have shown

that the surface polariton backward-propagation modes, plex

B T E1 ,

couple with the com-

B T E5

modes in the presence of the material loss; hence, instead of two backward

propagating

T E1

and

T E5

modes, we have one coupled mode as shown in Fig. 3.3(a).

Also shown in Fig. 3.3, are the computed (b) normalized energy velocity, (c) total power ux, and (d) power ux in the core versus frequency. Fig. 3.3(b) indicates that

CHAPTER 3.

SLOW LIGHT IN MWS

62

the normalized energy velocity never exceeds the speed of light in vacuum and never goes to zero; the

vE

of the coupled backward mode is always greater than 0.001 c and

near the surface magnetic plasmon frequency (at 4.71 GHz) has the minimum value. We do not show the loss number here, but by looking at Fig. 2.5, we can conclude that the imaginary parts of the propagation constant is large whenever

vE

is small;

thus, the slow-light modes decay very fast before they can propagate an appreciable distance. From Fig. 3.3(d), we can see that the power density in the core is anoma-

lously large near the magnetic surface polariton frequency. The dimensionless power ux in the core has a maximum peak at the magnetic surface polariton frequency,

fms = 4.71 GHz,

which

10, 000

times greater than the dimensionless power ux at

frequencies away from

fms .

However, it is important to note that the total energy ux

is considerably smaller at this point because of the vortex-like property of a metamaterial waveguide. The power density in the core is partially canceled by the power density in the cladding; hence, the total power ux is much smaller than the power ux in the core. Consequently, the energy velocity has a small value at

fms

although

the energy propagates very rapidly in the core or in the cladding. The high-power density property comes from the phase matching condition ,

core + cladding = 0,

at

fms .

This property is also been seen in another metamaterial waveguide structure,

so-called super waveguide, proposed in Ref. [33, 34] . In Fig. propagating 3.4, we study the energy velocity and the power ux for the forward-

T E1 and T E5 modes.

The behavior of these modes is completely dierent First, we note that no coupling occurs

to that of the backward-propagating ones. between the

F T E1

and

F T E5

modes, which are shown with the red and blue curves,

CHAPTER 3.

SLOW LIGHT IN MWS

63

5.5

5.5

(a)
f (GHz)

4.5

f (GHz)

TEF 1

(b)
5 4.5 4 4 10 5.5

TEF 5
4 0 5.5 1
(1/cm)

10

10

10

10

vE/c

(c)
5
f (GHz) f (GHz)

(d)
5 4.5 4

4.5 4 3.5 0 0.25 0.5 0.75 1

<Sz total>/|E0|2

<Score>/|E0|2 z

Figure 3.4: (a) The dispersion curves of the lossy, dispersive metamaterial waveguide for the forward propagating parameters used in Fig. respectively. 3.

T E1
(b).

(red curve) and

T E5

(blue curve) modes with the

The normalized energy velocity versus frequency.

(c) and (d) indicate the normalized total power ux and power ux in the core,

respectively, in Fig. 3.4. Second, the high-power transmission property disappears in these types of modes. Fig. 3.4(d) indicates that the maximum power density

in the core for the forward-propagation modes at the magnetic surface polariton frequency is at least 1000 times smaller than the backward-propagating modes. From the previous chapter, we know that the forward-propagating modes have the lower cuto frequency, where the dispersion curve intersects with the light-line. At this point, the energy velocity is equal to the speed of light in the cladding, and monotonically decreases with increasing frequencies as shown in Fig. 3.4(b). However, the energy velocity is never zero.

CHAPTER 3.

SLOW LIGHT IN MWS

64

Next, we calculate the energy velocity and the total power ux of the In Fig. 3.5(a), we plot the real part of the propagation constant, mode.

T E2

mode.

Re[],

for the

T E2

The normalized power ux with respect to the eld intensity is shown in

Fig. 3.5(c). As expected, the total power ux has a positive value for the forwardpropagation mode and a negative value for the backward-propagating mode at all frequencies. Fig. 3.5(d) indicates that the power density in the core has a peak at the magnetic surface polariton frequency. However, the magnitude is small in comparison with the power density in the core for the coupled

T E5

and

T E1

modes. Hence, the

high-power transmission property is only seen for the coupled modes.

(a) 5
f (GHz)

(b) 5
f (GHz)

4.5 4 0

TEB 2 TE2
F

4.5 4

(1/cm)

10 (c)
f (GHz)

|vE|/c

10

10

5
f (GHz)

5 4.5 4 3

(d)

4.5 4 0.5

0.5
Stotal/E2 z 0

1
Score/E2 z 0

B Figure 3.5: (a) The dispersion curves of the backward propagating T E2 mode (blue) F and the forward propagating T E2 mode (red). (b) log-plot of the normalized energy velocity versus frequency. (c) The normalized total power ux and (d) the power ux
in the core for the

T E2

modes versus frequency.

We calculate the energy velocity over a broadband range of frequency for

T E2

CHAPTER 3.

SLOW LIGHT IN MWS

65

modes (Fig.

3.5(b)).

Although our calculations conrm that the energy velocity

cannot go to zero in a lossy metamaterial waveguide, the energy velocity is extremely small near the resonance frequency (smaller than

106

c ). However, at this resonance

frequency, the propagation loss is massive. To have a better insight, we re-plot the energy velocity for the backward-propagation

T E2

mode and the propagation loss

versus frequency in Fig. 3.6. As can be seen, when the group velocity is larger than 0.01 c, the loss [20Im[]/ log(10)] is smaller than 0.01 dB/cm. However, the loss is greater than 150 dB/cm at the resonance frequency. Thus, at resonance, we have an ultra-slow-light mode with extremely high loss!

10

(a) |vE|/c
10
2

10

10

4.5

Loss (dB/cm)

(b)

150 100 50 0 3.5

4.5

5.5

f (GHz)

Figure 3.6: (a) The normalized energy velocity and (b) the propagation loss for the backward-propagation

T E2

mode versus frequency.

CHAPTER 3.

SLOW LIGHT IN MWS

66

3.5 The role of intrinsic loss on slow-light modes


In a recent Nature Letter [35], Tsakmakidis et al. proposed a metamaterial waveguide structure that theoretically achieves stopped light under single-mode operation and over a broadband range of frequencies, forming what they call a trapped rainbow. Their work has generated considerable interest recently; for example, papers have been written examining a trapped rainbow on metallic grating structures [39], stopped-light in an anisotropic waveguide [41, 42], and a trapped rainbow in graphene [44]. While an interesting theoretical study, Tsakmakidis et al. made a bold assumption, namely that their proposed metamaterial was completely lossless. In this section, we revisit the predictions proposed in [35], and critically analyze the important inuence of essential metamaterial loss on the slow-light modes. The basic analysis of Tsakmakidis et al. builds on the work of I.V.Shadrivov et al., which predicted that the vortex-like energy ux property of a metamaterial waveguide leads to stopped light at some critical thickness. In other words, the power ux

direction inside the metamaterial layer is opposite to the one in the dielectric layers; thus, when the power ux in the core is completely canceled by the power ux in the cladding, light can be stopped. Previously, in our dispersion curves calculations for a lossless metamaterial waveguide with a xed thickness (see e.g. Fig. 2.5(a))

we have shown that at some critical frequencies the TE or TM modes have zero group velocity. We could also plot the propagation constant,

versus thickness for

a xed frequency. Hence, we can nd the zero-group velocity modes by changing the thickness of a waveguide. Using this approach, Tsakmakidis et al. have proposed

a wedge-shaped metamaterial structure, which can stop various frequencies of light altogether at dierent points, forming a trapped rainbow. It is interesting to note

CHAPTER 3.

SLOW LIGHT IN MWS

67

that the relation between the frequency of a light beam and the thickness of the dispersionless waveguide is dierent than in the dispersive waveguide, as waves with larger frequencies are stopped at larger core thickness in a dispersive metamaterial waveguide. We show the schematic of the wedge-shaped waveguide in Fig. 3.7(a). The green light is stopped at larger core thickness while the yellow light, with smaller frequency, is stopped at smaller core thickness.

(a) (a)
-d
Region(I) Region(II)

(b)

d
Region(I)

2d

Region(I) Region(II) Region(III)

Figure 3.7: (a) The schematic of the wedge-shaped metamaterial waveguide proposed in [35]. The green light with higher frequency is stopped at higher core thickness while the yellow light is stopped at smaller core thickness. (b) The schematic of the metamaterial waveguide structure used in our calculation.

To avoid complications, we consider an asymmetric waveguide structure with a xed thickness rather than a wedge-shape waveguide, in which a lossy metamaterial layer is surrounded by two regular dielectrics as shown in Fig. 3.7(b). The relative permittivity and permeability of the three waveguide layers (top to bottom) are and

i ()

i (),

where

i = {1, 2, 3}.

As in Ref. [35], the permittivity and permeability of are lossless and dispersionless,

the cladding layers

while for the lossy metamaterial we use the standard Drude and Lorentz medium models, Eqs. (2.22,2.23). In all calculations, we use and we take the metamaterial slab thickness to be

(1 = 2.56 3 = 2.25 1 = 3 = 1) 1

F = 0.6, p =

6c , o =

9c ,

d/2 = 0.55c/c ;

these parameters

CHAPTER 3.

SLOW LIGHT IN MWS

68

have been chosen so as to yield identical results to those presented in [35] for when

= c

= 0.

In our loss model, we take

= 0.01c .

We investigate the rst three TM modes, and calculate the complex waveguide dispersion curves in the presence of metamaterial loss and dispersion by using transcendental equations (2.15,2.16) for the proposed asymmetric waveguide structure. In Fig. 3.8, we plot (a) the real part and (b) the imaginary part of the normalized propagation constant versus the normalized frequency with respect to

c . It is impor-

tant to note that we mainly use dimensionless units in Fig. 3.8; therefore, converting to any frequency is trivial, and our essential results applies to all frequencies. The characteristic model of an asymmetric waveguide and a symmetric waveguide structure are similar. However, some minor dierence can be observed between them; for example, in comparison with a symmetric waveguide (e.g. see Fig. 2.5(b) and Fig. 8), there are two light-lines, which correspond to

and

3 ,

in Fig. 3.8. Moreover,

the TM modes are perturbed at two frequencies, which correspond to the magnetic surface polariton frequency for

and

3 . The T M1

mode (blue curve in Fig. 3.8), be-

cause of its surface polariton characteristic, distinctly changes near these two points,

ms1 1.28c & ms2 1.35c .

However, the other TM modes are not considerably

changed at the magnetic surface frequencies because the phase matching condition,

core +cladding = 0, is not satised simultaneously at both waveguide interfaces.


the guidance condition,

Since

|n2 | > max{n1 , n3 },

only is satised at the frequency below

1.14 c ,

we re-plot the dispersion curves of the proposed structure for the frequency

range of Fig. 3.9. In light of dispersion curves of Fig. 3.9, we can conclude that, rst, it is impossible to have single-mode operation ; second, there is no point with zero group velocity at any frequency. Therefore, the claims in Ref. [35] are not valid when

CHAPTER 3.

SLOW LIGHT IN MWS

69

loss is included. To obtain better insight, we highlight the critical frequency region for

T M2

modes in Fig. 3.9, and compute the lossless model (purple curve) for

T M2

mode in this region. Using lossless model, we recover the key result of [35] and obtain a point of zero group velocity for

= c . c

However, for the realistic case, where loss is

included, the mode dispersion near

changes dramatically: the complex mode splits

in two and the normal modes pulls away from its lossless counterpart; rather than having a normal mode and a complex mode, one obtains complex forward and backward propagation modes. The inuence of material loss on mode structure is highly non-perturbative near

c ,

where even a tiny amount of loss produces unusually large

changes in the dispersion such that the group velocity is never zero.

1.4 1.3

1.4 1.3 light line

1.2 TM3 1.1 1 0 15

TM1

1.2 1.1

TM2
30 45 60

1 0 15

Metamaterial light line 30 45 60

2Re[]c/c

2Im[]c/c

Figure 3.8: Dispersion curves of the proposed waveguide structure in Fig. 3.7(b) with

= 0.01c

versus dimensionless frequency for TM modes.

(a) The dimensionless

normalized real part of the propagation constant, of the propagation constant,

Re[],

and (b) the imaginary parts

Im[].

CHAPTER 3.

SLOW LIGHT IN MWS

70

1.02 1.13 1.09 1.05 1 TM3 0.95 6 12 18 24 0.99 12 metamaterial light line TM2 light line TM1 1.01 =0 =0.01c

/c

/c
1 vE = 0 18 24

2c/c

2c/c

Figure 3.9: Dispersion curves of the proposed waveguide structure in Fig. 3.7(b) for TM modes with the

= 0.01c .

