Laszlo Solymar, Ekaterina Shamonina - Waves in Metamaterials-OUP (2009) PDF
Laszlo Solymar, Ekaterina Shamonina - Waves in Metamaterials-OUP (2009) PDF
Laszlo Solymar, Ekaterina Shamonina - Waves in Metamaterials-OUP (2009) PDF
L. Solymar
Imperial College, London
E. Shamonina
University of Erlangen-Nürnberg
1
3
Great Clarendon Street, Oxford OX2 6DP
Oxford University Press is a department of the University of Oxford.
It furthers the University’s objective of excellence in research, scholarship,
and education by publishing worldwide in
Oxford New York
Auckland Cape Town Dar es Salaam Hong Kong Karachi
Kuala Lumpur Madrid Melbourne Mexico City Nairobi
New Delhi Shanghai Taipei Toronto
With offices in
Argentina Austria Brazil Chile Czech Republic France Greece
Guatemala Hungary Italy Japan Poland Portugal Singapore
South Korea Switzerland Thailand Turkey Ukraine Vietnam
Oxford is a registered trade mark of Oxford University Press
in the UK and in certain other countries
Published in the United States
by Oxford University Press Inc., New York
c L. Solymar and E. Shamonina 2009
The moral rights of the author have been asserted
Database right Oxford University Press (maker)
First Published 2009
All rights reserved. No part of this publication may be reproduced,
stored in a retrieval system, or transmitted, in any form or by any means,
without the prior permission in writing of Oxford University Press,
or as expressly permitted by law, or under terms agreed with the appropriate
reprographics rights organization. Enquiries concerning reproduction
outside the scope of the above should be sent to the Rights Department,
Oxford University Press, at the address above
You must not circulate this book in any other binding or cover
and you must impose the same condition on any acquirer
British Library Cataloguing in Publication Data
Data available
Library of Congress Cataloging in Publication Data
Data available
Printed in Great Britain
on acid-free paper by
CPI Antony Rowe, Chippenham, Wiltshire
ISBN: 978–0–19–921533–1
1 3 5 7 9 10 8 6 4 2
To Oscar and Sasha
This page intentionally left blank
Preface
3 Plasmon–polaritons 75
3.1 Introduction 75
3.2 Bulk polaritons. The Drude model 77
3.3 Surface plasmon–polaritons. Semi-infinite case,
TM polarization 79
3.3.1 Dispersion. Surface plasmon wavelength 79
3.3.2 Effect of losses. Propagation length 83
3.3.3 Penetration depth 85
3.3.4 Field distributions in the lossless case 88
3.3.5 Poynting vector: lossless and lossy 90
3.4 Surface plasmon–polaritons on a slab: TM polarization 92
3.4.1 The dispersion equation 92
3.4.2 Field distributions 97
3.4.3 Asymmetric structures 100
3.5 Metal–dielectric–metal and periodic structures 102
3.6 One-dimensional confinement: shells and stripes 103
3.7 SPP for arbitrary ε and µ 106
3.7.1 SPP dispersion equation for a single interface 106
3.7.2 Domains of existence of SPPs
for a single interface 108
3.7.3 SPP at a single interface to a metamaterial:
various scenarios 110
3.7.4 SPP modes for a slab of a metamaterial 116
Contents xiii
A Acronyms 325
References 349
Index 381
Basic concepts and basic
equations
1
1.1 Introduction 1.1 Introduction 1
1.2 Newton’s equation and
We shall make an attempt in this chapter to introduce most of the basic electrical conductivity 2
concepts needed later in the book and in a form upon which we can 1.3 Maxwell’s equations,
build in later chapters. Some will be very basic indeed, like fields and fields and potentials 3
potentials, some others will be closer to practical cases, e.g. the inci- 1.4 The wave equation and
dence of a plane wave upon a dielectric slab. Some will represent useful boundary conditions 4
artifices like electric and magnetic dipoles, leading from there to the 1.5 Hollow metal waveguides 6
introduction of the concept of polarizability: how large are the dipole 1.6 Refraction at a boundary:
moments excited by various fields. Boundary conditions and boundary Snell’s law and the Ewald
circle construction 8
refraction problems must of course be mentioned because they will ap-
1.7 Fermat’s principle 11
pear repeatedly in the analysis of metamaterial properties. The Ewald
1.8 The optical path and lens
circle construction will then come in as a useful technique. Dispersion design 12
is again one of the basic concepts. In any wave phenomenon one must 1.9 The effective ε in the
know the relationship between frequency and wave number.1 Talking presence of a current 14
about dispersion necessitates the introduction of forward and backward 1.10 Surface waves 14
waves. The latter were hardly mentioned in the past, their properties 1.11 Plane wave incident
rarely emphasized in studies of electromagnetic waves, but they have upon a slab 16
lately risen to fame due to their role in metamaterials. Another de- 1.12 Dipoles 17
parture from usual introductions is the prominence of circuit theory, a 1.13 Poynting vector 18
subject often neglected in physics syllabuses. Very often we shall have 1.14 Radiation resistance 18
to look at phenomena both from the point of view of fields and cir- 1.15 Permittivity and
cuits. Sometimes the two explanations reinforce each other, sometimes permeability tensors 19
we shall resort to only one of them because the other one may be too 1.16 Polarizability 20
unwieldy. Acquaintance with Fourier analysis is of course a basic re- 1.17 Working with tensors 21
quirement whether one is concerned with the temporal or the spatial 1.18 Dispersion: forward
regime. In more general terms it means working in reciprocal space but and backward waves 21
we shall do that only sparingly, mostly in the one-dimensional context. 1.19 Mutual impedance and
The subject of metamaterials can be highly mathematical, particu- mutual inductance 23
larly when radiation effects are included. Our aim, as mentioned in 1.20 Kinetic inductance 24
the Preface, is to offer a treatment that can be happily studied by a 1.21 Four-poles: impedance
and chain matrices 25
final-year undergraduate so we shall try to keep it simple. Some mathe-
matics is of course unavoidable. There are Maxwell’s equations to start 1.22 Transmission line
equations 26
with, which involve a fair amount of vector analysis. So we shall present
1.23 Waves on four-poles 28
1.24 Scattering coefficients 30
1 Engineers prefer to call it propagation coefficient, a usage we shall sometimes prac-
1.25 Fourier transform and
tise when it sounds more appropriate in the context.
the transfer function 31
2 Basic concepts and basic equations
the equations but they will not be called upon very often. The wave
equation and some simple theory of differential equations is of course
necessary. Some tensors will also be mentioned, but only at an elemen-
tary level: no fancy co-ordinate transformations, only a description in
terms of matrices. There is obviously not enough space to derive all the
equations needed. The question is which equations to derive, or at least
give hints about the derivation, and which ones to postulate. We hope
the compromises chosen will be acceptable.
The aim of this chapter is to serve partly as an introduction to con-
cepts and equations, and partly as a reference library to which the reader
can turn when the need arises.
J = σE , (1.4)
where
σ0 N e2 τ
σ= , σ0 = , (1.5)
1 + j ωτ m
and σ0 is the electrical conductivity. Note that we have abandoned here
the time-honoured notation for conductivity that is simply σ instead of
σ0 . The reason is that we wish to reserve σ, a complex quantity, for
giving the relationship between the current density and the electric field
that will be often needed in the form of eqn (1.5).
1.3 Maxwell’s equations, fields and potentials 3
∇×H = J + j ωD , (1.6)
∇×E = −j ωB , (1.7)
∇·D = ρ, ∇· B = 0, (1.8)
D = εE , B = µH . (1.9)
The electric and magnetic fields may also be expressed in the form
∂A
E = −∇ϕ − and H = ∇ × A. (1.10)
∂t
In eqns (1.6)–(1.9) the quantities printed with bold letters are vectors, H
is the magnetic field, B is the magnetic flux density, D is the electric flux
density, also called the dielectric displacement, J is the current density, ρ
is the charge density, A is the vector potential, ϕ is the scalar potential,
µ and ε are the material constants, permeability and permittivity. ∇ is
a differential operator defined as
∂ ∂ ∂
∇= ix + iy + iz , (1.11)
∂x ∂y ∂z
where ix , iy and iz are unit vectors in the x, y and z directions.
Two further relationships often referred to in the book, are Faraday’s
and Ampere’s laws, the former relating the temporal derivative of Φ, the
magnetic flux, to V , the induced voltage, as
V = −j ωΦ , (1.12)
and the latter relating the line integral of the magnetic field to I, the
enclosed current:
Z
Hds = I , (1.13)
ε = ε0 εr and µ = µ0 µr , (1.15)
4 Basic concepts and basic equations
where εr and µr are the relative permittivity (also called dielectric con-
stant or dielectric function) and the relative permeability. In the early
times when Gaussian units were in fashion ε0 and µ0 were taken as
unity. With the emergence of SI (System International) units they have
the numerical values of
P = Np and M = Nm , (1.17)
where p and m are the electric and magnetic dipole moments defined as
D = ε0 E + P (1.19)
and
B = µ0 H + M . (1.20)
The usual problem, both for natural materials and metamaterials, is
to find the relationship between the fields and the polarizations and
determine from that the values of εr and µr . In natural materials this
is a serious problem because, even in the linear case, we have only a
very approximate idea of what is going on at the atomic or molecular
level. For man-made materials the difficulties are somewhat easier to
overcome because we have a better knowledge of the properties of the
elements inserted.
∇2 E + k 2 E = 0 , (1.21)
where
k 2 = ω 2 µε . (1.22)
In Section 1.2 we started with the assumption of harmonic time varia-
tion. We add to it now harmonic variation in space as well. The solution
of eqn (1.21) is assumed in the form
E = E0 e −j k · r , (1.23)
where k is the wave vector that indicates the direction of propagation, r
is a position vector and E0 is a constant vector perpendicular to k. This
is called a plane wave because none of the variables change in a plane
perpendicular to k.
The magnitude of the wave vector k = |k| is known as the wave
number by physicists and as the complex propagation coefficient by en-
gineers. Substituting eqn (1.23) into the wave equation and separating
the variables we find that a solution exists when the equation
k = β − jα, (1.25)
and they will be called the propagation coefficient, β, and the attenua-
tion coefficient, α. Assuming propagation in the z direction and taking
the electric field as a scalar, eqn (1.23) may be written as
E = E0 e −j βz e −αz , (1.26)
i.e. the wave declines exponentially in the direction of propagation as
may be expected. Although losses are indispensable if we want an ac-
curate answer, very often we can find the main features of a physical
phenomenon by neglecting losses. In those cases we shall make no dis-
tinction between k and β.
In this plane-wave solution the magnetic field is perpendicular both
to the electric field and to the wave vector, and its magnitude is equal
to
r
E µ0
H= , η0 = , (1.27)
η0 ε0
where η0 is the free-space impedance equal to 120π ohm. Note that the
vectors E, H and k are a right-handed set. It has become fashionable to
refer to media where this relationship is satisfied as right-handed media
in contrast to left-handed media to be discussed in Section 2.11.
6 Basic concepts and basic equations
where subscripts t and n stand for the tangential and normal compo-
nents. If medium 2 is a metal, it is nearly always (surface plasma waves
are again exceptions) taken as having infinite conductivity, which means
that the tangential component of the electric field must vanish in medium
2.
5
∂Hz ∂Hy When the electric field has compo-
− = j ωε0 Ex , (1.32) nents only in the transverse plane while
∂y ∂z
propagating in a waveguide, the wave
∂Ex is referred to as a transverse electric or
= −j ωµ0 Hy , (1.33)
∂z TE wave.
∂Ex
= j ωµ0 Hz . (1.34)
∂y
Eliminating Hy and Hz from the above equation we end up with the
wave equation in the form
∂ 2 Ex ∂ 2 Ex
2 + + k02 Ex = 0 , (1.35)
∂y ∂z 2
√
where k0 = ω ε0 µ0 . The propagation along the waveguide in the z di-
rection may be described by the function exp(−j kz z) as for plane waves.
But there is a new feature now. The wave is inside a waveguide. As we
learned in the previous section the tangential component of the electric
field must vanish at a metal boundary. These boundary conditions are
clearly satisfied on the x = 0 and x = b surfaces since the chosen electric
field is perpendicular to those surfaces. But the Ex component must
vanish at y = 0 and y = a. It would vanish if those boundaries coin-
cided with the nodes of a standing wave. Hence, it is logical to assume
that the variation in the y direction will be of the form sin(ky y), and if
we choose the nearest nodes then we need to take ky = π/a. Hence, the
solution for the electric field is
Ex = E0 sin(ky y) e −j kz z . (1.36)
Substituting the above equation into eqn (1.35) we find the relation-
ship
π 2
kz2 = k02 − . (1.37)
a
2
Clearly, there is propagation in the z direction when kz is positive, and
exponential decay when kz2 is negative. The limiting case is when kz = 0,
which occurs when
2
2π π 2
k02 = = or λ = 2a . (1.38)
λ a
8 Basic concepts and basic equations
1 ∂Hy1
Ex1 = −
j ωε1 ∂z
6 This terminology is not a fortunate one. For plane waves both the electric and the magnetic fields are in the transverse
plane, perpendicular to the direction of propagation. They are called, and should be called, transverse electromagnetic or TEM
waves. This is in agreement with the definition in the previous section where a mode in a waveguide is called a TE mode,
provided it has only transverse electric fields. Calling waves, as we do in this section and many times later, TE or TM waves
depending which field is perpendicular to the plane of incidence can only lead to confusion, but the notation is so widespread
in the literature that we reluctantly accept it. There is actually another set of notations, used often by physicists, which is
based on the direction of the electric field vector, E. It is called p polarization when E is parallel and s polarization when it is
perpendicular to the plane of incidence (s coming from the German word senkrecht).
1.6 Refraction at a boundary: Snell’s law and the Ewald circle construction 9
kz1 h
A e −j (kz1 z + kx1 x) − B e j (kz1 z − kx1 x) .
i
=
ωε1
(1.41)
B B
P
P Q
k1b
k1
k2 kɫ k1a
O O
ș1 ș 2 k2a
B’ B’
(a) (b)
Fig. 1.4 Wave vector diagram at the boundary of two media. (a) k1 < k2 and (b) k1 > k2
Boundary
Medium 1 Medium 2
xB B Land Sea
P ș2
ș1
zA
xP
A
xA
zA z=0 zB
(a) (b)
Fig. 1.5 Fermat’s principle for the refraction problem. (a) Geometrical scheme. (b) Practical example ‘damsel in distress’
dT xP − xA 1
=0 = q
dxP v
2 land
(xP − xA )2 + zA
xB − xP 1
−q , (1.53)
2 vsea
(xB − xP )2 + zB
all the optical paths are equal when the plane wavefront is reached.
Traditionally it is the job of a lens to realize this conversion. So, let us
find the contour of a lens that will do it.
We shall assume a 2D geometry with the rays emanating from point O
and work out the optical path of a general ray travelling in the θ direction
(see Fig. 1.6(a)) and that of the axial ray when they reach the P1 P2
wavefront. The former one propagates in free space hence its optical
path is OP1 , whereas the latter one propagates in free space from O to
C and in the lens from C to P2 . Hence, the equality of optical paths
demands
whence
d(1 − n)
r= . (1.59)
1 − n cos θ
This may be recognized as a hyperbola in polar co-ordinates. It has
an asymptote at
1
cos θ = . (1.60)
n
Transformation to normalized Cartesian co-ordinates,
r r
z= cos θ and x= sin θ (1.61)
d d
yields the more familiar equation of a hyperbola8 8
2D lens contours are generally known
as circles. In fact, the above equation
2 of a hyperbola for low angles is approx-
x2
n 1 imately the same as that of a circle.
z− − 2 = . (1.62)
n+1 n −1 (n + 1)2
For n = 1.5 and 2 the contours are plotted in Fig. 1.6(b). As usual, it
has a convex shape.
14 Basic concepts and basic equations
N e2
ωp2 = (1.67)
ε0 m
is known as the plasma frequency. Equation (1.66) is the dielectric
constant of an ideal plasma. Interestingly, it is negative below the plasma
frequency and positive above the plasma frequency.
Note that eqn (1.22) still applies. We only need to replace ε by εeff .
When εeff is complex, k will also be complex, implying both propagation
and attenuation. When εeff is a negative real number then k is purely
imaginary. There is no propagation. The wave declines without any
change in the phase.
that the waves can propagate along the boundary but their amplitudes Boundary
decline exponentially away from the boundary as shown in Fig. 1.7.
Medium 1 Medium 2
This can happen when kx is sufficiently high so that kz1 and kz2 are
imaginary and must be replaced by −j κ1 and −j κ2 , where κ1 and κ2
are real. Hence, the propagation coefficient in the x direction is obtained x
Field amplitude
as
z
kx2 = k12 + κ21 = k22 + κ22 . (1.68)
We are looking for a wave that can exist on the surface without an
Fig. 1.7 Variation of a surface wave
input.9 Hence, we can take A = 0 but B is taken as finite. It is the amplitude at a boundary
amplitude of the wave that declines away from the boundary in the
negative z direction. The equations for the magnetic and electric fields 9
In more pretentious language it means
are then that we are looking for eigensolutions.
Hy1 = B e κ1 z e −j kx x ,
−κ1
Ex1 = B e κ1 z e −j kx x ,
j ωε1
−kx
Ez1 = B e κ1 z e −j kx x (1.69)
ωε1
in medium 1 and
Hy2 = C e −κ2 z e −j kx x ,
κ2
Ex2 = C e −κ2 z e −j kx x ,
j ωε2
−kx
Ez2 = C e −κ2 z e −j kx x (1.70)
ωε2
in medium 2. In order to satisfy the boundary conditions we need to
match Hy and Ex at z = 0, which yields10 10
Note that the boundary condition
ε1 Ez1 = ε2 Ez2 is automatically satis-
κ1 κ2 fied, provided eqn (1.71) holds.
B=C and − = . (1.71)
ε1 ε2
Note that we have taken both κ1 and κ2 to be positive. Hence,
eqn (1.71) can be satisfied only when ε2 is negative. This is a con-
dition for the existence of a surface wave. It is interesting to note at this
stage that eqn (1.71) is equivalent to the condition that
ζe = −1 . (1.72)
Another condition to satisfy is given by eqn (1.68). We may then
substitute κ1 and κ2 from eqn (1.68) into eqn (1.71) and obtain, after
some algebraic operations, a relationship between kx and ω,
r
ω εr1 εr2
kx = . (1.73)
c εr1 + εr2
We want the wave to travel along the surface, hence kx must be real.
With εr2 < 0 we may then write the condition to satisfy as
16 Basic concepts and basic equations
B 2j (1 − ζe2 ) sin(kz2 d)
R= = , (1.75)
A (1 + ζe )2 e j kz2 d − (1 − ζe )2 e −j kz2 d
F 4ζe
T = = , (1.76)
A (1 + ζe ) e z2 − (1 − ζe )2 e −j kz2 d
2 j k d
C 2(1 + ζe )
= , (1.77)
A (1 + ζe ) e z2 d − (1 − ζe )2 e −j kz2 d
2 j k
1.12 Dipoles 17
D 2(ζe − 1)
= , (1.78)
A (1 + ζe ) e z2 d − (1 − ζe )2 e −j kz2 d
2 j k
1.12 Dipoles
The microscopic electric and magnetic dipole moments, p and m, were
introduced in Section 1.3 for determining the macroscopic electric and
magnetic polarizations, P and M. However, dipoles are useful in many
other contexts partly because they can provide good approximate models
to describe complicated electromagnetic phenomena but also because
they can form the basis of actual devices like the electric dipole antenna.
We shall present here the equations relating the dipole moments to
fields in a compact vectorial form that physicists prefer. It might look
unfamiliar to electrical engineers:
3(r · p)r − r2 p 2
1 2 r p − (r · p)r
E= (1 + j k0 r) + k0 e −j k0 r ,
4πε0 r5 r3
(1.79)
with k0 = 2π/λ. The relationship between the magnetic dipole moment
and the magnetic field takes an entirely analogous form: p should be
replaced by m and ε0 in front of the square bracket by µ0 .
One works usually in spherical co-ordinate systems (z, θ, ϕ) and the
usual assumption is that the dipole moments (whether electric or mag-
netic) point in the z direction. To find the components of the electric
field we need then the unit vector in the r and θ directions:
Substituting eqns (1.80) and (1.81) into eqn (1.79) we obtain the electric-
field components in the form
2p cos θ
Er = (1 + j k0 r) e −j k0 r , (1.82)
4πε0 r3
p sin θ
Eθ = (1 + j k0 r − k02 r2 ) e −j k0 r , (1.83)
4πε0 r3
Eϕ = 0, (1.84)
j ωp sin θ
Hϕ = (1 + j k0 r) e −j k0 r . (1.85)
4πr2
The analogous equations for a magnetic dipole are
18 Basic concepts and basic equations
m cos θ
Hr = (1 + j k0 r) e −j k0 r , (1.86)
2πµ0 r3
m sin θ
Hθ = (1 + j k0 r − k02 r2 ) e −j k0 r , (1.87)
4πµ0 r3
Hϕ = 0, (1.88)
B = µ0 µr H , (1.98)
where µr denotes now the tensor in eqn (1.97). Analogously, the per-
mittivity may also be a symmetric tensor in both natural materials and
in metamaterials.
1.16 Polarizability
Let us take an element and ask the question whether electric or magnetic
dipole moments can be induced in it by an electric or magnetic field.
Assuming the relationships to be linear we may write them as
p = αe E and m = αm H , (1.99)
where αe and αm are the electric and magnetic polarizabilities. Equa-
tion (1.99) implies that the dipole moments will point in the same direc-
tion as the fields that induce them. We can, however, write more general
relationships if we allow a field in one direction to induce a dipole mo-
ment in another direction. Then, the polarizabilities become tensors and
eqn (1.99) may be rewritten as
p = αe E and m = αm H , (1.100)
where αe and αm are symmetric tensors. However, this is still not general
enough because an electric field might induce a magnetic dipole moment
and, vice versa, a magnetic field might induce an electric dipole moment.
Acknowledging the possibility of such cross-polarization we write the
general relationship as
where new notations have been introduced. Now the electric and mag-
netic polarizability tensors are denoted by the superscripts ee and mm,
respectively, indicating that the electric field induces an electric di-
pole and the magnetic field induces a magnetic dipole. The cross-
polarizability tensors αem and αme are related to each other (the so-
called Onsager relations) as
py = αem
yz Hz . (1.104)
There are also other notations in the literature but they don’t seem to
be simpler either.
∂Ey
= j ωµ0 µxx Hx , (1.106)
∂z
∂Ey
= −j ωµ0 µzz Hz , (1.107)
∂x
∂Hx ∂Hz
− = j ωε0 εyy Ey , (1.108)
∂z ∂x
from which the wave equation can be derived as
∂ 2 Ey ∂ 2 Ey ω2
µzz 2 + µxx 2 + 2 εyy µxx µzz Ey = 0 , (1.109)
∂z ∂x c
which is the same kind of second-order differential equation as the wave
equation (1.21).
ω 2 = k 2 c2 + ωp2 . (1.110)
It follows from this equation that there is no propagation for ω < ωp and
that for large enough ω we again obtain the linear relationship ω = kc.
The phase velocity is the velocity with which a single-frequency wave
travels. But of course a single-frequency wave does not carry any infor-
mation. The velocity of a group of frequencies that do carry information
is the group velocity defined as
refractive index
dω
vg = . (1.111)
dk
A wave with positive group velocity is called a forward wave (phase
and energy move in the same direction) and one with a negative group
velocity is a backward wave. The distinction between forward and back-
ward waves is quite fundamental and particularly important in the the-
frequency ory of metamaterials, as we shall see later in the book.
Under the heading of dispersion it is desirable to mention anomalous
Fig. 1.9 Anomalous dispersion dispersion as well. It is a term from the nineteenth century when ab-
sorption spectra of various materials were studied in the optical region.
The variation of refractive index close to the absorption peak took the
shape shown in Fig. 1.9. There is clearly a region where dn/dω < 0 tak-
ing its smallest value at the resonant frequency. Does it mean that the
corresponding group velocity is negative? Not necessarily. Considering
the definition of the refractive index in Section 1.4 as n = ck/ω we find
dn c k dk
= − + , (1.112)
dω ω ω dω
whence
dω c
= , (1.113)
dk dn
n+ω
dω
i.e. the condition for the group velocity to be negative is not only that
dn/dω must be negative but also the slope of the n versus ω curve must
be sufficiently large.
1.19 Mutual impedance and mutual inductance 23
I1 I1
loop 1
loop 1 loop 2
loop 2
(a) (b)
Fig. 1.10 Magnetic coupling for two loops is positive in the axial case (a) and negative in the planar case (b)
Φ2 = M21 I1 . (1.114)
Note that the mutual inductance can be complex if the distance be-
tween the elements becomes comparable with the wavelength. For most
of the book we shall be concerned with physical situations in which the
mutual inductance is real, but it may still be positive or negative. A sim-
ple illustration in Figs. 1.10(a) and (b) shows the difference between the
two cases. The mutual inductance is positive if the magnetic field lines
24 Basic concepts and basic equations
cross the two loops in the same direction (we call this the axial config-
uration), and the mutual inductance is negative (planar configuration)
when the magnetic field lines are in the opposite direction.
lm lm
Lk = and Rk = . (1.119)
N e2 S N e2 Sτ
Looking at the expression of the kinetic resistance it becomes clear
that it is nothing else but the ordinary resistance. On the other hand,
the expression for the kinetic inductance is entirely new. As the cross-
section of the conductor gets smaller the kinetic inductance becomes
comparable with the magnetic inductance. It could even become the
dominant inductance, as for example in circuits containing nanorods.
Since the plasma frequency (defined by eqn (1.67)) will often appear in
this book we shall rewrite the expression for the kinetic inductance in
the form
l
Lk = , (1.120)
πr ε0 ωp2
2
Iin Iout
Vin 1 2 N Vout
Zu
I(z) I(z + dz)
d2 V
+ Yu Zu V = 0 . (1.132)
dz 2
Comparing the above equation with eqn (1.21) we find that the two
are identical if we take for the propagation constant
k 2 = Yu Zu . (1.133)
Now let us take the values
of course not a plane wave but Fig. 1.15 and eqn (1.135) are also capable
of providing the characteristics of a plane wave if we introduce
Lu = µ0 and Cu = ε0 , (1.138)
which would lead again, in a different manner, to eqn (1.137).
eqns (1.135) and (1.144) are different. The reason is that in Fig. 1.15
we already assumed that the line is continuous, whereas our chain of
four-poles implies a set of discrete elements. We can, of course, still
affect the conversion from a discrete to a continuous line by assuming
that ka ≪ 1, in which case eqn (1.144) reduces to
(ka)2 = ω 2 LC , (1.145)
which is identical with eqn (1.135) considering that Lu = L/a and Cu =
C/a.
We have now proven that our last two models may lead to the same
result. We should notice, however, that we have now a wave solution
for any kind of four-poles. For our second example we shall take the
four-pole of Fig. 1.17(a) where
1 1
Z= and Y = . (1.146)
j ωC j ωL
The corresponding dispersion equation is
1
cos ka = 1 − 2 . (1.147)
2ω LC
plotted in Fig. 1.17(b) for ka in the range 0 to π. The phase velocity,
ω/k, is always positive and the group velocity, dω/dk is always negative.
This is our first example of a backward wave.
Our third example is quite a different four-pole, shown in Fig. 1.18(a).
We have separated here the mutual inductance, M , and the self-induc-
tance, L, as it leads to simpler mathematics. Its chain matrix can be
found quite simply in the form
0 −j ωM
b= 2 ,
(1.148)
1 L ω
1 − 02
−
−j ωM M ω
√
where ω0 = 1/ LC. The dispersion equation derived from the above
parameters, solving this time for ω is
30 Basic concepts and basic equations
Fig. 1.18 (a) An equivalent circuit for magnetoinductive waves and (b) its dispersion curves for positive and negative coupling
ω0
ω= r . (1.149)
2M
1+ cos(ka)
L
The corresponding dispersion curves for 2M/L = 0.1 and −0.1 are
shown in Fig. 1.18(b). It may be seen that we have a forward wave for
positive M and a backward wave for negative M . Both waves will often
appear in this book under the name of magnetoinductive waves.
the values that those more interested in what is going on will compare
with theoretical results.
Z
F (g) = G(fx , fy ) = g(x, y) e −j 2π(fx x + fy y) dxdy . (1.153)
Z
g(x, y)|z=z2 = T (fx , fy )G(fx , fy ) e j 2π(fx x + fy y) dfx dfy . (1.154)
and z2 then the transfer function of each one of them must be multiplied
together similarly to chain matrices in the previous section.
As an example, let us find the transfer function of a wave in free space
propagating at an angle θ1 relative to the z axis. As we have seen before,
the phase varies in the z and x directions as exp[−j (kz z + kx x)] where
p
T1 (kx ) = e −j ω 2 µ0 ε0 − kx2 d1 , (1.158)
4ζe
T2 (kx ) = , (1.159)
(1 + ζe ) e j
2 kz2 d2 − (1 − ζe )2 e −j kz2 d2
p
T3 (kx ) = e −j ω 2 µ0 ε0 − kx2 d3 . (1.160)
the details of a spatial function the higher the spatial frequencies that
can reproduce those details. Note, however, that eqn (1.155) must be
satisfied. When the spatial frequency is high enough to satisfy the in-
equality
√
kx > ω µε , (1.161)
then, according to eqn (1.155), kz , the wave number responsible for
propagation in the z direction, becomes imaginary. The wave declines
exponentially, as we saw for surface waves in Section 1.9, but the ex-
pressions derived are still valid.
This page intentionally left blank
A bird’s-eye view of
metamaterials
2
2.1 Introduction 2.1 Introduction 35
2.2 Natural and artificial
The aim of this chapter is to provide an introduction to the subject. It is materials 35
called a bird’s-eye view because it looks at the subject of metamaterials, 2.3 Determination of the
admittedly superficially, from above, pausing at certain views that the effective permittivity/
authors like to share with the reader and making some relevant com- dielectric constant
in a natural material 40
ments without going into very much detail. Physical concepts, most of
2.4 Effective plasma frequency
them already introduced in Chapter 1, will be further developed. The of a wire medium 42
mathematics will be kept to an absolute minimum. It will mostly be
2.5 Resonant elements
simple algebra. for metamaterials 44
Who might benefit from this survey? Well, the authors have certainly 2.6 Loading the transmission
benefited from writing it because they were forced to pronounce judg- line 45
ment on the relative simplicity of the various parts of the subject and 2.7 Polarizability of a
had to decide on where to start, how to proceed and what to include. For current-carrying resonant
a beginner, who first comes into contact with the subject, there might be loop: radiation damping 50
too many new concepts. Even then it is hoped that some concepts will be 2.8 Effective permeability 52
picked up, stored in the brain, and will later act as catalysts facilitating 2.9 Dispersion equation of
magnetoinductive waves
the absorption of further information. For a research student who has derived in terms of
already acquired some familiarity with the fundamentals it might serve dipole interactions 55
as a reinforcement of existing knowledge. For unbelievers who question 2.10 Backward waves
the correctness of all new ideas wherever they come from, until properly and negative refraction 56
checked, this chapter might offer new things to worry about. For those 2.11 Negative-index materials 58
who are thoroughly familiar with all the basic tenets of modern research 2.12 The perfect lens 66
in metamaterials this might still be suitable for bedtime reading. 2.13 Circuits revisited 73
This chapter makes no claim to rigour, which will be sacrificed at
the altar of simplicity. We shall not be concerned with priorities either.
When we just want to get through a large number of different concepts
it is not worth pausing and telling the reader who did what and when
and it is particularly difficult to reconcile rival claims to priority. We
shall therefore quote relatively few references, but that would not, of
course, apply to other chapters.
a are amorphous, meaning that all those elements are heaped upon each
other in a random manner, others are crystalline, which means that they
arrange themselves into some regular periodical pattern.
Our main interest is in the interplay of waves and materials restricted
to classical physics. The key parameter is a/λ, where a is the distance
between elements in the material and λ is the free-space wavelength.
For simplicity, let us assume that the elements arrange themselves in a
regular cubic lattice, the same in all three directions, as may be seen in
Fig. 2.1. We may now look at two cases: the wavelength is comparable
with a or much larger than a. In the first case the Bragg effect comes
into play. The simplest example is shown schematically in Fig. 2.2. An
electromagnetic wave may be seen to be incident perpendicularly at a
lattice. The wave propagating then from row 1 to row 2 will cover a
Fig. 2.1 Cubic lattice path a. The part of the wave that is reflected by row 2 will have covered
a an additional distance a when arriving back at row 1. When a happens
to be equal to one half of a wavelength then the waves reflected by all
the rows will have the same phase and will reinforce each other. If there
are many rows, and there are indeed many of them in a crystal, then
incident most of the incident power may be reflected. This effect is at the basis
wave of X-ray and electron diffraction in crystals.
When the wavelength is much larger than the lattice period then no
such dramatic effect occurs, but it is nonetheless significant. There may
not be major reflection or diffraction but the electromagnetic wave is still
considerably affected when it enters a material. We may then ignore the
details and pretend that there is no discrete structure: the material is
1 2 3 homogeneous and continuous. The aim is then to find some effective
parameters like electric permittivity and magnetic permeability. This is
Fig. 2.2 Electromagnetic wave incident
known as the effective-medium approximation. Summarizing, there is
normally at a lattice the Bragg effect, when the distance between the elements is comparable
with the wavelength, and there is effective-medium response when that
distance is much less than the wavelength.
Now let’s think of artificial materials in which atoms and molecules
are replaced by macroscopic, man-made, elements. Let’s not worry for
the moment how the elements remain in their allotted space. That may
not be always obvious but we can safely assume that we have complete
freedom in choosing both the elements and the distance between them.
Now, all dimensions are bigger than in natural materials but the division
into the above two categories is still valid. When the separation between
the elements is comparable with the wavelength we have the Bragg effect,
and when the separation is much smaller than the wavelength we can
resort to effective-medium theory. In the former case we talk about
photonic bandgap materials and in the latter case about metamaterials.
Can we have a better definition of metamaterials? Not easily. There
is broad agreement on what the subject is about but not about all the
details. It would need a fairly long description accompanied by a number
of examples to be more precise. We shall give here two definitions in
current use.
1. Metamaterials are engineered composites that exhibit superior prop-
2.2 Natural and artificial materials 37
erties not found in nature and not observed in the constituent materials.
2. A metamaterial is an artificial material in which the electromagnetic
properties, as represented by the permittivity and permeability, can be
controlled. It is made up of periodic arrays of metallic resonant elements.
Both the size of the element and the unit cell are small relative to the
wavelength.
Definition 1 is too general, whereas definition 2 is not general enough:
neither of them mention applications and do not even pay homage to
negative refraction. We shall make no attempt here to give a more com-
prehensive definition. Perhaps definition 2 could be made more general
by adding that control, among other things, means that it is possible to
achieve simultaneously negative permittivity and negative permeability
at the same frequency, which will then lead to negative-index media and
to negative refraction. It is not easy to find a definition that would sat-
isfy everybody. For a discussion of the difficulties of a proper definition
and for many other ideas on the subject see Sihvola (2007).
We have talked about natural and artificial materials and their rela- a
tionship to transverse electromagnetic waves. But electromagnetic waves
n-1 n n+1
are not the only ones that should be considered. In a crystal, for ex-
ample, the atoms and molecules may move relative to each other. They
cannot move far away because there are some restoring forces. One of
the manifestations of these motions and of the forces opposing them is
the emergence of acoustic waves. xn-1 xn xn+1
Let us take a one-dimensional chain of atoms in which the elements are
at a distance a from each other at rest. Figure 2.3 shows schematically Fig. 2.3 1D chain of atoms showing
the positions of three of the atoms at xn−1 , xn and xn+1 . Three elements displacement from the quiescent posi-
are usually sufficient when we can get away with an approximation that tion
takes into account only nearest neighbours (see, e.g., Brillouin 1953;
Dekker 1965). Note that all the displacements are in the longitudinal
direction, which makes the problem conceptually simpler. How can we
work out the net force? If xn+1 > xn then there will be a force on the
atom xn wanting to move it to the right. Conversely, if xn > xn−1 then
Za
there will be a force to the left. A proper mathematical formulation
using Newton’s equation (eqn (1.1)) followed by the assumption of a
wave solution will yield the dispersion equation for acoustic waves (Fig.
2.4). It shows the relation between frequency, ω, and wave number,
k. The uppermost frequency ωa at which acoustic waves can propagate
occurs at ka = π or λa = a/2, where λa is the acoustic wavelength. At
frequencies above ωa acoustic waves of the kind we investigated cannot
propagate. The band up to ωa is the pass band and above it, where the 0
0 S
wave cannot propagate, is the stop band. ka
To give another example plasma waves may also propagate in a natural
material. Take sodium for example. It is a metal in which numerous Fig. 2.4 Dispersion curve of acoustic
electrons float in a pool and are compensated by positive ions. In the waves
simplest case all the electrons undergo simple harmonic motion in the
direction of their propagation. If the frequency is high enough, above the
plasma frequency somewhere in the ultraviolet, these electrons can move
quite freely. We should also mention spin waves in magnetic materials
38 A bird’s-eye view of metamaterials
In = I0 e −j nka . (2.3)
We obtain the dispersion equation for MI waves as
1 This is now the second time in this section that we mention nearest-neighbour interaction. It is not something we would
readily associate with waves. It makes no sense for water waves and even less for electromagnetic waves in vacuum. There are
no neighbours in vacuum. On the other hand, it would not be difficult to develop a physical picture of wave propagation based
on nearest neighbours if we give a little thought to it. Imagine, for example, a large number of houses next to each other in a
street in which nearest neighbours can talk to each other over the fence. An interesting piece of news could certainly reach the
last house in the street by propagating along the row of houses via nearest-neighbour interaction.
2.2 Natural and artificial materials 39
Z0
2M
1 L
Z
Z0
Z0
2M
1 L
S/2 S Fig. 2.7 Dispersion of magnetoinduc-
tive waves with positive magnetic cou-
ka pling
ω2
L 1 − 02 + 2M cos(ka) = 0, (2.4)
ω
where we took into account that
1
Z0 = j ωL + . (2.5)
j ωC
Note that eqn (2.4) is identical with eqn (1.149). The corresponding dis-
persion equations have already been plotted in Fig. 1.18. It is replotted
in Fig. 2.7 for 2M/L > 0 for a more pervasive examination.
It shows some similarity to the dispersion curve of acoustic waves
at least in the sense that the group velocity, dω/dk, is always positive
and at the band edge the group velocity is zero, as it is for all waves
on discrete structures. Note also that there is a lower cutoff frequency
below which the MI wave cannot propagate. The pass band is within
the range
ω0 ω0
r <ω< r . (2.6)
2M 2M
1+ 1−
L L
There is no reason of course that the coupling between the metama-
terial elements has to be magnetic. It can be electric. Well before the
advent of metamaterials an experiment, shown in Fig. 2.8(a), was per-
formed by Shefer (1963), using a set of metallic rods (Fig. 2.8(b)). One
of the horns is a transmitter of microwaves, the other horn is a receiver,
and a wave travels along the rods from one horn to the other horn due
to electric coupling. Typical dimensions in the experiments were l = 12
mm, d = 1 mm and a distance between the rods of 5 mm. They found
good transmission between the horn antennas at around the frequency
of 1.2 GHz. A more recent example of wave propagation by rods on a
substrate (Hohenau et al., 2005), at a frequency five orders of magnitude
higher (360 THz), is shown in Fig. 2.8(c), where the element dimensions
are 800 nm × 80 nm × 50 nm. The distance between the elements is
320 nm.
40 A bird’s-eye view of metamaterials
Height (nm)
topography
y (µm)
0
l
í
0 2 4
z (µm)
(a) (b) (c)
Fig. 2.8 Electric coupling on a chain of rods. (a) Schematic of a setup and (b) dimensions of the rod used in the experiment
by Shefer (1963). Copyright
c 1963 IEEE. (c) Topography of the setup used by Hohenau et al. (2005). Copyright
c 2005
EDP Sciences
D = ε0 E + P , (2.7)
as has already been given in eqns (1.9) and (1.19).
The central question is the relationship between E and P in a ma-
terial in which electric dipoles appear in response to an electric field.
Let us consider an element inside a cubic material at the centre of a
rectangular co-ordinate system (0, 0, 0) and apply an electric field, Eext ,
to the material parallel to the z co-ordinate. What will be the electric
field at our chosen element? One’s first thought is that the electric field
will be equal there to the external field. This would indeed be the case
if the electric field would not cause the positive and negative charges to
separate. But the charges do separate. Each dipole will then contribute
to the electric field at (0, 0, 0). We shall call the field there Eloc , the
local field, which is equal to
eqns (1.82) and (1.83). When we sum up the contributions at the point
(0, 0, 0) in a cubic lattice from every element inside the sphere it yields
zero electric field as shown in Appendix B. So, after all, the electric
field is equal there to the external field. Is it? Not really, because this
was the contribution of the elements inside the sphere. We still need
to consider the effect of the elements outside the sphere. Now we can
no longer say that the effect will be zero because for elements far away
we also need to take into account the slowly declining radiation field.
It is a tremendously difficult problem. For a recent analysis see Belov
and Simovski (2005a). Here, we shall disregard all those difficulties
and follow the time-honoured method of accounting for the effect of
the elements far away by charges induced on the surface of the sphere.
The electric field at the centre may then be determined by a simple
integration from the charges far away as equal to P/3ε0 , which gives for
the local field
P
Eloc = Eext + . (2.9)
3ε0
But the polarization P at (0, 0, 0) is proportional to the local electric
field there
P = N αe Eloc , (2.10)
where αe is the atomic/molecular polarizability discussed in Section 1.16.
It tells us how effective the local electric field is in producing an electric
polarization. From eqns (2.9) and (2.10) we obtain
N αe
P = Eext . (2.11)
N αe
1−
3ε0
We may now rely on the definition of εeff as
Ea
I= , (2.16)
Zw
where
Zw = Rw + j ωLw (2.17)
is the impedance of the wire. Next, we shall find the average current
density in the unit cell that has an area of a2 . It is
E
Jav = . (2.18)
(Rw + j ωLw )a
2.4 Effective plasma frequency of a wire medium 43
S21 [dB]
Having found the relationship between the electric field and the cur-
rent density we can follow the method outlined in Section 1.9 to find the
effective dielectric constant as
1
εr = 1 + ε0 (Rw + j ωLw )a . (2.19)
jω
Defining now
1
ωp2 = , (2.20)
ε0 aLw
with ωp being an effective plasma frequency,3 we may rewrite eqn (2.19)
as
ωp2 eff
εr = 1 + , (2.21)
jω
ω2 −
τw
where losses are characterized by the time constant
Lw
τw = . (2.22)
Rw
The expressions for the resistance and inductance (see, e.g., Grover 1981)
are
3 Itis always an advantage to look at a phenomenon from different angles. The effective plasma frequency in eqn (2.20) is
written by Pendry et al. (1996) in the form
e2 Neff
ωp2 = .
ε0 meff
It is the same as eqn (1.67) but instead of the actual electron density, N , and actual electron mass, m, the physical quantities
appearing in the above equation are Neff , the average electron density in the unit cell, and meff , the effective mass of the
electron defined so that eqn (2.20) should be satisfied. According to Pendry et al. the reduction of plasma frequency may be
interpreted as due partly to a decrease in the effective electron density and partly to an increase in the effective mass, and that
increase may amount to four orders of magnitude. It is an interesting proposal but there is some inconsistency in it. If the
mass of the electron has increased reducing thereby the plasma frequency then the increased mass must have also reduced the
conductivity. A decrease in conductivity by four orders of magnitude would make the losses enormous and there is no evidence
of that. For a different criticism of the introduction of an effective mass see also Pokrovsky and Efros (2002c).