The dashed curves show the light-lines and metamaterial

light-line. The right plot shows the dispersion curves near the critical frequency for with

T M2 mode for the lossless model (purple curve) and the lossy model (green curves) = 0.01c .

In Fig. 3.10(a), we rotate Fig. 3.9 and plot frequency versus normalized propagation constant near

c .

From Fig. 3.10(a), it is apparent that the group velocity

does not go to zero; but, as we mentioned before, the energy velocity is the correct measure of transport in the lossy media, not the group velocity; therefore, we plot the energy velocity for

= 0.01 c

in Fig. 3.10(b), which shows that the energy velocity,

in marked contrast to the lossless case, is not zero at any frequency. Although the energy velocity goes as low as

0.003c, the imaginary part of the propagation constant

is large whenever the energy velocity is small, which leads to very high propagation losses. In Fig. 3.9(b), we also plot the propagation loss (=

20 Im[]/ ln[10])

in prac-

tical units of dB/cm for a range of frequencies in the optical (c /2 Wherever

= 385 THz).

vE < 0.01c,

the loss is greater than 300,000 dB/cm. To put this number

CHAPTER 3.

SLOW LIGHT IN MWS

71

in proper context, if this waveguide were used as a delay line, then for a delay, equal to 1000 times the fundamental period of the eld (e.g., for

D ,

c /2 = 385 THz,

D = 2.6 ps) the loss would be at least 200 dB. It is interesting to note that high losses
are also observed in slow-light band-edge modes in photonic crystal waveguides when there is even a tiny amount of surface-roughness [74] . In both cases, the large loss is a direct consequence of the large time that the slow light spends near the potential absorbers and/or scatterers.

30

TMB 2 vE = 0

(a)

2c/

25 20 15

(b)
10
6

10

10

10

10

0.98

1.02

1.04

10

Dimensionless frequency (/c)


Figure 3.10: Dispersion curves of the proposed waveguide structure for with

Loss factor (dB/cm)

10 0 10

TMF 2 1 1.02 1.04

10

|vE|/c

T M2

mode

= 0 (red curve) and with = 0.01 c

(blue and green curves). (b) The energy

velocity and propagation loss versus frequency for the loss model with = 0.01 c B F for T M2 (blue) and T M2 (green) modes. The propagation loss is calculated for a critical frequency in the optical (c /2

= 385 THz).

However, our general ndings

and conclusions scale to GHz and THz. For example, at a frequency of 1 THz, the loss is approximately 1,000 dB/cm or greater; this is completely impractical as a delay element, given the corresponding energy velocity.

The loss number,

= 0.01 c ,

used in our calculations, is 30 times better than

CHAPTER 3.

SLOW LIGHT IN MWS

72

has been experimentally achieved to date near optical frequencies and comparable to state-of-the-art at GHz frequencies [19, 20]. Therefore, it is expected that results would be even worse in a real metamaterial structure due to a higher loss number. We consider again the proposed metamaterial waveguide structure of Fig. 3.10; but this time with velocity with 3.11(a).

= 0.0001 c . The blue and red curves indicate the normalized energy
and

= 0.01 c

= 0.0001 c c ,

for the

T M2

mode, respectively in Fig.

Below the critical frequency,

the energy velocity has similar behavior

for the both cases; however, the group velocity is 100 times smaller in the latter one above the critical frequency. The essential point is that the energy velocity is never zero even if we introduce very small amount of loss into the metamaterial structure. Therefore, the assumption that the metamaterial metallic elements can be treated as perfect, lossless, conductors in the GHz domain is, of course, misleading. The neglect of loss is certainly not acceptable in a regime of slow light in metamaterials. In Fig. 3.11(b), we plot the loss propagation and the energy velocity with

Re[]/Im[] ratio versus the T M2


mode. The loss

= 0.01 c

and

= 0.0001 c

for the

propagation is very high or the

Re[]/Im[]

ratio is very low whenever the energy

velocity is smaller than 0.01 c. Fig. 3.11(b) indicates that the slow-light modes are extremely lossy even in the presence of small amount of loss

( = 0.0001 c ).

Regarding the possibility of compensating for the loss by introducing gain into the metamaterial: it has been rigorously shown in [48] that is impossible to eliminate loss in a negative-index material (except possibly at a single frequency). This is the result of the basic requirement of causality. Thus, we conclude that it is not possible to stop light over a rainbow of frequencies even in an active metamaterial waveguide structure.

CHAPTER 3.

SLOW LIGHT IN MWS

73

(a) 10 10 |v |/c 10 10 10
E

0 1 2 3 4

TMF
2

~ = 0.01 ~ = 0.0001

TMB 2

(b) 10 r/i ratio 10 10 10

1 1.05 1.1 1.15 Normalized frequency (/o) 10 10 10 10


4 3 2 1 0

TMB 2

10 10 10 10 10 Normalized energy velocity (|vE|/c)


Figure 3.11: (a) The energy velocity of the proposed waveguide structure with

Loss factor (dB/cm)

TMF 2

0.01

(blue curve) and

= 0.0001

(red curve) versus frequency, where

= /c .

=
(b)

The solid curves indicate loss propagation versus the energy velocity and the dashed curves indicate

r /i

ratio of the proposed structure in Fig. 3.7(b). 1

3.6 Concluding remarks


In this chapter, we have examined the slow-light mode properties of a metamaterial waveguide structure in the presence of absorption and dispersion. To study this property, we have derived the energy ux, energy density, and energy velocity. We have shown that to ensure positive energy density and to satisfy the causality principle, metamaterials should be dispersive and therefore should be lossy at some frequencies.

CHAPTER 3.

SLOW LIGHT IN MWS

74

Next, we have derived the correct form of the energy density and energy velocity in a dispersive and lossy metamaterials. We have shown that the intrinsic loss nonperturbatively aects the slow-light modes. In conclusion, our calculation for various modes, TE and TM modes, and structures, a symmetric and asymmetric structure, indicates that it is impossible to stop light in metamaterial waveguides such as those proposed in [35].

Chapter 4

Excitation of the modes of metamaterial waveguide structures

4.1 Introduction
In the previous chapters, we studied the properties of the modes of a metamaterial waveguide structure. Since the modes are almost all below the light-line, excitation of the modes is dicult unless a special technique for phase-matching is employed. One of the well-known techniques is the attenuated total reection (ATR) method, which has been commonly used to couple surface polariton modes in metals with an incident beam using a prism (see Fig. 4.1). In 2000, when Ruppin investigated the properties of surface polariton modes in metamaterials, he used the ATR method in his numerical analysis [21]. Then Smith et al., based on the Ruppin idea, employed the ATR method to excite surface polariton modes in a metamaterial slab [22]; and Park
et al. extended Ruppin's work to the bulk and surface modes [78]. However, these

authors only made the connection between lossless dispersion curves of a metamaterial 75

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

76

waveguide and the minimum reectivity of the ATR method with the angle xed while varying the frequency. In this chapter, we develop the ATR method to study a lossy waveguide structure at a xed frequency while varying the angle. Also, we study the limitations of the ATR method and suggest how the accuracy of the ATR method might be improved by optimizing the air gap and the refraction index of the prism. However, before we study the ATR method, we rst derive the generalized transfer matrix formalism [84, 85, 86], which is applicable to a lossy multilayer system with a negative-index slab. The chapter closes with a brief look at the eect of the excited modes on the lateral Goos-Hanchen beam shift [87, 88].

4.2 Transfer matrix formalism for lossy metamaterial multilayer structures


The transfer matrix method is a standard technique to calculate the reectance and transmittance of multilayer structures. Since the combination of a prism and a waveguide in the ATR method form a multilayer system (see Fig. 4.1), we can employ the transfer matrix method to calculate the reectance of that system. It is important to note that the standard transfer matrix formalism needs to be slightly modied to be applicable for lossy metamaterial structures. In this section, we derive the appropriate transfer matrix formalism that is applicable for metamaterials and absorbing media.

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

77

The angle varied

incident nprism > n1 n1 = 1

prism reflected air gap z x

Surface wave

n2 < 0

metamaterial

Figure 4.1: Schematic geometry of the ATR method. A plane wave of frequency is incident on the boundary of a prism and air, with the angle of incident, dip in the reection at certain angles.

varied,

which couples with a surface wave on the metamaterial slab. Hence, it can lead to a

4.2.1

Theory

We consider a plane wave incident on a stratied isotropic medium with boundaries at

x = dm , with the permittivity, m , and the permeability, m . In each region, m and


in general can be negative or complex (see Fig. 4.2). The rst (0 , 0 ) and last

(t , t ) regions are semi-innite, and a plane wave is incident from the rst region with wave vector k and the angle

as shown in Fig. 4.1. The electromagnetic wave has a

plane of incidence parallel to the x-z plane; thus, all eld vectors are independent of
y. We separate the Maxwell equations in each region into those for the TE and TM

modes. From Eq. (2.3), we know that the wave equation for the electric eld has the

(fig 1). The region (t) is semi-infinite . the permittivity and permeability in each region are m & m , the plane wave is incident from region 0 and has the plane of incident parallel to the y-z plane. All field vectors are dependant on y and z only and independent of x. since = 0 , the Maxwell equations in any region can be separated into TE and x TM components governed by Em and Hm.
CHAPTER 4. EXCITATION ON THE MODES OF MWS 78

0 , 0

1, 1

2 , 2

3 , 3

4 , 4

l 1 , l 1

l ,l

z x
d1 d2 d3 d4 dm dl

Figure 4.2: The schematic of a one-dimension multilayer structure.

The TE waves are completely determined by,


form

2 2 + 2 + 2 m m 1/(r)0 , {1/(r) E(r)} = 2 /c2 E(r). ) Emx = y 2 z 1 H my = Emx We assume that all the regions of the aforementioned structure are i m z (
the electric eldE each region for the modes is described by in H =
mz mx

(4.1)

isotropic; hence,

1 im y

2 2 2 + 2 E m (x, z) = 2 m m E m (x, z), z 2 x c


and the magnetic eld in each region is obtained from

(4.2)

E m = i m H m . c

(4.3)

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

79

Using Eqs. (4.2) and (4.3), the electric and magnetic elds in each region for the TE modes are

+ E m y = eiz (Em eikm x + Em e+ikm x ),

Hm z = Hm x =

ckm iz e (Em eikm x m

+ Em e+ikm x ),
(4.4)

c iz e (Em eikm x m

+ + Em e+ikm x ),

H m y = E m x = E m z = 0,
where the amplitude

+ Em

and

Em

represent all wave components in each region m,

which propagate along the positive and negative

direction, respectively.

From

Chapter 2, we know that the component of wave vector in in

is

and the component

x, k m ,

in each region m is obtained from

km =

m m

2 2 = 2 c c

m m 0 0 sin2 ,

(4.5)

which satises the phase matching conditions and the dispersion relation.

Finally,

using the fact that the tangential components of the electric and magnetic eld should be continuous at

x = dm+1 ,

to satisfy boundary conditions, and Eq. (4.4), we have

+ + Em eikm dm+1 + Em e+ikm dm+1 = Em+1 eikm+1 dm+1 + Em+1 e+ikm+1 dm+1 ,

km (Em eikm dm+1 m

+ Em e+ikm dm+1 ) =

km+1 (Em+1 eikm+1 dm+1 m+1

+ Em+1 e+ikm+1 dm+1 ).

(4.6)

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

80

Choosing coordinates in which we get

dm+1 = 0 and expressing Eq.

(4.6) in the matrix form,

+ + Em Em+1 = M m,m+1 , Em Em+1


where

(4.7)

1 M m,m+1 = 1
k m m km m

1 x
to

1
km+1 m+1
where

x,

1
k m+1 m+1

(4.8)

We then transform the coordinates from

x = x dm+1 ,

and obtain

E m+2 (x) = E m+2 (x dm+1 ) = eikm+1 (xdm+1 ) eikm+1 (xdm+1 )

+ Em+2 Em+2

;
(4.9)

hence,

+ Em+2 Em+2

e =

ikm+1 dm+1

0 e
ikm+1 dm+1

+ Em+2 Em+2

.
(4.10)

Using (4.7), we have


where

+ Em Em

= T m,m+1

+ Em+1 Em+1

,
(4.11)

T m,m+1 = 1 (1 + 2 (1

km+1 m ikm dm )e km m+1 km+1 m +ikm dm )e km m+1

(1 (1 +

km+1 m ikm dm )e km m+1 km+1 m +ikm dm )e km m+1

.
(4.12)

The transfer matrix formalism can be used to determine the electromagnetic eld amplitudes in any region in terms of those in any other region. In particular, the

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

81

transmission, t, and reectance, r, coecients of the proposed l-layer structure in Fig. 4.1, with respect to a plane wave amplitude, can be calculated with

(4.13)

1 t = T 1,l , r 0
where

l T11 T12 T 1,l = = T m,m+1 . T21 T22 m=1


Then, Using the z-component of the Poynting vector ,

(4.14)

Sz = 1/2 Re[E H ] z ,

the

power of the incident, reected and transmitted plane wave can be expressed as

Pi = Re

k0 c 2 E , 20 0

PR = Re

k0 c 2 2 r E0 , 20

PT = Re

kl c 2 2 t E0 . 2l
(4.15)

We normalized the incident power to unity; thus, the conservation of the power yields

r2 + Re

kl 0 k0 l

t2 = 1 Ploss ,

(4.16)

where

Ploss is the absorption power.

Using Eqs. (4.13, 4.14, 4.16), the power transmis-

sion coecient of the aforementioned structure for the TE modes can be expressed as

k T = Re l 0 k0 l
where is

1 T11

(4.17)

T11

is the element of the transfer matrix

T , and the power reection coecient


2

R=

T21 T11

(4.18)

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

82

We emphasize that Eqs.

(4.17, 4.18) are applicable for any multilayer structure

containing metamaterial with intrinsic loss for the TE modes. We also note that the angle of incident

is implicitly contained in Eqs. (4.17, 4.18) in the denition of

km .

The transfer matrix for the TM modes can be obtained from duality by changing

in Eq. (4.12), and the transmission coecient for the

T M modes

is

T = Re

k l 0 k 0 l

1 T11

(4.19)

4.3 The Attenuated Total Reection technique (ATR)


The dispersion curves of a metamaterial waveguide structure cannot be excited directly by a light beam because, apart from the complex modes, they are conned below the light-lines; hence, indirect methods, such as grating techniques or the attenuated total reection (ATR) technique, need to be used to couple into the normal modes with an external light beam. In this section, we rst review the basic principles of the ATR method and then show how the ATR method can be used to obtain the dispersion curves of a metamaterial waveguide structure. We then study the effect of the intrinsic loss, the thickness of the air gap (the distance between a prism and a metamaterial slab), and the refraction index of the prism on the dispersion curves obtained from the ATR method. We close this section by nding the optimum conditions to obtain

R=0

in the ATR method.

4.3.1

Basic principles

In the previous chapters, we have shown that the dispersion curves of a metamaterial waveguide structure for the normal and surface modes are always below the light-line

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

83

(for example, see Fig. 2.5). Thus, the wave vector of a beam of light, k, is always smaller than the propagation constant,

of the normal and surface modes, and

phase-matching cannot be achieved even at normal incidence ( if we add a prism into the system (see Fig.

> k sin ).

However,

4.1), a beam of light will have

kz =

prism k sin ,

which can be sucient to excite the modes with

< kz .

Thus, the

normal and surface modes with propagation constants

between the air light-line and

prism light-line can be coupled with a light beam. In Fig. 4.3, we show schematically the allowed region in a (, f )-plane. The dark and light blue lines represent the lightline and prism light-line with

prism = 9.

The prism light-line can be shifted to the

right or left by changing the index refection of the prism; thus, we can excite the normal and surface modes with any propagation constants with an appropriate refraction index. In the ATR method, two forms of experimental scan are possible: an angle scan with the frequency xed, and a frequency scan with the angle xed. The trajectory of the former is a straight-line parallel to the

by choosing a prism

axis and the trajectory of the latter is

a line between the air light-line and the prism light line as shown schematically with the red dashed lines in Fig. 4.3. In both forms, the mode is excited where the scan line crosses the dispersion curve; however, the dispersion curves measured by angle scan and frequency scan are dierent. The angle scan measures the dispersion curve in the real- and complex- space while the frequency scan measures the dispersion curves in the real- and complex- space. This is discussed in more detail in Chapter 5. Fig. 4.1 shows the schematic geometry of the ATR method with the angle scan. A beam of light is incident through one side of a prism with

prism > air ,

where

is

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

84

5.2 Air Lightline

= const Prism Lightline

frequency (GHz)

4.7

fms

f = const 4.4 Metamaterial Lightline prism = 9 2 3 4 5

4 0

air = 1 1

fres

(1/cm)

Figure 4.3: Dispersion curves and prism coupling. Only the modes, with the propagation constant between the air light-line and the prism light-line, can be excited by the ATR method. The blue line shows the prism light-line with prism light-line can be shifted to higher

prism = 9.

The

by increasing the index of refraction of the

prism. Two forms of experimental scan (frequency xed and angle xed) in the ATR are shown schematically with red dashed-lines.

the permittivity. The beam passes through the prism and is incident on the second interface at an angle

If the incident angle,

is larger than the critical angle for , the beam undergoes total internal

total internal reection,

> c = sin1

air prism

reection and exists at the third side of the prism. But, because of the presence of the metamaterial slab near the second interface, the beam is not totally reected. The decaying evanescent wave of the light beam on the air side of the prism can couple to the normal and surface modes of the metamaterial waveguide; thus, the reectivity is reduced below unity, and a dip is observed in the reectivity at the angle at which the parallel component of the wave vector of the light beam matches with the propagation constants

of the modes. It is important to note that the minimum in the reectivity

does not yield the exact value for

of the normal and surface modes of the waveguide

because the waveguide structure is perturbed by the presence of the prism. However,

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

85

by controlling the air gap and the index of the prism, we can minimize the error. In the following section, we numerically nd the optimum conditions for the ATR method.

4.3.2

The optimum conditions for the ATR method

Using the ATR method, we can excite the normal and surface modes of a metamaterial waveguide structure; however, the main challenge in the ATR method is to nd an optimal distance between the prism and the slab. If the air gap is too small, the

system is over-coupled and the prism massively perturbs the waveguide system; thus, the excited mode will not be observed. On the other hand, if the air gap, d, is

too large, the system is under-coupled; thus, the mode is only weakly excited and only a small reectivity dip can be observed. We show numerically the eects of

the air gap width on the reectivity in Fig. 4.4. In our numerical results presented in Fig. 4.4, we consider a lossy metamaterial slab with the following parameters: in the Lorentz-Drude model. The

F = 0.56, p = 10 GHz, 0 = 4 GHz, = 0.1 GHz 2 2

thickness of slab is 2 cm, and we choose a prism with

prism = 6

(the schematic

geometry of the system is shown in Fig. 4.1). We present the calculation dispersion curves for the

T E1

mode of the proposed structure in Fig. 4.4(a). Since the mode

is calculated in the complex- space, we use the angle scan with the frequency xed at 4.8 GHz in the ATR method. The scan line, the red line parallel to the crosses the forward and backward propagating

axis,

T E1

modes at

= 27

and

= 70 ,

respectively. Hence, we expected dips appear in the reectivity at these angles. In Fig. 4.4(b), we show the calculated reectivity of the proposed system by using the transfer matrix method for four dierent air gap widths,

d = /2, /4, /8, and /16,

where

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

86

is the free-space wavelength of the incident light beam (here,

= 30/4.8 = 6.25
because the

cm). When

d = /2,

only a small reectivity dip is seen near

= 27

forward propagating mode is only weakly coupled with the beam, and the air gap is too large to excite the backward propagating mode. the air gap, However, when we reduce

d = /4,

the eectiveness of coupling is enhanced and results in almost

zero reectance at the minimum for the forward propagating mode; moreover, the backward propagating mode now can very weakly couple with the light beam. The very shallow dip can be observed near

= 65 ,

which corresponds to the excitation

of the backward propagating mode. As the air gap reduces further,

d = /8,

the dip

of the backward propagating mode becomes deeper and the width of the dip becomes smaller; however, the dip of the forward propagating mode disappears. It is important to note that the minimum reection seen at total internal reection,

= 23

is below the critical angle of the

c = 24 . Thus, it does not correspond to the excited forward d = /16,


the minimum reection goes up and

propagating mode dip. Finally when

the width of the dip becomes broader and tends to disappear; moreover, the place of the dip slightly shifts to the left. This means that the dierence between the

angle of minimum reection and that obtained from the dispersion curve starts to increase when the air gap reduces more than a specic value. In a later section, we attempt to nd numerically the optimal value for the air gap; but, as the rule of thumb

d /2 /4

for the forward propagating mode and

d /4 /8

for the

backward propagating mode. With the simple numerical analysis used to produce Fig. 4.4, we have shown

that the choice of the optimal air gap is a challenging part of the ATR method. In the literature, several schemes have been developed to improve the accuracy of the

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

87

4.9

(a)
4.88 4.86 0.5 0

d=/2

(b)

Reflectance |R|

0.5 0

d=/4

F (GHz)

4.84 4.82

TE

d=/8
0.5 0

= 70

4.8 4.78

= 27
0.5 0 1

d = / 16

4.76

(1/cm)

20

30

40

(degree)

50

60

70

80

Figure 4.4: Eects of the air gap width, d, on the reectance. (a) The

T E1 mode of the

waveguide proposed in 2.5. At 4.8 GHz, the forward and backward propagating T E1 modes are excited with the ATR technique at = 27 and = 70 , respectively (red lines). The prism we used has respectively, where

prism = 6 and the air gap is d = , /4, /8, and /16,

= 6.25

cm.

ATR method [81, 82, 83].

For example, the dispersion curves obtained from the

ATR method under the condition where there is a zero refraction in respect with the air gap, is studied with the exact solutions. Using the zero-reection condition, the theoretical expression for the optimal air gap has already been derived in the literature (see for example [81]); however, in this section, we numerically analyzed the optimal air gap for the aforementioned structure to give physical insight about the relationship between the optimal air gap, optimal refraction index of a prism,

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

88

intrinsic loss and the modes. Again, we consider the waveguide structure with a prism, which was previously studied in Fig. 4.4, and calculate the reectivity at 4.8 GHz under the zero-reection condition. The optimal air gap and the minimum refraction angle can be obtained by using the transfer matrix method and zero-refraction condition. The optimal

air gaps for the forward and backward propagating modes are 1.89 cm and 0.83 cm, respectively. However, Fig. 4.5 indicates that the dips corresponding to the backward and forward propagating modes are located at smaller angles than obtained from the dispersion relation. The dierence between the exact value and that determined

by the minimum reection are around loss,

and

7 ,

respectively.

The large intrinsic

= 0.1

GHz in the proposed structure, and the index of the prism are the To study the eect of the intrinsic loss, we recalculate

potential sources of errors.

the reectivity under the zero-refraction condition for the aforementioned structure with the material loss a hundred times smaller than the previous calculations. The results are shown with the red curves in Fig. 4.5. As shown, in this case, there is a very good agreement between the solutions obtained from the ATR method and the dispersion equation. Therefore, we can conclude that the accuracy of the ATR method is reduced by the intrinsic loss even under the zero-reection conditions. In a very lossy waveguide structure or for the complex modes, this dierence becomes so signicant that the ATR method under the zero-reection condition cannot be used to calculate the dispersion curves. To obtain better insight, we calculate the dispersion curves of the waveguide structure for the

T E1

mode by using the ATR method for two dierent material loss pa-

rameter values in Fig. 4.6. As shown in Fig. 4.6(a), there is an excellent agreement

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

89

a
|R|
0

1 0.5 = 0.1 = 0.001

1 0.5

b
|R|

0 1 0.8 0.6 0.4 0.2 0 20 30 40 50 60 70 80

(degree)

Figure 4.5: Reectance with an optimal air gap for dierent material loss. (a) The optimal air gap for the forward propagating cm with

T E1

mode is

d = 1.89

cm and

d = 6.05

= 0.1

GHz and

= 0.001

GHz, respectively. All the other parameters cm with

are the same as for Fig. 4.4. (b) The optimal air gap for the backward propagating

T E1

mode is

d = 0.83

cm and

d = 1.76

= 0.1

GHz and

= 0.001

GHz,

respectively.

between the dispersion curves for the normal

T E1

mode determined by the ATR

method under the zero-reection condition (red curves) and those calculated directly (blue curves) when the intrinsic loss is small (

= 0.001GHz) for the propagation con-

stant above the prism light-line. Nevertheless, the dierence between them for the complex modes becomes larger when the frequency increases. Hence, we can conclude that the ATR method only works when the imaginary component of the propagation constant

is small.

In other words, the ATR method fails for the complex mode

at higher frequency even in the very low loss waveguide structure because at that

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

90

frequency the imaginary component of

is not negligible. We also calculate the dis-

persion curves by using the ATR method for the waveguide with the moderate loss (

= 0.1GHz)

in Fig. 4.6(b). As mentioned before, the dierence between the exact

and calculated dispersion curves are signicant in this case as we cannot observe the splitting of the complex mode in the dispersion curves calculated by the ATR method. Hence, the ATR technique needs to be improved in order to be applicable for all the modes in lossy waveguide structures.