44 A bird’s-eye view of metamaterials
a µ0 a 2a 3
Rw = 2 and Lw = ln − . (2.23)
πrw σ0 2π rw 4
As an example, let us take a = 6 mm, rw = 0.03 mm, and σ0 = 5.8 ×
107 S/m for copper. The resultant plasma frequency may be calculated
from eqns (2.20) and (2.23) to be 8.73 GHz, well in the microwave region.
For copper at room temperature the time constant is τw = 2.24 × 10−8 s,
which makes the factor ωτw = 1230. For these parameters it is large
enough to ignore losses.
For a detailed experimental study see Gay-Balmaz et al. (2002). With
the above choice of rw = 0.03 mm and a = 6 mm the transmission
4
(S21 ) through N layers (each layer consisted of 39 wires clamped in
As mentioned before, a metamaterial
a groove in a substrate) as a function of frequency is plotted in Fig.
can be regarded a proper material only
if there is a sufficient number of ele- 2.10(b). The transition occurs at about 9.2 GHz, which is in fairly good
ments within a free-space wavelength. agreement with the theoretical value of 8.73 GHz calculated above. The
At 9.2 GHz the wavelength is 3.26 cm long dashed line in Fig. 2.10(b) is the reference level measured when no
so there are about 5 layers per wave-
length. Is that enough? From this set
wires are present. The curves with dot-dash, dashed, dotted and solid
of experiments we could conclude that lines correspond to N = 5, 10, 15 and 20 layers. It may be seen that
yes, five is enough but that is of course 5 layers already cause considerable attenuation but in order to have a
not a general proof. We believe, how- sharp transition 15 layers are needed. Increasing the number of layers
ever, that such a number would suffice
for many practical applications.
to 20 makes hardly any difference.4
be very compact, and for the inductance a loop that can couple to a
magnetic field whether it comes from an incident electromagnetic wave
or from currents flowing in neighbouring elements.
Are there other resonant elements in the same category? There are
plenty of them starting, probably, with the re-entrant cavity used in
klystron amplifiers going back more than half a century (see, e.g., Hansen
1939). We shall present a gallery of small resonators in Chapter 4. For
the time being we shall concentrate on one resonator, a member of the
family of split-ring resonators (SRRs), which has become very popular in
the last decade (Pendry et al., 1999). It consists of two concentric split
rings with gaps on opposite sides. Two realizations with small pipes and
in printed circuit form are shown in Figs. 2.12(a) and (b). The third one
(Fig. 2.12(c)) is the so-called complementary split-ring resonator where
metal replaces air and vice versa (see Baena et al. 2005a).
At first sight the physical phenomena governing the operation of a
SRR are quite complicated. Each ring has a self-inductance, there is Fig. 2.12 Split-ring resonators (a) as
a mutual inductance between them, there is a capacitance between the pipes, (b) in printed circuit form, and
rings and there are gap capacitances at the splits. If one wants to take (c) as a complementary variety
into account all these factors then it is difficult indeed to determine its
properties. It turns out, however, that a simplified physical picture can
lead to an excellent approximation (Marques et al., 2002c). First, ig-
nore the gap capacitances on the basis that they are small and they are
unlikely to have a major influence on the flow of currents. Secondly, ig-
nore the mutual inductance. In the third place, take the self-inductance
equal to the average self-inductance of the two rings. In the fourth place,
consider the two inter-ring capacitances between the splits as being con-
nected in series. We may then put these assumptions into mathematical
form. Take the average radius of the SRR to be equal to r0 , the average
inductance of the two rings equal to L and the inter-ring capacitance per
unit length equal to Cpu . Then, the capacitance of a half-ring is equal
to
πr02 µ0
M’ Lt M′ = . (2.27)
a
Ct The equivalent circuit of the loaded transmission line in terms of four-
poles may be seen in Fig. 2.14.
We can find the dispersion equation from there by the method outlined
previously. First, we need to find the chain matrix of the ‘load’, i.e. that
Fig. 2.14 The four-pole equivalent of a of the LC circuit coupled by M ′ to the transmission line. A little algebra
transmission line loaded by a resonant yields
loop
0.1
RHS of eqn (2.29)
1.4
0
0
Z0 Zq Z/Z 1.2
Zq
1 Z0
0.95 1 1.05 1.1 0 0.05 0.1
Z/Z 0 ka/S
(a) (b)
Fig. 2.15 Loading by rings. (a) Plot of the RHS of eqn (2.29) against ω/ω0 and (b) dispersion curve
2 2
ka ω 2 ω − ωq
4 sin2 = (1 − q 2 ) 2 2 , (2.29)
2 ωt ω − ω02
where
M ′2 1 ω02
q2 = , ωt2 = and ωq2 = . (2.30)
LLt Lt Ct 1 − q2
The left-hand side can vary only between 0 and 4, corresponding to
ka = 0 and ka = π. Hence, there is solution of eqn (2.29) only for those
values of the right-hand side that vary within the same limits. For values
of q 2 = 0.1, and ω0 /ωt = 0.1 we plot the right-hand side as a function
of ω/ω0 showing the range for which a solution exists. It may be seen
(Fig. 2.15(a)) that with good approximation there is solution between 0
and ω0 and again above ωq . There is no solution for ω between ω0 and
ωq . The corresponding dispersion curve is shown in Fig. 2.15(b). We
can say that there are pass bands between 0 and ω0 and above ωq up to
a frequency comparable with ωt , and there is a stop band between ω0
and ωq .
If we want to regard the medium made up by these resonant magnetic
elements as a continuous one then we can replace sin(ka/2) by ka/2.
Equation (2.29) then takes the form
2 2
ω 2 ω − ωq
k 2 = (1 − q 2 ) . (2.31)
c2 ω 2 − ω02
The main, and obvious, change is that in the continuous limit the sepa-
ration of the elements, a, no longer appears in the dispersion equation.
It is also clear now that the changes from pass band to stop band and
vice versa occur exactly at the frequencies ω0 and ωq . It may also be
seen that the frequency range extends to infinity.
48 A bird’s-eye view of metamaterials
Lt 1.0
Z/Zp
Ct Lw
0.7
0 0.05 0.1
ka/S
(a) (b)
Fig. 2.16 Loading by rods. (a) Four-pole equivalent and (b) dispersion curve
!
1 ωp2 1
Y = j ωCt + = j ωCt 1− ; ωp = p . (2.32)
j ωLw ω2 Lw Ct
ka ω 2 − ωp2
4 sin2 = . (2.33)
2 ωt2
Clearly there is no q
solution when ω < ωp but a solution exists up to
the frequency, ω = 4ωt2 + ωp2 . The corresponding dispersion curve is
shown in Fig. 2.16(b).
2.6 Loading the transmission line 49
1.4
L C
Zp
Z/Z0
1.2
M’ Lt
Zq
Ct Lw 1 Z0
0 0.05 0.1
ka/S
(a) (b)
Fig. 2.17 Loading by resonant loops and rods. (a) Four-pole equivalent. (b) Dispersion curve
ka ω 2 − ωp2 ω 2 − ωq2
4 sin2 = (1 − q 2 ) . (2.34)
2 ωt2 ω 2 − ω02
The above equation is very similar to eqn (2.29). The only difference
is that in the first term ω 2 should be replaced by ω 2 − ωp2 . What kind
of dispersion characteristics could we expect from eqn (2.34)? It will
depend on the relative values of the three characteristic frequencies, ωp ,
ω0 , and ωq . We know that ωq is larger than ω0 but ωp could be anywhere
depending on the parameters of the rods chosen. Let us choose it to be
larger than ωq . Then, for ω < ω0 we have a stop band, for ω0 < ω < ωq
we have a pass band that happens to be a backward-wave region, for
ωq < ω < ωp we have a stop band, and finally for ω > ωp we have again
a pass band. The dispersion curve for ωp = 1.2 ωq is plotted in Fig.
2.17(b).
The appearance of pass and stop bands is summarized in the diagram
of Fig. 2.18 for the three cases: Rods alone, resonant magnetic elements
alone, rods and magnetic elements combined. Let us just concentrate
our attention on the frequency region between ω0 and ωq in the three
cases. In case 1 that frequency region is a stop band. In case 2 it is
again a stop band, but in case 3 it is a pass band. This is a significant
conclusion. For rods alone or for magnetic elements alone it is a stop
band, but when the two are combined the stop band turns into a pass
band. The rules are quite easy and quite interesting. Two pass bands
50 A bird’s-eye view of metamaterials
make a pass band, a pass band and a stop band make a stop band,
and two stop bands make a pass band. According to these rules we can
assign +1 to a pass band and −1 to a stop band. Then 1 × 1 = 1,
1 × (−1) = −1, (−1) × 1 = −1 and (−1) × (−1) = 1. These are in
fact the same rules that will arise later when discussing negative-index
materials.
j ωµ20 S 2 H
m = µ0 SI = − , (2.37)
Z
whence according to the definition of eqn (1.100)
j ωµ20 S 2
αm = − . (2.38)
Z
If we want to think in more general terms then we can regard the above
polarizability as the αmm
zz component of the polarizability tensor relating
the z component of the magnetic dipole moment to the magnetic field
applied in the z direction.
Note that in the absence of losses αm is real. It becomes complex when
R is added to the impedance. However, this is not the only source of loss.
Power can turn not only into heat but it can also be lost by radiation.
In Section 1.14 we asked the question whether we can take into account
the radiated power by assigning to the element a resistance that we
called the radiation resistance. The derivation was done for an electric
dipole but the value of the radiation resistance for a magnetic dipole was
also quoted (see eqn (1.96)). A small loop is of course equivalent to a
magnetic dipole, hence we can take care of the radiation loss by adding
4
π 2πr0
Rs = η0 (2.39)
6 λ
to the ohmic resistance. For most metamaterials we do not need to
consider the radiation resistance because all dimensions are small relative
to the wavelength. But when the length of the line is larger than the
free-space wavelength, as will be the case in Section 8.2, then radiation
effects must be taken into account, which we can do by adding the
radiation resistance to the ohmic resistance. Under these conditions it
is preferable to work in terms of the inverse of magnetic polarizability,
which will then take the form
1 1 1 1
= + + , (2.40)
αm αm lossless αm ohmic loss αm radiation loss
where
ω02
1 fr L
= − ; fr = 1 − , (2.41)
αm lossless µ20 S 2 ω2
1 L ωL
= −j ; Q= , (2.42)
αm ohmic loss Qµ20 S 2 R
where Q is the quality factor. For the third term a little algebra will
yield
k3
1
=j 0 . (2.43)
αm radiation loss 6 πµ0
52 A bird’s-eye view of metamaterials
k3
1
=j 0 . (2.44)
αe radiation loss 6πε0
Mm = N m = N αm H . (2.45)
We shall now find the effective relative permittivity of the medium per-
pendicular to the plane of the element (according to the notations of
Section 1.16 this is the µzz component of the permeability tensor) from
the definition
B µ0 H + Mm Mm
µr = = =1+ . (2.46)
µ0 H µ0 H µ0 H
With the aid of eqn (2.45), and the definition of magnetic polarizability,
eqn (2.38), we find
µ N S2
µr = 1 − 0 (2.47)
j
L fr −
Q
or
F
µr = 1 − , (2.48)
j
fr −
Q
which is often regarded as the standard form in the literature. F is
defined as
µ0 N S 2
F = . (2.49)
L
Another form of eqn (2.47) for the lossless case may be obtained with a
little algebra as
2.8 Effective permeability 53
ω 2 − ωF2
µr = (1 − F ) , (2.50)
ω 2 − ω02
√
where ωF = ω0 / 1 − F . It may be seen from eqn (2.50) that in the
absence of losses there is a pole at ω0 and a zero at ωF . The variation
of µr with frequency is plotted schematically in Fig. 2.20.
The interesting thing is that between the pole and the zero the per-
meability is negative. How wide is the range of negative permeability?
It can be easily calculated when F is small. Then,
Pr
1 F
∆ωneg.perm. = ω0 √ − 1 ≃ ω0 . (2.51)
1−F 2
In our second model we shall take into account that all the other elements
will also contribute to the flux at element n. The total flux is then due 1
to the applied field plus the flux provided by all the other elements,
Z0 ZF Z
X
Φ = µ0 SH + I Mnn′ , (2.52)
where Mnn′ is the mutual inductance between element n and n′ and the
current I is assumed to be the same in all the elements. The correspond-
ing current must then satisfy the equation Fig. 2.20 Frequency variation of per-
meability
−j ω X
I= µ0 SH + I Mnn′ , (2.53)
Z
whence we may determine the current, from the current the magnetic
dipole density, and from that the effective permeability. We obtain fi-
nally
F
µr = 1 − , (2.54)
j
fr + ∆fr −
Q
where
1X
∆fr = Mnn′ . (2.55)
L
The difference between our first and second model is that a new term,
∆fr , enters the denominator that involves all the mutual inductances.
How can we find that term for a cubic lattice? We have met this problem
in a somewhat different form in Section 2.3. There, we were concerned
with the total electric field at an element due to all other elements. That
sum was shown to be zero in Appendix B. The same applies here. If
the total magnetic field is zero then the mutual inductance must also be
zero. For other lattice configurations the additional term represents a
small shift in the position of the pole where the effective permeability
tends to infinity.
Our third model for finding the effective permeability is based on
eqn (2.31) that gives the dispersion equation for loading by magnetic
elements in the continuous limit. Clearly, the dispersion equation and
the effective permeability are closely connected to each other. For a
54 A bird’s-eye view of metamaterials
ω 2 − ωq2
µr = (1 − q 2 ) , (2.57)
ω 2 − ω02
which is of the same form as eqn (2.50) but q 2 appears instead of F .
The two expressions would be exactly the same if F were to agree with
q 2 . In fact, if we look at eqns (2.27), (2.30) and (2.49) we shall find that
they do agree. Hence, our third model gives the same result as the first
one.
Our fourth model is the oldest of them all, the one that has been
applied to finding the material constants for natural materials for well
over a century, the Clausius–Mossotti model. We derived it in Section 2.3
for the effective permittivity, but the same equation is valid of course
in the magnetic case. We only need to substitute magnetic polarization
for electric polarization in eqn (2.14). We find for the lossless case
2F
ω2 1 − − ω02
3
µr = . (2.58)
2 F 2
ω 1+ − ω0
3
It may be seen that the shape of the µr curve has hardly changed but
the region of negative permeability has shifted p towards lower frequen-
cies.√The pole hasp moved from ω0 to ω0 / 1 + F/3 and the zero from
ω0 / 1 − F to ω0 / 1 + 2F/3. It follows from the above relations that
for small values of F the width of the negative permeability region has
not changed. It is ω0 F/2.
The fifth model is that of Gorkunov et al. (2002) who essentially repeat
the derivation of the Clausius–Mossotti equation but include the effect
of mutual inductances. Then, as we may guess from our second model
the additional term ∆fr appears. To be exact we need to add ∆fr both
to 2F/3 and to F/3 in eqn (2.58).
Losses. Although losses were taken into account in deriving eqn (2.48)
we disregarded them later for simplicity. In the general case when µr is
complex we use the notation
µr = µ′ − j µ′′ . (2.59)
This is plotted in Fig. 2.21 in the vicinity of the resonant frequency for
Q = 100, 1000 and 10 000 from eqn (2.48) with F = 0.1. As may be
expected, the real part of the permeability no longer tends to infinity
and the maximum µ′ achievable is reduced to 2, 3.6 and 6.1 for the three
values of the quality factor. The imaginary part of µr can be quite large.
Its maximum may be larger than that of the real part.
2.9 Dispersion equation of magnetoinductive waves derived in terms of dipole interactions 55
r r
0 0 0
0.8 1 1.2 0.8 1 1.2 0.8 1 1.2
ZZ 0 ZZ 0 ZZ 0
(a) (b) (c)
Fig. 2.21 Frequency variation of the real and imaginary parts of permeability for (a) Q = 100, (b) Q = 1000, (c) Q = 10 000
m0 = αm H , (2.60)
where H comes from all the other dipoles. The element positioned at a
distance na away for n > 0 will give a contribution
X
Htotal = H(n > 0) + H(n < 0) = (mn + m−n ) f (n) . (2.63)
56 A bird’s-eye view of metamaterials
mn = m0 e −j nka , (2.64)
which, substituted into eqn (2.63), leads to the equation
(1 + j k0 na) e −jk0 na
∞
1 1 X
= 3 cos(kna) . (2.65)
αm πµ0 a n=1 n3
The right-hand side of eqn (2.65) is known as the interaction function.
We shall denote it by IF . Hence, the general form of the dispersion
equation is
αm IF = 1 . (2.66)
In what sense is eqn (2.66) a dispersion equation? In the same sense
as the previous dispersion equations. IF depends on ω and k and the
polarizability depends on ω, hence eqn (2.66) relates ω and k to each
other.
This is now the third time that we derive a dispersion equation for a
set of resonant loops. The derivation in Section 1.23 was based purely
on circuit concepts, whereas the one in Section 2.2 was based on nearest-
neighbour coupling and Kirchhoff’s law. It was also assumed there that
all dimensions are small relative to the free-space wavelength. Hence, if
we want eqn (2.66) to reduce to eqn (2.4) we have to take the lossless
case, assume that k0 a ≪ 1 and secondly that the infinite sum should be
reduced to n = 1. We then obtain
ω2 πr4
L 1 − 02 + µ0 30 cos ka = 0 . (2.67)
ω a
Now we are nearly there. Remember that in the derivation used in
6
the present section we assumed magnetic dipoles instead of loops and
For nearest neighbours and small
we found the magnetic field due to a loop in the dipole approximation.
wavelength the radial magnetic field
takes the form Thus, if we further note that in this approximation the mutual induc-
m
tance between two loops a distance a apart in the axial configuration
H=
2πµ0 a3
. (2.68) is6
Considering further that m = µ0 πr02 I,
that the flux Φ is µ0 Hπr02 and the mu-
µ0 πr04
M= , (2.69)
tual inductance is defined as M = Φ/I 2a3
we can find M in the form of eqn (2.69).
then eqns (2.67) and (2.4) may be seen to be identical.
vg
vp
Q P R
B’
(a) (b)
vg2
whereas for negative ε and µ we have a left-handed set. The wave vector
k tells us the direction of the phase velocity, the Poynting vector tells us
the direction of the group velocity. If the two are in opposite directions
we have a backward-wave material with all that implies. Thus, negative
refraction at the boundary of two materials, one having positive material n2 = 0.17
constants and the other negative ones, follows immediately. But there n2 = 0.2
is an alternative explanation. We may argue that the square root in n2 = 0.3
eqn (2.70) can be positive or negative. It is sensible to take it positive n2 = 1
when the material constants are both positive and take it negative when
both material constants are negative. But that will have an influence on n2 = -1
ș1=10°
Snell’s law (eqn (1.46)), n2 = -0.3
n2 = -0.2
n1 sin θ1 = n2 sin θ2 . (2.72) n2 = -0.17
Let us take now medium 1 as free space, n1 = 1, and see the direc-
tion of the refracted wave (Fig. 2.23) when medium 2 has a large range
Fig. 2.23 Refraction at a boundary for
of refraction indices, which may be smaller than 1 and may take any various values of n2
negative value. The angle of refraction is 90◦ when n2 = sin θ1 (if n2 is
even smaller then total internal reflection occurs in medium 1). As n2
increases from this value below unity up to infinity the refracted angle ș1 ș1
declines from 90◦ to 0◦ . Note that the angle of refraction is the same
for n2 = −∞ as for n2 = ∞. Now, as n increases from minus infin-
ity to − sin θ1 the angle of refraction declines from 0◦ to −90◦ . If n2
is between − sin θ1 and 0 then there is again total internal reflection.
ș1 ș1
Clearly, negative n2 implies negative refraction. It may be worth reiter-
ating at this stage that we got negative refraction in two different ways.
In Section 2.10 from the properties of backward waves and in the present z
section from the concept of negative refractive index.
The most striking example of what we can do with a negative-index 0 d/2 3d/2 2d
material is Veselago’s flat lens. For n = −1 the angle of refraction is
equal to the negative angle of incidence hence all the rays emanating
Fig. 2.24 Negative refraction for n =
from a line source will be refocused inside the material and brought −1 (Veselago’s flat lens)
to another focus outside the material as shown in Fig. 2.24. The lens
thickness is d. A point source at a distance d/2 in front of the lens is
ȝr
reproduced at a distance d/2 behind the lens.
ENG DPS
2.11.2 Terminology
İr
Having realized that negative material constants may be interpreted
as having negative refractive index, and that the consequence is a left- DNG MNG
handed E, H, k relationship, Veselago called these materials left-handed.
In the many papers that followed Smith et al.’s (2000) rediscovery quite
a number accepted this terminology and referred to these materials as
Fig. 2.25 Terminology for materials
left-handed media (LHM) and to those with positive material constants with various signs of permittivity and
as right-handed media or (RHM). There is, however, a vocal minority permeability after Engheta et al. (2005)
unhappy with this description. They argue that left-handedness and
right-handedness had been terms widely used before in chiral materials
referring to the direction of chirality.
60 A bird’s-eye view of metamaterials
Fig. 2.29 Contours of lenses with internal foci. n = −0.2, −0.8, −1.2, −2, −4 (a)–(e)
d1 + nd2 = 0 . (2.76)
Then, the equation in polar co-ordinates reduces to
2
r1 r1
(n + 1) − 2n cos θ + (n − 1) = 0 . (2.77)
d1 d1
It may again be converted into normalized Cartesian co-ordinates yield-
ing the equation
2
n 1
x− + y2 = , (2.78)
n+1 (n + 1)2
which is clearly that of a circle centred at (n/(n + 1), 0). In the limit
of n → −1 it becomes a straight line, i.e. we obtain the flat lens as a
special case.
Next, let us find the contours for various values of n. Clearly, the
condition of zero optical path posed by eqn (2.76) can only be satisfied
if n < 0. Thus, the range of interest is for n between zero and −∞.
When n is small and negative the radius of the circle is close to unity,
but the internal focus is quite far away from the front surface because
negative index material
d2 = −d1 /n. The contour and the focusing mechanism are shown in
Fig. 2.29(a) for n = −0.2. The internal focus is there but the lens is
far too thick. As n tends towards −1 the lens flattens, as may be seen
in Fig. 2.29(b) for n = −0.8. At n = −1 the lens is flat. At this point
F2 F1 P the centre of the circle switches from minus infinity to plus infinity. As
n decreases below n = −1 the radius of the circle decreases, as may be
seen in Fig. 2.29(c) for n = −1.2. For even lower values, n = −2 and
−4 the full lens is shown in Figs. 2.29(d) and (e). It may be seen that
the lens gets smaller as n declines. As n tends towards minus infinity
conducting cylinder
the lens becomes a circle but its radius declines to zero.
An interesting variation on this theme is the reflector of Lagarkov
Fig. 2.30 A reflector that brings a par- and Kissel (2001) also in a 2D geometry. The reflector consists of a
allel beam to a focus. From Lagarkov layer of negative-index material upon a conducting cylinder as shown in
and Kissel (2001). Copyright
c 2001 Fig. 2.30. An incident parallel beam refracts in the negative direction
Springer Science + Business Media reaching the conducting (and therefore reflecting) cylinder at P from
2.11 Negative-index materials 63
which it is reflected, going through the focal point F1 inside the material
and after a further negative refraction on the air/negative-index material
boundary it comes to a second focus at F2 . A possible application of this
reflector is for a microwave antenna. Inverting the path of propagation
a line source positioned at F2 would produce a parallel beam.
We have looked in this section at another family of lenses made possi-
ble by the existence of negative-refractive-index materials. It gives new
options for designing lenses.
0
(roughly) 4.7 and 5.2 GHz. In the stop band the attenuation increases –10
from 2 dB to about 35 dB. When the rods are also included then the
–20
stop band turns into a pass band. The attenuation declines from 50 dB
to about 32 dB. Note that the attenuation is very high even in the pass –30
band because there is considerable power absorption by the rods. How- –40
ever, the stop band turning into a pass band proves that a material with –50
both material constants negative can propagate electromagnetic waves. 4.5 5 5.5 6 6.5 7
We shall present here another set of experimental results by Li et al. Frequency (GHz)
(2003). The negative-index material consisted of a similar combination
of SRRs and rods with the difference that there were two rods in the unit
Fig. 2.32 Transmission as a function
cell. The sample was placed at the focal point of a lens-compensated of frequency for SRRs (solid line) and
horn antenna with a similar antenna as the receiver. The experimental for SRRs + rods (dotted line). From
setup is shown in Fig. 2.33(a) and the results in Fig. 2.33(b). This time, Smith et al. (2000). Copyright
c 2000
by the American Physical Society
the S-parameters were measured (see Section 1.24 for their definition),
i.e. the reflection in the transmitter and the transmission in the receiver.
We find again that a pass band appears in the frequency range (from
64 A bird’s-eye view of metamaterials
1.0
B D
0.6
Experiment
M Simulation
0.4
W C
H L S
0.2
A
0
10 12 14 16 18
Frequency (GHz)
(a) (b)
Fig. 2.33 (a) A negative-index material between two lens-compensated horns. (b) Experimental results (solid lines) for
transmission (A) and reflection (B). Numerical simulations (dotted lines) for transmission (C) and reflection (D). From Li et al.
(2003). Copyright
c 2003 American Institute of Physics
Normalized power (linear scale)
1
Microwave absorber LHM
Teflon
Detector 0.8
0.6
0.4
T
0.2
Sample 0
–90 –60 –30 0 30 60 90
Fig. 2.34 (a) Experimental setup for measuring refraction, (b) Left-handed material realized by SRRs and rods, (c) measured
c 2001 AAAS
results. From Shelby et al. (2001a). Copyright
13.2 to 14.2 GHz) where both material parameters are negative. Curves
A and B show the measured values of S21 and S11 , whereas curves C
and D are derived from numerical simulation. The package used was
7
High Frequency Structure Simulation, HFSS,7 a frequency domain Maxwell’s equations solver. The sample
Ansoft Inc., Pittsburgh, 2002 was simulated by a unit cell with periodic boundary conditions. The
agreement may be seen to be extremely good.
The crucial experiment concerning negative refraction was done by
Shelby et al. (2001a). The sample made up by a combination of SRRs
and rods and both produced on printed circuit boards is shown in Fig.
2.34(a). The experimental setup with which they measured the angle
of refraction may be seen in Fig. 2.34(b). The arrays of SRRs and rods
were arranged in the shape of a prism. Microwaves at a frequency of
10.5 GHz were confined vertically by parallel metal plates and laterally
by absorbers. They were incident upon the back of the prism as shown
2.11 Negative-index materials 65
(a) (b)
Fig. 2.35 Numerical simulations of the refraction of a finite beam by a wedge, (a) εr = 2.2, µr = 1, (b) εr = −1, µr = −1.
From Kolinko and Smith (2003). Copyright
c 2003 Optical Society of America. For coloured version see plate section
by the thick black lines. After propagating through the prism the waves
were refracted at an angle of −61◦ corresponding to an effective refrac-
tive index of −2.7. As a control experiment the prism was replaced by
another prism of the same shape but made of Teflon, which has a refrac-
tive index of 1.4. It duly refracted in the positive direction with an angle
corresponding to n = 1.4. The measured beam profiles for Teflon (dot-
ted lines) and for the composite material (solid lines) are shown in Fig.
2.34(c) normalized to unity. The actual measured peaks are of course
quite different because the composite material has much higher losses.
We shall show here simulation of negative refraction by Kolinko and
Smith (2003). A finite beam is incident on a wedge-shaped material that
in the first case (Fig. 2.35(a)) has material parameters εr = 2.2, µr = 1,
and in the second case (Fig. 2.35(b)) εr = −1 and µr = −1. Similarly to
the experimental results shown in Fig. 2.34(c) there is positive refraction
in the first case and negative refraction in the second case. The numerical
package used was HFSS. The finite elements were tetrahedrons. For
convergence, as many as 100 000 elements were required.
perpendicular incidence
1 1 j kd j 1
= e = cos(nkd) + η+ sin(nkd) , (2.79)
T′ T 2 η
where η is the impedance of the medium. We also find that
R 1 1
=j η− sin(nkd) . (2.80)
T′ 2 η
√
p of ε and µ we have here the intermediate variables n = εµ and
Instead
η = µ/ε. The technique is then to express cos(nkd) with the aid of R
and T ′ and then find ε and µ from that. The expressions can be found
in the paper by Smith et al. (2000). The main problem is that cos(nkd)
is multivalued, which makes the problem a little unwieldy, but soluble.
The authors were capable of recovering both negative ε and µ, although
their meaning for an insufficient number of elements per wavelength is
controversial. For a discussion see Appendix C.
flat. Entirely flat. It is the same for every spatial frequency component.
Is that possible? Not really. A limit will be set, if by nothing else, then
by the period of the negative-index material (Haldane, 2002). If we can
make metamaterial elements of the size of 100 nm and if the distance
between them is also 100 nm then there would be a chance of making
a lens with a resolution approaching 100 nm. At the time of writing
it does not seem to be likely that such a lens could be made and such
resolution could be obtained, but it is possible in principle. Another
chance is obtained with a material in which only the dielectric constant
is negative. That will not yield a flat transfer function but it would be
flat enough for many purposes, and it would have the great advantage
that natural materials (e.g. silver) with that property exist. The period
in that material will be of the order of one tenth of a nanometer, thus,
at least on that account, the resolution could be extremely high. We
shall discuss the details of this mechanism in Chapter 5. In the present
section we shall show only a few representative examples.
Thus, the values of kz are identical in all three media. As a result, from
eqn (1.49)
ζe = −1. (2.83)
Substituting these values into eqns (1.158)–(1.160) we find
j kz d j kz d
− −
T1 = e 2 , T2 = e j kz d , T3 = e 2 , (2.84)
and, consequently,
T1 T2 T3 = 1 . (2.85)
The total transfer function is flat, but not only flat it is actually unity
for all values of the spatial frequency for which there is propagation.
The physics is quite simple. The amplitudes are always the same
because we neglected losses. The phase goes forward in medium 1 by
kz d/2, it goes backward in medium 2 by kz d (remember, it is a backward
wave) and it goes forward again in medium 3 by kz d/2. The total phase
change is zero. Thus, the transfer function is flat for all propagating
waves. This is remarkable but it is still far from the perfect lens. A
perfect lens should reproduce not only those spatial frequencies for which
there is propagation but also those for which the waves are evanescent.
This occurs when the spatial frequency is larger than k0 , i.e. the details
of the object are smaller than the free-space wavelength. This is when
we talk about subwavelength imaging.
The waves are evanescent in all three media when kx > k0 . Then kz1
and kz2 become imaginary (as in Section 1.10) and should be replaced
by −j κ1 and −j κ2 . Then, ζe takes the form
εr2 κ2
ζe = . (2.86)
εr2 κ1
We can then still use the condition ζe = −1, in which case eqn (2.84)
modifies to
κ1 d κ1 d
− −
T1 = e 2 , T 2 = e κ1 d , T3 = e 2 . (2.87)
Remarkably, once more,
z=0 T1 T2 T3 = 1 . (2.88)
The variation of Hy is plotted in Fig. 2.36. It may be seen that the
magnetic field is the same at z = 2d as at z = 0, and that applies to every
Fig. 2.36 Evolution of an evanescent
component single spatial frequency. The conclusion is that the transfer function is
unity whether the wave is propagating or evanescent. However large kx
is, a slab having the material constants εr2 = −1 and µr2 = −1 will
perfectly reproduce the object.
It is easy to understand the physics of phase cancellation. It is also
easy to understand (one might even say it’s trivial) that an evanescent
2.12 The perfect lens 69
5 7 9 11
z=0 6d
Fig. 2.37 Multilayer superlens
multilayer lens gives a much better image than a single-layer lens of the
same total thickness. This is of great practical significance, as will be
further discussed in Section 5.5.
z/O
z/O
í
z/O
(c)
Fig. 2.38 (a), (b) Poynting vector streamlines for a flat lens with d/λ = 500 taking only propagating components into account.
Object half-width is (a) one wavelength and (b) five wavelengths. (c) Evolution of the field across the lens with d/λ = 0.5 and
for various number of evanescent components included. Object half-width is 0.1 λ
RL
RL
LR
ka S ka S
Fig. 2.39 Circuit analogue of a left-handed medium (a). Its dispersion curve when ωRL 6= ωLR (b) and ωRL = ωLR (c)
transmission line.
The two models are of course identical, provided we change the nota-
tions to
CL = Cs , LL = Lsh , CR = Ct a , LR = L t a . (2.91)
The dispersion equation may be obtained with the aid of eqn (1.142) by
noting that
1 1
Z = j ωLR − and Y = j ωCR − , (2.92)
ωCL ωLL
yielding
ka (ω 2 − ωRL
2
)(ω 2 − ωLR
2
)
4 sin2 = −ZY = 2 2 , (2.93)
2 ω ωRR
where
1 1 1
ωRR = p ; ωRL = p ; ωLR = p . (2.94)
CR LR CR LL CL LR
The corresponding dispersion curve is plotted schematically in Fig.
2.39(b). There are two branches: the upper one is a forward wave and
the lower one is a backward wave. An interesting special case arises
when ωRL = ωLR , i.e. the series resonant circuit and the shunt resonant
circuit have the same resonant frequencies. In that case the upper and
lower curves have a common point at ka = 0, ω = ωRL (see Fig. 2.39(c))
where the group velocities are non-zero.
For some practical realizations of the equivalent circuits in the mi-
crowave region see Chapter 6.
Plasmon–polaritons
3
3.1 Introduction 3.1 Introduction 75
3.2 Bulk polaritons.
The Drude model 77
Plasmas is an old and respectable subject. It was started by Langmuir1
3.3 Surface plasmon–polaritons.
in the 1920s. Its properties have been studied ever since but not always
Semi-infinite case,
with the same vigour. Topics rose in response to new applications, and TM polarization 79
declined on reaching saturation or realizing that the chances of the envis- 3.4 Surface plasmon–polaritons
aged solution are fast receding. The development of radio broadcasting on a slab: TM polarization 92
led to the discovery of the Earth’s ionosphere that reflects radio waves 3.5 Metal–dielectric–metal
and is responsible for reception of radio signals when the transmitter and periodic structures 102
is over the horizon. Hannes Alfven’s theory of magnetohydrodynam- 3.6 One-dimensional
ics (1940) that treated plasma as a conducting fluid has been employed confinement: shells
and stripes 103
to investigate sunspots, solar flares, the solar wind, star formation and
3.7 SPP for arbitrary ε and µ 106
other topics in astrophysics. The interest in thermonuclear fusion came
in the wake of the hydrogen bomb. At one time it was believed that
fusion, assisted by plasmas, is round the corner. The belief turned out
to be not well founded. Later, the invention of lasers opened the new
chapter of laser plasma physics with some promise of fusion but that has 1
A transparent liquid that remains
not been realized either. High-energy physicists hope to use plasma ac- when blood is cleared of various corpus-
celeration techniques to dramatically reduce the size and cost of particle cles was named plasma (after the Greek
word that means ‘formed, jelly-like’) by
accelerators. Among the applications that have come off are microscopy
a Czech medical scientist Purkinje. The
and sensing (see, e.g., Yeatman 1996; Homola 2003; Barnes 2006). Nobel prize-winning American chemist
We have enumerated only a fraction of plasma phenomena that could Irving Langmuir first used this term to
be discussed. The subject is not only old and respectable: it is also very describe an ionized gas in 1927: Lang-
muir was reminded of the way blood
big. Our interest is limited to plasmas on surfaces, the kind that was plasma carries red and white corpus-
presented very briefly in Section 1.10. They are important for meta- cles by the way an electrified fluid car-
material applications that belong to two categories. One is their role ries electrons and ions. This analogy
in the subwavelength manipulation of images. If µr = 1 and εr = −1 is, however, slightly odd: blood plasma
is blood that is free from blood cor-
then the plasma resonances will limit the range of spatial frequencies puscles; electric plasma would lose its
for which the transfer function is approximately flat. This was already properties if we were to get rid of free
briefly discussed in Section 2.12.4. The second relationship with meta- charges in it.
materials is via elements that may exhibit plasma resonances at high
enough frequencies. They are mainly nanosized rods and rings and their
combinations, reminiscent of electric dipole and loop antennas operating
at radio frequencies.
There is, however, a third category that hardly exists at the moment.
Nanoscale geometry, for example, allows us to change plasma properties
and understanding of this mechanism would allow us to design metama-
terial elements with tailored properties. This is just one example that
may or may not turn out to be feasible but there are many others. We
76 Plasmon–polaritons
ωp2
εr = 1 − , (3.2)
ω2
known as the Drude model for a free-electron gas. In the presence of
losses the above expression modifies to
ωp2
εr = 1 − , (3.3)
ω (ω − jγp )
with the damping constant, γp = 1/τ , vanishing if the collision time,
τ , becomes infinite. In most metals, the plasma frequency is in the
ultraviolet,4 making them shiny in the visible range as all the incident 4
Taking, for instance, the electron den-
light is reflected. In doped semiconductors, the plasma frequency is sity in a typical metal as Ne = 6 × 1028
m−3 we can calculate that ωp = 2π ×
usually in the infra-red.
2.2×1015 s−1 , which corresponds to the
The asymptotic and limiting cases for eqn (3.2) are wavelength λp = 2πc/ωp = 136 nm,
which lies in the ultraviolet.
εr → −∞ as ω→0
εr (ωp ) = 0 . (3.4)
εr → 1 as ω→∞
What are the implications of eqn (3.2) for electromagnetic wave propaga-
tion in such a medium? An electromagnetic wave of the form
exp [j (ωt − k · r)] no longer propagates without dispersion. Its wave
number has to satisfy the condition
s s
2
ω√ ω ω p ωp2
k= εr µr = 1 − 2 = k0 1 − 2 , (3.5)
c c ω ω
resulting in a dispersion equation
3.5
3
light line
2.5
bulk Z/k=c
Z/Zp
2 plasmon−
polariton
1.5
1
plasmon
0.5 Z=Zp
2
ωL2 − ωT
2
ωT
εph = ε∞ + 2 (ε0 − ε∞ ) = ε∞ 1 + 2 , (3.8)
ωT − ω 2 ωT − ω 2
6
Note that ε0 in eqn (3.8) is not the with ωT p the so-called TO (transverse optical) phonon frequency and
free-space permittivity but the value of ωL = ωT ε0 /ε∞ the LO (longitudinal optical) phonon frequency (see,
the dielectric constant at ω = 0. e.g., Kittel 1963).6 The asymptotic and limiting cases here are
3.3 Surface plasmon–polaritons. Semi-infinite case, TM polarization 79
εph → ±∞ as ω → ωT
εph (ωL ) = 0 . (3.9)
εph → ε∞ as ω→∞
√
The dispersion equation is then k = ω εph µ. The frequency at which
k = 0 is known as the Reststrahl frequency (Fox, 2001). Note further
that in the frequency range ωT < ω < ωL the dielectric constant is
negative7 and the wave vector is imaginary, meaning a stop band for 7
An example is SiC, a polar material.
the bulk phonon–polariton. It may be instructive to note that formally It was used for subwavelength imaging
in the negative dielectric constant re-
the plasmon–polariton dispersion equation (3.5) is a special case of the gion (see Section 5.4.6) by Korobkin
phonon–polariton dispersion equation with ε∞ = 1, ωL = ωp and ωT = et al. 2006a; Korobkin et al. 2006b.
0.
which is another way of saying that for a surface mode to exist dielec-
tric constants in the two media must have different signs. The second
condition of eqn (3.10) follows from eqn (1.73), which is given below
r
εr1 εr2
kx = k0 . (3.12)
εr1 + εr2
It ensures that kx is real so that the mode can propagate along the
surface. It is a rather innocent-looking equation but it contains a large
amount of information about the properties of SPPs at a single inter-
face. Since εr2 is a function of frequency, eqn (3.12) is the dispersion
equation, i.e. it provides the relationship between the frequency and the
wave number. Inserting eqn (3.2) from the lossless Drude model into
eqn (3.12) we find
2 2
Z/k=c
bulk light line bulk
1.5 plasmon− Z/k=c 1.5 plasmon- 1/ 2
polariton polariton Z/k=c/Hr1
Z=Zp
Z/Zp Z=Zp
Z/Zp
1 1
Z=Z
s
Z=Z
s
0.5 surface 0.5
plasmon− surface
polariton plasmon-
polariton
0 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
k/kp
k/k
p
(a) (b)
Fig. 3.3 Dispersion of SPPs (solid line) at the boundary vacuum–metal (a) and at the boundary glass–metal (b). Also shown
is the dispersion curve of the bulk mode with ky = kz = 0 (thin solid line), the light line in vacuum and in the dielectric
(dashed lines), the dispersions of bulk (dotted line) and surface plasmons (dashed-dotted line)
−6
10
light
line
−7
10 −7 −6
Fig. 3.4 Dispersion curve for the 10 λp λ 10
s
metal–vacuum interface free space wavelength λ [m]
Z Z
1 s p
−2
permittivity H’, − H"
−H"
−4
H’
−6
−8
of hybrid wave, but call them surface plasma waves for large kx . Alas,
logic is not an influential factor when it comes to terminology. We shall
accept the majority view, accept the term surface plasmon–polariton,
and keep on referring to them as SPPs. This may also be the place
where we can mention the electrostatic limit, which means obtaining
the dispersion equation by solving Laplace’s equation instead of the full
apparatus of Maxwell’s equations. This is permissible for large kx . For
more details see Appendix F.
ωp2 ωp2 γp
ε′ = 1 − , ε′′ = . (3.22)
ω 2 + γp2 ω ω 2 + γp2
The corresponding values of kx′ and kx′′ may be determined from
eqn (3.12) by substituting into it the complex value of the dielectric
84 Plasmon–polaritons
0.7 0.7
0.6 0.6
Re(k ) Re(k )
x x
0.5 Im(k x) 0.5
Im(k x)
0.4 light line 0.4 light line
p
p
Z=Z Z/Z
Z/Z
s Z=Z
s
0.3 0.3
02 0.2
0.1 0.1
0 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
k /k kx/kp
x p
(a) (b)
Fig. 3.6 Dispersion for lossy metal. γp /ωp = 0.01 (a) and 0.1 (b)
constant. The real and imaginary parts of εr2 are plotted in Fig. 3.5 as
a function of frequency for γp = 0.01ωp.
The dispersion curves in the presence of losses may now be calculated
with the aid of eqn (3.12). They are plotted in Figs. 3.6(a) and (b). The
normalized horizontal co-ordinate is kx′ c/ωp = kx′ /kp for the propagation
coefficient (solid lines) and kx′′ c/ωp for the attenuation. For the lossless
case γp = 0 there is no attenuation in the pass band between 0 and
ωs . For γp /ωp = 0.01 the attenuation becomes significant just below
ωs but it declines sharply as the frequency decreases (Fig. 3.6(a)). The
same is true for γp /ωp = 0.1 but the effect is more substantial (see
Fig. 3.6(b)). All the information about propagation and attenuation is
properly contained in Figs. 3.6(a) and (b) but this is not the way most
people in the art of plasmas like to talk about attenuation. They prefer
to use the measure of propagation length, the distance the SPP travels
before its intensity is reduced by a factor of e. The relationship between
the propagation length and the attenuation coefficient is LSPP = 1/2kx′′
(the factor 2 is there because propagation length refers to intensity,
whereas we have used kx′′ for the attenuation of the wave amplitude).
For this reason we also plot the propagation length in Fig. 3.7 for the
same set of loss parameters. It may be seen that in the visible region
(with our choice of ωp ) it falls roughly between 0.2ωp and 0.4ωp. The
largest value of the propagation length for γp = 0.01 ωp is about 10 µm,
a very small length. On the other hand, it is still large relative to both
the free-space wavelength and the SPP wavelength. So there would be
about 20 wavelengths available for manipulation of information and that
might be enough for practical application.