4.9 (a) Frequency (GHz) 4.85 prism lightline Frequency (GHz)

4.9 (b) prism lightline

TE1

4.85

TE1
4.8

4.8

1.5

2 2.5 (1/cm)

3.5

1.5

2 2.5 (1/cm)

3.5

Figure 4.6: The dispersion curves of the system proposed in Fig. 4.4 calculated by the dispersion equation (blue solid curves) and by the ATR method with an optimal air gap (red dashed curves). Although (a) there is a good agreement between the solutions of the two methods for the small loss, between them are signicant for

= 0.001

GHz, (b) the dierence

= 0.1

GHz.

Fig. 4.6 The dispersion curves of the system proposed in Fig. 4.4 calculated by the dispersion equation (blue solid curves) and by an ATR method with an optimal of the prism can be controlled and using a prism with the appropriate index of refracair gap (red dashed curves). Although (a) there is a good agreement between the tion is an integral part of the experiment. If prism is = 0.001 GHz, (b) the dierence solutions of the two methods for the small loss, small, only a small propagation between them are signicant for = 0.1 GHz. constant range can be excited by the prism. On the other hand, if is too large,
prism
the waveguide structure is considerably perturbed by the prism with a high index

In the ATR experiment, not only the prism air gap, but also the index of refraction

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

91

refraction and the properties of the waveguide structure destroyed. In the following, we try to numerically nd the optimal air gap and refraction index of a prism to yield agreement between the dispersion curves calculated by the ATR method and the exact solutions.

4.88

prism = 3 lightline

prism = 12 lightline

4.86 Frequency (GHz)

4.84

4.82

4.8

TE1
4.78

4.76

1.5

2.5 (1/cm)

3.5

Figure 4.7: The dispersion curves of the waveguide structure proposed in Fig. indicate those with

4.4

using the ATR method with an optimal air gap. The green and red dashed curves

prism = 3 and prism = 12 ,

respectively; and the blue solid curves

show the exact solution of the dispersion equation.

In Fig. 4.7, we show the dispersion curves using the ATR method under the zerorefraction condition with two dierent prisms. The prism with low-index,

prism = 3,

(the green curve in Fig. 4.7) is suitable to use for exciting the forward propagating

T E1

mode since the

of this mode has a small value; while the high-index prism,

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

92

prism = 12,
agating

(the red curve) is more suitable to use for exciting the backward propmode. We notice that neither the low-index nor the high-index prisms

T E1

could correctly demonstrate the complex mode properties, as the splitting property of the lossy complex modes cannot be seen with the dispersion curves obtained from the ATR method. Since the imaginary component of the propagation constant

of the complex mode is considerably larger than the normal modes, the dispersion curves obtained by the ATR method depart from the exact solution more rapidly in the complex modes.

4.9

4.9

(a)
4.85 4.85

(b)

f (GHz)

TEF 1
4.8

TEB 1

f (GHz)
4.8 3 4 4.75 0

4.75

(1/cm)

4.9

d (cm)

(c)
4.85

f (GHz)

4.8

4.75

10

nprism
Figure 4.8: Optimum air gap and

nprism

for the

T E1

mode of the waveguide proposed

in Fig. 4.4. (a) The red and blue curves show the forward and backward propagating

T E1

mode. (b) The optimum air gap and (c) The optimum index of refraction of a

prism for those modes.

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

93

Finally, we derive the optimal air gap and the refraction index of the prism to yield agreement between the dispersion curves calculated by the ATR method and the exact solutions (see Fig. 4.8). To nd the optimal air gap and the refraction

index, we insert the propagation constants

calculated directly by the dispersion

equation, into the transfer matrix formalism, and solve the zero-refraction equation for the air gap and the refraction index of the prism. The optimal air gap and index of the prism are shown in Fig. 4.8(b) and Fig. 4.8(c), respectively. It can be seen that the range of the optimal air gap is between 0.5 cm and 3.5 cm (

/12 /2,where
1
is valid

= 6.25 cm), and it is larger for the forward propagating mode than for the backward
propagating mode. It seems that the simple rule mentioned previously,

here. However, the result for the optimal refraction index is more interesting. For the forward propagating mode, the optimal index starts just a little above the vacuum index refraction (1) and slowly increases with increasing frequency.

We can explain

this result by the fact that the prism perturbs the waveguide structure. If the index of the prism is close to that of the waveguide cladding, then the eect of the prism perturbation can be minimized. However, to satisfy exciting the mode with propagation constant as the

/c < nprism ,

the condition for

, the index of the prism should increase

of the mode is increased. This is the reason why the backward propagating

modes need a high-index prism to be excited. The point that is not clear for us in Fig. 4.8(c) is why is the optimal index of refraction of the prism is so large for the complex modes. For future work, the excitation of the complex modes with the ATR method needs to be studied further.

1d

mode. 2 increases rapidly for the complex modes.

/2/4 for the forward propagating mode and d /4/8 for the backward propagating

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

94

4.4 Anomalous Goos-Hanchen shift due to surface mode excitation


As an application of the excitation of the waveguide modes, we close this chapter by looking at the giant Goos-Hanchen shift due to the excitation of a mode in the metamaterial waveguide structure. First, we briey introduce the Goos-Hanchen shift particularly in metamaterials; then, we numerically show that the Goos-Hanchen shift can be large when the mode excitation occur. When a light beam is totally internally reected from a plane interface between two media, the reected beam is displaced forward (positive) or backward (negative) by a distance

parallel to the interface (see Fig.

4.9). This phenomenon is known as

the Goos-Hanchen (GH) shift [72] and happens because the interference of dierent transverse wave vectors are undergoing dierent phase changes. The GH shift was rst predicted by Newton; however, it took two centuries before Goos and Hanchen could experimentally proved this lateral beam shift. In a dielectric medium the GoosHanchen shift is usually positive; however, it could be negative in a lossy multilayer dielectric structure. In 2002, Berman showed that the GH shift is negative if one of the interfaces is a metamaterial [36]. Actually, the negative sign of the permeability and permittivity in metamaterials lead to the negative GH shift. After Berman's

paper, much attention was drawn to the GH shift in metamaterials (for example [37, 87, 88]). Chen et al. showed that the GH shift can be negative as well as positive in a metamaterial slab and Kiveshar et al. [37] predicted that the GH shift can be giant

in a multilayer structure containing a metamaterial.

Usually, the GH shift is much less than the beam width and cannot be easily

3 It is comparable or even larger than the wavelength of a beam (see Fig. 4.10).

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

95

Incident beam

prism nprism > n1

Reflected beam

Incident beam nprism > n1

Reflected beam

n1 = 1

GH shift

air gap z x

n1 = 1

GH shift

air gap z x

n2 < 0

metamaterial

n2 < 0

metamaterial

Figure 4.9: Schematic of the positive and negative lateral beam shift (Goos-Hanchen shift).

measured in practice; however, a large GH shift can be seen in a multilayer structure whenever surface modes of a structure are excited. Since the excited mode can transfer energy along the interface, the lateral GH beam shift signicantly intensies. In the following, we calculate the GH shift for the structure determined in Fig. 4.4. We have

1 shown that the forward propagating

T E1

mode is coupled with a prism with

=6

under the zero-refraction condition near

= 25 .

Next, we calculate the normalized

GH shift in respect to the wavelength of the light beam versus the angle between 20 to 30 degree in Fig. 4.10. The Goos-Hanchen shift can analytically be calculated as [87]

=
where

, 2

(4.20)

is the wavelength of an incident light beam, and

is the phase of the

reection coecient.

By using the transfer matrix formalism, we can obtain the

phase of the reectivity,

and then the GH shift.

As predicted, our calculation

shows the large negative GH shift,

/ 19,

near the minimum reectivity angle,

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

96

= 25 .

Moreover, there is another small peak near

= 24 ,

which corresponds to

the critical total internal reection angle. It is important to note that the GH shift is sensitive to the prism air gap. In our calculation, we derived the optimal air gap using the zero-reection condition, which is 1.89 cm for the forward propagating mode.

T E1

5 c 0

5 GH/ |R|

10

0.5

15

0 20

40 60 (degree)

80

20 20

22

24 26 (degree)

28

30

Figure 4.10: The lateral Goos-Hanchen (GH) shift for the reected beam of the system described in Fig. 4.4. Inset shows the reectance with the optimal air gap (d cm).

= 1.89

4.5 Concluding remarks


In this chapter, we have derived the generalized matrix transfer formalism, which is applicable for a lossy multilayer containing metamaterials. Also, we have discussed the ATR method for exciting the normal and surface modes of a metamaterial waveguide structure. We have mentioned that the careful choice of an optimal air gap is an

CHAPTER 4.

EXCITATION ON THE MODES OF MWS

97

integral part of the ATR method and it can be found by using the zero-reection condition. Although the modes obtained from the ATR method under the zero-

reection condition give a good agreement with the exact modes in a structure with a small intrinsic loss, the modes obtained by the ATR method considerably depart from the exact modes in a waveguide structure with a moderate loss. We have suggested that in a lossy waveguide structure and for the complex modes, the index of refraction of the prism also needs to be optimized. Finally, we have shown that a large GoosHanchen shift can be seen at the excited mode angle as the sign and magnitude of the GH shift is sensitive to the air gap and index of refraction of the prism.

Chapter 5

Dispersion curves of complex frequency modes of metamaterial waveguide structures

5.1 Introduction
The dispersion curves of a waveguide structure can be obtained from the dispersion equation,

D(, ) = 0,

where

is the angular frequency and

is the wave vector.

The modes, in general when intrinsic loss is introduced in a structure, are stated with complex

and complex

however, they can be represented by two subspaces,

= {(, ) : R} and = {(, ) : R} , in a four-dimensional complex complex

hyperspace. In the previous chapters, we have investigated the dispersion

curves of a metamaterial waveguide structure by looking for the modes with a complex and

and a real

( ).

However, it is not self-evident whether to make

complex

real, or vice-versa, or even to make both of them complex [89, 90, 91, 92]. In 98

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

99

this chapter, we calculate the dispersion curves of a metamaterial waveguide for real

and complex

( )

for several metamaterial waveguide structures; and compare subspace. We nd a signicant dierence between the dis-

with the diagrams in

persion curves calculated in

with those calculated in

for the complex modes.

The dierence is understandable since these two approaches solve fundamentally different problems. The real frequency approach

( )

describes solutions where the

wave decays in space, while the real wave vector approach tions that decay away in time.

( )

studies the excitacan be realized

The real frequency approach

( )

experimentally when the structure is excited with monochromatic light, and can be computed using frequency-domain methods, such as the vectorial eigenmode expansion technique [89]. However, the real wave vector approach

( )

can be realized

when the structure is excited with highly-collimated ultra-short pulse, and can be calculated using frequency-dependent time-domain method computationally. Next, we show that the dispersion curves of both

and

approaches can be

reproduced by using the ATR method, but with dierent experimental conditions. For the

approach, we should measure the ATR reectivity with the frequency

xed and the angle varied, while for the latter approach, we should measure the ATR reectivity with the angle xed and the frequency varied [79]. Thus, one can measure the dispersion curves of both types of modes experimentally. Finally, we calculate the dispersion curves of a lossy metamaterial waveguide structure employing perturbation theory for both these approaches. We show that there is some disagreement between the solutions obtained by perturbation theory with the exact solutions near the stopped-light mode regions in the

approach; however,

the perturbation theory can work well in the other regions. As we have mentioned in

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

100

the previous chapters, the inuence of the material loss on mode structure is highly non-perturbative near the stopped-light mode regions, so this disagreement is not surprising.

5.2 Comparison between and approaches


CHAPTER 5. DISPERSION CURVES OF MWS WITH COMPLEX- 127

In this section, we numerically investigate the dispersion curves of the systems that have been studied in 2.5, both in complex- and complex- approaches

( and ).

Light line
5

TE3

(a)
Frequency (GHz)

TE
5

Light line

(b)

Frequency (GHz)

TE2

Metamaterial light line

TE2 TE

Metamaterial light line

TE1

0 0

0.5

1.5

(1/cm)

2.5

3.5

0 0

0.5

1.5

(1/cm)

2.5

3.5

Figure 5.1:

Dispersion curves of the dispersionless and lossless metamaterial slab

in free space with

= = 4 (a) for the complex- approach ( ), and (b) for the complex- approach ( ) . The thick and the thin curves indicate the real and imaginary parts of for (a) and frequency for (b), respectively.
First, we consider a 2-cm-thickness metamaterial slab with

= = 4

in free

space. Fig. 5.1 indicates the calculated normal and complex modes of the proposed structure for (a) complex- and (b) complex- states. Since the proposed model is lossless, the calculated normal modes of the metamaterial waveguide structure for the

approach are identical to those calculated for the

approach. However, we nd

Figure 5.1: Dispersion curves of the dispersionless and lossless metamaterial slab in the free space with imaginary parts of the complex- approach

= = 4 (a) for the complex- approach ( ), and (b) for ( ) . The thick and the thin curves indicate the real and for (a) and frequency for (b), respectively.