3.3 Surface plasmon–polaritons. Semi-infinite case, TM polarization 85
−2
10
propagation length [m]
−4
10
−6
10 Jp/Zp= 0.01
Jp/Zp= 0.1
free O
OSPP
−8
10 −6 −5
10 10 Fig. 3.7 Propagation length. fp =
free space wavelength O [m] ωp /2π = 1200 THz
δair δair
−4 −4
10 δ 10 δ
metal metal
skin depth skin depth
free space λ free space λ
penetration [m]
λ
SPP penetration [m] λ
SPP
−6 −6
10 10
−8 −8
10 10
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
ω/ωp ω/ωp
(a) (b)
λp λ2
δm ≃ , δd ≃ . (3.27)
4π 4πλp
The penetration depth in the metal in the low-frequency limit is prac-
tically frequency-independent, whereas the penetration depth in the di-
electric is strongly frequency-dependent and can be quite large in the
infra-red (Abeles, 1986).
Barnes et al. (2003) in their review pointed out the importance for the
SPP-based nanocircuitry of these three characteristic length scales, the
propagation length, the penetration depth into the dielectric and into
the metal. The propagation length of the SPP mode, LSPP , is usually
determined by the loss in the metal. For a low-loss metal, for example,
silver, at a wavelength of 500 nm it is as large as 20 mm. The prop-
agation length sets the upper size limit for any photonic circuit based
on SPPs. The penetration depth into the dielectric material, δd , is typi-
cally of the order of one half the wavelength of light involved and dictates
the maximum height of any individual features, and thus components,
that might be used to control SPPs. The ratio of LSPP /δd thus gives
one a measure of the number of SPP-based components that may be
integrated together. The penetration depth into the metal, δm , deter-
mines the minimum feature size that can be used; this is between one
and two orders of magnitude smaller than the wavelength involved, thus
highlighting the need for good control of fabrication at the nanometer
scale.
Another important point is that the penetration depth into the metal
gives an idea of what thickness is required to allow coupling between the
modes on the two surfaces of a thin metallic film. As shown later in Fig.
3.17(b) for a certain set of parameters, and as will be discussed in much
3.3 Surface plasmon–polaritons. Semi-infinite case, TM polarization 87
1 1
10 10
penetration δ/λ
λ λSPP
SPP
δ δ
−1
air −1
air
10 10
skin depth skin depth
δmetal δmetal
−2 −2
10 10
−3 −3
10 10
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
ω/ωp ω/ωp
(a) (b)
Fig. 3.9 Penetration depth normalized to the free-space wavelength for (a) γp = 0, (b) γp = 0.1ωp
1 0
10
0.8
0.6 Ez1 −2
10
Ez2
0.4 Ex
Ez1
−4
Ez2
0.2 10
Ex
0 −6 −5 −4 −6 −5 −4
10 10 10 10 10 10
free space wavelength O [m] free space wavelength O [m]
(a) (b)
1
Ex
Ez
normalized eld
Hy
−1
−200 0 200 −100 0 100 −20 0 20
z [nm] z [nm] z [nm]
(a) (b) (c)
Fig. 3.11 Field amplitudes as a function of z for (a) ω/ωp = 0.25, (b) ω/ωp = 0.5, (c) ω/ωp = 0.7
s
ωp2 − ω 2
r
Ez1 −εr2
=j =j . (3.28)
Ex εr1 εr1 ω 2
−1 −0.5 0 0.5 1
250 100 20
z [nm]
z [nm]
z [nm]
0 0 0
Fig. 3.12 Magnetic-field (contour plot) and electric-field (field lines) distribution at the metal–air boundary for (a) ω/ωp = 0.25,
(b) ω/ωp = 0.5, (c) ω/ωp = 0.7
30 20
200 80 15
20
10
100 40
10
5
z [nm]
z [nm]
z [nm]
z [nm]
0 0 0 0
−5
−10
−100 −40
−10
−20
−80 −15
−200
−30 −20
Fig. 3.13 Poynting vector in the vicinity of the metal–air interface for (a) ω/ωp = 0.25, (b) ω/ωp = 0.5, (c) ω/ωp = 0.685, (d)
ω/ωp = 0.7. Lossless case
kx kx
Sx1 = |B|2 e 2κ1 z and Sx2 = |B|2 e −2κ2 z , (3.30)
2ωεr1 2ωεr2
3.3 Surface plasmon–polaritons. Semi-infinite case, TM polarization 91
30 20
200 80 15
20
10
100 40
10
5
z [nm]
z [nm]
z [nm]
z [nm]
0 0 0 0
−5
−10
−100 −40
−10
−20
−80 −15
−200
−30 −20
Fig. 3.14 Poynting vector in the vicinity of the metal–air interface for (a) ω/ωp = 0.25, (b) ω/ωp = 0.5, (c) ω/ωp = 0.685, (d)
ω/ωp = 0.7. Lossy case
where Sx1 and Sx2 are the x components of the Poynting vector in the
dielectric and in the metal, respectively. Taking εr1 = 1 the distribution
of the Poynting vector is shown in Figs. 3.13(a)–(d) both in air and in
the metal for ω/ωp = 0.25, 0.5, 0.685 and 0.7. It is interesting to note
that practically all the power is flowing in air at ω/ωp = 0.25. The
corresponding wave vector is close to the light line. As ω/ωp increases
some of the power flows in the metal, oddly enough in the opposite
direction. In the vicinity of the asymptote at ω/ωp = 0.7 the flow of
power flowing in the positive and negative directions is nearly the same.
There is very little net power flowing. If we want to find the net power of
the SPP moving in the positive x direction we need to integrate Sx1 for
z from −∞ to zero, and Sx2 from z = 0 to z = ∞. The integration can
be easily performed. The final result after some algebra can be found as
kx ε2r2 − ε2r1
Net power = |B|2 . (3.31)
4ωκ1 ε2r1 ε2r2
When εr1 = |εr2 | there is no net power flow at all. Otherwise, the net
power is in the forward direction. We have a forward wave as follows
anyway from the dispersion curve.
Why is there no Poynting vector component in the z direction? Both
Ex and Hy are non-zero hence the normal component of the Poynting
vector
1
Re Ex Hy∗
Sz = (3.32)
2
should exist. It does not for the reason that Ex and Hy are 90 degrees
out of phase, and their product has zero real part.12 However, in the 12
The same picture is known to occur
presence of losses (γp = 0.01ωp) Sz must be finite. The Poynting vector in the case of total internal reflection.
92 Plasmon–polaritons
−1 −0.5 0 0.5 1
500 200 40
z [nm]
z [nm]
z [nm]
0 0 0
Fig. 3.15 Magnetic-field (contour plot) and electric-field (streamlines) distribution at the metal–air boundary. Lossy case
must cross the boundary between the dielectric and the metal in order to
replenish the absorption occurring in the metal. This is shown in Figs.
3.14(a)–(d) for the same set of frequencies. An obvious consequence of
losses is that the wave attenuates as it propagates in the x direction.
Figures 3.15(a) and (b) show again (as in Fig. 3.12) the contour plots of
Hy in the xz plane, and the field lines of the electric field. The strong
decline of the magnetic field at ω = 0.7ωp may be clearly seen. There
is a bright spot between x = −80 nm and −40 nm but one period away
the intensity has so much declined that the maximum is hardly visible.
Note also that the field lines, just as the streamlines of the Poynting
vector, are slanted.
2
1 10
kp d=0.25
0.8
0.6 1
Hr =-1
kx /k p
10
Z/Z p
0.4
kp d=0.5
kp d=1
02
0 Hr =-2
10
0
0 2 4 6 8 10
0 50 100 150
kx /k p
d [nm]
(a) (b)
Fig. 3.17 (a) Dispersion curves for thin slabs for kp d = 0.25, 0.5 and 1. (b) kx versus thickness for two different values of ω
corresponding to εr = −1 and −2
(1 + ζe ) = ± e −κ2 d (1 − ζe ) , (3.34)
representing two modes guided by the metallic slab (in which the wave,
now aware of both surfaces, travels with one single velocity), one sym-
metric and one antisymmetric with respect to their field distributions.
As far as the dispersion curve is concerned it may be expected that the
unperturbed dispersion curve (i.e. the one for the single interface) will
split. The two branches of the dispersion equation given by eqn (3.34)
can be expressed in more concise form as
κ2 d
ζe = − tanh , (3.35)
2
and
κ2 d
ζe = − coth . (3.36)
2
These transcendental equations can be easily solved numerically. As
shown above there are two solutions, first reported by Oliner and Tamir
(1962). We shall denote the upper branch of the solution by ω (+) and
94 Plasmon–polaritons
0.8
0.6
Z/Z p
0.4
0.2
0
Fig. 3.18 Electrostatic approximation 0 2 4 6 8 10
kx/k p
and full solution for a slab. kp d = 0.25
the lower branch by ω (−) . In Fig. 3.17(a) we plot the upper and lower
branches for εr1 = 1 and for three different thicknesses (kp d = 0.25, 0.5
and 1) together with the solution for the single interface (d → ∞) that
is plotted with the dotted line. The meaning of the adjectives ‘upper’
and ‘lower’ make sense because they are above and below the dotted
line. For large kx both curves tend towards ωs , the same limit as for the
single interface. For small kx both types of dispersion curves tend to
the light line. We also find, not unexpectedly, that the split between the
modes becomes smaller as the slab becomes thicker. The propagation
coefficient is plotted in Fig. 3.17(b) as a function of slab thickness for
two different values of ω corresponding to εr = −1 and −2. It may be
seen that for a large enough thickness only one mode may exist, or even
14
It makes good sense that the surface none. The two surfaces are then uncoupled.14
modes that exist on a single interface at It is always nice to find analytical approximations. They give a sense
a given ω will split when the two sur-
faces are coupled. It is, however, less
of security. We know what is going on. The first one presented is
obvious that in the presence of coupling the electrostatic approximation that we have already come across in
the so far prohibited territory between Section 2.12.4. It is described in Appendix F. For the present case we
ωs and ωp may also be inhabited by find the two branches of the dispersion equation in the form
a new mode, the upper branch of the
dispersion curve. Physical considera-
ωp2
1 ± e −kx d .
tions would suggest that such a mode is
ω2 = (3.37)
permissible because the basic condition 2
εr2 < 0, is still valid, allowing charges
to accumulate at the surface. For thin enough slabs the electrostatic approximation turns out to be
extremely good for the lower branch, as may be seen in Fig. 3.18. It is
good too for the upper branch until it reaches the light line but then
it fails miserably. It cannot possibly follow the light line. This is of
course not surprising. When the wave velocity is close to the velocity of
light then nobody would expect an electrostatic approximation to work.
Interestingly, the approximation for the lower branch is still quite good
in the vicinity of the light line, although for sufficiently low values of kx
the approximation is bound to fail. How the approximation deteriorates
for the lower branch as well as for low values of kx may be seen in Figs.
3.19(a)–(c) for three different thicknesses. The approximate curve is
denoted by dashed lines. It may be seen that firstly the approximation
crosses the light line and secondly that the approximation becomes worse
as the thickness increases.
3.4 Surface plasmon–polaritons on a slab: TM polarization 95
0.5
Z(−) (electrostatic) light
0.4 light line
light
line Z(−) (electrostatic) line
0.3 Z(−) (electrostatic)
Z/Z p
(−)
Z
0.2 Z(−)
Z(−)
0.1
0
0 0.2 0.4 0.6 0.8 10 0.2 0.4 0.6 0.8 1 0 02 0.4 0.6 0.8 1
kx/k p kx/k p kx/k p
Fig. 3.19 (a) Electrostatic approximation and full solution for a slab. kp d = 0.25, 0.5, 1 (a)–(c)
ω2 ω4
kx2 = + . (3.42)
c2 c2 ωp2
Note that eqn (3.41) would further simplify when kp d/2 ≪ 1 to
ω2 4 ω4
kx2 = 2 + 2 4 . (3.43)
c d ωp
It can be immediately seen that the smaller d is the farther is the dis-
persion equation from the light line for a given ω. Next, we shall look
for an approximation for the upper branch. We shall assume again that
96 Plasmon–polaritons
)
Z)
Z
w
w
(lo
(lo
1
)
Z (+
Z | d
Z(+)
ne
0.8
li
Z|
ht
d
lig
0.6
Z)
Z/Z p
() (low
Z Z( )
0.4
0.2
0
Fig. 3.20 Full solution and various ap-
0 0.5 1 1.5 2
proximations for a slab. kp d = 0.5
kx/k p
kx is small and close to the light line so that we can start again with
eqn (3.38). Then, using the same technique we find
ω kp d
δ= tanh , (3.44)
ωp 2
leading to
" #
ω2
kp d
kx2 = k02 1 + 2 tanh2 , (3.45)
ωp 2
or after expanding the tanh function to
ω2 ω 4 d2
+ kx2 =. (3.46)
c2 4c4
It may now be seen that the smaller d is the nearer is this branch of the
dispersion equation to the light line.
Using the above-derived approximations for ω (−) and ω (+) we plot
in Fig. 3.20 the approximate and exact curves for kp d = 0.5. The ap-
proximation is very good for small kx up to about ω/ωp = 0.35 and
deteriorates afterwards.
So far we have looked at the lossless case but of course attenuation is
equally important if we have practical applications in mind. We need
to find not only the ω versus kx′ curve but also ω versus kx′′ . How will
kx′′ enter our equations? It will be through κ2 , which in the presence of
losses will take the form
Z(+)
Z(−)
0.6
k d=0.25
p
0.4 k d=0.5
p
p
Z/Z
kpd kpd
kpd=1 kpd=1
0.2
k d=0.5
p
kpd=0.25
−5 0 −5 0
10 10 10 10
kx"/kp k "/k
x p
(a) (b)
Fig. 3.21 Attenuation coefficient versus frequency for ω (+) -branch (a) and ω (−) -branch (b)
0
10 d=10nm 20nm
60nm
propagation length [m]
−2 100nm
10
d
−4
10
free-space O f
−6
10
Fig. 3.22 Propagation length versus
0 2 4 6 8 10
O f [Pm]
free-space wavelength
Ez Ez
Ez Ez
Hy Hy Hy
Hy
Sx Sx Sx
Sx
(a) (b)
Ez Ez
Ez Ez
Hy Hy
Hy Hy
Sx Sx
Sx Sx
(c) (d)
Fig. 3.23 Field distributions of ω (+) and ω (−) -branches for four typical frequencies (a)–(d) ω/ωp = 0.3, 0.5, 0.7, 0.735
deals with the symmetries of the charge distribution and therefore with 16
Maier et al. (2005) use a notation
the longitudinal components of the electric field. By now it has been somewhat at variance with the usual
largely accepted16 that the mode in which the normal component of the notions of symmetry and antisymme-
electric field is symmetric is denoted by s and that in which the normal try. They refer to the ω (+) -mode
as ‘antisymmetric (sb ) mode’ and to
component is antisymmetric is called a. In addition, when the waves are the ω (−) -mode as the ‘symmetric (ab )
bound the modes are referred to as sb and ab and when they are leaky17 mode’.
as sl and al .
If the metal is lossless the modes are lossless. If the metal is lossy 17
Leaky waves, as the name implies,
the modes are lossy, so much is clear. But which one is lossier? The leak out power, which is good for anten-
calculations show that the sb mode is less lossy, meriting the description nas but not for guiding the wave. Here,
we are concerned only with bound
of ‘long-range surface plasmon’. We shall show curves later but first let modes so, with regret, we shall not dis-
us try to give a physical explanation. cuss the properties of leaky waves.
One explanation may be based on the fact that the ω (+) -mode is nearer
to the light line. If it is nearer it resembles more an electromagnetic wave.
Electromagnetic waves travel in vacuum with no attenuation, hence the
upper branch will be the one that has the lower attenuation.
100 Plasmon–polaritons
H3
3.4.3 Asymmetric structures
Dielectric
We use here the terms symmetric and asymmetric in yet another sense,
H2 Metal d referring this time to structures. A symmetric structure is when the
H1 Dielectric dielectrics on the two sides of the metal slab are identical and it is
asymmetric when the dielectrics are different (see Fig. 3.24).
The asymmetric structure was investigated in some of the early papers
Fig. 3.24 Metallic slab surrounded by (Sarid, 1981; Wendler and Haupt, 1986; Burke et al., 1986; Yang et al.,
different dielectric media 1991) but received much less attention than the symmetric case. If there
3.4 Surface plasmon–polaritons on a slab: TM polarization 101
103 Ag
O=632.8 nm a=10 nm
17 nm
30 nm
102 50 nm
Propagation length L (cm)
101
100
10–1
10–2
Fig. 3.25 Propagation length of the
long-range SPP vs. dielectric constants
εr1 of the superstrate and εr3 of the
1.9 2.0 2.1 2.2 2.3 2.4 substrate for different thicknesses a of
the Ag layer. From Wendler and Haupt
H r1 H r3 = 2.1211) H r1 H r3 H r3 H r1 = 2.1211) (1986). Copyright
c 1986 American
Dielectric constants Institute of Physics
is only a small difference in the dielectric constant then we can still talk
about quasi-symmetric and quasi-antisymmetric modes but we find that
the field confinement is very asymmetric. The quasi-symmetric mode is
localized at the boundary between the metal and the dielectric with the
lower dielectric constant, whereas the quasi-antisymmetric mode is lo-
calized at the higher dielectric constant boundary. The quasi-symmetric
mode inherits the mantle of the sb mode: it is still the long-range mode,
it still has a lower attenuation and, as may be expected, the attenua-
tion decreases as the slab thickness decreases. The new phenomenon is
that the quasi-symmetric mode has a cutoff thickness below which no
propagation is possible. This is bound to occur (Berini, 2001) because
the sb mode no longer has the chance quietly to convert into the TEM
electromagnetic wave as d → 0 because the different dielectrics on the
two sides do not permit it. The quasi-antisymmetric mode remains the
short-range mode, its attenuation increases with decreasing slab thick-
ness, it has not got a cutoff thickness.
For detailed calculations of the propagation length in asymmetric
structures see Wendler and Haupt (1986). They investigate theoreti-
cally thin silver slabs at the He-Ne wavelength of 632.8 nm with a di-
electric constant of εr2 = −18−j 0.47 obtained from Johnson and Christy
(1972). They first assume that both dielectrics have identical dielectric
constants of εr1 = εr3 = 2.1211, then change the dielectric constant of
one of them within a range of about 20%. They find that the propaga-
tion length first remains unchanged and then suddenly increases. Their
102 Plasmon–polaritons
2 " 2 # " ω 2 k d 2
#
2 ω ω p
4
p p
kx2 = 1− − − . (3.50)
d ωp ω ω 2
Z)
(low−
1 1
Z (−)
Z (+) (e
l.-stat.) Z (+) (e
l.-stat.)
e
e
0.8 0.8
light lin
light lin
w−Z) Z (+)
Z |d at .)
l.-st Z (+) (lo
− Z)
(−) (e
Z Z(−)
/
/
Z (−)
0.4 0.4
0.2 02
0
0 2 4 6 8 10 0 2 4 6 8 10
kx/kp kx/kp
(a) (b)
Fig. 3.26 Dispersion curves for a metal–dielectric–metal structure. Full solutions and approximation for (a) kp d = 0.25 and
(b) kp d = 0.5
2
2 k
ck
C
2
p II a
2
Metal
Dielectric p
I II b
p/2
Cmax
IV
Cmin
III
O kp k
(a) (b)
Fig. 3.27 Multilayer metal–dielectric structure. (a) Sketch of the configuration. (b) Dispersion relations. From Economou
c 1969 by the American Physical Society
(1969). Copyright
z
H x
t H y
w
(a)
0
10
3.6
sa0b , aa0b
10 –1
3.2
aa 2b , sa1b , a b
EE0
DE 0
ss0b sb
2.4 aa 2b sa1b a
b
10 –3
sb
2 1 as2b sb
ss0b as 0b ssb as0b ss1b asb2
10 –4
0 0.04 0.08 0.12 0.16 0.2 0 0.04 0.08 0.12 0.16 02
Thickness t (Pm) Thickness t (Pm)
(b) (c)
Fig. 3.28 Metallic stripe. (a) Sketch of the geometry. (b), (c) Dispersion with thickness of the first six modes. Also shown the
ab and sb modes for w → ∞ for comparision. From Berini (2000b). Copyright
c 2000 by the American Physical Society
which they cannot propagate and some of the modes also have a cutoff
thickness.
For a stripe width of w = 0.5 µm the normalized propagation and
attenuation coefficients are plotted in Figs. 3.28(b) and (c) for the first
six modes as a function of stripe thickness. The sb and ab modes of
the 2D structure are also shown for comparison. As may be expected,
the antisymmetric modes have higher attenuation than the symmetric
modes. The most relevant conclusion is that the 1D stripe geometry
is in no way worse than the amply analyzed 2D geometry. In fact, it
may be much better, at least in theory. The attenuation of two of the
higher modes may be seen to decrease rapidly before they reach cutoff.
Presumably, it happens at the expense of confinement. The ss0b mode
has probably more practical significance. If we compare it with the sb
mode at the lowest thicknesses shown, below 20 nm, we find that the
attenuation of the stripe mode is very considerably below that of the 2D
planar mode, and the attenuation can be further reduced by decreasing
the stripe width.
Stripes having different dielectrics for substrates and superstrates have
also been investigated by Berini (2001). Small differences in the upper
and lower dielectric constants have been found to lead to larger differ-
ences in propagation properties. There has been no reply given as yet to
106 Plasmon–polaritons
the question posed by the very large propagation lengths (see Fig. 3.25)
obtained theoretically by Wendler and Haupt (1986), whether they can
be realized in stripe form.
ω < ωp . (3.56)
Having done the derivation for the case when the permittivity could
take negative values it is an easy task to derive the dispersion equations
for the surface waves in the general case. The approach is the same.
The results are though more far reaching and will have significant impli-
cations for the imaging mechanism of the ‘perfect’ lens to be discussed
in Chapter 5.
We disregard losses for simplicity. They can be included later if
needed. Looking for surface-wave solutions declining exponentially we
require the z components of the k vector to be purely imaginary so that
both
q
κ1 = kx2 − εr1 µr1 k02 (3.57)
and
q
κ2 = kx2 − εr2 µr2 k02 (3.58)
are to be positive and real. The dispersion equation is formally the same
as in the case of a metal. It may be obtained from the poles of the field
solutions. For the TM polarization it is
(kx2 − εr2 µr2 k02 )ε2r1 = (kx2 − εr1 µr1 k02 )ε2r2 (TM case) , (3.63)
4
TM
2
µ=1
1
µ=µ /µ
2
0
TE
−2
−4 TE
Table 3.1 Numerical values for the frequencies at which surface modes start or stop.
F = 0.56
4 ε
µ
εµ=1
2
µ=0
ε, µ
0
ω
0
ε=0
−2 µ=−1
ε=−1
−4
Fig. 3.30 Frequency dependence of
0.2 0.6 1 1.4 1.8
ω/ωp
permittivity and permeability for F =
0.56, ω0 = 0.4 ωp
ωp2
,
εr2 = 1 − (3.82)
ω2
and that the permeability has a resonant behaviour that can be described
as (see Section 2.8)
F ω2
µr2 = 1 − . (3.83)
ω 2 − ω02
These forms of εr2 and µr2 were frequently used in describing the re-
sponse of metamaterials comprising rods (Pendry et al., 1996) and SRRs
(Pendry et al., 1999). The choice of the parameters, of the plasma fre-
quency, ωp , of the magnetic resonance frequency ω0 and of the filling
factor, F , has a crucial effect on the behaviour of the metamaterial, in
112 Plasmon–polaritons
1.6
6
ne
1.4
li
ht
lig
12 4
(i) TM
1 H=0
2
p
1
Z/Z
P=P /P
0.8 (iii) TM (iii) TM
2
H=−1 Q
C
0.6 P=0 0
HP=1 (ii) TE
P=−1
1 (ii) TE B
0.4 Z A
(i) TM 0
−2
02
0 P
0 0.5 1 1.5 2 2.5 3 −4
−8 −6 −4 −2 0 2
kx/kp H=H2/H1
(a ) (b )
Fig. 3.31 SPP dispersion for single interface. F = 0.56, ω0 = 0.4 ωp . (a) ω–kx diagram. (b) µ–ε diagram
and
We will now give examples of four quite different scenarios for SPP
modes using typical sets of parameters. In all examples we keep F =
0.56 and choose for ω0 /ωp values of 0.4, 0.6, 0.8 and 1. In all four
cases,
√ the frequency at which εr = −1 is the surface plasmon frequency,
ωp / 2 ≃ 0.71ωp and the frequency at which εr = 0 is of course the
plasma frequency, ωp . Table 3.1 gives numerical values for the frequen-
cies of interest from eqn (3.85) in the four cases.
3.7 SPP for arbitrary ε and µ 113
1.6 6
ne
1.4
li
ht
lig
4
1.2
(i) TM
1 H=0
P=0 2
1
Z/Zp
P=P /P
0.8
2
H=P=−1, HP=1 Q
0.6 Z 0
0
(i) TM
0.4 A
−2
0.2
P
0 −4
0 0.5 1 1.5 2 2.5 3 −8 −6 −4 −2 0 2
k /k
x p H=H /H
2 1
(a ) (b )
Fig. 3.32 SPP dispersion for single interface. F = 0.56, ω0 = 0.6 ωp . (a) ω–kx diagram. (b) µ–ε diagram
Case (a) ω0 /ωp = 0.4. The variations of ε(ω) and µ(ω) as a function of
frequency are shown in Fig. 3.30. The question we are asking now is what
kind of surface modes may exist for this set of parameters. For simplicity
we shall take εr1 = µr1 = 1, although the equations are valid for the case
of a general lossless dielectric. It can be shown from eqns (3.76)–(3.79)
that there are three SPP modes, as may be seen in Fig. 3.31(a). The
first one is a TM mode that exists in the range 0 < ω < ω0 . Thus, the
presence of the magnetic resonance limits the range of the TM mode.
Instead of moving up to ωs it stops now at ω0 . In other words, the upper
frequency limit occurs where µr2 turns negative instead of εr2 = −1. The
second mode is a TE one that starts at the light line at a value of ω
where εr2 µr2 = 1 and tends asymptotically to the value of ω where
µr2 = −1. It is a backward wave. The third mode is again a TM one.
It starts at the same point as the TE mode but it moves upwards to
tend asymptotically to the εr2 = −1 line. Not shown in the figure is the
bulk mode that propagates for the range of frequencies for which both
εr2 and µr2 are negative.
Let us now return to our diagram of Fig. 3.29 showing the areas in
the εµ plane in which surface modes are possible. Each of our three
modes may be represented by a curve in this plane replotted in Fig.
3.31(b). For the lower TM mode µ is positive and ε is negative. As ω
varies from 0 to ω0 the curve denoted by (i) is described. As ω tends to
ω0 we know that the permeability tends to infinity and then suddenly
changes to minus infinity. Up to µ = ∞ there is a TM mode, but when
the permeability switches to −∞ the surface mode is no longer there.
As the frequency increases further the permeability now increases from
minus infinity. At a certain value of the frequency it reaches the point
P and then traverses a number of regions between P and Q. In the PA
region, as may be seen from the shading, there cannot be surface modes.
114 Plasmon–polaritons
1.6 6
ne
1.4
li
ht
lig
P=0 4
1.2
(i) TM
1 H=0
P=−1 2
1
HP=1 (iii) TE
Z/Zp
P=P /P
0.8 Z0
2
(ii) TM Q
H= − 1
0.6 0
(i) TM
B
0.4
−2
0.2 (ii) TM A (iii) TE
0 P
−4
0 0.5 1 1.5 2 2.5 3 −8 −6 −4 −2 0 2
k /k
x p H=H /H
2 1
(a ) (b )
Fig. 3.33 SPP dispersion for single interface. F = 0.56, ω0 = 0.8 ωp . (a) ω–kx diagram. (b) µ − ε diagram
The TE region is entered at A and this mode exists until the frequency
is high enough to reach B. This is the mode we have denoted by (ii).
Increase ω again and we move from B to C, which is in the TM mode
region corresponding to mode (iii). The curve continues beyond C as ω
increases but no further surface modes are possible. It is interesting to
note that this novel presentation in terms of geometrical loci in the ε,µ
plane is not only capable of accounting for all the modes but also makes
it clear how the various surface modes arise.
Next, let us consider what happens when ω0 increases, i.e. the mag-
netic resonance moves up in frequency. It may be shown that both the
TE mode and the upper TM mode we have seen in Fig. 3.31(b) become
flatter and then disappear altogether. For case (b), when ω0 /ωp = 0.6
is taken, only the lower TM mode survives, as shown in the dispersion
curve of Fig. 3.32(a). The geometrical locus of this TM mode in the
εµ plane resembles the one in Fig. 3.31(b), but has shifted to the right,
as illustrated in Fig. 3.32(b). What happened to the other two modes?
Could we get our answer from the geometrical loci? What happened is
that the PQ curve also moved to the right. At this particular value of
ω0 /ωp = 0.6 it crosses the ε = −1, µ = −1 point at A without ever
crossing the area in which either surface mode exists.
When ω0 increases further there is an inversion of the upper TE and
TM modes. First comes a backward TM mode and further up (although
starting again at the same point) a forward TE mode. These changes
are indicated by the next set of curves taken for ω0 /ωp = 0.8 (Fig.
3.33). Note that ω0 > ωs , hence the magnetic resonance does not affect
the lower TM mode. Its range is, as in the case when µ is frequency-
independent, between ω = 0 and ω = ωs . The dispersion curves of the
three modes are plotted in Fig. 3.33(a) and the corresponding curves in
the ε,µ plane in Fig. 3.33(b). It may be clearly seen that in the interval
3.7 SPP for arbitrary ε and µ 115
1.6 6
P=0
ne
1.4
li
ht
lig
(iii) TE 4
1.2 P=−1
H=0 (i) TM
1 Z
0
2
1
Z/Zp
P=P /P
0.8
2
Q
H=−1
(i) TM 0
0.6
A
0.4
−2
0.2 (iii) TE
0 P
−4
0 0.5 1 1.5 2 2.5 3 −8 −6 −4 −2 0 2
kx/kp H=H /H
2 1
(a ) (b )
Fig. 3.34 SPP dispersion for single interface. F = 0.56, ω0 = ωp . (a) ω–kx diagram. (b) µ–ε diagram
1 H
line
(+)
(iii) TM Z
light
0.8
H
(−)
(+) Z
0.6 (ii) TE Z
p
ZZ
(a)
P
Z(−)
0.4 Z0
(+)
Z
(−)
(i) TM Z
02
0
0 1 2 3 4 5
kx/kp
1 H
line
light
0.8 (+)
(iii) TM Z
H
Z(−)
0.6
(+) (b)
ZZp
(ii) TE Z
(−) P
Z
0.4 Z0
Z(+) (i) TM Z(−)
0.2
1 H
light line
(+)
0.8 (iii) TM Z
(ii) TE
Z(+)
H P
Z( ) (−)
Z Z0
0.6
ZZp
(+) (a)
Z
0.4 (i) TM Z
(−)
02
0
0 2 4 6 8 10
kx/kp
1 H
light line
(ii) TE (iii) TM
}
Z(+) (b)
0.4 Z(−)
}
(i) TM
02
kp d = 0.25 and 1.0 are shown in Figs. 3.35(a) and (b) and Figs. 3.36(a)
and (b). Comparing them to the unperturbed solutions for the single
interface shows some striking features in case (b). Apart from the low-ω
TM, which splits into two, two further modes, a backward and a forward
one, appear that asymptotically approach the miraculous frequency at
which εr = µr = −1. The higher kx is, the closer are the two modes (one
above and one below) to the single interface curve. Due to the different
character of the frequency dependence of ε and µ, the split between the
TE modes looks different from the split between the TM modes (it is
actually larger for TM). We will return to this case later when discussing
subwavelength imaging properties of the perfect lens.
The asymmetric structure when the two media surrounding the slab
are different were first considered by Ruppin (2001) and later generalized
by Tsakmakidis et al. (2006).
We finish this section by mentioning an idea due to Oulton et al.
(2008). The authors note that, due to losses and stringent fabrication
requirements, practical SPP waveguides have not succeeded in produc-
ing field confinement beyond that of dielectric waveguides. The new idea
118 Plasmon–polaritons
SEGMENT CAVITY
l
C
1
C ZL RB
S [
D
D0 Zn Z1
(a) (b)
Fig. 4.3 Shielded symmetrical slotted-tube line (a). Matching and tuning of the slotted tube resonator (b). From Schneider
and Dullenkopf (1977). Copyright
c 1977 American Institute of Physics
contributions from other parts of the resonator as well as follows from (a) (b)
the field pattern of Fig. 4.2(b). The inductance is due to the horizontal
current flowing in the cylinder.
Our first two examples were on microwave tubes. Interestingly, the
need for small resonators also arose in the detection of nuclear mag-
netic resonance as the resonant frequency moved towards the hundreds
of MHz region. A resonator with bouncing waves would have been too
large, whereas the traditionally used solenoid plus lumped capacitor no Fig. 4.4 Loop gap resonators
longer worked satisfactorily. The solution proposed by Schneider and
Dullenkopf (1977) is shown schematically in Fig. 4.3(a). It is a cylin-
drical structure called the slotted-tube resonator. As may be seen, it
bears a strong resemblance to the magnetron cavity. The main differ-
ence is that the inner tube is split so that the upper and lower parts of
the inner tube can serve as a transmission line similarly to a strip line.
The outer cylinder provides a shield. Combining these field concepts
with circuit considerations the authors used a matching capacitance C1
122 Small resonators
r0 t
w
l
(a) (b)
Fig. 4.5 Split-ring resonator. (a) the simplest version and (b) the double-ring version. From Hardy and Whitehead (1981).
c 1981 American Institute of Physics
Copyright
where t is the gap width. They found the above expression to be accurate
within about 20%. For an extension of the frequency range of the split-
ring resonator to 4 GHz see Momo et al. (1983).
The aims of Kostin and Shevchenko 1993; Kostin and Shevchenko 1994
were somewhat different. They wanted to realize an artificial medium (a) (b)
using metallic elements that may yield a high value of the imaginary
part of the permeability. They wanted high magnetic losses and at the
same time they wished to minimize the amount of metal, which made
them use thin films. Their first design was a non-resonant metallic loop.
Later ones were resonant solutions combining in slightly different ways
(c) (d)
capacitances and inductances, as shown in Figs. 4.6(a)–(c).
An element combining an electric and a magnetic dipole was proposed
by Saadoun and Engheta (1992). The corresponding media were called Fig. 4.7 Ω-particles. From Engheta
by the authors omega or pseudochiral media. The reason for the first et al. (2002)
designation is obvious, the elements are shaped in the form of the Greek
letter, Ω, as may be seen in Fig. 4.7(a) (some related elements are shown
in Figs. 4.7(b)–(d)). The second description is due to the fact that
although the element possesses no handedness, it couples to each other
electric and magnetic fields with a phase difference of 90 degrees. The
general relationship is formulated3 by the authors as 3
This is a somewhat different form of
the tensorial relationships given in Sec-
1 tions 1.15 and 1.16.
D = εE + Ωem B and H= B + Ωme E , (4.4)
µ
where ε, µ, Ωem and Ωme are the permittivity, permeability and coupling
tensors, respectively. ε and µ are diagonal tensors, whereas
Fig. 4.8 Effective permittivity and permeability of an Ω-particle medium. From Saadoun and Engheta (1994)
z z
P a y a y
(2004). It is a rather curious fact that the omega particle, in spite of its
obvious suitability as a metamaterial element, has been eclipsed by the
SRR plus rod combination.
A chiral element with actual handedness was analyzed by Bahr and
Clausing (1994). The one-turn helix (Figs. 4.9(a) and (b)) was either
right-handed or left-handed. A further chiral element similar to that of
z0 Saadoun and Engheta (1994) was investigated by Tretyakov et al. (1996).
The analysis was based on the available theory of both linear and loop
antennas (King, 1969). Their basic element is shown in Fig. 4.10. It
y0 consists again of a combination of an electric and a magnetic dipole but
A this time the perpendicular to the loop is in the direction of the electric
a f
A1
x0 dipole. The excitation of dipoles (electric and magnetic) in terms of fields
l (electric and magnetic) mediated by polarizability tensors is summarized
in Table 4.1. It shows which tensors play a role for particular excitations.
The terminology is the same as that presented in Section 1.16.
Let us take as an example a plane wave incident upon such an ele-
Fig. 4.10 Geometry of a chiral particle. ment with electric-field polarization in the x direction and magnetic-field
From Tretyakov et al. (1996). Copy- polarization in the z direction. The corresponding tensor elements are
c 1996 IEEE
right
αmm em me ee
zz , αzz , αyz , and αxx . The corresponding physics is quite straight-
forward. A magnetic field in the z direction sets up a current flowing
in the loop that gives rise to a magnetic moment in the z direction and
4.2 Early designs: a historical review 125
Table 4.1 Effective dipole moments for different exciting field directions. From
c 1996 IEEE
Tretyakov et al. (1996). Copyright
pz
m
E αee
zz
py αme
zz
αee
yz
pz
m
H αem
zz
py αmm
zz
αem
yz
pz
E m
αee
yy
py αme
zy
αee
zy
E αee
px xx
a a
CM
c
ext
CG LC
Fig. 4.14 Complementary split-ring
d resonator (a) and its equivalent cir-
cuit (b). From Beruete et al. (2006).
Copyright
c 2006 American Institute
CSRR of Physics
(a) (b)
0 0
–5 –5
Transmission (dB)
Transmission (dB)
–10 –10
3.4 3.5 3.6 3.7 3.8 3.9 4.0 4.1 4.2 3.4 3.5 3.6 3.7 3.8 3.9 4.0 4.1 4.2
Frequency (GHz) Frequency (GHz)
Fig. 4.16 Transmission spectra of individual SRRs with d, the width of the gap, as a parameter, obtained by (a) experiment
and (b) simulation. From Aydin et al. (2005)
(a) (b)
0 0
Transmission (dB)
Transmission (dB)
–5 –5
–10 –10
t = 0.2 mm –15 t = 0.2 mm
–15
t = 0.3 mm t = 0.3 mm
t = 0.4 mm t = 0.4 mm
–20 t = 0.5 mm –20 t = 0.5 mm
–25 –25
3.4 3.6 3.8 4.0 4.2 4.4 4.6 4.8 3.4 3.6 3.8 4.0 4.2 4.4 4.6 4.8
Frequency (GHz) Frequency (GHz)
Fig. 4.17 Transmission spectra of individual SRRs with t, the inter-ring separation, as a parameter, obtained by (a) experiment
and (b) simulation. From Aydin et al. (2005)
g l
w Fig. 4.18 Double-ring SRR. g = t =
t
w = 0.2 mm; metal depth = 0.2 mm.
The SRR side l = 1.8 mm. From Kafe-
c 2005
saki et al. (2005). Copyright
k IOP Publishing Ltd
T
outer
inner
Fig. 4.19 Transmission (T , in dB) ver-
double
sus frequency (in GHz) for the double-
ring SRR shown in Fig. 4.18(a), and
10 20 30 40 50 60 70 10 20 30 40 50 60 70 for its isolated outer and inner ring res-
Frequency (GHz) Frequency (GHz) onators (b). The background mater-
ial is air. From Kafesaki et al. (2005).
(a) (b) Copyright
c 2005 IOP Publishing Ltd
SRR (cf. the physical arguments advanced in Section 2.5 for the resonant
frequency of a SRR).
We may then ask the question what the third resonance at 70 GHz
(Fig. 4.19) is due to? It seems very likely to be an electrical reso-
nance akin to that of an electric dipole antenna. It needs to be recalled
that a centre-fed (or short-circuit) electric dipole antenna has resonances
when its length is approximately equal to an integral multiple of half-
wavelengths. More accurately, the resonant length of an infinitely thin
electric dipole is a little below half-wavelength and the resonant length
is further reduced as the diameter to length ratio is increasing. 70 GHz
corresponds to a free-space wavelength of 4.3 mm. Thus, a side of the
SRR in Fig. 4.18 being equal to 1.8 mm is about 0.42 wavelength. So
the electrical resonance hypothesis is plausible.
Kafesaki et al. (2005) also determined the transmission of a plane wave
for three further orientations of the SRR, as illustrated in Fig. 4.20(a).
In configuration (A) the electric field is perpendicular to the gap. The
corresponding resonances are shown in Fig. 4.20(b). Four resonances
may now be seen. The two magnetic resonances are at the same fre-
quencies, as may be expected, since the magnetic field is perpendicular
to the plane of the SRR. The physical mechanism for the two resonances
130 Small resonators
E
In-plane incidence Normal incidence
(A)
k
H
E
T
(B)
k
H (A) (B)
E (C)
(C)
k Frequency (GHz) Frequency (GHz)
H
Fig. 4.20 Plane-wave transmission. (a) Configurations (a). Transmission spectra for in-plane incidence (b) and for normal
c 2005 IOP Publishing Ltd
incidence (c). From Kafesaki et al. (2005). Copyright
(a) (b)
0 0
Transmission (dB)
–5 –5
–10 –10
–15 –15
–20 SRR –20 SRR
–25 CRR CRR
–25
3 4 5 6 7 8 9 10 11 12 13 3 4 5 6 7 8 9 10 11 12 13
Frequency (GHz) Frequency (GHz)
Fig. 4.22 Frequency response of an individual SRR and CRR (full ring) obtained from (a) experiments and (b) simulations.
From Aydin and Ozbay (2007b). Copyright
c 2007 Wiley-VCH Verlag GmbH & Co. KGaA
–10 0
30
Transmission (dB)
60
–20 90
E
H
–30
k
–40
Fig. 4.23 Transmission spectra for
oblique incidence. From Aydin and
3 4 5 6 7 Ozbay (2007b). c 2007
Copyright
Frequency (GHz) Wiley-VCH Verlag GmbH & Co. KGaA
dc g
w
Lc d
D w r2
b h r3 r1
g r4
f
b s t e dc
(a) (b) (c) (d) (e)
w t
r r
l s
g
(f ) (g) (h) (i) (j)
Fig. 4.24 A variety of resonators (a) by O’Brien and Pendry (2002), copyright
c 2002 IOP Publishing Ltd, (b) and (c) by
Guo et al. (2005), copyright
c 2005 IEEE, (d) by Hsu et al. (2004), copyright
c 2004 American Institute of Physics, (e) by
c 2006 Optical Society of America, (f) and (g) by Radkovskaya et al. (2007b), copyright
Bulu et al. (2005a), copyright
c 2007
Wiley-VCH Verlag GmbH & Co. KGaA, (h), (i) and (j) by Kafesaki et al. (2005), copyright
c 2005 IOP Publishing Ltd
p C2
C4
I C1
C6
C6
z
y C3
p C5
x
C1
Fig. 4.27 Isotropic resonators. (a)
From Gay-Balmaz and Martin (2002).
Copyright
c 2002 American Institute
of Physics.(b) From Chen et al. (2006).
Copyright
c 2006 American Institute
(a) (b) of Physics
Fig. 4.28 Capacitively loaded resonators. (a) From Wiltshire et al. (2003b). (b) From Syms et al. (2006b)
respectively.
A simple way of building a small resonator to which magnetic fields
have access is to take a metallic loop and insert a lumped capacitance.
This was done by Wiltshire et al. (2003b) who wound two turns of 1
mm diameter copper wire on a dielectric rod (9.6 mm diameter) and
tuned the frequency to 60 MHz by inserting a capacitor (nominally 100
pF) between the ends of the wire as shown in Fig. 4.28(a). Syms et al.
(2006b) used rectangular coils on one (design A) or both sides (design
B) of a PCB as shown in Fig. 4.28(b). A resonator that the authors
called an open split-ring resonator was used by Martel et al. (2004) as
a component in a microstrip line for realizing a filter. It is shown with
its equivalent circuit in Figs. 4.29(a) and (b).
It is also possible to augment the inter-ring capacitance in a SRR
by adding an external capacitor (Aydin et al., 2005). In one of their
examples the resonant frequency of the SRR was 3.63 GHz, which they
could change to 2.87 GHz and to 1.63 GHz by inserting capacitors of 0.1
134 Small resonators
OSRR
+ – – +
Top view
Ground plane window
Bottom view
(a)
R L C
d
Fig. 4.29 Open split-ring resonator.
From Martel et al. (2004) (b)
e pF and 0.8 pF, respectively, across the outer gap. In another publication
d Aydin and Ozbay (2007a) reported a reduction of the resonant frequency
from 3.82 GHz to 0.99 GHz by using a C = 2.2 pF capacitor in the
e gap. The authors also attempted to reduce the resonant frequency by
inserting capacitors between the rings. They succeeded, but found that
(a) it was a less efficient way of reducing the resonant frequency. Another
motivation for inserting a capacitance into a SRR is to tune the resonant
e frequency. This can be done with the aid of a varactor diode as shown
by Gil et al. 2004; Gil et al. 2006. They were able to tune their element
d from about 2.5 GHz to 3.1 GHz by applying a voltage to the diode in
the range of 0 to 30 V. The element is designed with the need in mind
to accommodate the diode and to facilitate its biasing. The two models
e g
considered, both elongated, are shown in Figs. 4.30(a) and (b).
All the elements discussed so far had anisotropic properties. If the
(b)
aim is to construct an isotropic medium then the unit cell should display
some 3D symmetry. Examples are the unit cube with six SRRs on its
Fig. 4.30 Capacitively loaded split surface (see Fig. 4.31(a)) as proposed by Pendry et al. (1999) and further
rings. From Gil et al. (2006). Copy- investigated by Baena et al. (2006), and the 3D symmetric element of
c 2006 IEEE
right
Gay-Balmaz and Martin (2002) (Fig. 4.31(b)). A fully symmetric design
by Soukoulis et al. (2006) and Padilla (2007) is based on quadruply split
4.3 A roll-call of resonators 135
200 nm
2.8
2.4
Wavelength [mm]
2.0
d
E
w l D
H
k (a)
600
500
400
Frequency (THz)
300
Fig. 4.34 The scaling of the simulated magnetic resonance frequency as a function of the linear size a of the unit cell for
the 1-, 2-, and 4-cut SRRs (solid lines with symbols). Up to the lower terahertz region, the scaling is linear. The maximum
attainable frequency is strongly enhanced with the number of cuts in the SRR ring. The hollow symbols as well as the vertical
line at 1/a=17.9 µm−1 indicate that µ < 0 is no longer reached. The non-solid lines show the scaling of the magnetic resonance
frequency calculated through the LC circuit model. From Zhou et al. (2005a). Copyright
c 2005 by the American Physical
Society. For coloured version see plate section
W do
z tAu-top
Au
H ZnS
ZnS Structure
tc
Substrate tAu Image
ZnS
(a) (b)
Fig. 4.35 Staple structure. (a) Schematic of the nanostructures. The light is incident from the top: for TM polarization, H
is in the y direction, perpendicular to the loop. (b) The equivalent LC circuit formed between the top staple structure and
its image in the metal. The dashed lines are the reference planes for the effective permeability calculation. From Zhang et al.
(2005c). Copyright
c 2005 by the American Physical Society
bold lines show contours by gold. The dielectric between the elements
and the reflector is ZnS. The authors call the resonators, quite aptly,
‘staples’. Taken together with their mirror image it is easy to see that
they possess inductances and capacitances. They were produced in 2D
arrays in the x and z directions.
The parameters for three structures realized are given in Table 4.2.
The experimental and simulation results for reflection at normal inci-
dence are plotted in Figs. 4.36(a) and (b). A large dip may be clearly
seen for configuration A at a frequency of about 45 THz for an incident
TM wave (magnetic polarization in the y direction). For configurations
B and C the dip (for the same polarization) occurs at higher frequen-
cies. The reason is clearly that for configuration B the height of the
‘staple’ H is smaller hence the inductance is smaller, and for configu-
ration C the flange is smaller, yielding reduced capacitance. Note that
the dip occurs only for the TM polarization when the magnetic field can
access the open loops. As may be seen in Fig. 4.36(a), the reflection is
frequency-independent for the TE polarization. The effective permeabil-
ity determined from their model is plotted in Fig. 4.36(c). A negative
permeability region for configuration C around 64 THz may be seen.
Clearly, the idea of having a reflector leads to interesting physics but
it makes it impossible to measure transmission. Absorption and effec-
4.3 A roll-call of resonators 139
100
C,TE
80
D,TM
60
(a)
40
A,TM C,TM
Reflectance (%)
20
Experiment
0
80
60
(b)
40
E H
k
d
b2
1
2b
Inductive Coupling
l1
φ
Fig. 4.38 An element of the bilay-
ered planar chiral metallic microstruc-
ture with inductive cross-layer cou- (a2 + b2) 1/2 a
pling. Two subwavelength metallic z
stripes of size d×∆ are mutually shifted d
in the xy plane on distance b and spaced l2
by a dielectric layer of thickness a (not
shown). From Svirko et al. (2001). y
Copyright
c 2001 American Institute ∆ x
of Physics
the frequency is high enough to be not too far from the metal’s plasma
frequency, and for the LC resonance the following simple argument can
be presented. The two rods of Fig. 4.39 have self-inductances, there is a
mutual inductance between them and there is a capacitance between the
I rods, analogously to a split-ring resonator. There are conduction cur-
rents (denoted by I flowing in the opposite directions, and displacement
I currents as well flowing between the rods.
Another illustration of the transition7 from the open resonator to the
Fig. 4.39 Conduction current in a short-rod pair due to Dolling et al. (2005) is shown in Fig. 4.40.
short-rod pair A set of short-rod pairs turn out to be suitable not only to provide
7
The U-shaped resonator, second from negative permeability but negative permittivity as well. This was shown
the left, is reminiscent of the hair- theoretically by Podolskiy et al. (2002) and is illustrated in Figs. 4.41(a)
pin resonator of Hong and Lancaster and (b). The corresponding refractive index is plotted in Fig. 4.41(c).
(1998). Experimental results by Shalaev et al. (2005) and Kildishev et al. (2006)
exhibited negative indices of refraction at a frequency of 200 THz but
other experimenters (Dolling et al., 2005; Garwe et al., 2006) did not
succeed in obtaining a negative index in that frequency range.
Negative index due to a structure of short wire pairs was found by
Fig. 4.40 Transition from the open res- Zhou et al. (2006) in a narrow frequency range near 14 GHz. A design
onator to the short-rod pair. From
c
by Zhang et al. (2005a), shown in Fig. 4.42 (top) relies, instead of short
Dolling et al. (2005). Copyright
2005 Optical Society of America rods, on short but wider surfaces facing each other. This structure is
then responsible for the magnetic resonance. Rods reappear to provide
4.3 A roll-call of resonators 141
Fig. 4.41 Permeability (a), permittivity (b) and refractive index (c) for the short-rod-pair medium. From Podolskiy et al.
(2002)
E k
a b
a
b
Fig. 4.44 Diagram (a) and SEM image
(b) of fabricated fishnet structure. The p
inset shows a cross-section of the pat-
tern taken at a 45◦ angle. From Valen- Ag
tine et al. (2008). Copyright
c 2008 1 µm
MgF 2
Nature Publishing Group
0
1,475 nm
–0.5
1,590 nm
–1.0
1,700 nm
–1.5
1,200 1,400 1,600 1,800
–5 0 5 10 –5 0 5 10
Wavelength (nm)
Position (mm) Position (mm)
(a) (b) (c)
Fig. 4.45 Experimental results and FDTD simulations. Fourier-plane images of the beam for the window (a) and prism sample
(b) for various wavelengths. The horizontal axis corresponds to the beam shift d, and positions of n = 1 and n = 0 are denoted
by the white lines. The image intensity for each wavelength has been normalized for clarity. (c) Measurements (circles wuth
error bars after four measurements) and simulation (solid line) of the fishnet refractive index. From Valentine et al. (2008).
Copyright
c 2008 Nature Publishing Group. For coloured version see plate section
and Martin 2001; Kottmann et al. 2001 both of regular and non-regular
shape. They concluded that there are many more resonances for a non-
regular shape, as shown in Fig. 4.46 where scattering cross-section is
plotted against wavelength for plane-wave incidence. The particle is a
right-angled triangle of 10 nm base and 20 nm side. The polarization of
the incident electric field is in the plane of the figure perpendicular to the
arrow in the inset. As many as five resonances may be seen, whereas an
ellipse has only one resonance. The electric-field distribution is shown
by a colour code in Fig. 4.47 at the main resonance of 458 nm. Normal-
ization is to the amplitude of the incident field. The amplitude remains
ten times that of the incident field at a 10 nm distance from the sharp
corner.
The effect of the coupling between two cylinders of 25 nm radii made
of silver was also investigated (Kottmann et al., 2000a) as a function of
the distance between cylinders. A plane wave in the direction of the axis
of symmetry was incident with the electric field perpendicular to the axes
of the cylinders. The resonant wavelengths were found to be 350 nm,
358 nm, 368 nm, 380 nm and 404 nm for distances between the cylinders
of 50 nm, 20 nm, 10 nm, 5 nm and 2 nm. A further radiation-induced
resonance was found by Kottmann and Martin (2001).
144 Small resonators
λ=458nm
Scattering cross section [nm]
λ=392nm
101
λ=369nm
100
λ=329nm
10–1
g1 d
r 10 G1 dM
L1 dM
I 1 (M) + - I 1 (M+dM)
~
(a) Ic
g2 V(M) M dM C dM V(M+dM) (b)
+ - ~
w I 2 (M) L dM G dM I 2 (M+dM)
h 2 2
w
y y
x x
Cg1
G1 dM
~
~ M dM ~
~ L1 dM
~ ~
G2 dM
L2 dM C dM
~
~
(c)
dM
~
~
~ Cg2 ~
~
~ ~
~
Fig. 4.48 Split-ring resonator. (a) A schematic view. (b) The equivalent circuit of a dϕ element. (c) The equivalent circuit
c 2004 American Institute of Physics, and Shamonin et al.
of the entire resonator. From Shamonin et al. (2004), copyright
(2005), copyright
c 2005 Wiley-VCH Verlag GmbH & Co. KGaA
placement current, Ic , flows between the outer and inner rings, (iv) the
inner and outer rings have inductances L1 and L2 and resistances R1 and
R2 , respectively, (v) the capacitances across the gaps in the outer and
inner rings are Cg1 and Cg2 , respectively, (vi) a spatially constant and
in time harmonically varying magnetic field will induce voltages both in
the outer and inner rings.
We shall now write Kirchhoff’s equations for a dϕ element of the SRR
shown in Fig. 4.48(b). The current equation for the upper line is
per unit radian. The current equation for the lower line is
I1 + I2 = I0 (4.10)
follows. I0 is a constant that will have to be determined from the bound-
ary conditions. We can similarly write the voltage equation in the form
F1 − F2
V (ϕ + dϕ) = V (ϕ) + I1 Z1 − j ωM I1 − I2 Z2 + j ωM I2 + , (4.11)
2π
where
dV F1 − F2
− = I1 (Z1 + Z2 − 2j ωM ) − I0 (Z2 − j ωM ) − . (4.14)
dϕ 2π
We have now two first-order differential equations in I1 and V from
which we may obtain a second-order differential equation with constant
coefficients in either variable that can be solved in terms of constants
and trigonometric functions. To continue we need the full equivalent
circuit shown in Fig. 4.48(c) which will help to formulate the boundary
conditions.
Next, let us look in more detail at what happens at ϕ = 0 (note
that in order to exploit symmetry our azimuthal co-ordinate system
runs from −π to 0, and from 0 to π. According to our model the gap
capacitor Cg1 is exactly at the position ϕ = 0 and has zero spatial
extent. There is, however, bound to be a voltage drop across it due to
the displacement current flowing through it. Consequently, the inter-ring
voltage cannot be continuous at ϕ = 0. The current must be continuous
but not necessarily differentiable. It follows then that we need to solve
4.4 A mathematical model and further experimental results 147
We have taken care of the currents and voltages but not, as yet, of the
excitation. For that we shall have to write Kirchhoff’s voltage law either
along the inner or the outer ring, which will include the excitation. The
condition for the total voltage to be zero along the outer ring may then
be written as
Z0 Zπ
I1 (0)
(Z1 +j ωM ) I1− dϕ + I1+ dϕ −2j ωM πI0 + = F1 . (4.17)
j ωCg1
−π 0
where
C C
κ2 = ω 2 Leq C , ,
γ1 = γ2 = ,
Cg1 Cg2
(L1 − M )2 (L2 − M )2
ν1 = , ν2 = ,
L L − M2 L1 L2 − M 2
√1 2
ν12 = ν1 ν2 , Leq = L1 + L2 − 2M . (4.19)
148 Small resonators
Cg1 1
ω2 = . (4.22)
Ct
Lt + Cg1 + Cg2
Cg2 4
The conclusion is now clear. The lowest resonant frequency of a SRR
is given by the equivalent circuit shown in Fig. 4.49, a generalization
Fig. 4.49 Equivalent circuit of that of Fig. 4.13(a). The inter-ring capacitance between two splits is
Ct /2. These two capacitances in series are then connected in parallel
with the two gap capacitances. Note that eqn (4.22) is a generalization of
the expression given by eqn (2.26), derived intuitively by Marques et al.
(2002b). It now includes the gap capacitances as well and, interestingly,
8
Having set up a model for a SRR we in a very simple manner.8
can easily extend the same principles
to the modelling of a singly split dou-
ble ring that also consists of two con- 4.4.2 Results
centric rings but differs from the SRR
by having only one of the rings split. The main interest is in the lowest resonant frequency because the aim
In fact, the inner ring being split is an is to produce a medium describable by effective material parameters.
old design due to Hardy and Whitehead The smaller the elements the more applicable effective-medium theory
(1981) that has already been shown in
Fig. 4.5. The analysis, quite similar to is. We shall now look at a few examples. A schematic drawing of the
that presented above, was performed by SRR has already been shown in Fig. 4.48(a). We shall now choose the
Shamonin et al. (2004). The conclu- dimensions as follows: Its height is h = 5 mm. The external radius of
sions were similar too. The main dif-
the outer ring is re1 = 11 mm and the wall thickness is 1 mm. The gaps
ference relative to the SRR was that
having only one split the resonance fre- are 1 mm both in the inner and in the outer ring. The wall thickness of
quency for the same parameters was the inner ring is 0.8 mm. We shall look at three examples in which the
higher. dimensions enumerated above are all identical but the external radius of
the inner ring takes three different values, namely re2 = 7.5, 8.5 and 9
4.4 A mathematical model and further experimental results 149
Table 4.3 Parameters and resonant frequencies of split-ring resonators with variable inner-ring
radius
1 1
current
current
0.5 0.5
0 (a) 0 (b)
1 1
voltage
voltage
0 0
−1 −1
−180 −90 0 90 180 −180 −90 0 90 180
M [deg] M [deg]
1 1
current
current
0.5 0.5
0 (c) 0 (d)
1 1
voltage
voltage
0 0
−1 −1
−180 −90 0 90 180 −180 −90 0 90 180
M [deg] M [deg]
4
current
current
2 5
0 0
−2 −5
1 (e) 1 (f)
voltage
voltage
0 0
−1 −1
−180 −90 0 90 180 −180 −90 0 90 180
M [deg] M [deg]
Fig. 4.50 Voltage and current distributions. fr = 50 MHz, 1.5, 2.5, 3.02, 3.88, 7.97 GHz (a)–(f)
A1 A2 B1 B2
B1 1.74 1.77
B2 2.59 2.67
(2005) in which four copper rings (Fig. 4.51) were used in different com-
binations. The dimensions of the rings may be recognized as those we
used in our earlier theoretical calculations with re2 = 7.5 mm. Out of the 11
For the same sets of resonators fur-
four rings it is possible to construct five resonators, as shown schemat- ther results (measuring the quality fac-
ically in Table 4.4. B1 A2 is the notation for the full ring A2 placed tor, Q) were obtained by Hao et al.
concentrically inside the split ring B1 . Similarly, A1 B2 means the split 2005b; Hao et al. 2005a.
ring B2 inside the full ring A1 , and B1 B2 means the smaller split ring
B2 inside the larger split ring B1 with splits at opposite sides. The last
one is of course a SRR. 12
MICRO-STRIPES is a registered
The purpose of the experiments was to measure the resonant frequen- trademark of Flomerics Ltd., Surrey,
cies of all four configurations.11 They are shown in column 3 of Table 4.4, UK.
whereas the resonant frequencies obtained by the MICRO-STRIPES12
numerical package are in column 5. As expected, B2 , the smaller split
152 Small resonators
ring has a higher resonant frequency than B1 . The smaller the rings
−2 −1 0 the higher the resonant frequencies. One would indeed expect simple in-
verse scaling of the resonant frequency with the radii. Note from Table
4.4 that the ratio of the inner radius of ring B1 to that of B2 is 1.49,
and the inverse ratio of the corresponding resonant frequencies is also
1.49. It may also be seen that in the configurations B1 A2 and A1 B2
the resonant frequency is determined by the size of the split ring. The
(a)
ratio of the corresponding resonant frequencies of 2.78 GHz and 1.89
GHz is 1.47, very close to the ratio of the radii. Thus, the influence of
the full inner and outer rings is merely to increase a little the resonant
frequency. The lowest resonant frequency is 1.44 GHz, obtained by the
SRR. The agreement between the experimental and simulation results is
quite close. The maximum deviation is for the B2 case, which amounts
to about 3%.
We have no analytic expressions for the resonance frequencies of the
single rings B1 and B2 . The characteristic equation (eqn (4.18)) makes
it possible to determine the resonant frequency, as follows from our dis-
tributed circuit model. The resonant frequency obtained from that is
(b) 1.50 GHz, in contrast to the measured 1.44 GHz. We have not shown the
analytic expressions for the singly split double rings of the A1 B2 and the
A2 B1 configurations, but they are available in Shamonin et al. (2004).
The resonant frequencies calculated from there are given in Table 4.4 as
2.17 and 2.89 GHz, which are still reasonable approximations.
For the resonant frequency of the SRR three more approximate ex-
pressions are available, that of eqn (4.22), and those of Pendry et al.
(1999) and Sauviac et al. (2004). We shall give here the expression of
Pendry et al. (1999) which is in the simple form of
r
c 3d
ωr = , (4.23)
(c) r π2 r
where d = ri1 − re2 , r = (ri2 − re2 )/2. The above equation yields a
resonant frequency of 1.62 GHz, far too high, whereas the expression
given by Sauviac et al. (2004) yields 1.18 GHz. On the other hand,
eqn (4.22) yields quite a good approximation of 1.35 GHz. For the
A1 B2 configuration there is also a simple analytical formula as given by
eqn (4.3). It gives a resonant frequency of 6.4 GHz, quite far from the
Fig. 4.52 Electric-field distributions at measured 2.78 GHz. The probable reason for the large discrepancy is
the fundamental (a), the first (b) and that they attributed too much importance to the gap capacitance, which
the second higher resonance (c). Po- was indeed quite large in their design shown in Fig. 4.5.
sitions of the voltage zeros are shown
by white dashed lines. From Hesmer
(2008). For coloured version see plate 4.4.3 A note on higher resonances
section
In Table 4.3 we gave the analytical results both for the fundamental
and for the higher-order resonances. We have also determined them by
13
CST MICROWAVE STUDIO is simulations using the numerical package CST MICROWAVE STUDIO13
a registered trademark of CST– (Hesmer, 2008). It is reasonable to expect good agreement for the fun-
Computer Simulation Technology AG,
Darmstadt, Germany
damental resonance, but the distributed circuit model is less likely to
be valid for the higher-order resonances. The analytical results from
4.4 A mathematical model and further experimental results 153
Table 4.3 are 1.50, 3.88 and 7.97 GHz. Our numerical simulations gave
1.46, 3.42 and 7.68 GHz, which shows larger discrepancies for the higher
orders.
The simulations also provided the field distributions at the resonant
frequencies. They are shown in Fig. 4.52. It may be immediately seen
that the field distribution has no zeros for the fundamental resonance,
two zeros for the first higher-order resonance and four zeros for the sec-
ond higher-order resonance. Looking at Figs. 4.50(e) and (f) we can see
that the number of zero points is the same as in the analytical solution.
We can also compare the positions of the zeros. The analytical calcu-
lations give (they are symmetric around the zero angle so we shall give
only the positive values) for f1 74.2◦ , and 26.2◦ and 147◦ for f2 . In
contrast the simulations yield 62◦ for the first and 19.5◦ and 145◦ for
the second higher resonance. Not a close agreement, but all are in the
same ballpark.
This page intentionally left blank
Subwavelength imaging
5
5.1 Introduction 5.1 Introduction 155
5.2 The perfect lens:
The subject of metamaterials started with Veselago’s (1967) introduc- controversy around
tion of negative refractive index and his simultaneous proposal for a flat the concept 157
lens. We have already discussed the flat-lens family in Section 2.11.4. 5.3 Near-perfect lens 159
Its characteristic feature is the existence of a focus inside the lens. The 5.4 Negative-permittivity
lens 176
basic principle of its operation has already been shown in Fig. 2.24; for
5.5 Multilayer superlens 188
convenience it is shown again in Fig. 5.1. The essential requirement is
that the index of refraction is n = −1. The ray trajectories follow then 5.6 Magnifying multilayer
superlens 192
in the geometrical optics approximation. A focal point is brought to a
5.7 Misconceptions 196
focal point and an object in the object plane is reproduced perfectly in
the image plane. This is an obvious advantage of the flat lens. It has no
optical axis. A simple illustration in Fig. 5.1 shows an object consisting
of three points in the object plane that is perfectly reproduced. At the
same time we need to mention a serious disadvantage of the flat lens.
It can work only at a single frequency because the n = −1 condition is
very strongly frequency-dependent. In traditional imaging language this
means that the flat lens has very large chromatic aberration.
As follows from the construction in Fig. 5.1, Veselago’s lens is perfect
when we consider propagating waves only. But, if the object is small
relative to the wavelength it will have both propagating and evanescent
components in its Fourier spectrum. The image is not perfect because
the subwavelength features have not been reproduced. It was shown by
Pendry (2000) that a flat lens in the Veselago geometry can reproduce
ș1 ș1
ș1 ș1
z
Fig. 5.1 Imaging three point sources
0 d/2 3d/2 2d with Veselago’s flat lens
156 Subwavelength imaging
1
There is some problem here with ter- subwavelength features on account of the evanescent waves growing1
minology. In Section 2.12 we have al- in the lens. This was already discussed in Section 2.12. Here, we shall
ready objected to the term ‘amplifica-
tion’ because there is no source of ad-
assume that the reader has already read that section, and is also familiar
ditional power there. Strictly speak- with surface plasmons discussed in Chapter 3. Later in this chapter we
ing we should not talk about growing shall of course look at subwavelength imaging from a number of different
evanescent waves either. To evanesce angles but here we shall start with a brief summary in the form of
means to fade away. A growing evanes-
cent wave is a contradiction in terms.
questions and answers.
So does it grow or does it evanesce? Firstly, what is the difference between Veselago’s flat lens and Pendry’s
It can’t do both. The correct descrip- perfect lens? The geometry is the same but not the parameters of the
tion would be the growth of the ampli- lens. Veselago’s lens will work whenever the index of refraction is n =
tude of the non-propagating waves in
the +z direction. Unfortunately, that’s −1. For Pendry’s perfect lens it is required for both material parameters,
quite a mouthful and goes against cur- εr and µr , to be equal to −1. Pendry’s lens is always matched. If
rent practice. Thus, with regret, we in Veselago’s lens the n = −1 condition is obtained by both material
shall often refer to growing evanescent parameters being equal to −1 then for a lossless lens there is no difference
waves.
between the Veselago and Pendry variety.
Secondly, why is the total phase change from object to image in
Pendry’s lens equal to zero? Because the wave in the lens is a back-
ward wave that has the same phase velocity as the forward waves from
object to lens, and from lens to image. Since the path in the lens is
equal to the path in air the total phase change is zero.
Thirdly, why is the total amplitude of a non-propagating space har-
monic from object to image equal to unity? This is because the rate of
increase in the lens is equal to the rate of decrease in air.
Fourthly, is a flat transfer function possible in practice? It is not
possible in practice and actually not in theory either because it implies
infinitely high fields.
In the fifth place, what will prevent the transfer function from be-
ing flat? Losses, because the growing non-propagating waves will be
attenuated and therefore perfect compensation of amplitudes becomes
impossible. Also, any deviation from the ideal conditions will cause the
appearance of a cutoff in the transfer function.
In the sixth place, why is there a chance to build a subwavelength lens
from a material in which only εr or only µr is equal to −1? It follows
from the properties of TM waves that their amplitudes depend strongly
only on ε and vice versa for TE waves.
In the seventh place, why does the perfect lens need both material
parameters to be −1 if with the right polarization the amplitudes depend
strongly only on one of the material parameters? The reason is that the
phase velocity of propagating waves and the rate of decay (or growth)
of non-propagating waves depend on both material parameters.
In the eighth place, why do we have growing non-propagating waves
in the epsilon-negative-only lens? Because surface plasmon–polaritons
(SPPs) are excited at the far boundary and they cannot be excited by
a propagating wave owing to their shorter wavelength.
In the ninth place, why is it necessary for the epsilon-negative-only
lens to be thin? This is because the SPP on the rear surface can only
be excited by an input wave if the front and rear surfaces are strongly
coupled, and they are strongly coupled only when they are near to each
5.2 The perfect lens: controversy around the concept 157
other.
In the tenth place, why is a multilayer lens preferable to a single-layer
lens? Because the SPP interaction between the front and rear surfaces is
still active for consecutive thin lenses, and at the same time the distance
between the object and image increases.
The above is a rather rough and very concise summary of what has
been said so far in this book about imaging by a flat lens. The rest of
the chapter will of course go into a lot of details. Section 5.2 will be
concerned with history, with the controversy that arose in the wake of
the publication of Pendry’s paper (2000). The most important part of
this chapter is probably Section 5.3, which will describe in great detail
what happens to the transfer function when the parameters deviate from
the ideal ones. Also in the same section, we shall establish relationships
between phenomena in the perfect lens and the physics of SPPs. In
Section 5.4 we shall investigate the quality of images under conditions
when only the permittivity is equal to −1. Section 5.5 is concerned
with the improvements a multilayer lens can offer. Section 5.6 discusses
magnification. Section 5.7 is an attempt to show and briefly discuss
some of the widely held beliefs that may not stand closer scrutiny.
changed: The goal was no longer solely to prove that the perfect lens
concept has flaws (it surely does, as nothing is ever perfect in nature),
but rather to understand the underlining physical mechanisms of sub-
wavelength imaging and to explore the possibilities offered by metama-
terial near-field lenses. Most researchers agree by now that the concept
of subwavelength imaging by a flat lens is both novel and useful.
differ slightly from the ideal constellation. In other words to what ex-
tent imaging beyond the diffraction limit is possible using a fabricated
material and to what extent the process can be simulated by numerical
methods that inevitably approximate the ideal situation.
To familiarize ourselves with the operation of a homogeneous meta-
material slab acting as a near-perfect lens we will from now on consider
only the stationary state. It follows from our previous discussion in
Section 5.2 that we can do it with a clear conscience: a steady state
would exhibit a high-frequency cutoff ensuring that all fields are square-
integrable.
We shall now investigate the functioning of the lens by making use
of various tools of analysis. Firstly, the transfer function provides the
information on how well (or how badly) the input field in the plane
z = 0 is reproduced in the image plane z = 2d. Secondly, the reflection
coefficient (identically zero in the perfect-lens idealization) when it is
different from zero indicates the existence of a surface mode at the front
surface. Thirdly, variation of individual Fourier components across the
slab helps to visualize the mechanism of what is going on inside the slab.
This section will be divided into a number of subsections. In Sec-
tion 5.3.2 we shall provide the equations needed. The effect of losses
on the transfer function, and in particular on the cutoff frequency, is
treated in Section 5.3.3, then in Section 5.3.4 we shall return to the loss-
less case and investigate the effect of deviations from the ideal values of
εr = µr = −1 followed by an investigation of the deviations from the
ideal refractive index of n = −1, and finally the case is examined when
ε and µ vary in such a manner that their product remains the same,
i.e. the refractive index does not change. In Section 5.3.5 the effect of
lens thickness is investigated in the presence of losses. The influence
of the various kinds of deviations is summarized in Section 5.3.6. An
interesting and important point is raised in Section 5.3.7 that may cheer
up those sceptical about the usefulness of numerical simulations. The
problem seems to be an inherent mismatch in the constitutive parame-
ters of the slab and the surrounding medium that may lead to incorrect
results.
A(kx ) e −j kx x .
X
Hy (x, z = 0) = (5.1)
kx
The field expansion of the full solution everywhere in the xz plane takes
the form
A e −j kz1 z + B e j kz1 z e −j kx x
X
Hy1 (x, z) =
kx
(medium 1) ,
C e −j kz2 z + D e j kz2 z e −j kx x
X
Hy2 (x, z) =
kx
(medium 2) ,
F e −j kz3 z − j kx x
X
Hy3 (x, z) =
kx
(medium 3) . (5.2)
A = A(kx ) and B = B(kx ) are the amplitudes of the incident and
reflected waves of the Fourier component kx in medium 1, C = C(kx )
and D = D(kx ) are the corresponding amplitudes of the transmitted
and reflected waves in medium 2 (slab) and F = F (kx ) is the amplitude
of the transmitted wave in medium 3. The corresponding components
of the electric field are
X kz1
A e −j kz1 z − B e j kz1 z e −j kx x
Ex1 =
ωε1
kx
(medium 1) ,
X kz2
C e −j kz1 z − D e j kz1 z e −j kx
Ex2 =
ωε2
kx
(medium 2) ,
X kz3
Ex3 = F e −j kz3 z − j kx x
ωε3
kx
(medium 3) (5.3)
and
X kx
A e −j kz1 z + B e j kz1 z e −j kx x
Ez1 = −
ωεr1
kx
(medium 1) ,
X kx
C e −j kz1 z + D e j kz1 z e −j kx x
Ez2 = −
ωεr2
kx
(medium 2) ,
X kx
Ez3 = − F e −j kz3 z − j kx x
ωεr3
kx
(medium 3) . (5.4)
162 Subwavelength imaging
We wish to see what impact minor deviations from the ideal perfect
lens constellation would have, therefore medium 1 and 3 are assumed to
be identical (in the simplest case with εr1 = εr3 = 1, µr1 = µr3 = 1). The
amplitudes B, C, D, F are given by eqns (1.75)–(1.78). In particular,
the transfer function for the distance 2d is
F −j kz1 d
T = e
A
4ζe e −j kz1 d
= , (5.5)
(1 + ζe )2 e j kz2 d − (1 − ζe )2 e −j kz2 d
B 2j (1 − ζe2 ) sin(kz2 d)
R= = . (5.6)
A (1 + ζe )2 e j kz2 d − (1 − ζe )2 e −j kz2 d
Note that the presence of a pole in the expressions for transmission
and reflection indicates excitation of a SPP mode at the frequency ω
considered.
Next, we shall go through a number of examples of minor perturba-
tions in the perfect-lens parameters.
4ζe e −kx d
T = . (5.9)
(1 + ζe )2 e kx d − (1 − ζe )2 e −kx d
Consequently, the first term in the denominator of eqn (5.5) is regarded
as being close to zero. If kx is not too large but still kx ≫ k0 then the
second term of the denominator in eqn (5.9) dominates, leading to a
transfer function
5.3 Near-perfect lens 163
4 e −kx d
T ≃ =1 (5.10)
4 e −kx d
that is entirely flat. But this is only true for this particular range of
not too large values of kx . The situation is quite different in the limit
kx → ∞. The second term in the denominator of the transfer function,
which declines exponentially with kx , becomes smaller than the first term
and eventually vanishes in the limit kx → ∞. The resulting transfer
function in the limit of high kx values is thus given by
4 e −kx d 4 e −2kx d
T ≃ ≃− . (5.11)
(1 + ζe )2 e kx d (ε′′ )2
Thus, in the presence of loss the transfer function declines exponentially
as kx approaches infinity.
The value of kx at which the transfer function (eqn (5.5)) stops being
flat and starts to decline with kx is called the cutoff frequency and can
be estimated from the condition that the first term in the denominator
(which is the dominant one at not too large kx ) and the second term in
the denominator (which dominates as kx → ∞) are equal (so that the
magnitude of the total transfer function reduces to 1/2),
(1 + ζe )2 ≃ (1 − ζe )2 e −2kx d , (5.12)
yielding the cutoff value of the spatial frequency at
1 2
kxcutoff ≃ ln ′′ . (5.13)
d ε
This approximate expression gives already an idea about how much
the deviation of the metamaterial parameters from the ideal situation
εr = µr = −1 and the thickness of the metamaterial slab as well affect
the cutoff frequency. Note that loss in permeability does not enter the
expression for the cutoff frequency in the electrostatic limit. It would of
course appear for TE polarization. We shall discuss this a little later.
A more general formulation can also be easily derived (see Podolskiy
and Narimanov 2005). Introducing
s
kx2
χ= − 1, (5.14)
k02
we rewrite κ1 and κ2 as
κ1 = k0 χ , (5.15)
ε′′ + µ′′
κ2 ≃ k0 χ 1 − j , (5.16)
2χ2
so that
164 Subwavelength imaging
1
10 1.1
0 1.05 −5
10
transfer function |T|
−1
1
10 −5
G=10
−3 0.95
−2 G=10
10 G=0.1 G=0.1 −3
0.9 G=10
without slab
−3
10 0.85
−4 0.8
10 0 1 0 1
10 10 10 10
k x /k 0 k x /k 0
(a) (b)
Fig. 5.2 Transfer function of a near-perfect lens. d/λ = 0.1, losses only in ε, δ = 10−5 , 10−3 and 0.1. (a) Exact solution
(solid curves) and electrostatic approximation (dotted curves) are shown together with the transfer function for the case when
the slab is removed and the total distance from object to image plane is still the same, 2d (dashed curve). Note that curves
obtained within the electrostatic limit coincide with the exact solution. (b) Detailed view of a part of the graph; only exact
solutions shown
1
10
0
10
transfer function |T|
−5
G=10
−1 −3
10 G=10
−2 G=0.1
10
without slab Fig. 5.3 Same as in Fig. 5.2(a), losses
−3
10 only in ε, but for d/λ = 0.5. Exact
solution (solid curves) and electrostatic
approximation (dotted curves). Elec-
−4
10 trostatic limit is good for large kx and
0 1
10 10 for small loss. It is less accurate for a
kx /k 0 loss of δ = 0.1
comparison, the transfer function for the case when the slab is removed
is plotted as well (dashed line). In Fig. 5.2(a) a thin slab is taken,
d/λ = 0.1. For a small amount of loss, δ = 10−5 , the transfer function
exhibits a cutoff at kxcutoff /k0 = 100. This means a slab of 0.1λ thickness
with εr = −1−j 0.00001 and µr = −1 would reproduce accurately objects
with a resolution of λ/100. That is pretty good. Increasing the loss
reduces the resolution. For δ = 10−3 the cutoff frequency reduces to
kxcutoff = 10k0 , and for δ = 0.1 to kxcutoff = 3k0 . The resolution of the
slab with δ = 10−3 is therefore λ/10 and for δ = 0.1, which is not an
unrealistic value, only λ/3.
Is such a device better than empty space? Certainly yes; the transfer
function is flatter and more evanescent components can make it to the
image plane even for such a high loss of δ = 0.1. The situation is quite
different for propagating components with kx < k0 . After propagation
166 Subwavelength imaging
1
10
0
10 −5
transfer function |T|
G=10
−1 −3
10 G=10
G=0.1
−2
10 without slab
1
10
−5
G=10
0
10 −3
G=10
transfer function |T|
G=0.1
−1
10 without slab
−2
10
−3
10
Fig. 5.5 Same as in Figs. 5.2(a), 5.3
−4
and 5.4, losses only in ε, but for d/λ = 10
0 1
2. Electrostatic limit is no longer a re- 10 10
liable approximation kx /k 0
1
10
0
10
transfer function |T|
−5
G=10
−1 −3
10 G=10
G=0.1
−2
10
without slab
−3
10
kxcutoff /k0 = 1.5 does not make the imaging device look feasible.
As far as the electrostatic limit is concerned, it becomes, as thickness
increases, less and less accurate (Figs. 5.3–5.5). For d/λ = 0.5 the elec-
trostatic limit can be seen to differ considerably from the exact solution
for δ = 0.1. For d/λ = 1 it occurs at δ = 10−3 and for d/λ = 2 the
electrostatic limit δ = 10−5 is no longer capable of reproducing the real
situation.
1 1
10 10
0 0
10 10
transfer function |T|
−3 −3
10 10
−4 −4
10 0 1 10 0 1
10 10 10 10
k x /k 0 k x /k 0
(a) (b)
Fig. 5.7 Effect of deviations in the real part of permittivity. (a) εr = −1 − δ, µr = −1. Only one resonance can be seen. (b)
εr = −1 + δ, µr = −1. Two resonances are excited. Cut-off is sensitive to variations in ε. The electrostatic limit is working
well
1 1
10 10
0 0
10 10
transfer function |T|
−4 −4
10 0 1 10 0 1
10 10 10 10
k x /k 0 k x /k 0
(a) (b)
Fig. 5.8 Effect of deviations in the real part of permeability. (a) εr = −1, µr = −1 − δ. Only one resonance can be seen. (b)
εr = −1, µr = −1 + δ. Two resonances are excited. Thickness d/λ = 0.1. Note that the electrostatic limit (dotted line) fails
completely
1
T ≃ 2 4 . (5.27)
δ k0 2kx d
− e +1
4 kx
" 2 #
1 4 kxcutoff
kxcutoff = ln . (5.28)
d |δ| k0
Again, if µr < −1 there will be only one surface mode that can exist,
and if µr > −1, there can be two resonances. Note that the effect of
variations in µ is weaker than the effect of variations in ε, in agreement
with the cutoff expressions (5.26) and (5.28).
An interesting feature of the transfer function is that its magnitude
can be larger than unity, not only at the two resonances but also every-
where in-between, the plateau of the transfer function means the evanes-
cent part of the spectrum gained in amplitude as compared to its values
coming from the object. The question is whether or not there is anything
wrong with energy conservation (see also Section 5.7). The electrostatic
limit fails as expected. It does not know about variations in µ.
Figures 5.9(a) and (b) show that a thicker slab fails to be a good lens;
the transfer function deteriorates too quickly. The cutoff is too short
and no reasonable resolution can be obtained, even with variations as
small as δ = 10−5 ! In addition, in the case of variations in ε only, the
electrostatic limit begins to fail as the cutoff moves to smaller values of
kx .
5.3 Near-perfect lens 171
1 1
10 10
0 0
10 10
transfer function |T|
−3 −3
10 10
−4 −4
10 0 1 10 0 1
10 10 10 10
k x /k 0 k x /k 0
(a) (b)
Fig. 5.9 (a) Same as in Fig. 5.7(b) but for thickness d/λ = 2. The only good of the electrostatic limit is that it also predicts
a cutoff, but it fails to give any reasonable quantitative agreement with the full solution. The cutoff is less than 2k0 even for
a deviation from εr = −1 as small as 10−5 . (b) Same as in Fig. 5.8(b) but for thickness d/λ = 2. The electrostatic limit fails.
The cutoff is less than 2k0 even for a deviation as small as 10−5 from µr = −1
1 1
10 10
0 0
10 10
transfer function |T|
−3 −3
10 10
−4 −4
10 0 1 10 0 1
10 10 10 10
k x /k 0 k x /k 0
(a) (b)
Fig. 5.10 Keeping refractive index n = −1. Deviations in ε and µ correlated. Thickness d/λ = 0.1 (a) and 2 (b). Electrostatic
limit works only for a thin slab (a); it is too inaccurate for a thick slab (b)
to the cases when only one material parameter is different from that of
the perfect lens? The transfer function for the propagating components
is, apparently, better than in previous examples. The reason is quite
obvious. The refractive index n being minus unity means merely that
the propagating components exhibit negative refraction at an angle equal
to the angle of incidence. So refraction is at the desired angle but the
reflection coefficient is no longer zero, as for the perfect transfer function,
because the impedance of the negative-index medium,
r r r
µ µ0 µ0
η= = (1 ± δ) 6= = η0 , (5.31)
ε ε0 ε0
differs from that of free space. The good news is that the electrostatic
limit seems to work. At least as long as the cutoff frequency is large
relative to k0 . Remember, the electrostatic limit implies that both κ2
and κ1 are equal and can be approximated by kx . In the case considered
κ1 and κ2 are indeed equal to each other. But the other assumption,
that they can be approximated by kx , is only valid for kx ≫ k0 . So
when the cutoff is too low—and this is, for instance, the case for the
curve with δ = 1, e.g. εr = −2, µr = −0.5—the electrostatic limit fails
yet again. This failure of the electrostatic limit is even more obvious for
a thicker slab (see Fig. 5.10(b)). As far as resolution is concerned, just
as in previous examples, it deteriorates logarithmically with |δ| and is
inversely proportional to the slab thickness.
4
Many authors like to refer to this fact
5.3.5 Near-perfect? Near-sighted!
as an ‘uncertainty principle’ using the
quantum-mechanical language to de-
Suppression of the high-frequency wave components limits the resolu-
scribe the fact that the narrower the tion of the lossy slab acting as an imaging system since the spatial size
object the broader is its Fourier spec- of the image (say, its half-width ∆) and the spectral width of the cor-
trum. responding spectrum, δ, are inversely proportional to each other.4 The
5.3 Near-perfect lens 173
12 12
exact exact
low-loss approx. low-loss approx.
10 10
−5
−5
H’’=10
P’’=10
0
k x /k 0
cut−o k x /k
8 8
−3 −3
P’’=10 H’’=10
6 6
cut−o
P’’=0.1
4 4 H’’=0.1
2 2
8 −5
H’’=P’’=10
6 −3
H’’=P’’=10
4 H’’=P’’=0.1
0 0.5 1 1.5 2
thickness d/O
(c)
Fig. 5.11 Near-sighted lens. Cut-off frequency versus slab thickness. Losses in µ only (a), in ε only (b) and in both ε and µ
(c). Solid line: exact calculation, circles: low-loss approximation from Podolskiy and Narimanov (2005)
for lossy µ than for lossy ε. As thickness increases, even small losses
either in ε (Fig. 5.11(a)) or in µ (Fig. 5.11(b)) or in both ε and µ (Fig.
5.11(c)) eventually bring the cutoff down to k0 . Note that the defini-
tion chosen for the cutoff frequency assumes that the transfer function
reaches the magnitude of 1/2. For any loss, there is a critical thickness
for which the definition of a cutoff becomes meaningless as |T | < 1/2
5
Note that the plots have been calcu- for any kx (noticeable in each of Figs. 5.11(a)–(c)5 as an abrupt end for
lated from the exact expressions but curves with the largest value of loss). An important conclusion follows:
the results calculated from the approx-
imation of Podolskiy and Narimanov
while a thin lens of d ≪ λ can provide subwavelength resolution, the
(2005) are practically identical. resolution of a lossy far-field lens with d ≥ λ does not perform better
than usual optical devices (Podolskiy and Narimanov, 2005). Thus, the
expression for the resolution limit of a LHM-based lens proves that the
area of its subwavelength performance is usually limited to the near-field
6
Podolskiy and Narimanov (2005) in- zone.6
troduced as ‘near-sighted’ the flat near- We can put it in a different way. The consequence of the extreme
perfect lens, a catchy name that has
often been used. Interestingly, it was
dependence of the resolution on the deviation from the perfect-lens con-
Narimanov’s group that later and in dition is that in order to achieve a cutoff frequency not lower than 10
parallel to Engheta’s group proposed k0 , a slab of 0.1 λ thickness would require a loss level of about 0.002,
the concept of a ‘far-sighted’ cylindri- but a slab of 0.67 λ would tolerate a loss of no greater than 6 × 10−19
cal lens described below in Section 5.6
(see also Jacob et al. 2006; Salandrino
(see also Smith et al. 2003; French et al. 2006).
and Engheta 2006).