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

101

that the complex modes are completely dierent between the 5.1(a) shows that the real part of the wave vector,

and

states. Fig.

of the complex modes starts at

the cut-o frequency, the critical point where the frequency has its minimum value and the group velocity is zero, and monotonically decreases. On the other hand, in Fig. 5.1(b) the real part of the frequency for the complex modes starts at the cut-o propagation constant, the critical point where the energy density is zero and and goes to zero when

vg ,

0.

It is interesting to note that the complex modes are

almost above the light-line in the complex- state, and look like the leaky modes in a conventional waveguide, while the complex modes in the complex- state are not restricted above the light-line. However, the complex modes of a dispersionless metamaterial waveguide in both approaches are the proper solution of the dispersion equation as we have previously shown for
CHAPTER 5. DISPERSION CURVES OF MWS WITH COMPLEX-

approach.

117

0.35 4.7 Re[f] (GHz) 4.6 4.5 4.4 4.3 4.2 4.1 0 1 TEproper 2 TEImproper
2

(a)

Real approach Re[f] (GHz)

4.7 4.6 4.5 4.4 4.3 4.2

(b)

Real approach TEproper 2

0.15 TEImproper 2 0 1 2 Re[] (1/cm) 3 4

0.05

2 (1/cm)

4.1

Figure 5.2: The dispersion curves of the dispersive, lossless metamaterial waveguide p for the T E2 mode with F = 0.56, = 10 GHz, and 0 = 4 GHz for the Lorentz2 2 Drude model of the metamaterial slab and with air claddings (a) for a real- , complex-

and (b) a real- , complex- approach. The blue and red curves indicate the proper

and improper solutions of the dispersion equation, respectively.

Next, we calculate the dispersion curves of a dispersive and lossless metamaterial


Figure 5.2: The dispersion curves of the dispersive, lossless metamaterial waveguide p = 10 GHz, 0 = 4 GHz for the for the T E2 mode with these parameters F = 0.56, 2 2 Lorentz-Drude model of the metamaterial slab and with the air claddings (a) for a real- , complex- and (b) a real- , complex- approaches. The blue and red curves indicate the proper and improper solutions of dispersion equation, respectively.

Im[f] (GHz)

0.25

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

102

waveguide structure in the complex- approach. Fig. 5.2 shows the

T E2

mode in the

complex- and the complex- subspaces; moreover, we calculate the improper

T E2

mode, the unphysical solution of the dispersion equation which violates the guidance condition (see section 2.4), in Fig. 5.2 (red curves). As expected, we have the identical dispersion curves for the normal

T E2 mode and dierent solution for the complex T E2

mode in the real- and the complex- states; however, compared with the previous case, (the dispersionless waveguide structure), we nd that the dierence between the complex- and the complex- states in the complex

T E2

mode is much more

striking. While analytically we have previously shown that the complex modes are always the proper solutions in the complex- state, Fig. 5.2(b) indicates that the

complex modes are the improper solutions in the complex- state. In other words, the complex modes of a dispersive metamaterial waveguide surprisingly disappear in the complex- approach. Finally, we compare the dispersion curves of a lossy and dispersive metamaterial waveguide in the complex- and complex- states. In Fig. 5.3, we show the

T E2
in

modes of the aforementioned lossy metamaterial waveguide with the

= 0.1 GHz

and

states. The calculated dispersion curves for the

state are quite

dierent from those calculated for the

state. As we have mentioned in Chapter 2,

the intrinsic loss modies considerably the dispersion curves in the complex- state; for example, the curves that bend back near the resonance frequency or stopped-lightmode regions are completely destroyed in the presence of material loss. However, the dispersion curves, astonishingly, remain almost unchanged by introducing the material loss for the complex- approach. Fig. 5.3(c) clearly indicates that the real part of dispersion curves of the lossy metamaterial waveguide with

= 0.1 GHz is similar to

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

103

(a)
frequency (GHz)

complex approach
4.4

frequency (GHz)

4.7

4.7

(b) complex approach

4.4

4 0 2 4 Re[] (1/cm) 6 7

4 0
3

4 Im[] (1/cm)

8.5

x 10

(c)
4.7 Re[f] (GHz) Im[f] (GHz)

(d) complex approach


9 9.5 10

4.4

complex approach

4 0 2 4 (1/cm) 6 7

10.5

4 (1/cm)

Figure 5.3: The dispersion curves of the dispersive, lossy metamaterial waveguide for

= 0.1 GHz. The red curve indicates (a) the real and (b) the imaginary part of in a subspace. The blue curve indicates (c) the real and (d) the imaginary part of frequency in a
the mode with the parameters given in the Fig 5.2, and with subspace.

T E2

those of the lossless metamaterial waveguide. We have repeated our calculation for the proposed waveguide with the higher loss in the complex- approach, and highlight that the real part of the dispersion curves are evidently insensitive to the material loss in this approach. Hence, the point with zero slope survives in the presence of intrinsic loss in the complex- approach; however, the mode is very lossy at this critical point (see Fig. 5.3(d)). It is important to note that the conventional denition of the power ux, energy density, and energy velocity are not correct in the complex- approach;

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

104

hence, the derivation of their correct forms are indeed an interesting problem for future work.

10

10

10 Loss (dB/cm)

10
1

10

10 10
0

10

4.3

4.32

4.34 4.36 4.38 Frequency (GHz)

4.4

10

Figure 5.4: The group velocity (blue) and the loss propagation (red) of the metamaterial waveguide proposed in Fig. 5.2 for the critical frequency,

T E2 mode in the complex- approach. At 4.4118 GHz, the group velocity becomes zero while loss is innitive.

We calculate the group velocity and the loss propagation of the aforementioned lossy metamaterial waveguide for the bound

T E2

mode in the complex- approach

|vg|/c

over a broadband range of frequencies in Fig. 5.4. At the critical frequency point,

f=

4.4118

GHz, the group velocity is zero; however, the loss propagation is enormously

large near the critical frequency as it goes to innity at

f = 4.4118
or

GHz. It is well

established that the group velocity denition in the complex

space is only Therefore,

meaningful when the loss propagation has comparatively small value.

the small or even zero group velocity value in the complex- approach is, indeed, unphysical, and it does not show the correct measure of transport. We now derive the relationship between the dispersion curves in the complex-

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

105

and the complex- approaches. We follow a similar analysis as in [89] . In general, both

and

have complex value in a transcendental dispersion equation,

D(r + ii , r + ii ) = 0.

(5.1)

We consider a point (0 , 0 ), where

and

are real. Using a Taylor expansion of

the dispersion equation, and keeping only a rst-order term, we have

D(, ) = D(0 , 0 ) + ( 0 )
So, using Eq. (5.2), we have

0 ,0

+ ( 0 )

.
0 ,0

(5.2)

D(0 , 0 + r + ii ) = D(0 , 0 ) + (r + ii )

D D

= 0,
0 ,0
(5.3)

D(0 + r + ii , 0 ) = D(0 , 0 ) + (r + ii )

= 0.
0 ,0

Hence,

vg

=
0 ,0

D D /

=
0 ,0

r + ii . r + ii

(5.4)

If the dispersion curves in the complex- and complex- coincide at a point (0 , 0 ), we have

i =

i . vg
and

(5.5)

Eq. (5.5) gives us the relationship between the imaginary part of this equation is only valid when the intrinsic loss is small. In Fig. 5.5, we show the imaginary part of

However,

of the lossy metamaterial waveguide

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

106

4.4

4.35 4.41 Frequency (GHz) 4.3 4.4 4.25 4.39 4.2 0.1 4.15 0.2 0.3

4.1

0.1

0.2

0.3 Im[] (1/cm)

0.4

Figure 5.5: The imaginary part of the propagation constant, space for the

in the complex-

T E2

mode of the waveguide proposed in Fig. 5.2. The solid blue curve

represents the exact solutions of the imaginary part of represents the imaginary part of highlighted in the inset.

and the dashed red curves

calculated from Eq. 5.5. The critical region, (slow-

light mode region), where there is some disagreement between the two solutions, is

for the

T E2

mode in the complex- space directly by using the transcendental disper-

sion equation (blue curve). Next, we derive the imaginary part of group velocity and the imaginary part of

by inserting the

in the complex- space (this result is given

by the red dashed curve). Fig. 5.5 demonstrates that the relationship in Eq. (5.5) holds well except near the critical frequency region where

vg 0.

Since the group

velocity has a small value and the intrinsic loss has non-perturbative inuence near the critical frequency region, it is expected that Eq. (5.5) will fail. It is interesting to note that the relationship in Eq. (5.5) is not valid in a photonic crystal medium near the band edge, where the group velocity also has a small value [89].

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

107

5.3 The experimental conditions of the ATR method


In this section, we used the ATR method to measure the dispersion curves of a lossy metamaterial waveguide structure in both complex- and complex- approaches. The experimental conditions of the ATR method in the complex- approach is slightly dierent from those in the complex- approach. In Chapter 4, we mentioned that

the real part of propagation constant can be found from the momentum conservation equation

Re[] = (nprism /c)x sin min .

(5.6)

We have assumed that the incident frequency is xed and that the incident angle is varied in the experiment. Since the minimum of the reectance is associated with

excitation of the modes, the dispersion curves in the complex- space can be found from Eq. ( 5.6). However, if we are interested in reproducing the dispersion curves in the complex- space by using the ATR method, we have to change the initial condition in the ATR experiment. This time, we x the incident angle and vary the incident frequency; hence, the

in the complex- approach can be obtained from

Re[] = (nprism /c)min sin x .

(5.7)

It is interesting to note that this issue was discussed extensively in 1970's. Though calculated dispersion curves of a lossy metallic waveguide demonstrated that the surface plasmon modes bend back at dispersion curve did not bend back at

fms , fms .

Otto [80] found that the experimental However, Arakawa et al. reported that

the curve did bend back in their experiment [93]. Several years after Otto's paper, Kovener et al. [79] found that the two diering results of Otto's experiment and

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

108

Arakawa's experiment are not contradictory, and both of them satisfy the Fresnel equations. However, Otto's method gives the dispersion curves in the complex-

space while the Arakawa's method gives the dispersion curves in the complex- space.

4.8 TE0

= 58 = 70

1 (a) 0.8 |R| 0.6 0.4 Akawaras method 40 f = 4.715 B f = 4.725 A 60 (deg) 80 100 (b)

4.75 f (GHz) A A B 4.7 |R| B C f = 4.725 fms

0.2 20 1

0.8

C B

np = 8 (c) np = 7

0.6
np=6 np=7 np=8

Ottos method 10 0.4 4.65 4.7

A np = 6 4.75 f (GHz) 4.8

4.65 0

4 6 (1/cm)

T E0 surface polariton mode of the metamaterial waveguide proRe[] bends back near fms in the complex- space (blue curve) while Re[] at fms in the complex- space (red curve). (b) Reectance versus the incident angle with nprism = 6 and dgap = 0.5 cm. As expected from (a), min 70 . (c) Reectance versus frequency with dgap = 0.5 cm and = 70 . The labeled letters, (A,B,C), represent the resonance frequencies for nprism = 6, 7, 8, reFigure 5.6: (a) The posed in Fig 5.2. The spectively.

The same scenario arises in a lossy metamaterial waveguide. The surface polariton modes bend back near

fms

in the complex- approach while they do not bend back In Fig. 5.6, we show this phenomenon by using the

in the complex- approach. ATR technique.

We consider a 2-cm-thickness metamaterial slab with dispersion

and absorption with air claddings. The blue curve in Fig. 5.6(a) indicates the

T E0

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

109

mode in the complex- space. The surface polariton

T E0

mode bends back near the

magnetic surface polariton frequency, 4.71 GHz, and couples with the However, in the complex- approach, the

T E4

mode.

T E0

mode does not bend (red curve) and

at fms . First, we simulate Arakawa's experiment with frequency xed and the
incident angle varied. We need a prism with an extremely high index of refraction to excite surface polariton modes at a region where the dispersion curve bends back. The dashed lines in Fig. 5.6(a) are the prism light lines at indices of refractions. excite the As shown in Fig.