60
Hr =−1−G, Pr =−1
Hr =−1−jG, Pr =−1
50 Hr =−1−G, Hr Pr =1
Hr =−(1+G)-1, Hr Pr =1
40 Hr =Pr =−1−G
0
Hr =−1, Pr =−1−G
cut−o k /k
x
30 Hr =−1, Pr =−1−jG
ulations. They find that the finite discretization mesh acts like imaginary
deviations from µr = εr = −1 and leads to a cross-over in the transfer
function from constant values to exponential decay around kxcutoff lim-
iting the attainable super-resolution. A qualitative model is proposed
that is capable of describing the impact of the discretization. kxcutoff is
found to depend logarithmically on the mesh constant, in qualitative
agreement with the transfer matrix method simulations.
|T|
x
1 (a)
0 0.1
10
1
|E |
|T|
x
1 (b)
0 0,1
10
1
|E |
|T|
x
1 (c)
0 0.1
10
1
|Ex|
|T|
1 (d)
0 0.1
−50 0 50 1 10 100
x [nm] kx/k0
Fig. 5.13 Imaging for a low-loss (ε′′ = 10−2 ) silver slab of various thicknesses. The object (dashed lines) and image (solid
lines) are shown in the left-hand column. The amplitude of the transfer function is shown in the right-hand column. d = 60,
40, 20, 10 nm (a)–(d)
|T|
x
1 (a)
0 0.1
10
1
|Ex|
|T|
1 (b)
0 0,1
10
1
|Ex|
|T|
1 (c)
0 0.1
10
1
|Ex|
|T|
1 (d)
0 0.1
−50 0 50 1 10 100
x [nm] kx/k0
(a) (b)
2 1
A: k =5k
x 0
|Ex(kx)|
1.5
A B 0.5
|T|
1
C
0.5 0
0 1
0 5 10 15 20 B: k =10k
x 0
k /k
|Ex(kx)|
x 0
0.5
1
C: kx=15k0
|Ex(kx)|
0.5
0
0 10 20 30 40
z [nm]
Fig. 5.15 Transfer function (a) and the evolution of individual Fourier components for a number of spatial frequencies indicated
by the points A, B and C (b)
1
full
2 input
10 ES
full
0.8
ES
transfer function |T|
air
2
intensity |Ex|
0 0.6
10
0.4
−2
10
0.2
−4
10 0
0 10 20 30 40 50 −80 −60 −40 −20 0 20 40 60 80
kx/k0 x [nm]
(a) (b)
Fig. 5.17 (a) The transfer function and (b) the image for the 5-18-5 nm configuration
our case it was merely by trial and error) how altering the geometry
affects the transfer function. We shall show here the object, the image
and the transfer function in Fig. 5.17 for the 5–18–5 configuration, all
dimensions in nm. Interestingly, the transfer function between the two
SPP resonances increases. If higher values of kx dominate that means
that the image is made up by higher harmonics and, consequently, it
may be narrower than the object. Compression has been achieved.
40nm
80nm
z-axis
(a)
Object Image
intensity - |V | 2 intensity - |V | 2
Image with
silver slab
Image without
silver slab
Fig. 5.18 Reproduction of Fig. 2 of Pendry’s paper on the perfect lens (2000). (a) Plan view of the new lens in operation. A
quasi-electrostatic potential in the object plane is imaged by the action of a silver lens. (b) The electrostatic field in the object
plane. (c) The electrostatic field in the image plane with and without the silver slab in place. The reconstruction would be
c 2000 by the American Physical Society
perfect were it not for finite absorption in the silver. Copyright
in Fig. 5.19(b). The same conclusion can be drawn from the transfer
functions plotted in Figs. 5.20(a) and (b). The one for the 40-nm lens
is much narrower. There is roughly a factor of two between the cutoff
frequencies.
The numerical mistake is unfortunate but that is not the point we wish
to make. It is the electrostatic approximation that we wish to examine
more closely. Pendry based the electrostatic approximation on the fact
that in the examples investigated all dimensions were much smaller than
the electromagnetic wavelength. The question is whether small dimen-
sions in themselves are sufficient for the accuracy of the electrostatic
approximation to the exact result. Pendry certainly came to the con-
clusion that for a thin lens the electrostatic approximation, which in the
lossless case predicts a flat transfer function and perfect reconstruction
of the image, is correct. This is suggested both in the figure caption of
his Fig. 2 (our Fig. 5.18) and in the text where he writes: ‘Evidently, only
the finite imaginary part of the dielectric function prevents ideal recon-
struction’. This statement can be checked. Comparing the electrostatic
approximation with the exact result in Fig. 5.19(a) we can see that the
plots are similar. The same is true if we look at the transfer functions:
182 Subwavelength imaging
0.1 0.18
"20−40−20" "10−20−10"
object object ES
0.08
full
ES
2
2
intensity |Ex|
intensity |E |
x
0.06
full
0.09
0.04
air
air
0.02
0 0
−100 −50 0 50 100 −100 −50 0 50 100
x [nm] x [nm]
(a) (b)
Fig. 5.19 The object (Pendry’s double step function shown as gray bars), the image from the full solution (solid line) and from
the electrostatic approximation (dashed line) and the image in the absence of the slab (dashed-dotted line). Silver slab with
εr = −1 − j 0.4 of the thickness d = 40 nm (a) and d = 20 nm (b)
the electrostatic approximation is quite close to the exact result, see Fig.
5.20(a). Thus, the claim that the electrostatic approximation is applica-
ble might be acceptable. However, if we think of the physical reason for
the growth of the non-propagating waves, namely the excitation of SPP
resonances, then this argument cannot be correct. It is incorrect on two
counts. In the absence of losses the transfer function will tend to infin-
ity at the value of kx that satisfies the resonance condition. Secondly,
the resonance is always followed by a cutoff. The cutoff will limit the
resolution but the resonance will do something worse: it will lead to an
image in which the spatial harmonic causing the resonance dominates.
In the next example we shall show what happens when the object is
the same and the thickness of the lens is the same but we shall drastically
reduce losses by taking ε′′ = 10−4 . The field distribution in the image,
and the transfer function are shown in Fig. 5.21. The electrostatic ap-
proximation yields a very good reproduction of the object, much better
than the one achieved for ε′′ = 0.4. Alas, the results calculated from
the exact solution show a much wider image caused not so much by the
cutoff but by the SPP resonance. There is also considerable discrepancy
now between the cutoff points of the approximate and exact results.
We may thus say categorically that the electrostatic approximation
fails when the losses are zero or very small. For high enough losses the
electrostatic approximation works reasonably well. High losses are good
because they blunt the resonances.
The mathematical reason for the failure of the electrostatic approx-
imation for ‘epsilon-negative-only’ materials (first pointed out by Sha-
monina et al. (2001)) was already given in Section 2.12.4. It comes down
to the case of which one of two small quantities is negligible.
Whether the electrostatic approximation works or not is not an im-
portant matter. One can always rely on the exact expression. The
5.4 Negative-permittivity lens 183
2 2
10 10
full full
ES ES
1 1
10 10
0 0
|T|
|T|
10 10
−1 −1
10 10
−2 −2
10 10
0 5 10 15 20 0 5 10 15 20
k /k k /k
x 0 x 0
(a) (b)
Fig. 5.20 (a) Transfer function for the same object for a slab of silver, εr = −1 − j 0.4. Thickness of lens (a) d = 40 nm and
(b) d = 20 nm
2
1.6 10
full
1.4 ES
1
1.2 10
full
1
|E |2
0
|T|
0.8 10
x
0.6
−1
0.4 10
0.2
ES −2
0 10
−200 −100 0 100 200 0 5 10 15 20
x [nm] kx/k0
(a) (b)
Fig. 5.21 (a) Image and (b) transfer function for a slab of silver, εr = −1 − j 10−4 . Lens thickness d = 40 nm
R
(a) Z (b)
L
Y R
30 X
4
L
2 R
HF oscillator
1 9 L
3
10.5
HF
receiver
90
Y
9 2 4
1
3
Fig. 5.22 Schematic representation of X
the experimental setup. From La- 105
c
garkov and Kissel (2005). Copyright
11
2005 by the American Physical Society
(a) (b) H
UV exposure E
Conformable mask
UV exposure
Tungsten Conformable glass
Conformable
Silver lens
mask Dielectric spacer (PMMA)
mask
mask support
Planar lens (silver)
Vacuum contact Dielectric spacer (PMMA or SiO2)
Resist
Resist stack
Vacuum contact
BARC
Resist
Resist stack
Silicon substrate BARC
z
Silicon substrate
Y
x
Fig. 5.23 Schematic representation of two experimental setups comparing hard-contact and flat-lens lithography. From Melville
c 2006 Elsevier Ltd
et al. (2006). Copyright
PMMA9 –Ag–SiO2 with thicknesses 25–50–10 nm. The wavelength of 9 poly-methyl metachrylate
operation was 365 nm. The patterns that were imaged included isolated
lines, line pairs, gratings and arrays of dots with some feature sizes
smaller than 100 nm. Gratings with periods down to 145 nm were
resolved.
In a later paper by Melville et al. (2006) the relative merits of hard-
contact lithography (often referred to as evanescent near-field optical
lithography) and flat-lens lithography were explored. The aim was to
project an image into the photoresist. Schematic drawings of the two
somewhat different experimental setups are shown in Figs. 5.23(a) and
(b). They found that above the diffraction limit both hard-contact litho-
graphy and the silver lens perform in a similar manner. Below the dif-
fraction limit the hard/contact lithography produces clearer images in
the resist. The proposed explanation is that it is due to losses in the sil-
ver and to (insufficient) quality of the silver deposition. Both single- and
double-layer lenses were examined by Melville and Blaikie (2006). They
found that for the same total thickness of silver, the resolution limit is
qualitatively better for the double-layer stack. However, pattern fidelity
is reduced in the double-layer experiments because of increased surface
roughness. The same group was also concerned with modelling the sil-
ver superlens (Blaikie and McNab, 2002; Melville and Blaikie, 2007) in
order to assist the experimental work.
In another set of experiments with a silver lens by Fang et al. (2005)
the object was an array of 60-nm wide slots of 120-nm pitch next to
the word NANO inscribed onto a chromium10 screen (see Fig. 5.24). 10
Note that the plasma frequency of
The illumination coming from below is at a wavelength of 365 nm. The chromium is sufficiently far away from
that of silver so as not to interfere with
silver lens is separated from the object by a layer of PMMA and the the superlensing effect.
image is projected into a layer of photoresist. Detection is achieved by
186 Subwavelength imaging
PR
Ag
PMMA
Quartz Cr
Fig. 5.25 Schematic representation of an imaging experiment with slits for detection. From Korobkin et al. (2006a). Copyright
c 2006 Optical Society of America
80 80 80
60 60 60
z [nm]
40 40 40
20 20 20
0 0 0
−40 0 40 −40 0 40 −40 0 40
x [nm] x [nm] x [nm]
1 1 1
0.2 02 0.2
0 0 0
−40 −20 0 20 40 −40 −20 0 20 40 −40 −20 0 20 40
x [nm] x [nm] x [nm]
Fig. 5.28 Superlens of the overall thickness of 80 nm. Top: Poynting vector streamlines. Bottom: object (dashed line) and
image (solid line). (a) Multilayer lens of four 10-nm silver layers separated by 10-nm layers of air. (b) Two 20-nm silver layers
separated by 20-nm air layer. (c) Single lens of 40-nm thick silver slab. εr = −1 − j 0.1
1
10 0.5
1 layer
|Ex|2
1 layer
0
2 layers 0.2
2 layers
|E |2
0
|T|
x
10
4 layers
0
0.06 4 layers
|E |2
x
−1
10 0
0 1 −100 −50 0 50 100
10 10 x [nm]
kx/k0
(a) (b)
Fig. 5.29 Superlens with one, two and four 10-nm thick layers of silver. Transfer function and image distribution
quality. For two silver layers of 20 nm each (Fig. 5.28(b)) the image
becomes wider, and even wider when there is only one single silver layer
of 40 nm thickness (Fig. 5.28(c)). The above example offers a clear proof
for the efficacy of the multilayer lens both by the Poynting vector picture
and by the improved fidelity of reproduction.
There can be no doubt, however, that the multilayer lens is not a
magic cure for all the ills of the silver lens. The image quality is bound
to deteriorate with the number of layers. We shall show here a further
example that considers not only the fidelity of reproduction but resolu-
tion as well, and look at the image after 1, 2 and 4 layers. The object
consists of the two step functions 50 nm apart shown already in Fig.
5.18(b). For the same amount of loss as in the previous example we plot
the transfer function and the image intensity in Figs. 5.29 and 5.30 for
d = 10 nm and 20 nm, respectively. For d = 10 nm there is a slight
deterioration of the image as the number of layers increases from 1 to 4.
For d = 20 nm the two peaks can be clearly resolved for 1 layer, there
is some deterioration for 2 layers and the object is unrecognizable when
there are 4 layers. The conclusions are the same as before. For good
imaging, thin layers are needed.
We have now covered the major principles upon which multilayer
lenses work. For further discussions see Feng et al. 2005; Feng and
Elson 2006; Dorofeenko et al. 2006; Webb and Yang 2006; Melville and
Blaikie 2007.
Another strand of thought that could lead to multilayer imaging comes
from traditional treatments of finding the effective dielectric constant
(for a review see, e.g., Bergman 1978). In its electrical engineering con-
text see the detailed analysis by Wait (1962). In optics they appear
in the classical work of Born and Wolf (1975). They are called strati-
fied media described by transfer matrices and used mainly as filters and
matching elements. For small modulation of the dielectric constant they
5.5 Multilayer superlens 191
1
10
0.12 1 layer
|E |2
x
0
1 layer
0.03 2 layers
2 layers
2
|Ex|
0
|T|
10
4 layers
0
0.01 4 layers
|Ex|2
−1
10 0
0 1 −100 −50 0 50 100
10 10 x [nm]
kx/k0
(a) (b)
Fig. 5.30 Superlens with one, two and four 20-nm thick layers of silver. Transfer function and image distribution
provide holographic filters (see, e.g., Solymar and Cook 1981; Solymar
et al. 1996), and they also appear in photonic bandgap materials in
their 1D version13 (Scalora et al., 1998). 13
Scalora et al. (1998) were the first
In the simple case of a multilayer made up by isotropic materials of to realize that the transparency of a
stack of metal–dielectric layers could
equal thickness the derivation of the effective dielectric constants is given be much higher than that of a single
in Appendix I. It results in an anisotropic material with longitudinal and one. They also introduced the term
transverse dielectric constants given as ‘resonance-enhanced tunnelling’.
1 −1 1
ε−1 ε1 + ε−1
z = 2 and εt = (ε1 + ε2 ) . (5.34)
2 2
The idea is that for sufficiently thin materials simple averaging works.
For εr1 = 1 and εr2 = −(1 + δ) the effective dielectric constants,
2
εz ≃ and εt ≃ −δ , (5.35)
δ
turn out to be of opposite sign. Taking εr1 = 1 and εr2 = −1 we find
the interesting outcome that
εz = ∞ and εt = 0 . (5.36)
The consequence is that an image pasted on the front surface will appear
at the back surface. It is an interesting concept worth mentioning but it
is a poor approximation. A layer thickness of 20 nm (for which imaging
is shown in Fig. 5.30) is already too thick for the approximation to work.
For further discussions see Wood et al. (2006).
14
The difference between the two methods may be best seen by looking Although we have a healthy scep-
ticism concerning the results of many
at the Poynting vector. The above effective-medium approximation gives simulations we are happy with this one
a straight line, whereas the one based on the transfer matrices leads to because it comes from a very reliable
the diverging-converging picture of Fig. 5.28. The result of a numerical source (our group) and provides further
simulation14 by CST MICROWAVE STUDIO (Tatartschuk, 2008) is proof for the correctness of the physical
picture presented in Fig. 5.28.
shown in Fig. 5.31.
192 Subwavelength imaging
−2.6
100
−2.8
80
−3
60
−3.2
Y [nm]
−3.4
40
−3.6
20
−3.8
0
−4
−20 −4.2
−60 −40 −20 0 20 40 60
X [nm]
Fig. 5.31 Numerical simulation of the streamlines and amplitude of the electric field for a multilayer silver lens. From
Tatartschuk (2008). For coloured version see plate section
400
−4.5
200
−5
0
Y [nm]
−5.5
−200
−6
−400
−6.5
−600
−7
−600 −400 −200 0 200 400 600
X [nm]
Fig. 5.32 Magnifying superlens. Numerical simulation of Poynting vector streamlines emanating from two subwavelength
objects. εr = −1 − j 0.4. From Tatartschuk (2008). For coloured version see plate section
(a) (b)
Fig. 5.33 (a) Schematic of magnifying optical hyperlens and numerical simulation of imaging of subdiffraction-limited objects.
(b) An arbitrary object ‘ON’ imaged with subdiffraction resolution. Line width of the object is about 40 nm. The hyperlens is
made of 16 layers of Ag-Al2 O3 . From Liu et al. (2007b). Courtesy of Prof. Xiang Zhang of University of California at Berkeley.
Copyright
c 2007 AAAS. For coloured version see plate section
0.247
Magnifying
superlens
‘‘Super’’ ~70 nm
Plasmon illumination
Z (a.u.)
images
Plasmon
Phase-matching structure Sample rays –0.041
Fig. 5.34 (a) Schematic of the magnifying superlens integrated into a conventional microscope. The plasmons generated by
the phase-matching structure illuminate the sample positioned near the centre of the superlens. The lateral distance between
the images produced by the alternating layers of materials with positive and negative refractive index grows with distance along
the radius. The magnified images are viewed by a regular microscope. (b) The cross-section of the optical image indicates
resolution of at least 70 nm or ∼ λ/7. Z, optical signal; X, distance. From Smolyaninov et al. (2007). Copyright
c 2007
AAAS
n1 d1 = −n2 d2 , (5.37)
where d1 and d2 are the widths of regions 1 and 2. The parameters were
so chosen by the authors that the above relationship is satisfied at the
wavelength of excitation.
The objects were 2 or 3 rows of PMMA dots in the radial direction
positioned near the centre of the lens. For the three-row case the distance
196 Subwavelength imaging
Regular microscope
objective
Dielectric droplet
acting as a mirror
Surface plasmon-
Sample produced image
Focal point
Gold film
Glass prism
Fig. 5.35 Magnifying superlens. From
Smolyaninov et al. (2005). Copyright
c 2005 by the American Physical So- Laser illumination
ciety
between them at the input of the lens was 70 nm. Upon illumination
they gave rise to three divergent SPP ‘rays’. The field distribution at
the output of the lens was then large enough (10 times as large as at
the object) to be further magnified by a conventional microscope. The
microscope output showing three distinct peaks is plotted in Fig. 5.34(b).
An entirely different approach to image magnification was proposed
by Smolyaninov et al. (2005) based on earlier work (Smolyaninov, 2003;
Smolyaninov and Davis, 2004). The essential ingredient is a parabolic-
shaped glycerine droplet on a metal film as shown in Fig. 5.35. The
waves involved are SPPs excited by laser light. In the experiments, light
of 502 nm wavelength was used at which the SPP wavelength is about
70 nm giving an effective index of refraction of about 7. Imaging is due
to reflection of the SPPs by the parabolic boundary. An object placed
in the vicinity of the focal point will have an image as shown in Fig.
5.35. The authors successfully imaged arrays of nanometric pinholes in
the gold film. The field distribution is then viewed from above by a
conventional microscope.
5.7 Misconceptions
Whatever our chosen field is we like to understand it in simple qualitative
terms. What can we say about the superlens? It gives a perfect image
but a perfect image cannot be obtained in practice due to losses, im-
perfections, time delay and the granular nature of the metamaterial. It
is easy to understand the consequence of granular structure. Surely, we
cannot obtain a resolution that is better than the period in the meta-
material. The influence of other effects is less obvious. Some of the
relationships are quite complicated and superficial references to them
may not be adequate. The misconceptions, quoted below, may only be
entertained by a small minority but we think it is worth including them
in this section.
Let us first look at the superlens with εr = µr = −1.
5.7 Misconceptions 197
(a) (b)
Fig. 6.1 (a) The SRR-loaded square waveguide. (b) Sketch of the experimental setup. From Marques et al. (2002a). Copyright
c 2002 by the American Physical Society
0 0
10 10
Backward wave
20 pass band
20
30 30
40 Waveguide Cut Off 40
S21(dB)
Waveguide cut off
S21(dB)
50 50
60 Stop band with negative transversal 60
70 permeability 70
80 80
90 90
100 100
6 7 8 9 10 11 12 13 14 15 6 7 8 9 10 11 12 13 14 15
Frequency (GHz) Frequency (GHz)
(a) (b)
Fig. 6.3 Measured transmission coefficient. There is (a) a stop band when the cutoff frequency of the waveguide is below
the elements’ resonant frequency; (b) a pass band when the cutoff frequency of the waveguide is above the elements’ resonant
frequency. From Hrabar et al. (2005). Copyright
c 2005 IEEE
∂Hz ∂Hy
− = j ωε0 Ex , (6.2)
∂y ∂z
∂Ex
= −j ωµ0 µyy Hy , (6.3)
∂z
∂Ex
= j ωµ0 Hz . (6.4)
∂y
The above set of differential equations are still linear with constant
coefficients, hence they can be solved without difficulty. Using again the
assumption of a wave solution and satisfying the boundary conditions
on the waveguide wall we end up with the relationship
π 2
kz2 = µyy k02 − . (6.5)
a
202 Phenomena in waveguides
This equation differs from that of eqn (1.37) by having µyy on the
right-hand side. But that makes all the difference. We can now explain
the experimental results shown in Figs. 6.2 and 6.3. Above the cutoff
frequency k 2 − (π/a)2 > 0, hence µyy being negative will result in kz2
being negative and, consequently, in a stop band. Below the cutoff
2
frequency k 2 − (π/a) < 0, hence µyy being negative results in kz2 being
positive, and, consequently, in a pass band.
The original explanation provided by Marques et al. (2002a) is some-
what different. They quoted Rotman (1962) who simulated plasmas
not only by metallic rods (discussed in Section 2.4) but also by cutoff
waveguides. The claim was that a cutoff waveguide behaves analogously
to a plasma below its plasma frequency. Hence, Marques et al. argued, if
a set of elements, capable of producing negative permeability, is inserted
into a cutoff waveguide that should lead to propagation, the situation
being equivalent to a medium having both negative ε and negative µ.
This argument can be made quantitative by assuming, with Rotman,
that a cutoff waveguide could be regarded as having an effective relative
dielectric constant of
2 ω 2
λ p eff
εr eff = 1 − = 1− , (6.6)
2a ω
where
cπ
ωp eff = (6.7)
a
is the effective plasma frequency.
In general, the propagation coefficient of an electromagnetic wave in
a homogeneous medium can be written as
ω2
kz2 = µr εr , (6.8)
c2
where the wave happens to propagate in the z direction.
Using the definition of the effective relative dielectric constant in the
form of eqn (6.6) we find that
" 2 #
λ π 2
2 2 2
kz = k0 µr 1 − = µr k0 − , (6.9)
2a a
which is of the same form as eqn (6.5). Thus, we may conclude that the
heuristic derivation in terms of the effective dielectric constant leads to
the same result as the one based on the field equations. The advantage
of the heuristic derivation is that it links the phenomena observed in
waveguides to those in left-handed materials allowing the possibility that
both ε and µ may be negative.
Note, however, that eqn (6.9) does not represent the whole truth. It
has µr instead of µyy . Looking at eqn (6.9) one might come to the con-
clusion that propagation is possible in a cutoff waveguide filled with an
isotropic negative-permeability material. And that would be the wrong
conclusion. Equation (6.5) suggests that in order to have propagation
6.2 Propagation in cutoff waveguides 203
(a) (b)
(c) (d)
Fig. 6.4 Transformation of the waveguide problem using the image principle. From Belov and Simovski (2005b). Copyright
c 2005 by the American Physical Society
in cutoff waveguides.
It should be mentioned that cutoff waveguides that can propagate
electromagnetic waves have potential for applications because their di-
mensions may be much smaller than waveguides working above cutoff.
However, devices in cutoff waveguides are in no sense new. In the 1960s
and 1970s it was a hot subject. The research resulted in a number of
microwave devices: not only broadband filters for which it is most suit-
able but also circulators, frequency mixers, multipliers, phase shifters,
etc. The main aim was to reduce weight and, often, trade weight for
loss. For a review see Craven (1972).
Finally, we need to face the undisputable truth that the explanation
of propagation in a cutoff waveguide in terms of negative transverse
permeability is not new either. The same explanation was suggested by
Thompson (1955) whose paper in Nature covered both theory and exper-
iment. The experiments were conducted in a circular metallic waveguide
loaded by a longitudinally biased ferrite rod. The phenomenon was fur-
ther investigated by Thompson (1963).
Hollow circular waveguides with perfectly conducting walls and loaded
along their axes by a periodic array of thin dielectric (ε > 0) and metallic
(ε < 0) layers, were investigated by Govyadinov and Podolskiy (2006)
in the infra-red region. Note that the anisotropy is this time in the
axial direction. The authors showed that it was possible to obtain both
positive and negative effective indices of refraction. When n < 0 then
propagation is only possible when the waveguide radius is sufficiently
small. Hence, if the radius of the waveguide (with the load inside) is
gradually reduced waves can still propagate and the result is high power
compression.
G
w Fig. 6.6 Layout of the SRR loaded
coplanar waveguide. From Martin et al.
(2003a). Copyright
c 2003 American
Institute of Physics
Fig. 6.7 Photograph of a four-element filter based on complementary SRRs. (b) Simulated and (c) measured S11 (thin line)
c 2004 IEEE
and S21 (thick line) coefficients. From Falcone et al. (2004a). Copyright
6.4 Tunnelling
Tunnelling is a well-known phenomenon in a cutoff waveguide. It means
that in spite of the wave being evanescent a certain proportion of the
input power can get through to the output. As we know, there can be a
growing wave in a negative-index material. Would the same thing apply
in a cutoff waveguide? The analysis was done by Baena et al. (2005c).
A schematic representation of the waveguide is shown in Fig. 6.8. As
may be seen, there are 5 sections. The waveguide dimensions are such
that in the empty sections, 2 and 4, the wave cannot propagate. In
sections 1 and 5, both taken as infinitely long, the waveguide is above
cutoff by choosing ε to be high enough. In section 3 there is a negative-
index material. The length of each section is li . The electromagnetic
wave is incident from section 1 in the TE10 mode. Since the size of the
waveguide is the same everywhere the TE10 mode remains unchanged
in all the other sections.
It is shown that in the lossless case for isotropic materials the condi-
tions for perfect tunnelling (100% transmission) are as follows:
l2 α2 + l3 α3 + l4 α4 = 0 (6.10)
and
Z2 = Z3 = Z4 , (6.11)
l3 = l 1 + l 2 . (6.12)
It may now be noticed that the conditions for perfect tunnelling are the
same as for perfect imaging.
6.4 Tunnelling 207
2.5 5
1 2 3 4 5 1 2 3 4
2.0 4
Field amplitude
Field amplitude
1.5 3
1.0 2
0.5 1
0.0 0
–10 –5 0 5 10 15 –10 –5 0 5 10 15
y (cm) y (cm)
(a) (b)
Fig. 6.9 Field distribution inside the structure of Fig. 6.8 when section 3 is filled by an isotropic metamaterial with εr = µr = −1.
(a) Perfect tunnelling when l3 = 2l2 = 2l4 . (b) Field distribution when l4 → ∞. Dashed lines show the field amplitude when
section 3 is empty. From Baena et al. (2005c). Copyright
c 2005 by the American Physical Society
The distribution of the electric field in the waveguide under these con-
ditions was calculated by Baena et al. (2005c). The geometrical dimen-
sions are a = 24 mm, l3 = 30 mm and the frequency is 5 GHz. We have
the conventional tunnelling situation when ε2 = ε3 = ε4 . Then, there
is a standing-wave pattern in section 1, and the wave declines beyond
that, as shown in Fig. 6.9(a) (dashed line). Under optimum conditions
(continuous line) there is now no standing-wave pattern and the input
amplitude may be seen to be the same as the output amplitude, the
sign of perfect tunnelling. Any deviation from the optimum conditions
leads to rapid degradation of the transmitted power. In Fig. 6.9(b) the
field distribution is shown for two different conditions: when section 3 is
empty (dashed line) and when section 4 extends to infinity (continuous
line). As may be expected, there is an exponential decay in both cases,
but in the latter one the growing wave in section 3 is still there.
A somewhat analogous situation was investigated by Alu and Engheta
(2003) in free space for a periodic array of ε-negative materials paired
with µ-negative materials. Among others they found conditions for per-
fect tunnelling.
Perfect tunnelling in free space through a negative-positive-negative
permittivity medium was investigated by Zhou et al. (2005b) both exper-
imentally and by simulation. They found that perfect tunnelling occurs
at two frequencies at which there is an increase in the magnetic field.
They showed also that the perfect tunnelling effect is insensitive to the
input angle of the incident wave.
A related problem, perfect tunnelling under conditions of frustrated
total internal reflection for pairs of metamaterials was investigated by
Zhou and Hu (2007).
208 Phenomena in waveguides
Shorted (via) stub inductor Fig. 6.11 Layout of the unit cell of
the microstrip left-handed transmission
line. From Caloz et al. (2004). Copy-
Unit cell c 2004 IEEE
right
other. For a detailed description of this coupler, for its design, theory
of operations and measurement results see Caloz et al. 2004; Caloz and
Itoh 2006.
2.5
2
one to two dimensions is quite straightforward. The two transmission
lines are then rectangular to each other and loading is introduced in
1.5 both directions as shown in Fig. 6.13, which is a direct generalization
1 of the 1D circuit of Fig. 6.10 to two dimensions. A number of aspects
0.5 of these 2D loaded transmission lines were investigated by Eleftheriades
and coworkers in a series of papers (Eleftheriades et al., 2002; Grbic and
0
0 5 10 15 20 25 30 Eleftheriades, 2003b; Grbic and Eleftheriades, 2003c; Grbic and Eleft-
Cell number (column) heriades, 2003d; Grbic and Eleftheriades, 2004; Iyer et al., 2003).
We shall describe here in some detail one of their imaging experiments
Fig. 6.15 The measured vertical elec- (Grbic and Eleftheriades, 2004) in which microstrip lines (dielectric con-
tric field above row 0 at 1.057 GHz. stant, 3.0, substrate thickness 1.52 mm) are loaded by 2-pF capacitors
The vertical dashed lines identify the in parallel, and by 18-nH inductors in series. The size of the unit cell
source (column 0) and image (column
is equal to 8.4 × 8.4 mm. There are 19 × 5 unit cells for the loaded
10) planes, while the vertical solid
lines identify the interfaces of the left- transmission line (see Fig. 6.14 for the experimental setup), and 19 × 12
handed planar slab. The growth of unit cells on each side for the bare (unloaded) transmission lines that
the evanescent waves in the left-handed are analogous to right-handed media. The effective propagation coeffi-
lens is clear. From Grbic and Eleftheri-
c 2004 by the
cients of the loaded and unloaded media were designed to be equal in
ades (2004). Copyright
American Physical Society magnitude but opposite in sign at 1.00 GHz. In the experiment the
first unloaded medium is excited by a vertical monopole fed by a coax-
6.7 Imaging in two dimensions: transmission-line approach 211
Fig. 6.16 The measured vertical electric field detected 0.8 mm above the surface of the entire structure at 1.057 GHz. The
plot has been normalized with respect to the source amplitude (linear scale). From Grbic and Eleftheriades (2004). Copyright
c 2004 by the American Physical Society. For coloured version see plate section
ial cable through the ground plane. The monopole attaches the inner
conductor of the coaxial cable to the unloaded medium, while the outer
conductor of the coaxial cable is attached to the ground plane. The
excitation is in the central row 2.5 unit cells away from the interface.
The image is then supposed to be 2.5 unit cells away on the other side
of the loaded medium. The vertical electric field is detected 0.8 mm
above the surface using a short vertical probe connected to a network
analyzer. Its value above row 0 is plotted in Fig. 6.15 at the optimum
frequency of 1.057 GHz. The source plane at column 0 and the image
plane at column 10 are denoted by dotted lines, whereas the boundaries
of the loaded line are denoted by thick continuous lines. The value of
the field may be seen to be the same at the source and at the image. A
two-dimensional plot of the electric field is shown in Fig. 6.16. As may
be expected, the field is highest at the output interface. The lateral size
of the image was found to be equal to 0.21 effective wavelengths, in con-
trast to the 0.36 wavelengths that is the diffraction-limited value. This is
the first time that the classical limit was beaten by a flat negative-index
lens. It was made possible by the small size of the loading inductors and
capacitors, by the fact that the non-resonant structure was less lossy
than its resonant counterparts and finally because it was relatively easy
to produce an isotropic backward-wave medium.
This page intentionally left blank
Magnetoinductive waves I
7
7.1 Introduction 7.1 Introduction 213
7.2 Dispersion relations 214
Magnetoinductive (MI) waves have come about as a by-product of the
7.3 Matching the
research on metamaterials. The magnetic elements used in the first transmission line 217
realization of negative refraction by Shelby et al. (2001a) were split-ring 7.4 Excitation 217
resonators that could be modelled (Marques et al., 2002c) as LC circuits, 7.5 Eigenvectors and
provided the dimensions are small relative to the free-space wavelength. eigenvalues 218
We shall also assume, as we have done so far, that the separation of 7.6 Current distributions 220
the elements is also much smaller than the wavelength. This is often 7.7 Poynting vector 224
referred to as the quasi-static approximation. 7.8 Power in a MI wave 225
The simplest realization of an LC circuit as a metamaterial element 7.9 Boundary reflection
is a capacitively loaded loop, shown schematically in Fig. 7.1(a). Two and transmission 225
such loops close to each other are coupled to each other due to the mag- 7.10 Tailoring the dispersion
netic field of one loop threading the other loop and inducing a current characteristics:
in it (Fig. 7.1(b)). The presence of such coupling leads to waves that biperiodic lines 227
were called MI waves by Shamonina et al. (2002a). They belong to the 7.11 Experimental results 231
category of slow waves that propagate at a velocity less than that of 7.12 Higher-order
interactions 236
light.
7.13 Coupled one-dimensional
The properties of magnetoinductive waves have received considerable lines 238
attention since they were first proposed (Shamonina et al., 2002a). A
7.14 Rotational resonance 242
detailed study of the dispersion characteristics was carried out in Sha-
7.15 Applications 243
monina et al. 2002b, some other aspects of the theory were treated in
Syms et al. 2005a; Syms et al. 2005b; Sydoruk et al. 2005, experimental
results and comparison between experiments and theory were given in
Wiltshire et al. 2003b; Wiltshire et al. 2004b; Sydoruk et al. 2006;
Syms et al. 2006b; Sydoruk et al. 2007b; Syms et al. 2007a, devices
in MI waveguides were reported in Shamonina and Solymar 2004; Syms
et al. 2005c; Syms et al. 2006a and review papers were published in
Shamonina and Solymar 2006; Shamonina 2008. (a) (b)
Further properties of MI waves concerned with 2D effects and retarda-
tion will be presented in Chapter 8. Note that we have already discussed Fig. 7.1 (a) Capacitively loaded loop.
some simple aspects of MI waves earlier in this book in Section 1.23, (b) Two magnetically coupled elements
among wave solutions on four-poles (Brillouin, 1953): an example was
given concerned with magnetic coupling within one resonant four-pole.
The result was a dispersion equation of the form
ω0
ω= , (7.1)
2M
1+ cos(ka)
L
where ω0 is the resonant frequency, L is the inductance and M is the
214 Magnetoinductive waves I
M M
In = I0 e −j kna , (7.4)
7.2 Dispersion relations 215
N
N-1
z
3
2 3 N-1 N
1 1 2
y r0 Zt
x x
r0
a a
V1 V1 Zt
(a) (b)
k = β − jα, (7.5)
where β is the propagation constant and α is the attenuation coeffi-
cient. The dispersion equation (7.1) may then be separated into real
and imaginary parts yielding
ω02
1− + κ cos(βa) cosh(αa) = 0, (7.6)
ω2
1
− κ sin(βa) sinh(αa) = 0, (7.7)
Q
where κ is the coupling coefficient equal to 2M/L and Q is the quality 1
The simplicity of derivation and the
factor of the resonant circuit equal to ωL/R. Note that M may be ease with which such arrays can be
constructed (see Section 7.4) makes it
positive or negative, as discussed in Section 1.23. It is positive when the
likely that MI waves will be included
array is axial (see Fig. 7.3(a)) and negative in the planar configuration1 in the near future in the undergradu-
(Fig. 7.3(b)). ate syllabus of both physics and elec-
If losses are small enough (Q is high enough) then trical engineering courses. It would also
help to demystify the concept of back-
ward waves. As things stand the only
cosh(αa) = 1 , sinh(αa) = αa , (7.8) place where backward waves make an
appearance in the physics syllabus is in
which means that the dispersion equation for the phase change per ele- the derivation and interpretation of the
ment remains the same, and the losses per element are given as optical branch of acoustic waves. But
even then the emphasis is on the di-
1 atomic lattice and on the interaction
αa = . (7.9) with incident electromagnetic waves.
κQ sin(βa)
The backward-wave character, if at all,
It may indeed be expected that losses decline as the coupling coeffi- is only casually mentioned. In circuit
cient and the Q of the circuit increase. Dependence on βa may also be theory backward waves may appear (as
anticipated, that attenuation should be minimum at the resonant fre- in Section 1.23) but they are usually as-
sociated with lumped inductors and ca-
quency and should increase towards the band edge where βa = 0 or π. pacitors and not as waves in which the
The approximation, however, breaks down right at the band edge where magnetic and electric fields vary along
the attenuation should not be infinitely large. We can get the correct the direction of propagation. A look at
result by solving numerically eqns (7.6) and (7.7). the MI transmission line and the corre-
sponding dispersion curves would con-
We have already plotted in Fig. 1.18 the dispersion equations for κ = vince students that there is nothing ex-
±0.1 for the lossless case. We shall replot them in Figs. 7.4 and 7.5 for otic about backward waves.
Q = 40, 100 and 1000 including both the phase change and attenuation.
216 Magnetoinductive waves I
1.1 1.1
1.05 1.05
Q=40
0
0
Q=100
Z/Z
1 Z/Z 1
Q=1000
0.95 0.95
0.9 0.9
0 0.5 1 0 0.1 0.2 0.3 0.4
Ea/S Da/S
1.1 1.1
1.05 1.05
Q=40
0
0
Q=100
Z/Z
Z/Z
1 1
Q=1000
0.95 0.95
0.9 0.9
0 0.5 1 0 0.1 0.2 0.3 0.4
Ea/S Da/S
It may be seen that (i) the attenuation sharply increases as the quality
factor declines, and (ii) both phase change and attenuation vary rapidly
near the band edge.
It needs to be noted that the coupling coefficient is not necessarily
small. In principle it may be as high as 2. In practice, the highest value
measured so far (Syms et al., 2006b) is 1.5 in the axial configuration in
which the elements can be quite closely packed and −0.7 (Syms, 2006)
in the planar configuration. For κ = 1.5 and Q = 100 the dispersion
curves look quite different, as may be seen in Fig. 7.6. There is still
a lower cutoff frequency but the bandwidth now extends to arbitrarily
high frequencies with an asymptote at
1
βa = arccos , (7.10)
κ
1.5
ZZ0
0.5
0
0 0.5 1 Fig. 7.6 Dispersion for an axial array
Ea/S in the case of a strong coupling. κ = 1.5
ZT = j ωM e −j ka . (7.12)
It turns out that the terminal impedance is not a real constant, as for a
coaxial line for example, but it is complex and frequency-dependent.
7.4 Excitation
The dispersion equation is derived on the assumption that there is no
external excitation. As it happens, it is quite easy to include possible
218 Magnetoinductive waves I
V = ZI . (7.13)
The above equation may also be regarded as a generalized Ohm’s law.
V and I are N -dimensional vectors
Z0 j ωM 0 0 ... 0 0
j ωM Z0 j ωM 0 ... 0 0
Z=
... ... ... ... ... ... ... .
(7.15)
0 0 0 . . . j ωM Z0 j ωM
0 0 0 ... ... j ωM Z0
This is clearly a tri-diagonal matrix. The main diagonal elements are all
Z0 and the elements next to them (left, right, up or down) are all j ωM .
If the line is terminated by an impedance ZT then the last element in
the matrix (N th row, N th column) should be replaced by Z0 + ZT .
If the voltage excitation vector is known the current may be obtained
by inverting the Z matrix,
I = Z −1 V . (7.16)
Very often, only the first element is excited. In which case only the first
component of the voltage vector is finite and all the others are zero.
nlπ
In(l) = I(N ) sin (n = 0, 1, 2, . . . N, N + 1). (7.17)
N +1
The value of I(N ) can be found from the orthonormality condition
7.5 Eigenvectors and eigenvalues 219
(l) (m) 0 if l 6= m
I ·I = , (7.18)
1 if l=m
yielding
2
I(N ) = I0 . (7.19)
N +1
The corresponding eigenvalues may be obtained from their definition
as
lπ
λl = Z0 + 2j ωM cos . (7.21)
N +1
Having got the eigenvectors and eigenvalues we can find, in closed
form, the response to a general excitation. A voltage vector V of dimen-
sion N can be expanded in terms of the eigenvectors as
N
X
V= µl I(l) . (7.22)
l=1
µl = V · I(l) . (7.23)
N
X µl
I= I(l) . (7.24)
λl
l=1
It may be seen from the above equation that in the general case all
the eigenvectors are excited. In order to excite a single mode the cor-
responding eigenvalue must be close to zero.2 From eqn (7.21) the lth 2
When the eigenvalue is exactly zero
eigenvalue is zero when the current amplitude will be infinitely
large. This is a consequence of neglect-
ing losses.
lπ
Z0 + 2j ωM cos = 0. (7.25)
N +1
But this is nothing else than our dispersion equation. A single mode,
i.e. a single value of ka, can be excited for a discrete set of frequencies.
For N = 5 the normalized set of eigenvectors are plotted in Fig. 7.7(a).
For the axial configuration and for κ = 0.1 the corresponding values of
ka are shown on the dispersion curve of Fig. 7.7(b).
220 Magnetoinductive waves I
1
0
−1
1
0
−1
n
1
current I
0
1.1
−1
1
1.05
0
0
1
/
−1
1
0.95
0
−1 0.9
1 2 3 4 5 0 0.5 1
Fig. 7.7 (a) Eigenvectors and (b) the element number n a/
corresponding values of ka for a 5-
element array (a) (b)
1 1 1
01 0.1 0.1
Im(In )
Im(In )
Fig. 7.8 Travelling wave. Complex amplitude of currents along the axial array, normalized to that at element 1, at ω/ω0 =
0.9757, 1 and 1.0262 (a)–(c)
1 1
0.1 01
0.01 0.01
Im(In )
Im(In )
Im(In )
Fig. 7.9 Travelling wave. Complex amplitude of currents along the planar array, normalized to that at element 1, at ω/ω0 =
0.9757, 1 and 1.0262 (a)–(c)
|I |
1
0
1
|I |
2
0
1
|I |
3
0
1
|I |
4
0
1
|I |
5
pattern for the currents. Standing waves that have an integral number of
0 half-periods correspond to the eigenvectors of the system (see eqn (7.17)
1 and Fig. 7.7) which manifest themselves as resonances. In general, a
line consisting of N elements can have up to N resonances at N discrete
|I2|
frequencies given by eqn (7.25). It is clear that all the resonances are
0
1 within the pass band of the MI waves since the argument of the cosine
function can never exceed π.
|I3|
illustrate this we consider again the five-element axial line (for which
0 the eigenvectors and the corresponding resonances are shown in Fig.