= 70

for three dierent cannot

5.6(a), a prism with

nprism < 6

T E0

mode where its

>6

(1/cm). Thus, we set a prism 0.5 cm above We then calculate the reectance for

the metamaterial waveguide with

nprism = 6.

the frequency xed at 4.4725 GHz. The minimum reectance is at is in good agreement with the calculated dispersion curve,

68 ,

which

exact = 70 ,

(where the

prism light-line and blue dispersion curve cross each other in Fig. 5.6(a)). We repeat our calculation for the frequency xed at 4.4715 GHz. The minimum reectance is now at

53 ,

(labeled B in Fig.

5.6(b)).

This value is slightly dierent with

the exact solution,

exact = 58 ,

because it behaves like a complex mode rather than

surface polariton mode at this frequency. However, this ATR technique veries that the dispersion curve bends back because the location of the dip goes to lower angles by decreasing frequency. Next, we simulate the ATR technique proposed by Otto to measure the

for the

aforementioned metamaterial waveguide structure. The distance between the prism and the waveguide is again xed to be 0.5 cm, and the beam propagates from the prism with the incident angle xed, indices of refractions,

x = 70 .

We plot the reectance for three dierent

nprism = 6, 7, 8;

their dips are labeled with 1,2,3, respectively.

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

110

As expected, the resonance frequencies decrease by increasing the index of refraction (see Fig. 5.6(a) and (c)). If the dispersion curve bents back, we should not observed the minimum reectance for

nprism = 7 or 8.

Hence, it is important to emphasize

that the Otto's ATR technique does not work well for the dispersion curves of a lossy waveguide in the complex- approach if the loss is too high. By looking at Fig. 5.6(c), we nd that the magnitude and the width of the dips are changed by modifying the index of refraction. Since the propagation loss (e.g. see Fig. 2.5(b)) signicantly

increases from point A to point C, the width of the dip labeled C is greater than the dip labeled A. Finally, by the same method, we can show that the bound modes bend back at the resonance frequency in the complex- approach while they do not bend back in the complex- approach. In Fig. 5.7, we show the role of material loss on the reectance spectrum. It

is well-known that we cannot use the ATR method to measure the dispersion curve of a lossless waveguide structure. If a waveguide is lossless, the reectance would The

be unity because no energy could be coupled to the surface or bound mode. insets in Fig. 5.7 show the with

T E2

mode of the aforementioned metamaterial waveguide

= 0.1

GHz, (blue solid curve), and

= 0.01

GHz, (red dashed curve) near

the critical frequency region, (stopped-light mode region), in both complex- and complex- approaches. We have previously shown that the dispersion curves in the presence of material loss drastically pull away from the lossless counterpart near the critical frequency in the complex- approach while they are almost unchanged in the complex- approach. incident frequency xed, In Fig. 5.7(a), we calculate the reectance with the GHz, by using Akawara's ATR method for the GHz, (blue), and

f = 4.4118

metamaterial waveguide with

= 0.1

= 0.01

GHz, (red). Fig.

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

111

1 0.9 0.8

(a)
4.41

0.9 0.8

(b)

4.42

0.6

4.4

Re[f]

0.7

0.7 0.6

4.4 1 1.5 2

|R|

0.5 0.4 0.3 0.2 0.1 0 20 30 40 50

|R|

1.5

0.5 0.4 0.3

Re[]

complex approach

0.2 0.1

complex approach

(deg)

60

70

80

0 4.38

4.4

4.42

4.44

f (GHz)

Figure 5.7: (a) Reectance versus the incident angle for the metamaterial waveguide

= 0.1 GHz, (blue solid curve), and = 0.01 GHz, (red dashed curve) with prism = 7, and dgap = 1.8 cm at f = 4.4118 GHz. The inset shows the dispersion curves of the proposed structure near the critical frequency region for the T E2 mode in the complex- approach. (b) Reectance versus frequency with the parameters used in (a) at = 37 . The inset indicates that the dispersion curves are almost insensitive to material loss in the complex- approach.
with 5.7(a) indicates that the value of

has considerable inuence on the location of the

reectance minima. The two reectance minima move toward each other when the value of

decreases.

Finally, they overlap when the loss parameter is too small.

On the other hands, the value of

has almost no inuence on the location of the 5.7(b). However, it

reectance minima in Otto's ATR method as shown in Fig.

aects the magnitude and the width of the reectance dip. This result is acceptable because Otto's ATR method gives the dispersion curves in the complex- approach, and we know that the dispersion curves in this approach is insensitive to the material loss. Hence, Otto's method is better to use when we are interested in reproducing the dispersion curves of a lossless waveguide.

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

112

5.2 5 Re[f] (GHz) 4.8 4.6 4.4 4.2 4 1 4 2 (1/cm) 3 4 2 3 1 = 45 0 (a) Reflectance |R|

1 (b) 4 3 0.5 2

0 1 0 4 4.2 4.4 4.6 Frequency (GHz) 4.8 5

5.2 5 Re[f] (GHz) 4.8 4.6 4.4 4.2 4 1 F

= 26 = 39

1 (c) Reflectance |R| (d)

f = 4.4 B

0.5

B F

2 (1/cm)

0 20

40

60 (deg)

80

Figure 5.8: The dispersion curves of the metamaterial waveguide structure proposed in Fig. 5.2 for

TE

modes (a) in the complex- approach and (c) in the complex-

approach. (b) Reectance versus frequency with the following parameters: prism = 9, dgap = 1 cm, and = 45 . The marked numbers represent the resonance frequencies for

T E0 , ..., T E4

modes. (d) Reectance versus the incident angle at 4.4 GHz for mode, respectively.

T E2

mode. The B and F marks represent the resonance for the backward and forward propagation

T E2

To obtain better insight into the dierence between Otto's ATR method and Arakawa's ATR method, we plot the complete dispersion curve spectrum of the aforementioned lossy metamaterial waveguide with

= 0.1

GHz for TE modes in both

complex- (Fig. 5.8(a)) and complex- (Fig. 5.8(c)) approaches. Although we have not optimized the air gap, a distance between the prism and the waveguide, and the prism index of refraction, there is a good agreement between the calculated curves in Fig. 5.8(a) and the resonance frequencies in Fig. 5.8(b). The marked numbers in Fig. 5.8(b) represent the order of TE modes. The small unmarked dips on the right corner

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

113

of the Fig. 5.8(b) correspond to the higher order TE modes, which are not plotted in Fig. 5.8(a); however, they are dense near the resonance frequency. It is important to highlight that the resonance frequencies obtained by the Otto's ATR method correspond to the dispersion curves of a lossy structure in the complex- approach and not to the dispersion curves of a lossless structure in the complex- approach. Since the dispersion curves of a lossy structure in the complex- approaches are similar to the dispersion curves of a lossless structure in the complex- approach (compare Fig. 5.8(a) with Fig. 2.5(a)), this lead to a misinterpretation of Otto's ATR method in the literature [21, 78, 79]. We note that under no conditions can the complex modes be detected with Otto's ATR method. This is acceptable because the complex modes are not supported in the complex- approach. However, if Otto's ATR method corresponds to the dispersion curves of a lossless waveguide in the complex- approach, the complex modes should be observed. As expected, the complex modes can be detected by using Arakawa's ATR method; however, the index of refraction of the prism must be high to detect them. We have discussed this issue previously in Chapter 4.

5.4 Perturbation theory


The traditional method to derive the dispersion curves of a lossy waveguide from the lossless one, is to use perturbation theory [50, 51]. However, in this section, we show that perturbation theory tends to fail in a lossy metamaterial waveguide near the critical frequency regions in the complex- approach. For the sake of completeness, we derive the generalized perturbation equation of

in the complex- approach.

When we introduce the small amount of loss into a system, the permeability and permittivity will be slightly changed (

+ , + ).

Hence, we can write

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

114

the generalized Hermitian eigenproblem (Eq. (A.13)) in the following form

(A
(0)

(0)

+ A) = (B

(0)

+ B) ,

(5.8)

where

(1)

+ ...

and

= (0) + (1) + ...

are expansions of the new

eigensolutions in powers of the perturbation we get

. If we keep terms up to the rst order,

(0) (0) (1) (0) (0) (0) (0) (1) A + A = (1) B + (0) B + (0) B .

(5.9)

By using the Hermitian property of the operators and taking the inner product on both side of Eq. (5.9), the generalized form of the perturbation theory is

(0)

(1)

(0)

(0) (0)

(0) (0) (0) B

(0)

(0)

(5.10)

For a small change of

and , z 0

we get

B = 0

and

A =

z +

(
t ) c

0
c t

z +

t )

(5.11)

Inserting Eq. (5.11) and Eq. (A.14) into Eq. (5.10), the perturbation equation for the TE modes is

(1)

= 4c

2 2 2 ([Ey (x) + Hy (x) + + Hz (x)]dx + 1 E 2 (x)dx (x) y

(5.12)

In Fig. 5.9, the blue curves show the real and the imaginary components of the

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

115

4.4 4.3 f (GHz) 4.2 4.1 4

(a)
f (GHz)

4.4

(b)
4.3 4.2 4.1

3 4 Re[]

0.1

0.2 0.3 Im[]

0.4

Figure 5.9: The blue and red curves represent the exact and perturbation solutions of the waveguide proposed in Fig. 5.3 for (a) the real part and (b) the imaginary part of the

T E2

mode in the complex- space.

dispersion curves of the aforementioned lossy metamaterial waveguide for the

T E2

mode in the complex- space, and the dashed red curves show the dispersion curves of that structure calculated by perturbation theory (Eq. (5.12)). Fig. 5.9 indicates that the dispersion curves calculated by perturbation theory agree with the exact dispersion curves solutions except near the critical frequency region. Moreover, the complex modes cannot be derived by perturbation theory because the denominator of Eq. (5.12) is always zero for the complex modes. From the standard perturbation theory [51, 58, 95], it can be shown that the perturbation equation in the complex- space is

(1)

(0) = 2

2 2 ([Ey (x) + Hy (x) + + 2 Ey (x)dx

H 2 (x)]dx + z

(5.13)

Figure 5.9: The blue and red curves represent the exact and perturbation sol of the waveguide proposed in Fig. 5.3 for the

T E2

mode in the complex- spac

CHAPTER 5.

DISPERSION CURVES OF MWS WITH COMPLEX-

116

However, Eq.

(5.13) is only valid for dispersionless structures.

Finding an ap-

propriate perturbation equation for a dispersive waveguide structure is an interesting problem for future work.

5.5 Concluding remarks


In this chapter, we have shown that one can study the modes of a lossy metamaterial waveguide structure using two dierent approaches: a complex propagation constant approach and a complex frequency approach. The results of these two approach

in general are not the same because these approaches fundamentally solve dierent problems. We have also shown that the dispersion curves of MWS in a complex

propagation constant approach and a complex frequency approach can be reproduced by using the ATR method but with dierent experimental conditions.

Chapter 6

Conclusions and suggestions for future work

In this thesis, we have investigated the optical properties of metamaterial waveguide structures. We have shown that the negative-index of refraction of a metamaterial leads to unusual optical properties, which cannot be seen in a conventional waveguide structure. We have calculated the dispersion curves of the metamaterial waveguide structure and demonstrated a rich variety of modes, such as surface polariton modes, complex modes, backward propagating modes, and even stopped-light modes that can be supported by metamaterial waveguide systems. Most of the unique properties of the metamaterial waveguide come from the fact that the energy ux in the core ows in the opposite direction to that in the cladding. Subsequently, we have examined these unusual electromagnetic properties by considering the role of dispersion and loss, which are inherent features of metamaterials. Our results demonstrated

that the intrinsic loss signicantly modies the dispersion curves and consequently the properties of the metamaterial waveguide structures. 117 Moreover we have found

CHAPTER 6.

CONCLUSIONS AND SUGGESTIONS FOR FUTURE WORK 118

several novel properties in the presence of loss, such as a splitting of the complex modes, bending back of the modes near the resonance frequency, and disappearance of the stopped-light modes. One of our more useful and important result is that we have explicitly shown that light cannot be slowed down in a realistic metamaterial waveguide structure. In summary, the dispersion curves of a lossless metamaterial

waveguide structure near the slow-light-mode domains are strongly modied when arbitrarily small amounts of loss are introduced into the system. We also studied the ATR method to excite the modes of the metamaterial waveguide structure. Although the ATR method is a useful tool to excite the normal modes in a waveguide with a low loss, this technique requires satisfying some non-trivial conditions in order to be applicable for complex modes or the normal modes in a lossy metamaterial waveguide. We also have shown that the excitation of the modes leads to anomalous lateral beam phase shifts. In summary, the ATR method is only a

useful tool to obtain the dispersion curves of a lossy metamaterial waveguide in the complex- approach and those of a metamaterial waveguide with very small loss in the complex- approach. Finally, we have shown that one can study the modes of a lossy waveguide using two approaches: a complex frequency approach and a complex propagation constant approach. The results of these two approaches are not the same; thus, one needs to obtain the dispersion curves of a waveguide system in both these approaches to fully study the optical properties of the system. Future work on metamaterial waveguide structures could include examining the eects of anisotropy and gain and further investigation of the complex- modes.

Bibliography

[1]

J. B. Pendry, Negative refraction makes a perfect lens, Phys. Rev. Lett. 85, 3966 (2000).

[2]

A. Alu, N. Engheta, Achieving transparency with plasmonic and metamaterial coatings, Phys. Rev. E 72, 016623 (2005).

[3]

U. Leonhardt, Optical conformal mapping, Science 312, 1777 (2006).

[4]

J. B. Pendry, D. Schurig, D. R. Smith, Controlling electromagnetic elds, Science


312, 1780 (2006).

[5]

A. Alu, N. Engheta, Guided modes in a waveguide lled with a pair of singlenagative (SNG), double-negative (DNG), and/or double-positive (DPS) layers,
IEEE Trans. Microwave theory Technol. 52, 199 (2004).

[6]

M. C. K. Wiltshire, J. B. Pendry, I. R. Young, D. J. Larkman, D. J. Gilderdale, J. V. Hajnal, Microstructured magnetic materials for RF ux Guides in magnetic resonance imaging, Science 291, 849, (2001).

[7]

U. Leonhardt, Invisibility cup, Nature Photonics 1, 207-208 (2007).

119

BIBLIOGRAPHY

120

[8]

J. C. Bose, On the rotation of plane of polarization of electric waves by a twisted structure, Proc. R. Soc. 63, 146 (1898).

[9]

W. E. Kock, Mettalic delay lenses, Bell Syst. Tech. J. 27, 58, (1948).

[10] D. R .Smith, W. J. Padilla, D. C. Vier, S. C. Nemat-Nasser, S. Schultz, Composite medium with simultaneously negative permeability and permittivity, Phys.
Rev. Lett. 84, 4184 (2000).

[11] J. B. Pendry, A. J. Holden, D. J. Robbins, W. J. Stewart, Magnetism from conductors and enhanced nonlinear phenomena, IEEE Trans. Microw. Theory
Tech. 47, No.11, 2075 (1999).

[12] V. G. Veselago, The electrodynamics of substances with simultaneosly negative values of

and

Sov. Phys. Usp. 47, 509 (1968).

[13] A. Schuster, An introduction to the theory of optics (Arnold, London, 1904).

[14] D. V. Sivukhin, the energy of electromagnetic waves in dispersive media, Opt.


Spectrosk 3, 308 (1957).

[15] R. A. Shelby, D. R. Smith, S. Schultz, Experimental verication of a negative index of refrection, Science 292, 77 (2001).

[16] G. Dolling, C. Enkrich, M. Wegener, C. M. Soukoulis, and S. Linden, Negativeindex material at 780 nm wavelength, Opt. Lett. 32, 53 (2007).

[17] V. M. Shalaev, W. Cai, U. K. Chettiar, H. Yuan, A. K. Sarychev, V. P. Drachev, and A. V. Kildishev, Negative index of refraction in optical metamaterials, Opt.
Lett. 30, 3356 (2005).

BIBLIOGRAPHY

121

[18] S. Zhang, W. Fan, N. C. Panoiu, K. J. Malloy, R. M. Osgood, Experimental Demonstration of Near-Infrared Negative-Index Metamaterials, Phys. Rev. Lett.
95, 137404 (2005).

[19] V. M. Shalaev, Optical negative-index metamaterials, Nature Photon. 1, 4148 (2007).

[20] C. M. Soukoulis, S. Linden, M. Wegener, Negative refractive index at optical wavelengths, Science 315, 47 (2007).

[21] R. Ruppin, Surface polaritons of a left-handed medium, Phys. Lett. A (2000).

277,

61

[22] J. Gollub,D. Smith,D. Vier,T. Perram,J. J. Mock, Experimental characterization of magnetic surface plasmons on metamaterials with negative permeability, Phys.
Lett. B 71,195402 (2005).

[23] I. Wu,T. M. Grzegorczyk, Y. Zhang, J. A. Kong, Guided modes with imaginary transverse wave number in a slab waveguide with negative permittivity and permeability, J. Appl. Phys. 93,9386 (2003).

[24] I. V. Shadrivov, A. A. Sukhorukov, Y. S. Kivshar, Guided modes in negativereftactive-index waveguides, Phys. Rev. E 67,057602 (2003).

[25] L. Zhou, C. Chan, Vortex-like surface wave and its role on the transient phenomena of metamaterial focusing, Appl. Phys. Lett. 86,101104(2005).

[26] P. Baccarelli, P. Burghignoli, F. Frezza, A. Galli, G. Lovat, S. Paulotto, Fundamental modal properties of surface waves on metamaterial grounded slab, IEEE
Trans. Microwave Theory Tech. 53, 1431-1442 (2005).

BIBLIOGRAPHY

122

[27] D. R. Smith, N. Kroll, Negative refractive index in left-handed materials. Phys.


Rev. Lett. 85, 2933 (2000).

[28] I. V. Shadrivov, A. A. Sukhorukov, Y. S. Kivshar, Complete Band Gaps in OneDimensional Left-Handed Periodic Structures, Phys. Rev. Lett. 95,193903(2005).

[29] J. Li, L. Zhou, C. T. Chan, and P. Sheng, Photonic Band Gap from a Stack of Positive and Negative Index Materials,Phys. Rev. Lett. 90, 083901 (2003).

[30] H. Jiang, H. Chen, H. Li, Y. Zhang, S. Zhu , Omnidirectional gap and defect mode of one-dimensional photonic crystals containing negative-index materials,
Appl. Phys. Lett. 83, 5386 (2003).

[31] J. P. Xu, Y. P. Yang, H. Chen, S. Y. Zhu, Spontaneous decay process of a two-level atom embedded in a one-dimensional structure containing left-handed material, Phys. Rev. A 76, 063813 (2007).

[32] J. Kastel, M. Fleischhauer, Suppression of spontaneous emission and superradiance over macroscopic distances in media with negative refraction, Phys. Rev. A
71, 011804 (2005).

[33] Q. Cheng, T. J. Cui, High-power generation and transmission through a lefthanded material, Phys. Rev. B 72, 113112 (2005).

[34] Y. H. Yao, T. J. Ciu, Q. Cheng, R. Liu, D. Huang, D. R. Smith, Realization of a super waveguide for high-power-density generation and transmission using right and left-handed transmission-light circuits, Phys. Rev. E 76, 036602 (2007).

[35] K. L. Tsakmakidis, A. D. Boardman, O. Hess, Trapped rainbow storage of light in metamaterials, Nature 450, 397 (2007).

BIBLIOGRAPHY

123

[36] P. R. Berman, Goos-hanchen shift in negative index media,Phys. Rev. E 66, 067603 (2002).

[37] I. V. Shadrivov, A. A. Zharov, Y. S. Kivshar, Giant Goos-Hanchen eect at the reection from Left-handed metamaterials, Appl. Phys. Lett. 83, 2713 (2003).

[38] J. He, S. He, Slow propagation of electromagnetic waves in a dielectric slab waveguide with a left-handed material substrate, IEEE microwave and wireless
components letters, Vol. 16, No. 2, (2005).

[39] Q. Gan, Z. Fu, Y. J. Ding, F. J. Bartoli, Ultrawide-bandwidth slow-light system based on THz plasmonic graded metallic grating structures, Phys. Rev. Lett.
100, 256803 (2008).

[40] L. V. Alekseyev, E. Narimanov , Slow light and 3D imaging with non-magnetic negative index systems, Optics Express, Vol. 14, Issue 23, pp. 11184-11193 (2006).

[41] Y. J. Huang, W. T. Lu, S. Sridher, Nanowire waveguide made from extremely anisotropic metamaterials, Phys. Rev. A 77, 063836 (2008).

[42] T. Jiang, Y. J. Feng, Slow and frozen waves in a planar air waveguide with anisotropic metamaterial cladding, CMMT 450775 (2008).

[43] J. He, Y. Jin, Z. Hong, S. He , Slow light in a dielectric waveguide with negativerefractive-index photonic crystal cladding, Optics Express, Vol. 16, Issue 15, pp. 11077-11082 (2008).

[44] L. Zhao, S. F. Yelin, 'Trapped rainbow' in graphene, arxiv:0804.2225 (2008).

BIBLIOGRAPHY

124

[45] J. B. Pendry and D. R. Smith, Reversing Light With Negative Refraction, Phys.
Today 57(6), 37 (2004).

[46] J. Valentine, S. Zhang, T. Zentgraf, E. Ulin-Avila, D. A. Genov, G. Bartal and X. Zhang, Three-dimensional optical metamaterial with a negative refractive index,
Nature 455, 376 (2008).

[47] A. Cho, Bizarre 'Metamaterials' for Visible Light in Sight?, Science 321, 900 (2008).

[48] M. I. Stockman, Criterion for negative refraction with low losses from a fundamental principle of causality, Phys.Rev.Lett. 98, 177404 (2007).

[49] A. Reza, M. M. Dignam, S. Hughes, Can light be stopped in realistic metamaterials?, Nature 455, E-10 (2008).

[50] A. W. Synder and J. D. Love,"Optical Waveguide Theory "(Chapman&Hall, London,1983).

[51] C. Vassallo, "Optical Waveguide Concepts" (Elsevier,Amesterdam,1991).

[52] R. E. Collin, "Field Theory Of Guide Waves" (IEEE Press,Piscataway,1991).

[53] A. Peacock, N. Broderick, Guided modes in channel waveguides with a negative index of refraction, Optics Express, Vol. 11, No. 20, 2502 (2003).

[54] K. L. Tsakmakidis, A. Klaedtke, D. Aryal, O. Hess, Single-mode operation in the slow-light regime using oscillatory waves in generalized left-handed hetrostructures, J. Appl. Phys. 89, 201103 (2006).

BIBLIOGRAPHY

125

[55] W. Shu, J. M. Song, Complete mode spectrum of a grounded dielectric slab with metamaterials, PIER 65, 103-123 (2006).

[56] Q. Sui, F. Li, Complex guided wave soloutions of grounded dielectric slab made of metamaterials, PIER 51, 187-195 (2005).

[57] R. Chern, C. C. Chang, Surface and bulk modes for periodic structures of negative index materials, Phys. Rev. B 74, 155101(2006).

[58] M. Skorobogatiy, M. Ibanescu, S. G. Johnson, O. Weisberg, T. D. Engeness, M. Soljacic, S. A. Jacobs, Y. Fink, Analysis of general geometric scaling fundamental connection between

perturbations in a transmitting waveguides:

polarization-mode dispersion and group-velocity dispersion, J. Opt. Soc. Am. B


19 2867 (2002).

[59] S. G. Johnson, M. Ibanescu, M. A. Skrobogatiy, O. Weisberg,T. D. Engeness, M. Soljacic, S. A. Jacobs, J. D. Joannopoulos, Y. Fink, Low-loss asymptotically single-mode propagation in large-core Omniguide brs, Opt. Express 9, 748 (2001).

[60] K. Sakoda, Optical properties of photonic crystals (Springer 2005).

[61] J.l u, B. Wu, J. A. Kong, Guided modes with a linearly varying transverse eld inside a left-handed dielectric slab, J. Electro. Waves and Appl. 20, 689-697 (2006).

[62] P. Markos, C. M. Soukoulis, Transmission studies of left-handed materials, Phys.


Rev. B 65, 033401 (2001).

BIBLIOGRAPHY

126

[63] K. Halteman, S. Feng, P. L. Overfelt, Guided modes of elliptical metamaterials waveguides, Phys. Rev. A 76, 013834 (2007).

[64] A. Govyadinov, V. Podolskiy, Gain-assisted slow to superluminal group velocity management in nano-waveguides, Phys. Rev. Lett. 97, 223902 (2006).

[65] I. V. Shadrivov, A. A. Sukhorukov, Y. S. Kivshar, Nonlinear surface waves in left-handed materials, Phys. Rev. E 69, 016617 (2004).

[66] G. V. Eleftheriades, K. G. Balmain, Negative-refraction metamaterials, (Wiley, Hoboken, 2005).

[67] P. W. Milonni, Fast light, slow light and left-handed light, (IOP publishing, London, 2005).

[68] T. J. Cui, J. A. Kong, Time-domain electromagnetic energy in a frequencydispersive left-handed medium, Phys. Rev. B 70, 205106 (2004).

[69] A. D. Boardman, K. Marinov, Electromagnetic energy in a dispersive material,


Phys. Rev. B 73, 165110 (2006).

[70] R. Ruppin, Electromagnetic energy density in a dispersive and absorptive material, Phys. Lett. A 299, 309 (2002).