1
7.7). Figure 7.10 shows the relative amplitudes of the currents in all five
|I5|
elements as a function of frequency for the case when the first current is
0 being driven by an external voltage. There are five resonant frequencies.
0.95 1 1.05 Note that the third resonance is missing in elements 2 and 4 and the
Z/Z0
second and fourth resonances are missing in element 3.
We can selectively exclude some of the resonances from being excited
Fig. 7.11 Resonances: symmetric exci- by choosing either a symmetric or an antisymmetric excitation. Figure
tation. Normalized current amplitudes 7.11 shows the currents for the case of a symmetric excitation, when
as a function of frequency for a five- the central loop is excited by an external voltage. Due to the symmetry
element axial structure with element 3 argument it is quite obvious why the second and the fourth resonances
being excited. Q = 100. κ = 0.1
that correspond to the antisymmetric eigenmodes are missing here.
Finally, Fig. 7.12 shows the currents for the case of an antisymmetric
7.6 Current distributions 223
1
|I1|
0
1
|I2|
0
1
|I3|
0
1
|I4|
0
1
Fig. 7.12 Resonances: antisymmetric
|I5|
1 1
n
n
current I
current I
0 0
excitation, when the first loop and the last loop are excited by external
voltages in antiphase. In this case the central loop is not excited at all,
and only the second and the fourth resonances are present.
(iii) Excitation outside the pass band
Excitation of an array outside the pass band results in evanescent
magnetoinductive waves. As follows from the dispersion equation (7.1)
there are two different branches of evanescent waves. For the branch
with β = 0 currents in all elements are in phase, while for the branch
with βa = π currents in neighbouring elements are always in antiphase.
We take as an example a lossless 41-element axial line (which is long
enough in order to disregard reflections from the unmatched ends) with
the central element being excited. Figure 7.13 shows the current distri-
butions for two values of the frequency, one from the lower and another
one from the upper stop band, which correspond to αa = 0.2. Obviously
224 Magnetoinductive waves I
1
0
P1
x/r
0 P2
−1
Fig. 7.14 Poynting vector streamlines
for a 5-element array. From Shamon- −1 0 1 2 3 4 5
c 2002
ina et al. (2002b). Copyright
z/r
American Institute of Physics 0
the current amplitudes decay strongly as we move away from the ele-
ment that is driven by an external source. This result has an important
consequence. When operating in the stop band, a magnetoinductive
line is able to replicate any pattern of excitation by the ‘pixel-to-pixel’
mechanism. This significant property is crucial for the design of the
magnetoinductive near-field lens (see Section 7.15.3).
In = j ωCVn . (7.27)
Assuming again a travelling wave and substituting the values of In−1
and In+1 in terms of In and making use of the dispersion equation the
expression for the stored energy simplifies to
ω02
W = L|In |2 . (7.28)
ω2
The group velocity may be obtained from the dispersion equation (7.1)
as
!
dω d ω0 ω0 a ω0 3
vg = = p = k sin(ka) , (7.29)
dk dk 1 + κ cos(ka) 2 ω
line 1 line 2
V1 Zt
Fig. 7.15 Two magnetoinductive lines
joined together a1 ab a2
and
e −j (ka)2 − e −j (ka)1
R = , (7.37)
e j (ka)1 − e −j (ka)2
2j sin [(ka)1 ]
T = . (7.38)
e j (ka)1 − e −j (ka)2
7.10 Tailoring the dispersion characteristics: biperiodic lines 227
This situation can arise only when the sole difference between the two
media is that the self-impedances (and hence the dispersion curves) are
different.
Note that eqn (7.37) appears in Tretyakov’s book (2003) but the un-
derlying physics is quite different in the two cases. Here, we are con-
cerned with the reflection and transmission of waves propagating in dif-
ferent periodic media, whereas Tretyakov’s expression is valid when a
plane electromagnetic wave is incident upon a periodic medium and
higher-order modes at the boundary can be disregarded. It is a coinci-
dence that the expressions are identical.
A further simplification can be obtained by considering the continuous
limit when (ka)1 , (ka)2 ≪ 1. Then, the reflection and transmission
coefficients reduce to
k1 − k2 2k1
R= and T = . (7.39)
k1 + k2 k1 + k2
The above expressions look quite familiar, occurring for example when
Schrodinger’s equation is solved for an electron wave incident upon a
potential barrier (see, e.g., Solymar and Walsh 2004). There are no
periodic media in that case, simply a wave incident from a medium with
propagation constant k1 upon a medium with propagation constant k2 .
For the reflection and transmission coefficients to make sense it is a
necessary condition that the power flow should be the same in lines 1
and 2. The power in lines 1 and 2 may be written as
1
1 − |R|2 ωM1 sin(ka)1 ,
P1 = (7.40)
2
1 2
P2 = |T | ωM2 sin(ka)2 . (7.41)
2
Substituting eqns (7.34) and (7.35) into the above equations for power
it may be shown, using a fair number of algebraic operations, that P1 =
P2 , i.e. the power across the boundary is conserved.
(2f/m) 1/2
frequency
(2f/M)1/2
M f m f M
0 0.5 1
2a propagation constant 2Ea/S
(a) (b)
Fig. 7.16 (a) Diatomic chain of atoms and (b) the phonon dispersion curve with two branches
coefficient 2β. The two waves are then in synchronism because they
propagate with the same phase velocity
ω 2ω
vp = = , (7.42)
β 2β
and the signal wave can be amplified by transferring power from the
pump wave.
In this section our aim is to show that we have some freedom over
the dispersion characteristics and, in particular, we can realize the syn-
chronism condition. The way to do it is suggested by the close analogy
between MI waves and acoustic waves. It is well known that the disper-
sion characteristics of diatomic solids differ greatly from those having
identical elements (see, e.g., Brillouin 1953). The consequence of having
a material in which two different atoms of different masses alternate (e.g.
NaCl) is the appearance of a new band, known as the optical branch, in
the dispersion characteristics. A sketch of the dispersion characteristics
of such a diatomic solid (Brillouin, 1953) is shown in Fig. 7.16.
We have now two pass bands. When it comes to acoustic waves in a
solid we have very little freedom. The two different masses are provided
by Nature and we have very little control over the way such masses
arrange themselves. However, when metamaterial elements are put next
to each other we can build every single element as we wish.
There are two obvious ways of achieving double periodicity: (i) To
change some parameter of the element (L or C resulting in a change of
resonant frequency) and (ii) to vary the distance between the elements,
which means that there will be two different mutual inductances. These
possibilities are shown schematically in Figs. 7.17(a) and (b) for the pla-
nar configuration and in Figs. 7.17(c) and (d) for the axial configuration.
Let us look at such a biperiodic line and apply Kirchhoff’s equations
to elements 2n and 2n + 1 (see Fig. 7.18). We find
(a) (b)
Fig. 7.17 Schematic representation of
planar (a) and (b) and axial (c) and
(d) biperiodic configurations. (a), (c)
Resonant frequency varies from element
to element. (b), (d) Distance varies
between neighbouring elements. From
Sydoruk et al. (2005). Copyright
c
(c) (d) 2005 American Institute of Physics
a1 a2
M1 M2
We shall assume now propagating solutions both for the even- and for
the odd-numbered elements
ZZ0
1
1.2
0 Ea/S 1
1
1.2
0 Ea/S 1
'acoustic’ branch
Fig. 7.20 Dispersion curve for the
ZZ0
2.8
(2k, 2ω )
0
2
0
ω/ω
1.5
Fig. 7.21 Dispersion curve for the
(k, ω ) biperiodic array permitting the propa-
0
1 gation of both signal (k, ω) and pump
(2k, 2ω) waves required for paramet-
0.8
ric amplification. From Sydoruk et al.
1/3 2/3 1 c 2005 American
(2005). Copyright
2βa/π
Institute of Physics
the pump wave as 2βa = 2π/3 at the frequency 2ω0 /2π = 127.74 MHz.
These requirements can be satisfied with an axial structure where the
distance between the elements is a = 0.5ro (M = 0.336L), C1 = 164 pF,
C2 = 56 pF, L = 33 nH. The corresponding dispersion curve is shown in
Fig. 7.21. It can be seen that the condition of synchronism is satisfied
and both the signal and the pump waves can propagate in the system.
So far we have talked about an infinitely long biperiodic line. Does a
terminal impedance exist that matches the line, that makes it possible to
have a single travelling wave? Since the line consists now of two different
kinds of elements it may be expected that we shall need two terminal
impedances, one to be inserted into the last, and the other one into the
last-but-one element. This may be shown to be the case. The values of (a)
the two terminal impedances may be found from a calculation similar
to that in Section 7.3 as
M1,2 Z01,02
ZT (1, 2) = − . (7.48) (b)
M1,2 + M e j k(a1 + a2 )
2,1
The first experimental results were obtained (Wiltshire et al., 2003b) not
long after the derivation of the dispersion equation (Shamonina et al.,
Fig. 7.22 Capacitively loaded loops.
2002a). The basic element of the line, a capacitively loaded loop was re- Photographs of single element (a) and
alized by winding two turns of 1-mm diameter copper wire on a dielectric fragments of an axial line (b) and a
rod (diameter 9.6 mm), and tuned to the desired frequency of 60 MHz by planar line (c). From Wiltshire et al.
inserting a capacitor (nominally 100 pF) between the ends of the wire, (2003b). Copyright
c 2003 IEE
as shown in Fig. 7.22(a). Two lines were assembled from these elements,
an axial line in which 32 elements were arranged along the axis of a di-
electric rod, spaced by their diameter (Fig. 7.22(b)) and a planar line in
which 15 elements were placed side-by-side (Fig. 7.22(c)). The resonant
frequency and the quality factor were determined as f0 = (61.4 ± 0.4)
MHz and Q = 48 ± 5.
232 Magnetoinductive waves I
12 0
phase/ S [rad]
amplitude [dB]
8 −20
4 −40
0 −60
10 20 30 10 20 30
element number element number
(a) (b)
Fig. 7.23 Phase (a) and amplitude variation (b) along axial structure at 61 MHz. From Wiltshire et al. (2003b). Copyright
c 2003 IEE
64 64
frequency f [MHz]
frequency f [MHz]
60 60
56 56
0 0.5 1 0 0.2 0.4
Ea/S Da/S
(a) (b)
Fig. 7.24 Dispersion relationship for axial line. (a) Propagation constant and (b) attenuation constant versus frequency. From
c 2003 IEE
Wiltshire et al. (2003b). Copyright
66 66
frequency f [MHz]
62 frequency f [MHz] 62
58 58
0 0.5 1 0 0.1 0.2
Ea/S Da/S
(a) (b)
Fig. 7.25 Dispersion relationship for planar line. (a) Propagation constant and (b) attenuation constant versus frequency.
c 2003 IEE
From Wiltshire et al. (2003b). Copyright
SSRRs
Microstrips
d1
d2
s
h1 h2 w Fig. 7.26 (a) Photograph and sketch of
Ground the planar MI transducer. (b) sketch of
Dielectric plane the element. From Freire et al. (2004).
substrate Copyright
c 2004 American Institute
(a) (b) of Physics
2.0
PCB thickness
1.5
1.0
κ
Single-sided
Double-sided
0.5
(a)
(b) (c)
Fig. 7.29 (a) Experimental arrangement of MI waveguides; (b) experimental frequency variation of S21 for waveguides based
on double-sided coils and a fixed number of elements (30) and different element spacing, and (c) corresponding result for a fixed
element spacing (2.5 mm) and different numbers of elements. From Syms et al. (2006b). Copyright
c 2006 IOP Publishing
Ltd
attenuation per element comes to the low figure of 0.12 dB per element.
The dependence on the number of elements of the S21 versus frequency
curve is shown in Fig. 7.29(c). It may be seen that it is feasible to set
up long lines.
All the experiments mentioned so far were on lines consisting of iden-
tical elements. We shall now report experimental results measured on
biperiodic lines by Radkovskaya et al. (2007a). The loop used in the
experiments, we shall call it a split pipe, is shown in Fig. 7.30(a). It
became a resonant element at a desired frequency when loaded by a ca-
pacitor. Two sets of elements were used, one loaded by a capacitor of
330 pF, and the other one by a capacitor of 680 pF, yielding resonant
frequencies of 46.21 MHz and 32.46 MHz. The two sets were interleaved
to produce a biperiodic line as shown in Fig. 7.30(b) for the axial and in
Fig. 7.30(c) for the planar configuration where the positions of the trans-
mitter and receiver coils are also shown. The measurement technique
was the same as described earlier in this section.
The measured and theoretical curves of ω versus βa are shown in Figs.
7.31(a) and (b) for the axial and planar configurations, respectively. The
236 Magnetoinductive waves I
Receiver coil
Transmitter
coil
w Receiver
coil
r
h
g
Transmitter coil
Fig. 7.30 Biperiodic arrays of capacitively loaded split pipes. Schematic representations of (a) element dimensions, (b) axial
and (c) planar configurations with measuring coils. From Radkovskaya et al. (2007a). Copyright
c 2007 IEE
50 50
frequency [MHz]
frequency [MHz]
45 45
axial array planar array
40 theory 40 theory
experiment experiment
35 35
30 30
0 0.5 1 0 0.5 1
Ea/ S Ea/ S
(a) (b)
Fig. 7.31 Dispersion characteristics for the axial (a) and planar (b) biperiodic structure. From Radkovskaya et al. (2007a).
c 2007 IEE
Copyright
agreement is remarkably good for the planar line and quite good for the
axial line. In particular, it should be emphasized that for the biperiodic
line there is hardly any difference between the dispersion curves of the
axial and planar configurations as predicted by eqn (7.47).
1.4
N=1
N=2
N=5
12
N=20
ZZ0
1.4
N=1
N=2
12 N=5
N=20
ZZ0
0.8
Fig. 7.33 Same as in Fig. 7.32 but as-
0 0.5 1 suming a quadratic decay of the cou-
Ea/S pling constant with distance
∞
X
Z0 In + j ω Mm (In+m + In−m ) = 0 , (7.49)
m=1
where Mm is the mutual inductance between two elements a distance
ma from each other. Assuming again a wave solution of the form of
eqn (7.4) the dispersion equation can be derived as
∞
ω02 X
1− + κn cos(nkd) = 0 , (7.50)
ω 2 n=1
where κn = 2Mn /L and only the lossless case is considered. How large
is the influence of higher-order couplings? Assuming cubic decay of the
mutual inductance with distance and κ1 = 0.5 the dispersion equation
is plotted in Fig. 7.32 for N = 1, 2, 5 and 20. The effect of higher
interactions may be seen to be small. For finite-size elements the cubic
decay is not a good approximation. The actual decay may be closer to a
quadratic one. In Fig. 7.33 we plot again the dispersion curve for N = 1,
2, 5 and 20 and κ1 = 0.5. For this lower decay, as expected, the effect
of higher orders is more significant.
Can we say anything in more general terms about the dispersion
curve? The answer is yes, when κn can be expanded into a series in
238 Magnetoinductive waves I
M2 M2
M4 M4
Fig. 7.34 Schematic representation of M M M
the coupling between two lines of reso-
nant magnetic metamaterial elements. M3 M3
Mutual inductances M1 , M2 , M , M3
and M4 between the nearest neighbours M1 M1
are shown by arrows. From Sydoruk
c 2006 by the
et al. (2006). Copyright
American Physical Society (n − 1)th nth (n + 1)th
∞
1h
Lim e j ka + Lim e −j ka .
X i
n−m cos(nka) = (7.53)
n=1
2
For our purpose the most important property of this special function
is that it is monotonic, from which it follows that the dispersion curve,
provided the expansion of κn is possible in the form of eqn (7.51), is also
monotonic.
The generalization to higher-order interactions is also straightforward
for the case when there is excitation by a set of voltages. The mathe-
matical form is still given by eqn (7.13) as
V = ZI , (7.54)
but Z is no longer a tri-diagonal matrix. The main diagonal elements
are still equal to Z0 but the off-diagonal elements are now equal to
55 55
frequency [MHz]
frequency [MHz]
50 50
45 45
40 40
0 0.5 1 0 0.5 1
Ea/S Ea/S
(a) (b)
Fig. 7.36 Dispersion characteristics of coupled planar lines with h = 20 mm (a) and 10 mm (b). Theory (solid lines) and
c 2006 by the American Physical Society
experiment (circles and squares). From Sydoruk et al. (2006). Copyright
50 50
frequency [MHz]
frequency [MHz]
46 46
42 42
0 0.5 1 0 0.5 1
Ea/S Ea/S
(a) (b)
Fig. 7.37 Dispersion characteristics of coupled planar–axial lines with h = 30 mm (a) and 15 mm (b). Theory (solid lines)
and experiment (circles and squares). From Sydoruk et al. (2006). Copyright
c 2006 by the American Physical Society
a Out a Out
h h
In In
(a) (b)
Fig. 7.38 Schematic representation of the (a) unshifted and (b) half-a-period shifted coupled lines. The first element of the
lower array is excited by a transmitting coil and the signal in the last element of the upper array is measured by a receiving
coil. From Radkovskaya et al. (2007b). Copyright
c 2007 Wiley-VCH Verlag GmbH & Co. KGaA
50 50 80
49 49
Frequency [MHz]
Frequency [MHz]
40
48 48
47 47
0
46 46
45 45 −40
44 44
0 1 2 3 4 0 1 2 3 4
Relative shift, '/a Relative shift, '/a
(a) (b)
Fig. 7.39 Contour plots of transmission between the split-pipe arrays as a function of frequency and shift, (a) experiment, (b)
theory. From Radkovskaya et al. (2007b). Copyright
c 2007 Wiley-VCH Verlag GmbH & Co. KGaA. For coloured version see
plate section
z [cm]
rotating
dipole
í
N −1
ω02 κN −j N ka X
1− 2 + e + κn cos(nka) = 0 , (7.58)
ω 2 n=1
Table 7.1 Rotational resonances ωn /ω0 for a 9-element ring. From Soly-
mar et al. (2006)
0 1.069 1.085
1 1.052 1.049
2 1.011 1.005
3 0.970 0.969
4 0.946 0.949
N
ω2 X
1 − 02 + κn cos(nka) = 0 , (7.59)
ω n=1
7.15 Applications
7.15.1 Introduction
Waves propagating on coupled resonant structures had applications in
microwave tubes and in linear accelerators (Bevensee, 1964). In the
metamaterial context at the time of writing we can only talk about po-
tential applications. To that category belongs the early work of Wiltshire
et al. (2001) who conducted, with the aid of swiss rolls, magnetic infor-
mation from an MRI machine to a detector. An application as a delay
line was envisaged by Freire et al. (2004). Their device was shown in Fig.
7.26(a). The information travelled from one strip line to the other strip
line via a MI wave. The maximum delay time measured was about 6 ns.
Phase shifters were designed by Nefedov and Tretyakov (2005). They
made good use of the fact that lines with positive phase shift (those
that support forward waves) and also with negative phase shift (those
that support backward waves) are simultaneously available.5 Near-field 5
That arrangement can lead to shorter
imaging was demonstrated by Freire and Marques (2005). Later, it was phase shifters. If, for example, a phase
shifter of -20◦ is required it can be a
shown by the same group (Mesa et al., 2005) that the imaging strongly short one if negative phase shift is avail-
depends on the characteristics of the receiver. able. On the other hand, if one has to
In this section we shall investigate three kinds of applications in a little rely on a positive phase shift then the
more detail. Section 7.15.2 will describe the signal processing aspects phase shifter must be 17 times larger to
offer a phase shift of 340◦ .
of MI waves that require various waveguide components. Section 7.15.3
will discuss potential applications for imaging with some experimental
results showing the quality of imaging obtained. Finally, Section 7.15.4
244 Magnetoinductive waves I
1.0 1.0
Transmitted
Normalized power
Fig. 7.41 Magnetoinductive waveguide mirror. (a) Variation of |R|2 and |T |2 with ω/ω0 for κ = 0.2 and different values of µ.
(b) Variation of |R|2 and |T |2 with µ for ω = ω0 . From Syms et al. (2006a). Copyright
c 2006 IEE
will resurrect the rotational resonance of the previous section and discuss
briefly its potential applications in magnetic resonance measurements.
µ2 − 1 2µ
R= and T = . (7.62)
µ2 + 1 µ2 + 1
7.15 Applications 245
–2 –1 0 1 2
Zm Z Zm Z Zm 1 Z Zm 2 Z Zm Z Zm
Transmitted
Incident + Reflected
(a)
1.0 1.0
Normalized power
0.8 0.8
Transmitted power
P (=P )
0.6 0.4 0.6
0.2
0.4 0.4
02 0.2
0 0
0.90 0 95 1.00 1.05 1.10 1.15 0.90 0.95 1.00 1.05 1.10 1.15
(b) (c)
Fig. 7.42 (a) Magnetoinductive waveguide Fabry–Perot resonator. (b) Variation of |T |2 with ω/ω0 for κ = 0.2 and different
values of µ. (c) Variation of |T |2 with ω/ω0 for two-loop cavity with κ = 0.2 and µ = 0.2. From Syms et al. (2006a). Copyright
c 2006 IEE
Zm Zm Zm 1 Zm 2 Zm 1 Zm 2 Zm 1 Zm 2 Zm Zm
Transmitted
Incident + Reflected
(a)
1.0 1.0
Reflected
0.8 0.8 Uniform
Normalized power
Reflected power
0.6 Tapered
0.6
0.4 0.4
02 0.2
Transmitted
0 0
0 2 4 6 8 10 12 14 16 18 20 0 90 0.95 1.00 1.05 1.10 1 15
Number of periods ZZ
(b) (c)
Fig. 7.43 (a) Magnetoinductive waveguide Bragg grating. (b) Variation of |R|2 and |T |2 at ω = ω0 , with the number of periods
for a Bragg grating with κ = 0.2 and µ1 = 1.1, µ2 = 0.9. (c) Variation of |R|2 with ω/ω0 for a similar grating containing
20 periods (full line) and corresponding result for grating with raised cosine taper (dashed line). From Syms et al. (2006a).
Copyright
c 2006 IEE
element 0. They can be obtained from eqns (7.63)–(7.65) and the an-
alytical solution may be found in Syms et al. (2006a). Choosing again
the coupling coefficient between nearest neighbours in the uniform lines,
κ = 0.2 and µ1 = M1 /M = µ2 = M2 /M equal to 0.2 and 0.4 the nor-
malized transmitted power is plotted in Fig. 7.42(b). Clearly, there is a
transmission peak around the resonant frequency.
If we insert additional resonant loops between elements 0 and 1 in
Fig. 7.42(a) then we have additional transmission peaks. The transmit-
ted power as a function of frequency for a two-loop cavity is shown in
Fig. 7.42(c) for the same set of parameters. There are now two narrow
transmission peaks.
Following further the optical analogy we should be able to have large
reflection in a certain frequency band if a large number of small re-
flections add coherently. These are called Bragg reflectors in optics. We
can achieve a large number of small reflections if the mutual inductances
undergo a small but periodic variation along the line (see Fig. 7.43(a),
where only one period is shown). The period is Λ = 2a. How many peri-
7.15 Applications 247
2
Zm –1 N+2
Port 3 (I3 , n) Zm Zm
Z Port 2 (I’n )
1 Z N+1 Z
Zm 0
Zm Zm
Z Z +1 Z
Transmitted N
Port 1 (I1 , n) R2 Zm Zm T2
Z m3
–2 –1 0 Z Zm Z Zm Z
Z m2 1 R1 Z Zm Z Zm Z
Transmitted Zm Zm T1
Incident + Reflected Z
2 Z Z
Zm Zm Zm
Port 2 (I2 , n) Z Port 1 (I n) Z
Z
Zm Zm Zm
(a) (b)
Fig. 7.44 (a) Magnetoinductive waveguide 3-port power splitter. (b) Magnetoinductive waveguide directional coupler. From
c 2006 IEE
Syms et al. (2006a). Copyright
ods do we need for large reflection? The answer is given in Fig. 7.43(b),
where the dependence of the reflection and transmission coefficients is
plotted as a function of the number of periods for µ1 = 1.1 and µ2 = 0.9.
It may be seen that to reach saturation (that is 100% reflection) about
20 periods are needed. Next, we calculate the reflected power as a func-
tion of frequency for the same 20 periods and the same parameters. It
may be seen in Fig. 7.43(c) that close to total reflection occurs within a
narrow band centred on the resonant frequency of the element.
Have we seen similar things before? If we look at the alternating values
of mutual inductance in Fig. 7.43(a) it should remind us of the biperiodic
lines of Figs. 7.17(a)–(d). In Section 7.10 we looked at this problem and
came to the conclusion (an infinite line was assumed there) that for
a biperiodic line there is a stop band in the middle of the dispersion
characteristics. In the present section we have a finite biperiodic line
but the conclusion is the same. Instead of saying that there is a stop
band we say now that we have nearly perfect reflection within a certain
band due to Bragg reflection—and that’s the same thing.
Another useful device may be seen in Fig. 7.44(a). It is shown in Syms
et al. (2006a) that by judicious choice of M1 , M2 and M3 any desired
power ratio between the two output lines can be achieved without any
reflection in the input line.
Our final example is a four-port device shown schematically in Fig.
7.44(b). The free parameters are M ′ and N (note that Fig. 7.44(b) shows
only one extra element in the coupling region), the mutual inductance
and the number of elements in the coupling region. The aim is to direct
desired powers to the two outputs. The analysis may be found in Syms
et al. (2006a).
248 Magnetoinductive waves I
1’ 2’ 3’ 4’
1 2 3 4
(a)
a receiver
h
... -3 -2 -1 0 1 2 3 ...
transmitter
(b)
7.15.3 Imaging
We shall discuss here a particular type of imaging under conditions when
all the dimensions are small relative to the wavelength. It is a pixel-by-
pixel imaging that simply translates the object. An example is shown in
Fig. 7.45(a) where the object, consisting of points 1, 2, 3, 4 is translated
along the dotted lines to 1′ , 2′ , 3′ , 4′ , a distance h away. In order to
simplify the problem we shall look at one object point only and represent
it by a small non-resonant transmitter coil at the point x = 0, y = 0. The
imaging is tested by moving a small, non-resonant receiver coil along the
line y = h. If the received power has a narrow maximum in the vicinity
of the point x = 0, y = h then we can regard it as an image. However,
if all we have is a small transmitter coil and a small receiver coil then
the maximum along the y = h line will be wide, corresponding to the
field distribution of the small transmitter coil. How could we make it
sharper? Let’s insert a MI waveguide between the transmitter and the
receiver, as shown in Fig. 7.45(b). We may now claim that the field at
the point x = 0, y = h will be higher due to the coupling of element
0 both to the transmitter and to the receiver. This claim, however, is
not necessarily correct. Inserting the MI waveguide will not, in general,
make the field opposite the transmitter more concentrated because there
will be a MI wave propagating in both directions away from element zero
spreading the power in the x direction.
The remedy is to have a MI wave in the y direction but suppress
it in the x direction. How can we do this? We have actually done
so in Section 7.13. When we have two coupled lines and the elements
are above each other then there is no power transfer along the coupled
lines, provided the coupling is high enough. This conclusion can be
7.15 Applications 249
0
60 60
í
Frequency f [MHz]
Frequency f [MHz]
í
í
40 40
í
30 30 í
í í
Normalized distance x/a Normalized distance x/a
(a) (b)
Fig. 7.46 Near-field imaging for the double lens with h = 10 mm. Magnetic-field distribution in the image plane versus
c 2007 American Institute
frequency (contour plot). Experiment (a) and theory (b). From Sydoruk et al. (2007b). Copyright
of Physics. For coloured version see plate section
drawn from Fig. 7.36(b), which shows that there is a stop band at ω0
for h = 10 mm, and also from Fig. 7.39(a), showing no power transfer
for the unshifted case. Thus, power can be transferred in the y direction
but now it cannot spread in the x direction.
The relevant experiment was done (Sydoruk et al., 2007b) with two
coupled lines consisting of split-pipes and separated from each other by
h/2. The elements are arranged as in Fig. 7.38 but the measurement is
done now by placing the transmitter below the element in the middle
at a distance h/4, while the receiver moves above the upper line at a
distance h/4, as shown schematically in Fig. 7.45(c). The experimental
and theoretical results are shown in Figs. 7.46(a) and (b) for a distance
of h = 20 mm. The horizontal axis shows the displacement of the re-
ceiver, whereas the frequency is on the vertical axis. The measured field
strength is colour-coded. There is remarkable agreement between theory
and experiment, both showing that an image exists in the vicinity of the
resonant frequency at which sideways propagation of MI waves is pro-
hibited. The image is translated in the present case by 2h. Translation
further away can be realized by inserting further lines. The essential
criterion is that sideways propagation of MI waves must be prohibited.
z [cm]
rotating
dipole
í
í
x [cm]
Fig. 7.47 Bilayered rotational res-
onator
y [cm] í í
I = Z−1 V. (8.1)
The impedance matrix as presented in eqn (7.15) is a tri-diagonal
matrix because the treatment there was restricted to nearest-neighbour
252 Magnetoinductive waves II
0.9
0.9
2
08
1.
k y a/S
1
1
1.08
1
y a í
2
Fig. 8.2 Capacitively loaded loops in 2
0.9
1
a .9
the planar configuration (a) geome- í 0
try, (b) normalized dispersion (isofre- x í í
quency) curves for a = 2.25r0 . From k xa/S
c 2005
Syms et al. (2005b). Copyright
(a) (b)
EDP Sciences
where In,m is the current in the element located at the nth row and
mth column, Mx and My are the mutual inductances in the x (horizon-
tal) and y (vertical) directions, respectively, and, as before, Z(ω) is the
self-impedance of the elements. For the planar configuration (see Fig.
8.2(a)), Mx = My and they are both negative. For the planar–axial case
shown in Fig. 8.3(a) Mx is of course still negative, but My is positive
and under the present conditions (element spacing being the same in
both directions) its value is smaller than |Mx |.
Assuming the current in the form
8.1 MI waves in two dimensions 253
1.05
9
1.0
1.0 1.01
1.01 5
0.97
k y a/S
.93 0.97
0 0.93
1.01 1.01
y a í
0.97 7
1.05 0.9
Fig. 8.3 Capacitively loaded loops
a 1.01 1.05 1.01
í in the planar–axial configuration (a)
x í í geometry, (b) normalized dispersion
k xa/S (isofrequency) curves for a = 2.25r0 .
From Syms et al. (2005b). Copyright
(a) (b)
c 2005 EDP Sciences
1
S= vg Es . (8.7)
2
254 Magnetoinductive waves II
The energy stored in any of the elements is given by the sum of the en-
ergies in the inductance, the capacitance and in the mutual inductances
relating to nearest neighbours. In terms of a single element (m, n) it is
given by
|In,m |2
1 2 ∗ ∗
Es = 2 L|In,m | + ωC
+ Mx In,m In−1,m + In+1,m
2a
∗ ∗
+ My In,m In,m−1 + In,m+1 . (8.8)
With our wave assumption for the current (eqn (8.4)), and using the
expression of the group velocity given by eqn (8.6), the power density
may be written as follows
1
S= ω|I0 |2 [Mx sin(kx a)ix + My sin(ky a)iy ] , (8.9)
2
i.e. it is independent of the circuit parameters L and C. The condition
for ω to be in the pass band must of course be satisfied.
medium 1 medium 2
m+1
m-1
Fig. 8.4 Two media consisting of ca-
pacitively loaded loops in the planar
configuration having different resonant
n=0 n=1 frequencies
medium 1 medium 2 medium 1 medium 2
k y a/S
k y a/S
í í
Fig. 8.5 Refraction at the boundary of two media. Planar configuration with ω02 /ω01 = 1.03. (a) The loci ω/ω01 = 1.11 and
corresponding phase velocity vectors; (b) directions for the group velocity vectors, assuming (kx1 a, ky1 a) = (−0.18, −0.12)π
and (kx2 a, ky2 a) = (−0.38, −0.12)π. From Syms et al. (2005b). Copyright
c 2005 EDP Sciences
velocities, very closely opposite to the chosen wave vectors, are shown in
Fig. 8.5(b). It may be seen from Fig. 8.5(a) that the construction can be
performed for all possible values of kx1 , i.e. at that particular frequency
a refracted wave exists for any incident wave.
In our next example we shall take the same planar configuration on
both sides of the boundary but now the resonant frequency is assumed to
be smaller in medium 2. We take ω02 = 0.97ω01. The ratio of the radii
is then reversed as shown in Fig. 8.6(a). Using the same construction as
before we find that kx2 > kx1 and refraction is now pointing away from
the perpendicular, as may be seen in Fig. 8.6(b), where the correspond-
ing group velocities are shown at the boundary of the two media. It may
also be seen in Fig. 8.6(a) that no refraction is possible for a range of
incident angles. This is the case of total internal reflection.
In our third example we choose planar–axial configurations on both
sides with all the parameters being the same but the orientation in
medium 2 is at right angle to that in medium 1, as shown in Fig. 8.7(a).
Note that the lattice in medium 2 is shifted by one half of the lattice
constant relative to that in medium 1, which ensures increased magnetic
256 Magnetoinductive waves II
medium 1 medium 2
medium 1 medium 2
k y a/S
k y a/S
í í
Fig. 8.6 Refraction at the boundary of two media. Planar configuration. ω02 /ω01 = 0.97. (a) the loci ω/ω01 = 1.08 and
corresponding phase velocity vectors; (b) directions for the group velocity vectors, assuming (kx1 a, ky1 a) = (−0.33, −0.19)π
and (kx2 a, ky2 a) = (−0.04, −0.12)π. From Syms et al. (2005b). Copyright
c 2005 EDP Sciences
medium 1 medium 2
m+1
m+1/2
m
m-1/2
m-1
n=0 n=1
x
(a)
1 1
medium 1 medium 2 medium 1 medium 2
k y a/S
k y a/S
0.5 0.5
0 0
í -0.5 0 0 0.5 1
k xa/S k xa/S
(b) (c)
Fig. 8.7 Refraction at the boundary of two media. Planar–axial configuration with different orientations of the loops in
medium 1 and medium 2. (a) The geometry of the loops; (b) the loci ω/ω01 = 1.01 and directions for group velocity vectors;
c 2005
(c) directions for the group velocity vectors for (kx1 a, ky1 a) = (−0.39, −0.42)π. From Syms et al. (2005b). Copyright
EDP Sciences
element number n y
Fig. 8.9 Contour plot of currents in a 2D array of resonators showing diffraction on a defect. For coloured version see plate
section
ZIn,m + j ωM (In,m−1 + In,m+1 + In−1,m Fig. 8.10 (a) Lattice of resonant el-
ements with hexagonal arrangement.
+In+1,m + In−1,m+1 + In+1,m−1 ) = 0, (8.10) (b) (a1 , a2 ) direct vectors of the hexag-
onal lattice and (b1 , b2 ) reciprocal vec-
where Im,n is the current in the (m, n) element, Z is the self-impedance tors
assumed again as lossless, M is the mutual inductance between nearest
neighbours. The position of an element is given by the radius vector
a1 · b1 = 1, a1 · b2 = 0,
(8.13)
a2 · b1 = 0, a2 · b2 = 1.
We shall now look for the solution of eqn (8.10) in the form
k = 2 π(f1 b1 + f2 b2 ), (8.15)
260 Magnetoinductive waves II
k y a/S
ω 1
=p , (8.16)
ω0 1 + κ[cos(2 πf1 ) + cos(2 πf2 ) + cos(2 πf1 − 2 πf2 )]
√
where ω0 = 1/ LC is the resonant frequency of the element and κ =
2M/L is the coupling coefficient. The ω/ω0 = constant curves as func-
tions of the wave vector are plotted in Fig. 8.11. The pass band is found
to extend from 0.93ω0 to 1.21ω0. Note that the waves are backward
waves with phase and group velocities in opposite directions.
In the examples to follow in the presence of an excitation we shall use
eqn (8.1) to determine the current distribution. First, however, we shall
attempt to find an approximate solution that will lead us to a mathemat-
ically familiar territory and offer a clear physical picture. We may expect
that the propagation of MI waves, similarly to the propagation of most
other waves, can be mathematically described by a second-order partial
differential equation. We shall therefore convert our difference equation
(eqn (8.10)) into a differential equation. We can do that when the wave
vectors are sufficiently small or in other words when the wavelength of
the MI wave is much larger than the element spacing. Consequently
we shall introduce the ν, µ co-ordinate system in the directions a1 and
a2 with the continuous variables ν = na and µ = ma, and replace the
discrete function In,m by the continuous function I(ν, µ). A change in
the subscript by unity would then be equivalent to a change of the con-
tinuous variable by a, which is then regarded as an elementary change.
To convert all the terms in eqn (8.10) into continuous variables we need
to expand the current into a Taylor series as follows
∂ ∂
I(ν + ∆ν, µ + ∆µ) = 1 + ∆ν + ∆µ
∂ν ∂µ
8.1 MI waves in two dimensions 261
∂2 ∂2
1
+ (∆ν)2 2 + 2∆ν∆µ
2 ∂ν ∂ν∂µ
2
∂
+(∆µ)2 2 I(ν, µ), (8.17)
∂ν
where ∆ν and ∆µ are small deviations from ν and µ. With the aid of
eqn (8.17) we may now convert eqn (8.10) into the differential equation
∂2 ∂2 ∂2 3 ∂2 ∂2
− + = + . (8.19)
∂ν 2 ∂ν∂µ ∂ν 2 4 ∂x2 ∂y 2
So, we end up with the familiar wave equation
∂2I ∂2I
2 + + k 2 I = 0, (8.20)
∂x ∂y 2
where
ω2
2 4 1
k = 2 1+ 1− 2 , (8.21)
a 3κ ω0
and k = |k|. It may be easily shown that the dispersion equation (8.16)
reduces to eqn (8.21) when ka ≪ 1.
A clear advantage of having the wave equation is that, at least for
certain geometries, we know the solutions from past experience. For
rectangular boundaries, possible solutions are
I = I0 J0 (kr), (8.23)
where J0 is the zero-order Bessel function of the first kind and r is the
distance from the centre of the circular structure of the elements.
If we choose a frequency and know which elements are excited (our
formulation allows all of them to be separately excited) we can determine
the current distribution from eqn (8.1). The result might be in the form
of odd-looking current distributions because several wave vectors may
coexist at a particular frequency. It is only at the spatial resonance
that no more than one wave vector survives. Spatial resonances are
characteristic to all wave phenomena. They exhibit the same behaviour
whether they occur in vibrating membranes, organ pipes or Fabry–Perot
resonators. The boundary condition to be satisfied in the present case
is that the current must vanish at the boundary.
262 Magnetoinductive waves II
0
20
y [cm]
0 20
20
-40
20 0 20 20 0 20 20 0 20
x [cm] x [cm] x [cm]
(a) (b) (c)
Fig. 8.12 Numerically obtained current distributions for circular boundary conditions at ω/ω0 = (a) 1.207, (b) 1.192, (c) 1.167.
Black dots show positions of the elements. From Zhuromskyy et al. (2005a). Copyright
c 2005 Optical Society of America.
For coloured version see plate section
kR = ρi , (8.24)
where ρi is the ith root of J0 .
We shall now look at the phenomenon of spatial resonance with the
aid of a few examples. At this stage we need to commit ourselves as
to the distance between the elements and the resonant frequency. The
loop radius, the wire diameter and the distance between the elements
are taken as 10 mm, 2 mm and 22.5 mm, and the resonant frequency
as 21.5 MHz. The inductance of the loop may then be determined from
standard formulae (Grover, 1981) that give L = 33 nH. The correspond-
ing capacitance can be determined from the resonant frequency as 1.66
pF. The mutual inductance between two neighbouring elements is 1.75
nH, yielding κ = −0.106. Owing to the hexagonal arrangement it is not
possible for all the elements to lie exactly on a circular boundary. With
our choice of 361 elements the deviation from the circular boundary is
quite small, as may be appreciated by looking at Figs. 8.12(a)–(c).
Knowing the geometry the frequencies of the first 3 spatial resonances
may be determined from eqn (8.24) as ω/ω0 = 1.207, 1.187, 1.155. Nu-
merical calculations based on the known values of the mutual impedance
matrix yield 1.207, 1.191, 1.165, a very good approximation. The ana-
lytical results are of interest because they give good approximation for
low values of k and, of course, they give an immediate idea of what
the current distribution looks like. For the general case, however, it is
more accurate to rely on the exact solution of the discrete problem that
is based on the inversion of the impedance matrix. Note that for the
numerical determination of the current distribution we need an excita-
tion. We assume that the central element out of the 361 is excited and
then proceed with the numerical solution. The numerically determined
current distributions are shown by a colour code in Figs. 8.12(a)–(c).
Normalization in each figure is to the maximum value within the figure.
8.1 MI waves in two dimensions 263
0
(a) (b)
40 40
y [cm]
y [cm] -20
20 20
-40
0 20 40 0 20 40
x [cm] x [cm]
0
(c) (d)
40 40
y [cm]
y [cm]
-20
20 20
-40
0 20 40 0 20 40
x [cm] x [cm]
Fig. 8.13 Numerically obtained current distributions for rectangular boundary conditions, (a) non-resonant distribution for
asymmetric excitation, (b)–(d) resonant excitation at ω/ω0 = 1.207, 1.202 and 1.201, respectively. Black dots show positions
of the elements. From Zhuromskyy et al. (2005a). Copyright
c 2005 Optical Society of America. For coloured version see
plate section
kx Dx = px π and ky Dy = py π, (8.25)
where px and py are integers.
We shall arrange now the hexagonal lattice in a 19 × 19 square geom-
etry. As in the previous example we shall compare the analytical and
numerical values for the frequencies of 3 spatial resonances (fundamental
plus two of the second order) shown in Figs. 8.13(b)–(d). The resonant
frequencies are ω/ω0 = 1.207, 1.202 and 1.201 from the numerical solu-
tions. There is some ambiguity in the analytical expression: due to the
jagged boundaries Dx cannot be exactly defined. The values we obtain
for the resonant frequencies from the analytical solution are 1.207, 1.202
and 1.202. The last two figures agree because in theory the two current
264 Magnetoinductive waves II
0
20 (a) (e) 20
y [cm]
0 0 -20
20 20
-40
20 0 20 20 0 20
0
20 (b) (f) 20
y [cm]
0 0 -20
20 20
-40
20 0 20 20 0 20
0
20 (c) (g) 20
y [cm]
0 0 -20
20 20
20 0 20 20 0 20 -40
0
Fig. 8.14 Magnetic-field distribution 20 (d) (h) 20
at four resonant frequencies: (a) and
(e) ω/ω0 = 1.206, (b) and (f) ω/ω0 =
y [cm]
0
10 (a) (b) 10
y [cm]
0 0
Fig. 8.15 The normal component of
magnetic field at ω/ω0 = (a) 0.98 and
10 10 (b) 1.01. From Zhuromskyy et al.
25 (2005a). Copyright
c 2005 Optical So-
10 0 10 10 0 10 ciety of America. For coloured version
x [cm] x [cm] see plate section
the axial (perpendicular to the plane of the elements) and radial com-
ponents of the magnetic field were measured on the other side. The
measured 2D distribution of these two components of the magnetic field
are given in their Figs. 4–7 for the first four spatial resonances. Unfor-
tunately, we have not been able to get hold of these figures. The reader
might want to look at the original publication. Our theoretical results
plotted in Fig. 8.14 display practically the same spatial variation for
both components of the magnetic field as in the experiments.
8.1.6 Imaging
We have already discussed imaging with MI waves (Sydoruk et al.,
2007b), both experimentally and theoretically, in Section 7.15.3 where
the ‘lens’ consisted of two coupled 1D lines made up by split pipes. It
was a pixel-by-pixel imaging that could be best described as channelling
the spatial information across the imaging device. The channelling oc-
curred under conditions when the propagation of the MI wave was for-
bidden along the length of the coupled lines. Much earlier, Wiltshire
et al. (2003a) successfully imaged an object with the aid of a hexag-
onally arranged array of swiss rolls consisting of 271 elements. The
authors excited the 2D resonant structure by placing an M-shaped wire
antenna below the structure and measuring the axial component of the
magnetic field on the top. The experimentally obtained image is shown
in their Fig. 4. The image we have calculated from our simple model (in
which we consider capacitively loaded rings instead of swiss rolls) may
be seen in Figs. 8.15(a) and (b) for ω/ω0 = 0.98 and 1.01, respectively.