[71] J. W. Dong, H. Z. Wang, Slow electromagnetic propagation with low group velocity dispersion in an all-metamaterial-based waveguide, Appl. Phys. Lett.
91, 111909 (2007).

[72] J. D. Jackson, Classical electrodynamics, (Wiley, New york, 1998).

BIBLIOGRAPHY

127

[73] R. W. Ziolkowski, Superluminal transmission of information through an electromagnetic metamaterial, Phys. Rev. E 63, 046604, (2001).

[74] S. Hughes, L. Ramunno,J. Young, J. E. Sipe, Extrinsic optical scattering loss in photonic crystal waveguides: role of fabrication disorder and photon group velocity, Phys. Rev. Lett. 94, 033903 (2005).

[75] A. Borde, L. H. Ford, T. A. Roman, Constraints on spatial distributions of negative energy, Phys. Rev. D 65, 084002 (2002).

[76] P. Yao, Z. Liang, X. Jiang, Limitation of the electromagnetic cloak with dispersive material, Appl. Phys. Lett. 92, 031111 (2008).

[77] R. Ruppin, Surface polaritons of a left-handed material slab, J. Phys. Condens.Matter 13,1811 (2001).

[78] K. Park,B. J. Lee,C. Fu,Z. M. Zhang, Study of the surface and bulk polaritons with a negative index metamaterials, J. Opt. Soc. Am. B /Vol. 22,No. 5 (2005).

[79] G. S. Kovener,R. W. Alexander,R. J. Bell, Surface electromagnetic waves with damping. I. Isotropic media, Phys. Rev. B 14, 1458-1464 (1976).

[80] A. Otto, Excitation of nonradiative surface plasma waves in silver by the method of frustrated total relection, Z. Phys. 216,398-410 (1968).

[81] H. Kitajima, K. Hieda,Y. Suematsu, Optimum condition in the attenuated total reection technique, Appl. Opt. 20, 1005-1009 (1981).

BIBLIOGRAPHY

128

[82] M. O. Petersen,B. S. Zhu,E. Dalsgaard,Extreme attenuation of total internal reection used for determination of optical properties of metals, J. Opt. Soc.
Am. A / Vol. 4,No. 9 (1987).

[83] H. Kitajima, K. Fujita, H. Cizmic, Zero reection from a dielectric lm on metal substrate at oblique angles of incidence, Appl. Opt. 23, 1937 (1984).

[84] C. J. Fu,Z. M. Zhang,D. B. Tanner, Energy transmission by photon tunneling in multilayer structures, J. Heat. Transfer 1046 / Vol.127 (2005).

[85] L. Gao, C. J. Tang, Near-eld imaging by amulti-layer structure consisting of alternate right-handed and leeft-handed materials, Phys. Lett. A 322, 390 (2004).

[86] J. Gerardin, A. Lakhtakia, Negative index of refraction and distributed bragg reections, Mic. Opt. Technol. Lett. 34, 409, (2002).

[87] L. G. Wang, S. Y. Zhu, Large positive and negative Goos-Hanchen shifts from a weakly absorbing Left-handed slab, J. Appl. Phys. 98, 043522 (2005).

[88] X. Chen, C. F. Li, Lateral shift of the transmitted light beam through a lefthanded slab, Phys. Rev. E 69, 066617 (2004).

[89] K. C. Huang, E. Lidorikis, X. Jiang, J. D. Joannopoulos, K. A. Nelson, Nature of lossy Bloch states in polaritonic photonic crystals, Phys. Rev. B 69, 195111 (2004).

[90] N. F. Declercq, J. Degrieck, Diraction of complex harmonic plane waves and the stimulation of transient leaky Rayleigh waves, J. Appl. Phys. 98, 113521 (2005).

BIBLIOGRAPHY

129

[91] A. Bernard, M. Deschamps, Comparison between the dispersion curves calculated in complex frequency and the minima of the reection coecients for an embedded layer, J. Acoust. Soc. Am. 107, 793 (2000).

[92] P. A. Sturrock, Kinematics of growing waves, Phys. Rev. 112, 1488 (1958).

[93] E. T. Arakawa, M. W. Williams, R. N. Hamm, R. H. Ritchie, Phys. Rev. Lett.


31, 1127 (1973).

[94] L. C. Andreani, M. Agio, Intrinsic diraction losses in photonic crystal waveguides with line defects, Appl. Phys. Lett. 82, 2011 (2003).

[95] C. Kottke, A. Farjadpour, S. G. Johnson, Perturbation theory for anisotropic dielectric interfaces, and application to subpixel smoothing of discretized numerical methods, Phys. Rev. E 77, 036611 (2008).

[96] V. M. Shalaev, Metamaterials:

A new paradigm of physics and engineering,

http://www.nanohub.org/courses/metamaterials (2008).

[97] S. A. Ramakrishna, Physics of negative refractive index materials, Rep. Prog.


Phys. 68, P. 476 (2005).

[98] R. W. Ziolkowski, A. D. Kipple, Causality and double-negative metamaterials,


Phys. Rev. E 68, 026615 (2003).

Appendix A

Orthogonality relations

In this appendix, we derive the orthogonality and normalization relations for a lossless metamaterial waveguide structure by using generalized Hermitian Hamiltonian formulation [58, 95]. The time-harmonic electromagnetic wave equations are

1 (r) 1 (r)
where

1 (r) 1 (r)

H(r) = E(r) =

2 H(r) c2 2 E(r), c2
We assume they are real.

(A.1)

and

are permittivity and permeability.

The

resultant eigensystem for Eq. (A.1) can be written as

LH H(r) LE E(r)
where

1 (r) 1 (r)

1 (r) 1 (r)

H(r) = E(r) =

2 H(r) c2 2 E(r), c2

(A.2)

LH

and

LE

are non-Hermitian operators [60]; hence, their eigenfunctions are

130

APPENDIX A.

ORTHOGONALITY RELATIONS

131

not orthogonal. We introduce the following set of eigenfunctions and Hermitian operators which satisfy Eq. (A.1) [60]

QH (r) QE (r) HH QH (r) 1 HE QE (r) QH (r)


and

(r)H(r) (r)E(r)

(A.3)

(r)

1 (r) 1 (r)

1 QH (r)
(r)

(A.4)

1 (r)

1 QE (r) (r)

QE (r)

can be shown to be Hermitian [60]. The new set of eigenfunctions

are orthogonal:

QE, (r), QE, (r) =


One should notice that the case that negative

dr |(r)| E (r) E (r) = ,

(A.5)

HH

and

HE

operators are only Hermitian in the spatial

and

are positive; however, in a metamaterial waveguide with both

and

, the above operators are not Hermitian and Eq. (A.5) is no longer

valid. Therefore, we need to derive generalized Hermitian Hamiltonian formulation to nd the eigenmodes of the waveguide [58, 59]. We start with Maxwell's equations for the time-harmonic electromagnetic elds

E = 0, H = 0,

E = i H, c H = i E. c

(A.6)

APPENDIX A.

ORTHOGONALITY RELATIONS

132

By introducing transverse and longitudinal components of the elds, Maxwell's equations take the form

t t

E t = E z , z

tE z tH z

E t z H t z

+ z( + z(

t t

E t ) = i ( H t + H z ), z c H t ) = i ( E t + E z ). z c
(A.7)

H t = H z , z

After eliminating the longitudinal components of the elds and moving all of the z derivatives to one side in Eq. (A.7), we arrive at

z 0

1 z

(
t ) c

0
c t

1 z

t )

Et = Ht

0 E t z i z . z 0 Ht
Next, we dene operators notation,

(A.8)

and

B,

and the electromagnetic eld pattern in Dirac

| ,as

z 0

1 z

(
t ) c

0
c

t z

1 z

t )

(A.9)

B 0

z , z 0

(A.10)

APPENDIX A.

ORTHOGONALITY RELATIONS

133

(A.11)

Et | . Ht
The following eigenproblem can be obtained

A | = i B | . z

(A.12)

We assume the waveguide is uniform along z; therefore, the electromagnetic elds has the form

eiz

where the wavenumber,

is the propagation constant of the modes.

Finally, by inserting

exp(iz) |

into Eq. (A.12), we have

A = B A B

(A.13)

In this form,

and

are Hermitian operators if both

and

have purely real

values. They satisfy the properties of the eigenproblems including the eigenvalues , , are real and satisfy the orthogonality relationship

= ,

[58].

Now, we can obtain the orthogonality and normalization relationship for the TE polarization elds as follows

APPENDIX A.

ORTHOGONALITY RELATIONS

134

dx

Et Ht

0 z

z 0

Et Ht

dx (E (x) H (x) + E (x) H (x)) z


(A.14)

+ 1 Re[ 2

z E (x) H (x). dx]


1 E (x)Ey (x) y

+ = 1 Re[ 2

(x)dx]

= P , ,
where

is the power ux. Eq. (A.14) shows that every mode plays its own part in

the power ux independent of the other modes. It is clear that in the lossy waveguide structure, since the energy is absorbed in the media, the power carried in each mode is not independent from the powers in the other modes anymore. We can also derive the generalized Hermitian Hamiltonian eigenproblem in terms of the frequency,

instead of wave number,

which leads to dierent orthogonality Eq. (A.7) can be re-written

relations. By introducing as

D t = E t ,

and

B t = H t ,the

APPENDIX A.

ORTHOGONALITY RELATIONS

135

Light line
5

Light line

Frequency (GHz)

Frequency (GHz)

2
TE2

1 0

TE2

TE1 TE0

Dielectric light line

TE1 TE0

Dielectric light line

0 0

0.5

1.5

(1/cm)

2.5

3.5

0 0

0.5

1.5

(1/cm)

2.5

3.5

Fig. A.1 The orthogonality relations (a) in terms of frequency and (b) in terms of wave number. with The blue curves are the dispersion curves of a dielectric waveguide

=4

and

=4

for the three lower order TE modes.

D t = Dz , z B t = B z , z

1 t Dz

1 D t z

+ z(

1 D t ) = i ( B t + B z ), z c

1 t Bz

1 B t z

+ z(

1 B t ) = i ( D t + D z ). z c

(A.15)

After some algebraic manipulations and assuming

| (x, z, t) =

D t (x) B t (x)

exp(iz it),

(A.16)

1 t

0
1 1 t t

Dt Bt

0 z z 0

Dt Bt

(A.17)

In the Eq. (A.17), the operators are Hermitian and thus have real eigenvalues,

, and

APPENDIX A.

ORTHOGONALITY RELATIONS

136

result in the following orthogonality relationship

| B | = W , + + z = 1 Re[ D (x) B (x). dx] = 1 (x)Ey (x)Ey (x)dx, 2 2


where

(A.18)

is the energy density. One should notice that the above equation only works

for non-dispersive media. Next, we normalize the amplitude of the electric eld,

E0 ,

for an asymmetric

metamaterial waveguide structure. By using the orthogonality relationship in terms of wave number, Eq. (A.14), we have

+ 1 1 Re[ Ey (x)Ey (x)dx] = , , 2 (x)


where assume arrive at

(A.19)

P = 1.

Following a substitution of Eq.

(4.12) into Eq.

(A.19), we

APPENDIX A.

ORTHOGONALITY RELATIONS

137

1 2

1 E (x)Ey (x)dx (x) y

=1 2d
0

2 cE0 2

dx dx

1 3

sin2 () exp[2k3 x]

dx

1 2

sin2 ( + k2 x)
(A.20)

2d

1 1

sin2 ( 2k2 d) exp[2k1 (x 2d)] = 1 22


2 1 k1
2 2 k1 +k2 2 2 2 k1 +2 k2 2 1

asy E0 =

2 c d +

2 3 k3

2 2 k3 +k2 2 2 2 k3 +2 k2 2 3

Following the same derivation as in Eq. (A.20), we can also normalize

E0

in terms of

frequency. By substituting Eq. (2.12) into Eq. (A.18), the normalized electric eld has the form

asy E0 =

2 d2 +
1 k1
2 2 2 2 k1 + 1 1 k2 2 2 2 k1 +2 k2 2 1

3 k3

2 2 2 2 k3 + 3 3 k2 2 2 2 k3 +2 k2 2 3

(A.21)

For a symmetric waveguide structure, Eqs. (A.20,A.21) respectively reduce to

sym E0 =

2 c d +

2
2 1 k1
2 2 k1 +k2 2 2 2 k1 +2 k2 2 1

,
(A.22)

sym E0 =

1 d2 +
1 k1
2 2 2 2 k1 + 1 1 k2 2 2 2 k1 +2 k2 2 1

Finally, the results for the transverse magnetic modes, TM, can be found from duality

APPENDIX A.

ORTHOGONALITY RELATIONS

138

by changing

and

E H .

Anda mungkin juga menyukai