The agreement with the experimental results of Wiltshire et al. (2003a)
is good. However, as the frequency increases beyond the resonant fre-
quency of the element the image quickly deteriorates. At ω/ω0 = 1.01
the object is unrecognizable.
Imaging with MI waves using two parallel planes of broadside-coupled
SRRs was carried out by Freire and Marques (2005). Their experimen-
tal setup is shown in Fig. 8.16(a). The two planes have areas of 7 × 7
cm2 . They are separated from each other by a foam slab of thickness
d = 4 mm. The substrate thickness is h = 0.254 mm and the dielec-
tric permittivity is εr = 10. The transmitting and receiving antennas
are square loops of 1 cm2 . The field strength was measured by moving
266 Magnetoinductive waves II
Fig. 8.16 (a) Experimental setup for imaging with magnetoinductive waves. (b) The magnitude of the transmission coefficient
between the input and output antennas at a frequency of 3.23 GHz. From Freire and Marques (2005). Copyright
c 2005
American Institute of Physics. For coloured version see plate section
the receiving antenna in the image plane. The experimental results are
given in Fig. 8.16(b). A focal region may be clearly seen where the field
strength is higher than in the surrounding region. The question whether
it is a proper focus is discussed by Mesa et al. (2005). A proper focus is
one that is measured by an ideal receiving device that does not interact
with the elements of the lens. The authors conclude that the focus mea-
sured by Freire and Marques (2005) was due to that kind of interaction
1
A generalization of the principle was i.e. not a proper focus.1 Further imaging work using MI waves in similar
presented by Sydoruk et al. (2007d). configuration was done later by Freire and Marques (2006), using this
The authors showed that such ‘im-
proper’ focus can be placed in any de-
time low-impedance receivers in order to avoid the interaction. They
sired position for any transmitter–lens showed both experimentally and theoretically (theory based on that of
configuration by appropriate matching Maslovski et al. (2004)) that good imaging can be achieved. A further
of the receiver. generalization of the problem was done by Marques and Freire (2005).
They showed theoretically that subdiffraction imaging devices cannot
produce focusing of power into three-dimensional spots of subdiffraction
size.
µ0 e −j k0 |r2 − r1 |
Z
A(r2 , t) = J(r1 ) dτ , (8.26)
4π |r2 − r1 |
where the current flows in point r1 and the vector potential is calculated
in point r2 . Hence, the mutual inductance between two loops must be
calculated from the above equation, increasing considerably the numer-
ical work.
If the size of the loop is comparable with the wavelength (say the
diameter of the loop is equal to λ/4, that may very well happen in
the optical region) then there are further complications. The current
distribution can no longer be regarded as uniform and one needs to resort
to the integral equations of antenna theory (see, e.g., King 1969) to find
the current as a function of azimuthal angle. We shall certainly not go
that far. In the present section we shall make a compromise: assume
a uniform current in the element but include retardation in calculating
the field quantities. As a result, the fields far away from the element
will decay as the inverse of the distance, which means that for a long
array it is necessary to include the effect of radiation when calculating
the interaction between any two elements in the array. The concept of
mutual inductance may still be used but it has to be calculated in a
different manner and, due to the phase delay it turns out to be complex.
In Section 8.2.2 we shall derive the dispersion equation for an infi-
nitely long line that now includes retardation, and compare the result
268 Magnetoinductive waves II
where now
ω02 k3
1 L
= 2 2 1− 2 +j 0 , (8.29)
αm µ0 S ω 6 πµ0
and
k03
1 j 1
f (n) = − − e −j k0 na . (8.30)
2 πµ0 nk0 a (nk0 a)2 (nk0 a)3
1.03
Ns = 10 Ns = 100 Ns = 1000
1.02
ZZ0
1.01
explanation? One may argue that if the line is infinitely long it cannot
radiate for the reason that there is no space left into which it could radi-
ate. However, this argument does not apply to the part of the dispersion
characteristics to the left of the light line. The imaginary part does not
cancel then: the dispersion equation predicts decay of amplitude due to
radiation. These waves belong to the category of leaky waves (see, e.g.,
Hessel and Oliner 1965; Marcuvitz 1956), an interesting subject with a
large literature, but beyond the scope of the present book.
Equation (8.28) is in an analytical form but the infinite series still
needs to be summed numerically up to a certain number of elements. It
can be shown that the series is convergent but converges rather slowly
in the vicinity of the light line. To demonstrate that convergence nu-
merically we need a few parameters, which we shall choose as L = 33
nH, ω0 /(2 π) = 0.96 GHz, a = 25 mm, r0 = 10 mm. The dispersion
curves corresponding to N = 10, 100 and 1000 elements are shown in
Fig. 8.17. Taking now N = 1000 we can next examine the difference
between the quasi-static and the full dispersion equation for the lossless
case. For the previous set of parameters it is shown in Fig. 8.18. The
main difference is that for the quasi-static case there is no dip at the
light line. For larger values of k there is hardly any difference.
N s= 1000
1.02
ZZ quasi-static
retarded
µ0 πr04
M= (1 + j k0 a − k02 a2 ) e −j k0 a . (8.31)
4a3
A further problem is the choice of resistances. We have decided to
regard the ohmic resistance as a parameter that follows from the assumed
value of Q, which we shall often vary within a wide range to illustrate the
effect of losses. What about the radiation resistance? Haven’t we proven
that the line does not radiate so we need not bother about radiation
resistance? Yes, that is true to the right of the light line and for an
infinite number of elements. Once the number of elements is finite we
are not entitled to disregard radiation. Hence, we shall include the
radiation resistance of a small loop that was given by eqn (1.96) and
repeated below
4
π 2 πr0
Rs = η0 . (8.32)
6 λ
For the parameters given above this radiation resistance turns out to
be 0.32 ohm. It must be added to the ohmic resistance so the effective
Q will decline.
Having got all the parameters of the impedance matrix we are ready to
determine the current distribution by inverting the impedance matrix.
Excitation by the first element. This means that in the voltage vector
V the first element is taken as 1 V and the further 499 elements as
zero. The current distribution may now be calculated for a given value
of frequency that we shall choose to be f = 1.01 f0.
The modulus and phase of the current flowing in the nth element are
plotted in Figs. 8.19(a) and (b) for Q = 100. The current, as may be ex-
pected, declines roughly exponentially. The phase variation looks rather
odd, quite different from phase variations one usually encounters. It
may be approximated by two straight lines: up to the first 15 elements
the phase increases linearly along the line and then the phase variation
reverses and declines linearly up to the end of the line. The increasing
variation is a sign of a backward wave, whereas the decreasing phase
272 Magnetoinductive waves II
0.2
current magnitude
10
phase [rad]
0.1
0
1 1
forward wave
normalized current
normalized current
backward wave
0
0
variation indicates a forward wave. The explanation for this rather odd
behaviour is that two waves are present: a backward wave and a forward
wave, both excited at element 1. The backward wave is excited with a
higher amplitude, hence it dominates for the first 15 elements. How-
ever, this backward wave is lossy and has a higher attenuation than the
forward wave. After propagating for 15 elements the two waves are of
about equal amplitude. After the 15th element the forward wave dom-
inates, and hence the decreasing phase variation. A rough calculation
would give for ka (i.e. for the phase change per element) −0.22 π for the
backward wave and 0.16 π for the forward wave. Note that the power
moves from the excited first element to the end of the line corresponding
to a positive group velocity, which applies both to the backward and to
the forward wave. But, as the name implies, the backward wave has a
phase velocity pointing towards the source.
An alternative way to gain an intuitive feel for the periodic variation
of the current is to plot its real and imaginary part. These are shown in
Figs. 8.20(a) and (b) for n = 1 to 20 and for n = 30 to 50, respectively.
These figures confirm the conclusions drawn from Fig. 8.19(b). At the
beginning of the line one can see only the backward wave. By element 15
it declines sufficiently so that the forward wave, which is less attenuated,
can be seen. Figures 8.20(a) and (b) may, clearly, provide the value of
8.2 MI waves retarded 273
0.16
|Fourier amplitude|
|Fourier amplitude|
0.8
0.12
0.4
0.08
1
Q= 5000
0.5
0
1
Q= 2500
0.5
amplit ude
0
1
Q= 1000
0.5
0
1
Q= 100
0.5
0 Fig. 8.22 Decay of current amplitude
0 100 200 300 400 500 as a function of element number. From
element number Zhuromskyy (2008)
600 Q = 5000
Q = 4000
Q = 3000
400 Q = 2000
Q = 1000
Q = 100
200
phase
−200
100, 1000, 10 000 and infinity. For a given value of Q we find a spread
in the corresponding k values. The amount of the spread depends, as
may be expected, on the value of Q. For Q = 10 000 and infinity the
lines representing the dispersion curve are quite narrow so they may
be taken as the ‘true’ dispersion curve. For Q = 10 000 there is a just
discernible curve in the range 1 < ka/π < 2. For Q = ∞ the curve
is clearly visible. The reason is that without a matched load there is a
nearly perfect reflection from the end of the array, hence, in the presence
of low ohmic losses, the k values are about the same in both directions.
The spread in the wave vector was shown in order to appreciate how
narrow (or wide) is the Fourier spectrum. Taking the maxima of the
spectrum, an unambiguous dispersion curve is obtained for each value
of Q as shown in Figs. 8.25(a)–(d). The new feature shown is the conver-
gence of the curve to the light-line for Q = 10 000 and ∞, which means
that another solution exists (represented by a local maximum) as well.
It could not be seen in Figs. 8.24(c) and (d) because their values were
too small. This is in line with the conclusions arrived at earlier in this
section. There are two waves excited, a backward wave and a forward
wave, the latter one propagating close to the velocity of light.
The dispersion curves shown in Figs. 8.25(a)–(d) have been found
by exciting the first element. Would we find the same curves if more
elements were excited? We tried exciting the first few elements (up to
5); the dispersion curves found were no different.
Imposing the wave number. In the example given above the temporal
frequency ω was imposed and we looked at the spread in the wave num-
ber. There is another obvious possibility: to impose the wave number
and look at the spread in the temporal frequency. We found similar
results to those of Figs. 8.24(a)–(d). We shall not show the spread here.
We shall give only one value of ω for one value of k. Our criterion for
finding that single value of ω is based on the concept of the input power
8.2 MI waves retarded 275
1.03 1 1.03 10
1 0.4 1 4
0.98 0 0.98 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
ka/π ka/π
(a) (b)
1.02 60 1.02 80
50
ω/ω0
ω/ω0
1.01 1.01 60
40
1 30 1 40
20
0.99 0.99 20
10
0.98 0 0.98 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
ka/π ka/π
(c) (d)
Fig. 8.24 Magnitude and distribution of the Fourier component plotted in the normalized ω − ka plane for various values of
the quality factor, (a)–(d) Q = 100, 1000, 10 000 and ∞. From Zhuromskyy (2008)
500
X
Pin = Vi Ii∗ . (8.33)
i=1
1.03 1.03
1.02 1.02
1.01 1.01
ω/ω0
ω/ω0
1 1
0.99 0.99
0.98 0.98
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
ka/π ka/π
(a) (b)
1.03 1.03
1.02 1.02
1.01 1.01
ω/ω0
ω/ω0
1 1
0.99 0.99
0.98 0.98
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
ka/π ka/π
(c) (d)
Fig. 8.25 Dispersion curves derived from the maximum of the Fourier spectrum, (a)–(d) Q = 100, 1000, 10 000 and ∞. From
Zhuromskyy (2008)
good to the right of the light line. The two curves coincide within the
thickness of the line. To the left of the light line they differ considerably.
We cannot say which is correct, they are just the results of different
excitations but there is no doubt that the one obtained from imposed
wave number is closer to the dispersion curve of the infinite line. The
information about the solution at the light line comes from the first
element excitation. The reason is technical. When ω is imposed we can
find a value of k close to the light line but we cannot impose k values
infinitely close to each other.
8.2.5 Conclusions
We cannot claim to be able to draw clear conclusions. When the line is
excited in a certain manner then we can predict what happens whether it
is ω or k that is imposed. We have shown that in the planar configuration
two waves are excited: a high attenuation backward wave and a low-
attenuation forward wave that travels practically at the velocity of light.
8.3 Non-linear effects in magnetoinductive waves 277
1.03
1.02
1.01
ZZ
0.99
1.03
1.02
1.01
ZZ
0.99
Fig. 8.27 Dispersion curves for Q =
1000 derived from the criterion of
0.98 Fourier spectrum maximum (solid
0 0.2 0.4 0.6 0.8 1
lines) and input power maximum (dot-
ka/S ted lines). From Zhuromskyy (2008)
We can find current distributions and from these, if needed, we can find
the radiation patterns. The ambiguity comes in when we consider the
dispersion curve for infinite number of elements. It is not something that
could ever be measured and for the lossy case its significance is rather
limited.
ωs2 1
2 = 1 + κ cos(β a) (8.35)
ω0 s
for the signal wave and
4ωs2 1
= (8.36)
ω02 1 + κ cos(2 βs a)
for the pump wave. We shall choose ωs /(2 π) = 63.87 MHz, and the
mutual inductance as M = 0.45L (i.e. κ = 0.9). We can then solve
8.3 Non-linear effects in magnetoinductive waves 279
(2f s ,2Es )
frequency f [MHz]
frequency f [MHz]
(2f s ,2Es )
(f s ,E s )
(f s ,E s )
eqns (8.35) and (8.36) for the resonant frequency and propagation con-
stant, giving, respectively, ω0 /(2 π) = 73.44 MHz and βs a = 0.38 π. For
elements with inductance L = 50 nH this value of the resonant frequency
can be achieved by loading the elements by capacitances of C0 = 94 pF.
The question arises whether we can have such a high coupling coeffi-
cient. According to recent experimental work (Syms et al., 2006b; Syms
et al., 2007a) this is indeed possible, provided the elements in the axial
configuration are sufficiently close2 to each other. So, the possibility is 2
In fact, coupling coefficients as high
there. If the sole purpose is to amplify a magnetoinductive wave then as κ = 1.5 have been reported (Syms
et al., 2006b).
such an axial array could do it. For the elements given above the disper-
sion curve is shown in Fig. 8.29(a). It can be seen that both the signal
and the pump waves can propagate and are phase matched.
An alternative solution is to rely on two coupled arrays (Sydoruk
et al., 2006). The dispersion curve for such coupled arrays has al-
ready been shown in Fig. 7.36 for the planar configuration. It consists
of two branches, hence the phase-matching condition can be satisfied
with small values of the coupling coefficient. We shall choose again
ωs /(2 π) = 63.87 MHz and L = 50 nH for the signal frequency and
the inductance of the loop. We take two identical arrays with sepa-
ration of h = 10 mm between them (see Fig. 7.35(a)). This time the
upper array is tuned to the signal frequency and the lower array to the
pump frequency. The distance between the elements within the same
array is a = 24 mm. Hence, the mutual inductances can be calculated3 3
Note that in Section 7.13 several
as M = 0.15L and M1 = −0.05L. These values are close to those mutual inductances were considered.
In the present section, for simplicity,
investigated experimentally in Section 7.13. For choosing the propaga- we use the approximation that only
tion coefficients we have quite a lot of freedom. We shall take them as nearest-neighbour coupling needs to be
βs = π/(3a) and βp = 2 π/(3a). Two parameters that are left undefined taken into account.
are the capacitances of the elements C1,2 . They are to be determined
from the condition that the two phase-matched waves satisfy the dis-
persion relation. For the two coupled waveguides it is given (Sydoruk
et al., 2006) as
280 Magnetoinductive waves II
M
L/2
L/2 L/2 L/2
... L/2 L/2
Fig. 8.30 An equivalent circuit of a sin- U
gle array. The capacitance is voltage-
dependent, leading to parametric inter-
action between the signal and the pump R R R ZT
1
j ωL + + R1 + 2j ωM1 cos ka
j ωC1
1
× j ωL + + R2 + 2j ωM1 cos ka = −ω 2 M 2 . (8.37)
j ωC2
The above equation must be satisfied both for ωs , βs and for 2 ωs , 2 βs .
From the solution we find then that C1 = 130 pF and C2 = 30 pF. The
corresponding resonant frequencies are ω01 /(2 π) = 62.49 MHz for the
upper array and ω02 /(2 π) = 129.15 MHz for the lower one. The two
branches of the dispersion curve showing phase matching are plotted in
Fig. 8.29(b).
1h
us (n) e j νs (n) + c.c. ,
i
Us (n) = (8.38)
2
1h
up (n) e 2j νs (n) + c.c. ,
i
Up (n) = (8.39)
2
where
d(CU )
In = . (8.41)
dt
The problem is now that the capacitor, being voltage-dependent (see
eqn (8.34)), the current will have a number of harmonics, even if we
can assume (as we did in eqns (8.39) and (8.40)) that only the signal
and pump voltages will be present across the capacitor. We shall follow
here the time-honoured method and simplify the problem by further
assuming that the currents will also have only those two frequencies:
1h
is (n) e j νs (n) + c.c.
i
Is (n) = (8.42)
2
and
1h
ip (n) e 2j νs (n) + c.c. .
i
Ip (n) = (8.43)
2
The relationship between the voltages and currents will be determined
by eqn (8.41). The linear part of the problem is trivial but the non-linear
part is less so. Let us evaluate the term γV 2 when the signal and pump
voltages are simultaneously present. We find
γh i2
γV 2 = us (n) e j νs (n) + up (n) e 2j νs (n) + c.c. . (8.44)
4
It can be immediately seen that the term varying at the pump frequency
will take the form
γ 2
u (n) e 2j νs (n) . (8.45)
4 s
It is a little more difficult to recognize the term varying at the signal
frequency. It is
γ
up u∗s (n) e j νs (n) , (8.46)
2
where the star is our other notation for the complex conjugate. There are
13 more terms in eqn (8.44) but none of them are at these frequencies.
Adding now the current due to the linear term in eqn (8.41) to those in
eqns (8.45) and (8.46) we find the signal and pump currents as
d dI(n)
M [I(n + 1) + I(n − 1)] + L + RI(n) + U (n) = 0 . (8.49)
dt dt
The above equation relates nearest neighbours to each other involving
the (n − 1)th, the nth and the (n + 1)th element. The equation needs
to be modified when it is written for the first and the last element in
the array that have only one neighbour. The technique is described in
Section 7.3 for finding the matching impedance. The present case is a
little more complicated because the line needs to be matched at both
the signal and the pump frequency so that
ZT (ωs ) = j ωs M e −j βs a
. (8.50)
ZT (2ωs ) = 2j ωs M e −2j βs a
Performing the rather tedious algebraic operations the final result may
be presented in matrix form (Sydoruk, 2007) as follows
(s) 1
Z −j E us + γZ(s) qs = us0
ωs C0
. (8.51)
(p) 1 γ
Z −j E up + Z(p) qp = up0
2ωs C0 2
Here, E is an N × N identity matrix,
R
L+ M e −j βs a ... 0
j ωs
R ..
M e βs a
j L+ ... .
j ωs
Z (s)
= j ωs
.. .. ..
. . . 0
.. .. ..
. . . M e −j βs a
R + ZT (ωs )
0 ... M e j βs a L+
j ωs
(8.52)
and
R
L+ M e −2j βs a ... 0
2j ωs
R ..
M e βs a
2j L+ ... .
2j ωs
Z(p)
= 2j ωs
.. .. .. .
. . . 0
.. .. ..
. . . M e −2j βs a
R + ZT (2ωs )
0 ... M e 2j βs a L+
2j ωs
(8.53)
8.3 Non-linear effects in magnetoinductive waves 283
dus,p (n)
z = (n − 1)a and us,p (n ± 1) = us,p (n) ± a . (8.54)
dz
As an illustration, let us convert the expression for the signal wave
j ωs M us (n + 1) e −j βs a + j ωs M us (n − 1) e j βs a (8.55)
into continuous form. We find that it comes to
dus dus
j ωs M e −j βs a + us (z) − a
us (z) + a e j βs a
dt dt
dus
= j ωs M us (z) + a sin(βs a) . (8.56)
dt
Following the same technique and assuming that losses are weak and the
non-linearity of the capacitance γ is small we end up (Sydoruk, 2007)
with the differential equations
dus (z)
= −j gs up (z)u∗s (z) − αs us (z)
dz
, , (8.57)
dup (z)
= −j gp u2s (z) − αp up (z)
dz
where
γ γ
gs = and gp = (8.58)
2aωs2 C0 M sin βs a 16aωs2 C0 M sin 2βs a
characterize non-linearity and
284 Magnetoinductive waves II
signal signal
amplitude [V]
amplitude [V]
pump pump
maximum position
R R
αs = and αp = (8.59)
2aωs M sin βs a 4aωs M sin 2βs a
are the attenuation coefficients. Note that eqns (8.57) are analogous
to the corresponding equations for parametric amplification and second-
harmonic generation well known in non-linear optics (Shen, 1984; Bloem-
bergen, 2005).
For a numerical example we shall take the parameters given above
and γ = 0.1 V−1 , up0 = 0.25 V and us0 = 0.5 mV. In the absence of
losses there is a periodic exchange of power between the signal and pump
waves, as shown in Fig. 8.31(a). Both waves decline in the presence of
loss as shown in Fig. 8.31(b) for Q = 250. In a practical case of course the
pump wave is much stronger than the signal wave. The amplification
against signal input amplitude for that case is shown in Fig. 8.32 for
us0 = 1 µV–100 mV. It may be seen that the amplification is constant
up to a certain signal amplitude and then starts to decline.
8.3 Non-linear effects in magnetoinductive waves 285
M1
L/2
L/2 L/2 L/2
... L/2 L/2
U R1 R1 R1 ZT1
M M M
C2 C2 C2 Fig. 8.33 Equivalent circuit of two cou-
pled arrays. The capacitance of the top
L/2 array is voltage-dependent, leading to
L/2 L/2 L/2 L/2 L/2 parametric interaction between the sig-
... nal and the pump. From Sydoruk et al.
V R2 R2 R1 ZT2
M1 (2007c). Copyright
c 2007 IOP Pub-
lishing Ltd
pump
amplitude [V]
pump
signal
amplitude [V]
(b)
high/low amplification
maximum gain [dB]
element number
amplitude [V]
no amplification
(c)
signal
max. gain pump
max. position
quality factor
(a)
amplitude [V]
(d)
Fig. 8.35 Maximum signal gain achieved (dotted line) and the position of the signal maximum (solid line) depending on the
amount of loss (a) and the distribution of the signal (solid lines) and pump (dashed lines) amplitudes along the waveguides for
(b) Q = 1000, (c) Q = 400, and (d) Q = 150. Other parameters are as in Fig. 8.34. From Sydoruk et al. (2007c). Copyright
c 2007 IOP Publishing Ltd
1h
us (n) e j (ωs t − nβs a) + c.c.
i
Us (n) =
2
1h
up (n) e 2j (ωs t − nβs a) + c.c.
i
Up (n) =
2
. (8.60)
1h j (ω s t − nβ s a)
i
Vs (n) = vs (n) e + c.c.
2
1h
vp (n) e 2j (ωs t − nβs a) + c.c.
i
Vp (n) =
2
8.3 Non-linear effects in magnetoinductive waves 287
Here, us (n) and up (n) are the amplitudes of, respectively, the signal and
pump waves in the upper array, and vs (n) and vp (n) are their amplitudes
in the lower array. The corresponding currents flowing through the
elements of the upper and lower arrays are, respectively,
(nl)
d[C1 (Us (n) + Up (n))]
In (n) =
dt
. (8.61)
d[Vs (n) + Vp (n)]
Jn (n) = C2
dt
Neglecting higher harmonics the currents can be written separately at
the signal and pump frequencies as
j ωs C1
Is (n) = [us (n) + γup (n)u∗s (n)] e j (ωs t − nβs a) + c.c.
2
j ωs C1
2up(n) + γu2s (n) e 2j (ωs t − nβs a) + c.c.
Ip (n) =
2
. (8.62)
amplitude [PV]
j ωs C2
Js (n) = vs (n) e j (ωs t − nβs a) + c.c.
2
The next steps are to substitute the currents and voltages into the dif- with amplification
ference equations of the magnetoinductive waves taking into account the without amplification
excitation and matched terminal impedances, turn them into differential
equations and solve the differential equations numerically. element number
A set of results showing the periodic exchange of energy between the
signal and pump waves is shown in Fig. 8.34. The gain against signal Fig. 8.36 Loss compensation by means
amplitude is plotted in Fig. 8.35(a), whereas Figs. 8.35(b)–(d) show of parametric amplification. The am-
the variation of signal and pump amplitudes as a function of element plitude of the signal in the ampli-
number. An interesting conclusion is that in the upper array the signal fied case (solid line) is almost con-
stant along the waveguide, whereas
amplitude may well exceed that of the pump. The pump amplitude is it declines exponentially in the non-
of course still high in the lower array. amplified case (dash-dotted) line. U0 =
A possible use of parametric amplification is for loss compensation. 1 µV, V0 = 50 mV, Q = 500, γ = 0.1
V−1 . From Sydoruk et al. (2007c).
Taking Q = 500 and an appropriate choice of parameters the signal c 2007 IOP Publishing Ltd
Copyright
amplitude in the presence and absence of loss is shown in Fig. 8.36. It
may be seen that the amplitude of the signal can be kept constant with
good approximation.
r r r2 r2 r r
(a)
Fig. 9.5 (a) Horizontal stack of a square lattice of thin wires. Length chosen so as to satisfy the Fabry–Perot condition of
resonance. (b) Image of letter P 2.5 mm in front of the object obtained by numerical simulation. (c) Image 2.5 mm behind
the lens obtained by numerical simulation. (d) Electric-field distribution in the image plane measured by near-field scanning.
From Belov et al. (2006a). Copyright
c 2006 by the American Physical Society. For coloured version see plate section
(a) (b)
Fig. 9.6 (a) Square lattice of capacitively loaded vertical wire structure, (b) view from above. From Belov et al. (2005).
c 2005 by the American Physical Society
Copyright
Fabri-Perot resonances
Normalized transverse component of wave vector, qt b/(2S)
evanescent modes
propagating modes
evanescent modes
Ey = E0 e −j (kz z + kx x) , (9.1)
and substituting it into the wave equation. The dispersion equation may
then be obtained as
k2
kz2 = µxx εyy k02 − x . (9.2)
µzz
The question of interest is whether the wave is propagating (kz2 > 0) or
evanescent (kz2 < 0). The boundary between the two cases may be easily
obtained from eqn (9.2). It occurs at kc , a critical value of the spatial
frequency, when
√
kx = kc = k0 εyy µzz . (9.3)
In eqn (9.2) there are two parameters, µxx εyy and µxx /µzz . We can
distinguish four cases:
(a) (b)
V = ZI , (9.9)
discussed several times in this book. Each element is then excited by
the current source and the currents in the elements may be determined
by assuming that each element is coupled to every other element. Such
an approach was attempted by Kozyrev et al. (2007). They were able
to find the focus in a narrow frequency band but no regular pattern at
different frequencies.
The general question of interest is when to use effective-medium theory
and when to resort to solution in terms of eigenwaves. The latter is more
accurate, but feasible only when the number of elements are not too large
(the limit posed by computational difficulties is around 104 ). For mag-
netically coupled elements like SRRs or capacitively loaded loops these
waves are magnetoinductive waves. There are of course other eigenwaves
as well, like the electroinductive waves of Beruete et al. (2006) and waves
propagating on an array of dipoles to be discussed in Section 9.6 con-
cerned with nanoparticles.
(a) 1 − ζe
R= , (9.10)
1 + ζe
where ζe is dependent on the propagation coefficients on both sides of
the boundary. Since all finite beams may be constructed by the super-
position of infinite beams this means that the reflection coefficient is
n1>0
different for each plane-wave component. When the reflected beam is
reconstructed from its elements it turns out that the beam still keeps its
n2<0<n1 negative lateral shift shape but it is laterally shifted in the positive direction. This applies to
the normal case, that is when the refractive index on both sides of the
(b)
boundary is positive.
When the beam is incident from a positive-index material upon a
Fig. 9.12 Schematic representation of negative-index material then the same model predicts a negative shift
(a) a positive and (b) a negative lat- (Fig. 9.12(b)). This was shown practically simultaneously by Berman
eral shift of an incident beam upon re- (2002) and Lakhtakia (2003) both of them formulating the problem an-
flection from (a) a positive- and (b) alytically. A similar analytical approach was used by Ziolkowski 2003a;
a negative-index material. From Zi-
olkowski (2003a). Copyright
c 2003 Ziolkowski 2003b who also did numerical calculations with the aid of the
Optical Society of America FDTD (finite-difference time domain) package. The frequency depen-
9.5 Gaussian beams and the Goos–Hanchen shift 301
(a)
dence of the permittivity and permeability, needed for this package, was
taken in the functional form of the Drude model.3 A Gaussian beam of 3
To become negative at a threshold
1-cm waist is incident at a frequency of 30 GHz from medium 1 upon frequency is a good approximation for
the permittivity. However, the known
medium 2. The material parameters of medium 1 are εr = 9 and µr = 1 negative-permeability materials do not
giving a refractive index of n1 = 3. Medium 2 is assumed to be a pos- follow this frequency dependence. The
itive index medium with εr = 3 and µr = 1 in the first case, and a permeability is negative within a band,
negative index medium with εr = −3 and µr = −1 in the second case. and not below a threshold value, as dis-
cussed many times in this book (see,
The incident angle is chosen to be 40◦ above the critical angle of 35.26◦, e.g., Section 2.8). Nevertheless, it
hence the beam in both cases will suffer total internal reflection. The re- seems likely that the model is good
flected and scattered beams yielded by the numerical package are shown enough to find the main features of the
in Figs. 9.13(a)) and (b), respectively. The direction of scattering may incident, scattered and reflected beams.
be immediately seen. For the positive-index material it is in the positive
direction and for the negative-index material it is in the negative direc-
tion. The actual shift is too small to appreciate it just by looking at the
figures. Ziolkowski (2003a) gives values of 32 cells for the positive shift
and 33 cells for the negative shift, both very close to the predictions of
his analytical model. The simulation space was 520 × 1040 cells. The
size of a cell was taken as 0.1 mm.
A large Goos–Hanchen shift was predicted theoretically by Shadrivov
et al. (2003) by having an extra layer between the positive- and negative-
index materials that can bring about the resonant excitation of surface
polaritons.
An interesting consequence of the negative Goos–Hanchen shift is that
in a negative-index waveguide, depending on the parameters of the three
media, propagation may be forward or backward, as shown in Figs.
302 Seven topics in search of a chapter
x 13
A B
2
C D
x 12
(a)
x 13
2 A B A
B
x 12
(b)
x 13
Fig. 9.14 A waveguide consisting of 2 A B
a negative-index material bounded by x 12
two positive-index materials. Parame-
ters are chosen so that (a) the wave
is forward, (b) the wave is backward,
and (c) the wave is stationary. From
Tsakmakidis et al. (2007). Copyright
c 2008 Nature Publishing Group. For (c)
coloured version see plate section
4 9.14(a) and (b) on the basis of a ray picture. It follows then (Tsak-
The authors called the resulting shape
an optical ‘clepsydra’, which according makidis et al., 2006; Tsakmakidis et al., 2007) that for some parameters
to the Oxford Dictionary is an ancient in between, the rays may bite their own tails: the propagation is neither
time-measuring device presumably sim- forward nor backward, it is stationary,4 as shown in Fig. 9.14(c). The
ilar to an hourglass. The pattern could
conclusion may then be drawn that such a device may stop and store
also be described as biconical.
light.
εr (ω) − 1
αe = d3 . (9.11)
εr (ω) + 2
304 Seven topics in search of a chapter
ωp2
εr (ω) = 1 − , (9.12)
ω2
we find that the resonant frequency is
ωp
ω0 = √ . (9.13)
3
The polarizability of the metallic sphere is then obtained by substituting
eqn (9.13) into eqn (9.11). We find
ω2
1 1
= 3 1− 2 . (9.14)
αe d ω0
Since in this section we intend to include radiation effects we should
add to the above equation the radiation damping term derived in Sec-
tion 2.7 for electric dipoles
k03
1
Im = . (9.15)
αe 6 πε0
How will this metallic sphere respond to an electric field? It will set
up an electric dipole. By definition, the dipole moment of the electric
dipole (see Section 1.16) is related to the electric field as
p = αe E . (9.16)
Let us now consider a linear chain consisting of N metallic nanospheres.
The aim is now the same as in Sections 2.9 and 8.2.2, where we derived
the dispersion characteristics of magnetoinductive waves based on the
concept of polarizability. We need two equations, the effect of the elec-
tric field on the polarization of the nth element, and the electric field
at the nth element produced by all the other elements known as IF , the
interaction function. The electric field due to an electric dipole has al-
ready been given in closed form in Section 1.12. Assuming that we have
N dipoles with an interelement spacing of a the field at the nth dipole
is given by summing up the individual contributions. From the require-
6
In eqn (9.17) we used our notation in- ment that αe IF = 1 the dispersion equation was obtained by Weber and
stead of that of Weber and Ford and
also changed to SI units. The disper-
Ford (2004) as
sion equation of Simovski et al. (2005)
is identical but, as mentioned in Sec-
k03 X
1 1 j 1
tion 8.2.2, they show that the imagi-
= − − e −j k0 na (9.17)
nary part of the equation cancels, and αe 2 πε0 nk0 a (nk0 a)2 (nk0 a)3
they also sum up some of the infinite
series.
for the electric dipoles transverse to the direction of propagation.6 It may
be seen that eqn (9.17) is entirely analogous to eqns (8.28)–(8.30) only
the magnetic polarizability should be changed to the electric polarizabil-
ity and µ0 to ε0 . This is not really surprising because in Section 8.2.2 we
used the magnetic dipole approximation to MI waves and in this section
the metallic nanospheres are regarded as electric dipoles.
9.6 Waves on nanoparticles 305
Transverse Modes
1.05
Light line
Quasistatic
1.00 lossless metal
Real(Z/Z0 )
0.95
Weber and Ford also proceeded to compare the quasi-static case with
that containing full retardation. This is shown in Fig. 9.15, which is
analogous to Fig. 8.18. It needs to be noted that Weber and Ford find
the retarded solution not from the infinite series but from a finite num-
ber of elements with an imposed wave vector and assuming complex ω.
For the transverse case, when dipoles are perpendicular to the direc-
tion of propagation, the dispersion curves have a dip at the light line
not present in the longitudinal case, as we have already noted in the
previous chapter.
Another attempt at finding the dispersion equation was made by
Koenderink and Polman (2006). They also solved the dispersion equa-
tion for k imposed and working in the complex ω region. Their calcula-
tions took into account both ohmic and radiation losses. Their results
differ considerably from those of Weber and Ford in the vicinity of the
light line as shown in Fig. 9.16. The two branches, to the right and to
the left of the light line do not cross, which is, apparently, not unusual
for polaritons. Their dispersion curve has zeros only at the band edges,
in contrast to those of Simovski et al. (2005) that yield zero group ve-
locity at a particular value of ka within the Brillouin zone. Since in
both models there is some numerical work involved it is too early to say
whether the two models contradict each other.
306 Seven topics in search of a chapter
5
Re( ) (1015 rad/s)
0
Fig. 9.16 Dispersion curves of
nanoparticle chains by Koenderink and 5
Polman (2006). Copyright
c 2006
by the American Physical Society.
Thin solid (dotted) line, quasi-static 0
dispersion curve of the longitudinal
(transverse) mode. Squares (circles), 5 ine
ht l
infinite chain dispersion curves of the
Lig
We have also seen that both ε and µ may have very small values so
we may end up with a very small refractive index. The consequences
were explored in a theoretical paper by Ziolkowski (2004). He assumed
that both ε and µ are zero but in such a manner that the characteristic
impedance η = (µ/ε)1/2 remains the same as for the adjoining medium.
He made the point that a zero-index medium has a static character in
space but the field magnitudes vary in time. He performed an FDTD
numerical study of a zero-index material flanked by two media of the
same characteristic impedance. He showed that during the rather long
transients the electric field does vary as a function of space. But when
the steady state is reached the field distribution in the adjoining finite-
index materials is the usual sinusoidal variation as a function of space
at a given moment in time but in the zero-index material there is no
spatial variation at all.
It would be difficult to produce a zero-index material, particularly
with the right characteristic impedance, but a low-index metamaterial
is well within the practical possibilities. The prominent effect would then
be on the refraction properties of the material. Let us remember Snell’s
9.7 Refractive index close to zero 307
ω
n1 sin θ1 = n2 sin θ2 and k = n . (9.19)
c 3º
Let us explore the implications of both relationships. We have already
had a good look at refraction in Section 2.11. We found that for an
incident angle of θ1 from medium 1 the refracted ray can propagate in
ș2
any direction from +90◦ to −90◦, depending on the refractive index of
medium 2 (see Fig. 2.23), and of course for certain values of n2 there
can be total internal reflection. Now, the more interesting case arises
when medium 2 has a refractive index of unity and that of medium 1 is
0
close to zero, say, 0.05. Then, for an incident angle of θ1 , the refracted 0 90º
angle is given by ș1
(a)
sin θ2 = 0.05 sin θ1 . (9.20)
d PEC walls
a2
h
Region 1
Region 3
PEC ring + dicl. gap
Region 2
s Fig. 9.19 Geometry of the 3D rectan-
gular metallic waveguide. The H-plane
width, s, is chosen so that region 3 be-
haves as a metamaterial with index of
refraction close to zero. Regions 1 and
3 have unity indices of refraction. From
Silveirinha and Engheta (2007). Copy-
right
c 2007 by the American Physical
Society
^r
^
w ^
u
a
ac
e, m Fig. 9.20 Cross-section of a spherical
scatterer composed of two concentric
layers of different isotropic materials.
ec, mc From Alu and Engheta (2005). Copy-
c 2005 by the American Physical
right
e0, m0 Society
7 -9 6 -3
x 10 x 10
6 5
5 a c = O /100 a c = O /10
Q s /O
2
Q s /O
4
2
4
3
3
2
2
1 1
0 0
0.5 06 07 08 09 10 05 06 07 08 09 10
a/a c a/a c
12
10
08 a c = O /5
Q s /O
2
06
04
02
00
00 02 04 06 08 10
a/a c
Fig. 9.21 Normalized scattering cross-section for a spherical element with three different sizes of the outer radius of the cover.
From Alu and Engheta (2005). Copyright
c 2005 by the American Physical Society
R1 R2
would not be able to see the object. Theoretical proof for this hypoth-
esis was provided by considering a spherical scatterer having materials
constants ε and µ. The scatter from this sphere is to be cancelled by
a coating with material constants εc and µc , shown in Fig. 9.20. The
mathematics is rather complicated: the scattered field must be writ-
ten as a sum of discrete spherical harmonics and it turns out that the
largest term (the dipolar one) can be cancelled. The final results can
be presented in a simple manner. In Fig. 9.21 the normalized scattering
cross-section is plotted as a function of a/ac , where a is the radius of the
sphere to be hidden and ac is the radius of the sphere with the coating
added. The parameters chosen are ε = 4ε0 , εc = −3ε0 , µ = µc = µ0 .
Note that the coating must have negative permittivity. It may be seen
from Fig. 9.21 that there is very good cancellation at a particular value
of the coating thickness when the object is well in the subwavelength
region (ac = λ0 /100 and λ0 /10, where λ0 is the free-space wavelength)
but much worse when the external radius is a fifth of the wavelength.
Thus, the idea works for subwavelength structures but, apparently, loses
its validity for larger objects. It is also true that, with scattering being
shape-dependent, each object requires a separate design for scattering
cancellation.
A different idea was proposed by Leonhardt (2006) and Pendry et al.
(2006). An object is made invisible because the rays of light do not
go through it but go around it, as shown schematically in Fig. 9.22 for
a simple case. Leonhardt (2006) starts with a disclaimer. He quotes
Nachman (1988) who proved that the inverse scattering problem has a
unique solution. If we measure the scattered field in all directions we
can uniquely determine the spatial variation of ε and µ that caused the
scattering. Hence, from our measurements we should always be able to
determine the distribution of the material parameters whether there is
a cloak there or not. Hence, there cannot be perfect invisibility. If our
measurements show that there is nothing there then the only possibil-
ity is that there is indeed no more than empty space there. However,
if we are allowed to talk about rays instead of waves then the situa-
tion changes. Waves diffract, rays are willing to bend if the index of
refraction varies spatially but they do not diffract. This is, of course,
the geometrical optics approximation. In that approximation, perfect
312 Seven topics in search of a chapter
(a) (b)
Fig. 9.23 Electric-field distribution for an incident plane wave in the vicinity of a perfectly conducting shell and a cloak. (a)
ideal parameters, (b) with a loss tangent of 0.1, (c) with an 8-layer approximation to the desired distribution of the material
parameters, (d) with a simplified cloak in which only µr is varying spatially. From Cummer et al. (2006). Copyright
c 2006
by the American Physical Society. For coloured version see plate section
(a) (b)
Fig. 9.25 Experimental field distribution for a copper cylinder (to be hidden) and a ten-layer cloak made up by resonating
c 2006 AAAS. For
elements. (a) in the absence and (b) in the presence of the cloak. From Schurig et al. (2006). Copyright
coloured version see plate section
dielectric host
metal wires
10.1 Introduction
Some births are more difficult than others. Pallas Athene is reputed
to have sprung out of Zeus’ head, and Aphrodite from the foam of the
sea. Other births have been known to be less instantaneous and to have
required the simultaneous efforts of a number of parents. Metamateri-
als belong to that category. It had a long gestation period and many
contributors. Before going into specifics and enumerating the various
prior influences it might be worth having a broader look at Physics that
has often been accorded a privileged position in the ranks of natural sci-
ence. We may even go further and quote Ernest Rutherford saying that
‘Physics is the only science: the rest is stamp collecting.’ This is not
quite true in the twenty-first century. Many other disciplines have been
using rigorous criteria for proving their theses. Nonetheless, we could
claim with good conscience that Physics is still the discipline farthest
away from the art of stamp collecting.
When we come to physics-in-the-making it is no longer a rational
subject. We do not know which hypotheses are useful and which are
useless. We do not know whether any particular approach to solving a
problem is feasible or not. We do not know the direction that further
research should take. We can only guess. It is necessary of course to
say nowadays that the new research we have in mind will benefit society
but we have no criteria to determine what benefits society and what
does not. Let us quote Rutherford again. In 1933, in his speech to the
British Association for the Advancement of Science, he claimed: ‘We
cannot control atomic energy to an extent which would be of any value
commercially, and I believe we are not likely ever be able to do so.’
Atomic energy provides as much as three quarter of French electrical
energy generation at the moment so, at least in the opinion of some
French authorities, the continuing research on atomic energy bore fruits.
316 A historical review
imaging, of being able to reproduce not only the travelling waves asso-
ciated with the object but all the evanescent waves as well. The four
papers taken together acted as a catalyst. Papers started to pour in.
Metamaterials was not an accepted term yet but everyone knew which
the fundamental papers were. People agreed and disagreed. The large
majority agreed, with a few dissenting voices.
Our aim in this chapter is not to review the birth of the subject in
any detail. It is rather to point out a number of crucial springboards
that kept up the momentum after the initial upsurge in popularity. The
seminal papers were a necessary condition for launching the subject,
the springboards were responsible for the continued interest. In the
next section we shall enumerate a selection of topics that were alive and
flourishing at the time of the upsurge and could be regarded as various
forerunners of the subject of metamaterials.
10.2 Forerunners
10.2.1 Effective-medium theory
An early success of materials science was to be able to describe in
terms of macroscopic quantities the response of a real material (one
that consists of atoms and molecules) to electric and magnetic fields.
Every textbook in solid state physics devotes some attention to early
attempts at homogenization and provides an expression for the permit-
tivity (or permeability) of a material, known as the Clausius–Mossotti
equation. These attempts, although made in the nineteenth century
(Mossotti, 1850; Clausius, 1879), still give inspiration today. The theory
of Gorkunov et al. (2002) relies on similar arguments. Another example
is the paper by Belov and Simovski (2005a) that discusses homogeniza-
tion in metamaterials including the radiation term. The arguments are
presented as generalizations of the Clausius–Mossotti equation.
It is also worth mentioning an early example by Lewin (1947) of the
calculation of the effective permittivity and permeability of a medium
loaded with spherical particles. The approach was taken up in the meta-
material context by Holloway et al. (2003).
10.2.4 Plasmon–polaritons
The interaction of plasmas with electromagnetic waves led both to a
new term, plasmon–polaritons and to a variety of phenomena both in
the bulk and on surfaces (see, e.g., Cottam and Tilley 1988; Mahan and
Obermair 1969). In fact, the so-called amplification of evanescent waves
described by Pendry (2000) is due to the excitation of a surface plasmon–
polariton on the far surface of the subwavelength lens. Recent studies
by Kempa et al. (2005) and by Belov and Simovski (2006) that have
their inspiration in the field of metamaterials are based on the earlier
work of Mahan and Obermair (1969).
on backward waves was done after the Second World War by electronic
engineers concerned with device applications. A number of successful
devices like backward-wave oscillators, amplifiers (Beck, 1958; Hutter,
1960) and antennas (Walter, 1965) made their appearance.
Veselago (1968) noted in his original paper the backward-wave charac-
ter of the electromagnetic waves in negative refractive index materials,
which he called left-handed media. Lindell et al. (2001) proposed that
to avoid confusion with chiral materials it would be more logical to in-
troduce the term backward-wave media. Unfortunately, their proposal
was not generally accepted. At the time of writing a plethora of terms
prevail.
Tretyakov, 2007).
10.2.15 Bianisotropy
The main questions arising are the properties of the elements and their
collective effect upon the propagation of electromagnetic waves; exactly
the same as those arising in the study of metamaterials. We could
actually argue that the subject of metamaterials simply swallowed that
of bianisotropy and related topics like chiral and bi-isotropic media. The
introduction of the omega particle by Saadoun and Engheta (1992),
treatment of chiral scatterers by Tretyakov et al. (1996) or more general
books like that of Lindell et al. (1994) have been very specific forerunners
of later studies. Examples are the study of the bianisotropic character
of split-ring resonators by Marques et al. (2002c) and the consideration
of chiral media for negative refraction by Pendry (2004).
3mz
Hx = sin θ cos θ cos ϕ , (B.2)
4πµ0 r3
3mz
Hy = sin θ cos θ sin ϕ , (B.3)
4πµ0 r3
mz
Hz = (3 cos2 θ − 1) . (B.4)
4πµ0 r3
We wish to find the magnetic field at (0, 0, 0) due to all the magnetic
dipoles within a sphere of radius R. For each component we need then to
sum the contribution from each element. In a number of textbooks (see,
e.g., Jackson 1967) the field at (0, 0, 0) is summed in a rectangular co-
ordinate system, which is the logical choice for a cubic lattice. We shall,
in a less rigorous manner, keep the spherical co-ordinates and instead of
328 Field at the centre of a cubical lattice of identical dipoles
summation integrate over the spherical volume. This method will easily
lead to the answer that all three components are zero. The integration
is of the form
1 R π 2π
Z Z Z
hHi i = Hi dV , (B.5)
V 0 0 0
where i = x, y, z, V is the volume of the sphere and
ωp2
ε0 En1 + ε0 εr2 En2 = ε0 En1 . (D.4)
ω2
The aim of this Appendix is to point out the different physical inter-
pretations of eqns (D.3) and (D.4). According to eqn (D.4) the difference
of the normal components of the dielectric displacement in the two me-
dia is equal to the surface charge density which may be shown to be
equal to ε0 (ωp2 /ω 2 )En1 . Equation (D.3) tells a different story. On the
left-hand side the dielectric constant used is the effective dielectric con-
stant and there is nothing on the right-hand side. Hence, we no longer
need to talk of surface charge.
Thus, all depends on the formulation of the problem. If we say that the
plasma’s relative dielectric constant is unity then we need to introduce
the surface charge in the boundary condition. If we use the effective
dielectric constant (as in eqn (D.2)) then we don’t need to worry about
surface charge at all.
This page intentionally left blank
The Brewster wave
E
Taking the positive sign in eqn (3.15) we find the dispersion curve of the
Brewster (Boardman, 1982; Welford, 1991) wave plotted in Fig. E.1(a)
together with
√ the surface plasmon dispersion curve that extends from 0
to ωs = ωp / 2. The upper branch looks similar but not the same as the
bulk plasmon dispersion curve. The curve intersects the y axis at the
same point, ωp , but the asymptote
√ is different. As ω → ∞ we find that
in the present case kx → k0 / 2, in contrast to kx → k0 for the bulk
wave.
The properties of this wave stem from the condition that it needs to
be incident at the Brewster angle. There is then no reflected wave and
we could look at the wave as moving along the dielectric–metal interface
with a propagation coefficient kx . Note that this is not a proper surface
wave. It radiates. It may be called a radiative wave or a leaky surface
wave. It refracts into the metal and propagates in the metal above
the plasma frequency. The relative dielectric constant of the metal is
between 0 and 1.
The condition for no reflection may be obtained from eqn (1.48) as
kz2 ε1
ζe = = 1. (E.1)
kz1 ε2
This is an unusual formulation of the condition of no reflection. Text-
books give it as
r
ε2
tan ϕ1 = , (E.2)
ε1
where ϕ1 is the angle of incidence. Since it is not obvious that the two
expressions are identical we shall show it below. Using the relationships
kz2 ε1 cos ϕ2
ζe = = , (E.4)
kz1 ε2 sin ϕ1
where ϕ2 is the angle of refraction. But from Snell’s law
r
ε1
cos ϕ2 = 1 − sin2 ϕ1 = sin ϕ1 , (E.5)
ε2
with which ζe = 1.
334 The Brewster wave
2 2 2
Brewster
mode
ZZp
1 1 1
SPP mode
0 0 0
0 1 2 3 4 0 1 2 3 4 0 1 2 3 4
kx /k p kx /k p kx /k p
(a) (b) (c)
Fig. E.1 Dispersion curve of the surface wave. (a) no loss; Brewster mode for ω > ωp and SPP mode for ω < ωs . (b) γp = 0.01.
√
(c) γp = 0.1. Asymptotes: the light line ω/k = c (dotted line) and ω/k = c 2 (dashed-dotted line)
It needs to be noted that kz1 and kz2 are now both positive and
real, and the same applies to the dielectric constants. Can the Brewster
wave be regarded as a surface wave? The justification is that the same
pattern moves along the surface undistorted by reflection. Alternatively
we may argue that the two different types of waves leading to the same
dispersion equation may be construed as one being obtained from the
condition that there is no incident wave (Section 1.10), and the other
one (in the present appendix) demanding that the reflected wave should
vanish.
Losses of surface plasmons have already been considered in
Section 3.3.2 and plotted in Fig. 3.6(a) and (b). We shall replot them in
Figs. E.1(b) and (c) for γp = 0.01ωp and 0.1ωp including this time the
upper branch. It may be seen that the effect of losses is to connect the
lower branch and the upper branch. The stop band between ωs and ωp
has now become a rather odd-shaped pass band (Arakawa et al., 1973;
Alexander et al., 1974; Kovener et al., 1976) showing a backward wave
between points where the group velocity gives the (physically impossi-
ble) value of infinity. The mode has very high attenuation, so it has no
practical significance. The upper branch lying to the left of the light
line is not much affected by loss but, as mentioned above, it is radiative.
We shall not consider it further since our interest is in guided waves.
The electrostatic limit
F
The electrostatic limit is sometimes invoked when a proper (meaning all
of Maxwell’s equations) formulation seems too daunting or one is after
a simple analytical solution. But very often it is part of a gradualist
approach. One may argue: An electrostatic approach is easier than a
proper one. If it leads to a solution that makes good sense physically and
if, in addition, it gives reasonable agreement with experimental results
then one might call off the chase for better results and has every right to
be contented. In the absence of experimental results there is, however,
a danger that what makes good physical sense is not necessarily true.
An example is Pendry’s contention that for the ‘perfect lens’ to operate
there is no need to have a negative permeability; it is sufficient to have
εr = −1. This is discussed in more detail in Chapter 5.
What do we mean by the electrostatic limit? It is a kind of hotch-
potch. We say that things vary with time but not a lot, so we are entitled
to ignore the time derivatives of the electric and magnetic fields. Having
neglected time derivatives the electric and magnetic fields got uncoupled
so we can rely on electric quantities alone. Our starting equations are
∇ × E = 0 , ∇ · D = 0 , D = εE . (F.1)
If the curl of the electric field is zero then it can be expressed as the
gradient of a scalar function
E = −∇ϕ . (F.2)
Then, if the dielectric constant is independent of space the equation to
solve is Laplace’s equation
∇2 ϕ = 0 . (F.3)
ϕ1 = A e −j kx x + κ1 z (F.4)
and
336 The electrostatic limit
ϕ2 = B e −j kx x − κ2 z . (F.5)
z The assumptions are the same sort as in Section 1.10 where we discussed
r2
x surface waves. The waves are declining away from the interface in both
r1 media. The difference is that instead of solving the wave equation we
now solve Laplace’s equation. The time variation in the form of exp(j ωt)
is tacitly assumed but made no use of.
Substituting our assumed solution into Laplace’s equation we readily
Fig. F.1 Two semi-infinite dielectrics obtain
r1
Next, we need to satisfy the boundary conditions at the interface that
z both the function ϕ and its derivative in the x direction must be con-
d r2 tinuous across the boundary, and that, in addition,
x
r1
dϕ1 dϕ2
εr1 = ε r2 . (F.7)
dz z=0 dz z=0
The result is that the electric fields decay nicely away from the interface
Fig. F.2 Thin dielectric slab between
at the same rate and the condition
two identical semi-infinite dielectrics
εr1 = −εr2 (F.8)
needs to be satisfied. This last condition may be seen to be much more
restrictive than those given in Section 3.3.1.
In order to find the dispersion equation ω(kx ) we have to substitute
1
This is, of course, a contradiction be- the frequency dependence of εr2 into eqn (F.8), which is1 equal to
cause this particular dependence on the
frequency comes about by taking into ωp2
account the time derivative of the elec- εr2 = 1 − (F.9)
tric field that we were supposed to have ω2
neglected. Nevertheless, these approxi- into eqn (F.8) leading to
mations very often lead to good results.
ωp
ω= √ , (F.10)
1 + εr1
the equation (3.17) quoted in Section 3.3.1. It is a straight horizontal
line that gives good approximation, provided kx is large enough.
ϕ1 = B e −j kx x + κz , (F.11)
ϕ2 = C e −j kx x − κz + D e −j kx x + κ(z − d) , (F.12)
ϕ3 = F e −j kx x − κ(z − d) . (F.13)
F.2 Symmetric slab 337
1 ± e −kx d
ω 2 = ωp2 , (F.15)
(εr1 + 1) ± (εr1 − 1) e −kx d
which in air simplifies to
ω 2 = ωp2 1 ± e −kx d ,
(F.16)
the expression quoted in eqn (3.37). The approximation is very good
for large kx . How good it is otherwise is discussed in Section 3.4.1 and
shown in Figs. 3.18 and 3.19.
This page intentionally left blank
Alternative derivation of
the dispersion equation for
SPPs for a
dielectric–metal–dielectric
structure: presence of a
surface charge
G
We derived the dispersion equation of a metal (plasma) slab by assuming
infinitely large transmission, i.e. taking the denominator of eqn (1.76)
equal to zero. In this appendix we derive the same dispersion equation
in a somewhat different manner. It is not a radically different approach:
it just puts the emphasis on the fields instead on the reflection and
transmission coefficients as it was done before. In addition, this approach
will easily lead to an analytic expression for the surface charge.
We have already given the field quantities both in the dielectric and in
the metal in Section 1.10. That was for a single interface. We shall write
them below for the case when the dielectric extends from z = −∞ to z H1
z = −d/2, and again from z = d/2 to z = ∞, the metal being between z d/2
x H2
then in the range z = −d/2 to z = d/2 (see Fig. G.1). We shall restrict z d/2
here generality and assume a solution in which the transverse electric H1
field is a symmetric function of z. The field quantities, as solutions of
Maxwell’s equations, may then be obtained for the TM mode as follows:
In the dielectric for z < −d/2
Fig. G.1 Dielectric–metal–dielectric
structure
d
κ1 z +
Hy1 = Hd e 2 e −j kx x , (G.1)
d
κ1 κ1 z +
Ex1 = − Hd e 2 e −j kx x , (G.2)
j ωε1
d
kx κ1 z +
Ez1 = − Hd e 2 e −j kx x , (G.3)
ωε1
where Hd is a constant. In the metal for −d/2 < z < d/2
340 Dispersion equation for SPPs: presence of a surface charge
ωp2
kz κ2 d
̺s = Hm 2 cosh . (G.12)
ω ωp − ω 2 2
The surface charge may also be obtained in a less formal manner by
invoking the underlining physics. Clearly, the surface charge will wax
when the transverse current feeds it and will wane when the transverse
current leads it away. Hence, the temporal rate of change of the surface
current density must be equal to the transverse current, or
341
j ω̺s + Jz = 0 . (G.13)
The relationship between the current density and the electric field
was one of the first few equations derived (see Section 1.2). Substituting
that into eqn (G.13) and solving it for ̺s we find that it agrees with that
obtained from eqn (G.12).
Talking of surface charge it is worth mentioning that if a transverse dc
voltage is applied across a metal–dielectric sandwich then a dc surface
charge will reside at the boundary. As a consequence there will be an
ac surface current as well which will alter the boundary conditions for
the magnetic field. For experiments and a discussion see Batke and
Heitmann 1984; Tilley 1988.
This page intentionally left blank
Electric dipole moment
induced by a magnetic field
perpendicular to the plane
of the SRR
H
A magnetic field will set up a current in the SRR, more correctly it will
set up currents both in the inner and outer rings that vary as a function
of angle. Consequently, there will be a voltage distribution between
the two rings that will lead to charges on the inner and outer surfaces.
Alternatively, we may argue that a necessary consequence of the inter-
ring capacitance are charges on the outer surface of the inner ring and
on the inner surface of the outer ring, as shown in Fig. H.1. The dipole
moment at a point ϕ is equal to qd, where q is the charge and d is the
separation of the rings. Due to symmetry only the x component of the
dipole moment will be different from zero, which is given by
px = qd sin ϕ . (H.1)
Hence, Px , the total dipole moment, is its integral between 0 and π
that needs to be multiplied by two because the other side of the ring y M
contributes the same amount (see Fig. H.1). Thus + _ + _
x + _ + _
z _ + _
Zπ +
Px = 2 px dϕ . (H.2)
0
Φ = πr02 µ0 Hθ . (J.2)
By definition
Φ
M= and m = µ0 πr02 I , (J.3)
I
where I is the current flowing in the loop. Hence, the mutual inductance
is
Baena, J. D., Marques, R., Medina, F., and Martel, J. (2004). Artificial
magnetic metamaterial design by using spiral resonators. Phys. Rev.
B , 69, 014402–1–5.
Bahr, A. J. and Clausing, K. (1994). An approximate model for artifi-
cial chiral media. IEEE Trans. Ant. Prop., 42, 1592–1599.
Bai, B., Svirko, Y., Turunen, J., and Vallius, T. (2007). Optical activity
in planar chiral metamaterials: Theoretical study. Phys. Rev. A, 76,
023811–1–12.
Barnes, W. L. (1998). Fluorescence near interfaces: The role of pho-
tonic mode density. J. Mod. Opt., 45, 661–669.
Barnes, W. L. (2006). Surface plasmon-polariton length scales: a route
to sub-wavelength optics. J. Opt. A: Pure Appl. Opt., 8, S87–S93.
Barnes, W. L., Dereux, A., and Ebbesen, T.W. (2003). Surface plasmon
subwavelength optics. Nature, 424, 824–830.
Batke, E. and Heitmann, D. (1984). Rapid-scan Fourier transform
spectroscopy of 2-D space charge layers in semiconductors. Infrared
Phys., 24, 189–197.
Beck, A. H. W. (1958). Space charge waves and slow electromagnetic
waves. Pergamon Press, New York.
Belov, P. A., Hao, Y., and Sudhakaran, S. (2006a). Subwavelength
microwave imaging using an array of parallel conducting wires as a
lens. Phys. Rev. B , 73, 033108–1–4.
Belov, P. A., Hao, Y., Sudhakaran, S., Alomainy, A., and Hao, Y.
(2006b). Experimental study of the subwavelength imaging by a wire
medium slab. Appl. Phys. Lett., 89, 262109–1–3.
Belov, P. A. and Silveirinha, M. G. (2006). Resolution of subwavelength
transmission devices formed by a wire medium. Phys. Rev. E , 73,
056607–1–9.
Belov, P. A. and Simovski, C. R. (2005a). On homogenization of elec-
tromagnetic crystals formed by uniaxial resonant scatterers. Phys. Rev.
E , 72, 026615–1–15.
Belov, P. A. and Simovski, C. R. (2005b). Subwavelength metallic
waveguides loaded by uniaxial resonant scatterers. Phys. Rev. E , 72,
036618–1–11.
Belov, P. A. and Simovski, C. R. (2006). Boundary conditions for
interfaces of electromagnetic crystals and the generalized ewald-oseen
extinction principle. Phys. Rev. B , 73, 045102–1–14.
Belov, P. A., Simovski, C. R., and Ikonen, P. (2005). Canalization of
subwavelength images by electromagnetic crystals. Phys. Rev. B , 71,
193105–1–4.
Belov, P. A., Simovski, C. R., and Tretyakov, S. A. (2002). Two-
dimensional electromagnetic crystals formed by reactively loaded wires.
Phys. Rev. E , 66, 036610–1–7.
Belov, P. A. and Zhao, Y. (2006). Subwavelength imaging at optical
frequencies using a transmission device formed by a periodic layered
352 References
Falcone, F., Martin, F., Bonache, J., Marques, R., and Sorolla, M.
(2004c). Coplanar waveguide structures loaded with split-ring res-
onators. Microw. Opt. Technol. Lett., 40, 3–6.
Fang, N., Lee, H., Sun, C., and Zhang, X. (2005). Sub-diffraction-
limited optical imaging with a silver superlens. Science, 308, 534–537.
Feng, S. and Elson, J. M. (2006). Diffraction-suppressed high-resolution
imaging through metallodielectric nanofilms. Opt. Exp., 14, 216–221.
Feng, S., Elson, J. M., and Overfelt, P. L. (2005). Optical properties of
multilayer metal-dielectric nanofilms with all-evanescent modes. Opt.
Exp., 13, 4113–4124.
Feng, Y., Zhao, J., Teng, X., Chen, Y., and Jiang, T. (2007). Sub-
wavelength imaging with compensated anisotropic bilayers realized by
transmission-line metamaterials. Phys. Rev. B , 75, 155107–1–6.
Fox, M. (2001). Optical properties of solids. Oxford University Press,
Oxford.
Fredkin, D. R. and Ron, A. (2002). Effectively left-handed (negative
index) composite material. Appl. Phys. Lett., 70, 1753–1755.
Freire, M. J. and Marques, R. (2005). A planar magnetoinductive lens
for 3D subwavelength imaging. Appl. Phys. Lett., 86, 182505.
Freire, M. J. and Marques, R. (2006). Near-field imaging in the mega-
hertz range by strongly coupled magnetoinductive surfaces: Experi-
ment and ab initio analysis. J. Appl. Phys., 100, 063105–1–9.
Freire, M. J. and Marques, R. (2008). Optimizing the magnetoinduc-
tive lens: Improvement, limits, and possible applications. J. Appl.
Phys., 103, 013115–1–7.
Freire, M. J., Marques, R., Medina, F., Laso, M. A. G., and Mar-
tin, F. (2004). Planar magnetoinductive wave transducers: theory and
applications. Appl. Phys. Lett., 85, 4439–4441.
French, O. E., Hopcraft, K. I., and Jakeman, E. (2006). Perturbation
on the perfect lens: the near-perfect lens. New J. Phys., 8, 271–1–12.
Friedrich, W., Knipping, P., and von Laue, M. (1912). Interferenz-
Erscheinungen bei Rontgenstrahlen. Sitzungsberichte der (Kgl.) Akad.
der Wiss. Bayer , 42, 303–322.
Froncisz, W. and Hyde, J. S. (1982). The loop-gap resonator: A new
microwave lumped circuit ESR sample structure. J. Magn. Res., 47,
515–521.
Gao, L. and Tang, C. J. (2004). Near-field imaging by a multi-layer
structure consisting of alternate right-handed and left-handed materi-
als. Phys. Lett. A, 322, 390–395.
Garcia, N. and Nieto-Vesperinas, M. (2002). Left-handed materials do
not make a perfect lens. Phys. Rev. Lett., 88, 207403–1–4.
Garcia-Garcia, J., Martin, F., Falcone, F., Bonache, J., Gil, I.,
Lopetegi, T., Laso, M. A. G., Sorolla, M., and Marques, R. (2004).
Spurious passband suppression in microstrip coupled line band pass fil-
References 357
merical simulations for a singly split double ring. J. Appl. Phys., 95,
3778–3784.
Shamonin, M., Shamonina, E., Kalinin, V. A., and Solymar, L. (2005).
Resonant frequencies of a split-ring resonator: analytical solutions and
numerical simulations. Microw. Opt. Technol. Lett., 44, 133–136.
Shamonina, E. (2008). Slow waves in magnetic metamaterials: history,
fundamentals and applications. phys. stat. sol. b, 245, 1471–1482.
Shamonina, E., Kalinin, V. A., Ringhofer, K. H., and Solymar, L.
(2001). Imaging, compression and Poynting vector streamlines for neg-
ative permittivity materials. Electron. Lett., 37, 1243–1244.
Shamonina, E., Kalinin, V. A., Ringhofer, K. H., and Solymar, L.
(2002a). Magneto-inductive waveguide. Electron. Lett., 38, 371–372.
Shamonina, E., Kalinin, V. A., Ringhofer, K. H., and Solymar, L.
(2002b). Magnetoinductive waves in one, two, and three dimensions.
J. Appl. Phys., 92, 6252–6261.
Shamonina, E. and Solymar, L. (2004). Magneto-inductive waves sup-
ported by metamaterial elements: components for a one-dimensional
waveguide. J. Phys. D: Appl. Phys., 37, 362–367.
Shamonina, E. and Solymar, L. (2006). Properties of magnetically
coupled metamaterial elements. J. Magn. Magn. Mater., 300, 38–43.
Shefer, J. (1963). Periodic cylinder arrays as transmission lines. IEEE
Trans. Microw. Theory Tech., 11, 55–61.
Shelby, R. A., Smith, D. R., Nemat-Nasser, S. C., and Schultz, S.
(2001b). Microwave transmission through a two-dimensional, isotropic,
left-handed metamaterial. Appl. Phys. Lett., 78, 489–491.
Shelby, R. A., Smith, D. R., and Schultz, S. (2001a). Experimental
verification of a negative index of refraction. Science, 292, 77–79.
Shen, J. T., Catrysse, P. B., and Fan, S. (2005). Mechanism for design-
ing metallic metamaterials with a high index of refraction. Phys. Rev.
Lett., 94, 197401–1–48.
Shen, J. T. and Platzman, M. (2002). Near-field imaging with negative
dielectric constant lenses. Appl. Phys. Lett., 80, 3286–3288.
Shen, Y. R. (1984). Principles of nonlinear optics. Wiley, New York.
Shvets, G. (2003). Applications of surface plasmon and phonon polari-
tons to developing left-handed materials and nano-lithography. Proc.
SPIE , 5221, 124–132.
Shvets, G. and Urzhumov, Y. (2004). Engineering the electromagnetic
properties of periodic nanostructures using electrostatic resonances.
Phys. Rev. Lett., 93, 243902–1–4.
Sihvola, A. (2007). Metamaterials in electromagnetics. Metamateri-
als, 1, 2–11.
Silin, R. A. and Sazonov, V. P. (1966). Slow wave structures. Sovetskoe
Radio, Moscow. in Russian.
372 References
materials with two resonant elements per unit cell,. Phys. Rev. B , 73,
224406–1–12.
Sydoruk, O., Shamonin, M., Radkovskaya, A., Zhuromskyy, O., Sha-
monina, E., Trautner, R., Stevens, C. J., Faulkner, G., Edwards, D. J.,
and Solymar, L. (2007b). A mechanism of subwavelength imaging with
bi-layered magnetic metamaterials: theory and experiment. J. Appl.
Phys., 101, 073903–1–8.
Sydoruk, O., Shamonina, E., and Solymar, L. (2007c). Parametric
amplification in coupled magnetoinductive waveguides. J. Phys. D, 40,
68796887.
Sydoruk, O., Shamonina, E., and Solymar, L. (2007d). Tayloring of the
subwavelength focus. Microw. Opt. Technol. Lett., 49, 2228–2231.
Sydoruk, O., Zhuromskyy, O., Shamonina, E., and Solymar, L. (2005).
Phonon-like dispersion curves for magnetoinductive waves. Appl. Phys.
Lett., 87, 072501–1–3.
Syms, R. R. A. (1986). Perturbation analysis of nearly uniform coupled
waveguide arrays. Appl. Opt., 25, 2988–2995.
Syms, R. R. A. (1987). Approximate solution of eigenmode problems
for layered coupled waveguide arrays. IEEE J. Quant. Electron., 23,
525–532.
Syms, R. R. A. (2006). private communication.
Syms, R. R. A., Shamonina, E., Kalinin, V., and Solymar, L. (2005a). A
theory of metamaterials based on periodically loaded transmission lines:
Interaction between magnetoinductive and electromagnetic waves. J.
Appl. Phys., 97, 064909–1–6.
Syms, R. R. A., Shamonina, E., and Solymar, L. (2005b). Positive and
negative refraction of magnetoinductive waves in two dimensions. Eur.
Phys. J. B , 46, 301–308.
Syms, R. R. A., Shamonina, E., and Solymar, L. (2006a). Magneto-
inductive waveguide devices. IEE Proc. Microw. Ant. Prop., 153, 111–
121.
Syms, R. R. A., Solymar, L., and Shamonina, E. (2005c). Absorbing
terminations for magneto-inductive waveguides. IEE Proc. Microw.
Ant. Prop., 152, 77–81.
Syms, R. R. A., Solymar, L., and Young, I. R. (2008). Three-frequency
parametric amplification in magneto-inductive ring resonators. Meta-
materials, 2, 122–134.
Syms, R. R. A., Sydoruk, O., Shamonina, E., and Solymar, L. (2007a).
Higher order interactions in magnetoinductive waveguides. Metamate-
rials, 1, 44–51.
Syms, R. R. A., Young, I. R., and Solymar, L. (2006b). Low-loss mag-
netoinductive waveguides. J. Phys. D: Appl. Phys., 39, 1945–1951.
Syms, R. R. A., Young, I. R., and Solymar, L. (2007b). Paramet-
rically amplified magneto-inductive ring resonators. In Proc. 1st Int.
References 375
Tsakmakidis, K. L., Klaedtke, A., Aryal, D. P., Jamois, C., and Hess,
O. (2006). Single-mode operation in the slow-light regime using oscil-
latory waves in generalized left-handed heterostructures. Appl. Phys.
Lett., 89, 201103–1–3.
Uslenghi, P. L. E. (2007). Invisible metamaterial trenches. In Proc. 1st
Int. Congr. on Advanced Electromagnetic Materials in Microwaves and
Optics (Metamaterials 2007), Rome, pp. 486–487.
Valanju, P. M., Walser, R. M., and Valanju, A. P. (2002). Wave refrac-
tion in negative-index media: always positive and very inhomogeneous.
Phys. Rev. Lett., 88, 187401–1–4.
Valentine, J., Zhang, S., Zentgraf, T., Ulin Avila, E., Genov, D. A.,
Bartal, G., and Zhang, X. (2008). Three-dimensional optical metamate-
rial with a negative refractive index. Nature, doi:10.1038/nature07247,
1–5.
Veselago, V. G. (1968). The electrodynamics of substances with simul-
taneously negative values of ε and µ. Sov. Phys. Usp., 10, 509–514.
(translated from Usp. Fiz. Nauk 92, 517, 1967).
Wait, J. R. (1962). Electromagnetic waves in stratified media. MacMil-
lan, New York.
Walter, C. H. (1965). Traveling wave antennas. McGraw Hill, New
York.
Wang, S., Tang, C., Pan, T., and Gao, L. (2006). Effectively nega-
tively reflective material made of negative-permittivity and negative-
permeability bilayer. Phys. Lett. A, 351, 391–397.
Webb, K. J. and Yang, M. (2006). Subwavelength imaging with a
multilayer silver film structure. Opt. Lett., 31, 2130–2132.
Webb-Wood, G., Ghoshal, A., and Kik, P. G. (2006). In situ exper-
imental study of a near-field lens at visible frequencies. Appl. Phys.
Lett., 89, 193110–1–3.
Weber, W. H. and Ford, G. W. (2004). Propagation of optical excita-
tions by dipolar interactions in metal nanoparticle chains. Phys. Rev.
B , 70, 125429–1–8.
Welford, K. (1988). The method of attenuated total reflection, Volume
Surface plasmon-polaritons of IOP Short Meetings Series, pp. 25–78.
Welford, K. (1991). Surface plasmon-polaritons and their uses. Opt.
Quant. Electron., 23, 1–27.
Wendler, L. and Haupt, R. (1986). Long-range surface plasmon-
polaritons in asymmetric layer structures. J. Appl. Phys., 59, 3289–
3291.
Whinnery, J. R. and Jamieson, H. W. (1944). Equivalent circuits for
discontinuities in transmission lines. Proc. IRE , 32, 98–114.
White, J., White, C. J., and Slocum, A. H. (2005). Octave-tunable
miniature RF resonators. Microw. Wireless Comp. Lett., 15, 793–795.
Whitney, A. V., Elam, J. W., Zou, S., Zinovev, A. V., Stair, P. C.,
Schatz, G. C., and van Duyne, R. P. (2005). Localized surface plasmon
References 377
Acoustic waves, 37–39, 76, 228, 321 Clausius-Mossotti, 42, 54, 303, 317, 327
Acronyms, 325 Cloaking, 290, 309–314
Ag, 101, 141–142, 185–187, 194 Complementary split-ring resonator, 45, 127, 205
Al2 O3 , 142, 194 Conductivity, 2, 6, 43
Anisotropic, anisotropy, 134, 191, 201–204, 289, 294–297, Composite right/left-handed structure, 73, 209
312, 318, 344–345 Compression, 179–180, 204
Anomalous dispersion, 22, 208 Confinement, 88, 100–106, 117–118, 322
Antisymmetric, 93, 97–105, 116, 171, 222–223 Continuous line, 29, 142, 207
Artificial materials, 35–40 Coplanar waveguide, 204–208, 320
Asymmetric structures, 100–101 resonant loops, 214
Attenuation coefficient, 5, 84, 97, 105, 206, 215, 298 Coupled
Au, 138, 142, 186–187 lines, 238–241, 248–249, 265, 279, 285, 291
Axial configuration, 24, 56, 216–219, 225–230, 234, nanoparticles, 321
253–257, 268, 279, 298, 347 oscillators, 159
resonant sheets, 290
Backward waves, 1, 10, 21–22, 56–59, 73, 208, 215, 230, resonators, 242, 321
239–243, 254–256, 318-319 surface plasmon–polaritons, 188
Backward-wave media, 60, 319 wave analysis, 142
Bandgap, 36, 191, 296, 322 waveguides, 279, 321
Bianisotropy, 132, 322 Coupling coefficient, 215–216, 227, 234, 246, 253, 257, 260,
Biperiodic lines, 227–229, 235, 247 278–279
Boundary conditions, 1, 4–9, 15, 64–67, 79, 104–106, Cross-layer coupling, 140
146–147, 201, 262–263, 331, 336–341, 345 Current vector, 219
Bragg Cutoff 169
condition, 295, 330 frequency, 173
effect, 36 waveguide, 201
grating, 246
reflection, 247, 322 Delay line, 243
reflector, 246 Descartes’ law, 9
Brewster wave, 80, 333–334 Diatomic chain, 228
Brillouin zone, 260, 305 Dielectric droplet, 196
Broadside-coupled split ring, 132, 265 Dielectric–metal–dielectric, 339
Build-up time, 159 Dipole, 17–19, 327
Bulk polaritons, 77–79 antenna, 290
approximation, 56
Capacitance, electric, 1, 20, 40, 51, 76, 124–125, 129–130, 203, 291,
gap, 45, 122, 127, 145–152 303–304
inter-ring, 45, 126–127, 133, 145–149, 343 field, 203
voltage dependent, 281, 285 field equations, 343
Capacitively loaded half-wave, 184
loop, 231 interactions, 55–56
rod, 293–294 magnetic, 1, 4, 19, 50–56, 123–124, 128, 203, 236,
split ring, 134, 236, 265 267–271, 327
Cavity, moment, 20, 51, 125, 304
magnetron, 120, 242, 352 nuclear, 250, 287
re-entrant, 45, 120, 319 rotating, 242, 250
resonator, 208, 242 Directional coupler, 209, 247
Chain matrices, 25–26, 32, 48 Discretization, effect of, 175–176
Chiral, chirality, 59, 123–124, 139–140, 319–322 Dispersion, 1, 21–23, 76–87, 92–96, 102–106, 159, 189, 208,
Chromium, 185, 194 217, 256, 298, 312
382 Index
characteristics, 49, 73, 77, 210–213, 227–231, 236–240, Gaussian beam, 301–302
247–250, 256–257, 266–269, 298, 304, 322 Generalized Ohm’s law, 218, 240, 271, 299
curve, 29–30, 39, 47–49, 73–84, 89–92, 98, 102–103, 116, Geometrical optics, 70–72, 155, 179, 192, 197, 311
193, 208, 214, 233–241, 252, 256, 260, 269–270, Gold, 69, 135–141, 187, 195–196
274–280, 298, 305–306, 333–334 Goos–Hanchen shift, 289, 299–302
equation, 28–29, 37–39, 46–49, 53–56, 74–83, 106–116, Group velocity, 22, 29, 39, 57–59, 69, 189, 225, 252–257,
193, 213–225, 231, 237–242, 251–254, 260–261, 272, 293, 298, 305, 334
266–270, 278, 297–298, 304–306, 334–341 Growing wave, 69, 206–207, 294
relation, 28, 103, 203, 214–217, 232–233, 279
Displacement current, 140, 144, 146 Hairpin resonator, 125, 140
Distributed circuits, 144 Helix, 124
Double circular elements, 123 Hexagonal lattice, 259–265
Double ring, 122, 126–129 Higher
quadruply split, 126–127 order interaction, 236–238
singly split, 122, 148, 152 resonance, 150–153
Drude model, 76–83, 106, 111, 301–304, 317 Historical review, 58, 120–126, 315–323
Hybrid modes, 76
Effective Hyperbolic
medium, 36, 148, 191–193, 299, 317 dispersion relation, 193, 298
permeability, 36, 52–55, 138–139, 200–201, 318 isofrequency curve, 298
permittivity, 14, 36, 40–42, 54, 123–124, 317 Hyperlens, 194
plasma frequency, 42–44, 202, 318
refractive index, 65 Imaging,
Eigensolution, 15, 93, 147 by channelling, 265, 291–294
Eigenvector, 214, 218–220, 222 near-field, 66, 249, 321
Electric resonance, 130, 136 of M-shaped wire, 265
Electroinductive waves, 299 of magnetic resonance, 229, 267, 320
Electrostatic approximation, 94–95, 102, 164–167 perfect, 66, 92, 159, 206, 323
failure of, 180-183, 197 subwavelength, 68–72, 79–81, 87, 117, 155–199, 290, 294,
Electrostatic limit, 70, 83, 162, 165–167, 170–171, 335 321
Equivalent circuit, 46, 74, 126–127, 133, 145–148, 209, 214, Impedance matrix, 218–220, 224, 251–252, 262, 271
280, 285 Indefinite media, 289, 296–298
Evanescent wave, 72, 156, 160, 176, 187–189, 197, 210, Inductance,
223, 290–294, 298, 317–318 kinetic, 24–25, 136
Ewald circle, 1, 8–11, 57, 254 mutual, 23-24, 29, 38, 45–46, 53–56, 73, 140, 144, 148,
Excitation of 209, 214, 225–228, 237–247, 252–254, 258–262,
loaded transmission lines, 210–211 267, 271, 278–279, 298, 327–328, 347
magnetoinductive waves, 214–224, 238, 251, 259–264, self-, 29, 45–48, 140
268-271, 276, 287 Inductive coupling, 140
surface plasmon–polaritons, 162, 168–171, 182, 188, Inhomogeneous, inhomogeneity, 157, 312, 320
195–197, 301, 318 Interaction,
higher order, 236–238
Fermat’s principle, 11–12 nearest-neighbour, 38, 203, 236–239, 252, 259, 266, 303,
Ferrite, 204, 318–320 321
Filling factor, 111 Interdigital, 209
Filters, 119, 125–126, 190–191, 199, 204–206, 320 Inter-ring capacitance, 45, 126–127, 133, 145–149, 343
Flat lens, 59–67, 71, 155–158, 179, 185–192, 290, 298, 316 Internal focus, 61–62, 72, 197
Focus, 59–63, 72, 155, 179, 186–189, 197, 210, 266, 289, Inverse scattering, 311, 321
298–299, 316, 321 Invisibility, 290, 309–314
Forward waves, 57, 156, 230, 243, 254–256 Isotropic resonator, 133–135
Fourier
analysis, 1, 31 Kinetic
coefficient, 273 inductance, 24–25, 136
component, 158–161, 179, 271–275 resistance, 25
plane, 143 Kramers–Kronig relations, 159
spectrum, 155, 172, 273–277 Kretschmann configuration, 81, 362
transform, 31–33
Four-poles, 25–30, 38, 45–49, 125, 213 Laplace’s equation, 83, 335–337
Full ring, 130–131, 151, 303 Leaky wave, 99, 269
Left-handed materials, 5, 59, 64, 73, 110, 157, 202
Gap capacitance, 45, 122, 127, 145–152 transmission line, 209–210
Index 383
Pass band, 37–39, 47–50, 63, 73, 84, 200–205, 222–223, Pump wave, 227–231, 250, 278–280, 284–288
227–228, 232–233, 239, 244, 254, 260, 278, 334
Penetration depth, 85–88, 100 Quality factor, 51–54, 151, 215–216, 229–233, 243,
Percolation, 139 272–275, 286
Perfect imaging, 66, 92, 159, 206, 323 Quartz, 186
Perfect lens, 32, 66–72, 81, 107, 115–117, 155–165, Quasi-static, 213, 266–270, 303–306, 327
169–181, 290, 297, 316, 321, 335
Perfect tunnelling, 206–207, 296 Radiation damping, 50–52, 268, 304
Periodic structure, 102–103, 158, 236, 290, 295, 319–321 Radiation loss, 51–52, 268–269, 305
Permeability Radiation resistance, 18–19, 51, 271
effective, 42, 52–55, 138–139, 200–201, 318 Reciprocity, 25–26, 30, 130
lossy, 167–168 Rectangular waveguide, 199
negative, 37, 53–54, 63, 106–108, 119, 126, 138–141, 202, Re-entrant cavity, 45, 120, 319
289–292, 296–297, 301, 316–320, 335 Reflection coefficient, 65, 160, 172, 244–245, 300
transverse, 204 Refractive index,
tensor, 19–20 deviation in, 169–174
Permittivity near zero, 307
effective, 40–42, 54, 123–124, 317 Resolution, 66–67, 72, 159, 165–176, 182–186, 190–197,
292–293, 321
lossy, 164–167
Reststrahl, 79, 186
negative, 37, 80, 119, 140, 141, 176–188, 207, 289–292,
Resonance,
296–297, 311, 316–318
electric, 130, 136
tensor, 19–20
enhanced tunnelling, 191
Phase shifter, 199, 204, 208–209, 243
higher, 150–153
Phase matching, 195, 278–280, 285
LC, 128, 140, 150
Phase velocity, 9, 22, 29, 57–60, 76, 156, 228, 250–256,
magnetic, 111–114, 121, 125–140, 229, 244, 249–250,
272, 278, 289, 312, 320
267, 320
Phonon, 76–78, 228
of triangle particle, 143–144, 303
Phonon–polariton, 76–79, 186
plasma, 75, 135–142, 303
Photonic
rotational, 242–244, 249–250, 288
bandgap, 36, 191, 322
spatial, 251, 259–265
circuit, 86
Resonant
crystals, 294–296 elements, 37, 44-45, 203, 242, 259, 277, 290–291, 303,
Photoresist, 185 319–323
Planar array/configuration/line, 23–24, 103–105, 140, loops, 46–51, 56, 214, 220, 243–246
215–217, 221–222, 228-240, 252–256, 268, Resonator,
276–279, 285–288, 347 complementary split ring, 45, 127
Planar–axial configuration, 240, 252–257, 298 hairpin, 125, 140
Plasma, 75 isotropic, 133–135
asymmetric structures, 100–101 open, 135
frequency, 14, 25, 37, 42–44, 76–78, 82, 102, 106, spiral, 126, 205
111–112, 137–140, 185, 202, 317–318, 322, 333 small, 119–154, 199
resonance, 75, 135–142, 303 split ring, 45, 63–65, 73, 111, 124–137, 145–152, 184,
simulation, 42 199–205, 233, 253, 265, 292, 298–299, 308, 313,
solid state, 318 329, 343–344
wave, 6, 37, 76–78, 83 Retardation, 3, 176, 213, 267–271, 305, 322, 347
Plasmon–polaritons, 76–118, 156–162, 168–169, 174–189, Right-handed, 5, 59, 73, 124, 209–210
195–197, 290, 294, 334, 339–340 Rods, 25, 39–40, 48–50, 63–64, 73–75, 111, 119, 139–140,
Polarizability, 1, 20–21, 41–42, 50–56, 124–125, 268, 304, 204–205, 293
343 Rotational resonance, 242–244, 249–250, 288
Polylogarithm, 238, 363
Poly-methyl metachrylate (PMMA), 185–186, 195 Scaling, 137, 152
Poynting vector, 18–19, 59, 72, 77, 90–92, 98, 190–193, Scattering, 193, 301, 311
197, 214, 224–225, 296 coefficients, 30–31, 43, 65, 200, 321, 329
optics, 189 cross section, 143–144, 310–311
streamlines, 71, 189, 311 inverse, 311, 321
Printed circuit board (PCB), 64, 234 Short-rod pairs, 139–141
Propagation constant (see also wave number), 27, 97, 215, Signal processing, 104, 234, 243, 303
220, 227–233, 279, 307 Signal wave, 227–230, 278, 283–287
Propagation length, 76, 83–88, 97–101, 106 Silver, 67–72, 82, 86, 101–103, 137, 141–143, 174–177,
Pseudochiral media, 123 181–197
Index 385
Silver lens, 82, 174, 181, 185–192, 197 Thin layers, 190, 345
Simulation, 42, 63–65, 120, 128–131, 137–143, 152–153, Total internal reflection, 11, 59, 207, 255, 300–301, 307
160, 175–176, 191–194, 205–207, 292–293, Transducer, 233
300–301, 308, 312 Transfer function, 31–33, 66–70, 75, 92, 156–183, 190–191,
Single interface, 77–80, 92–100, 106–117, 335–339 197
SiC, 79, 186–188 Transmission,
SiO2 , 185–187 coefficient, 8–10, 16, 30, 65, 200–201, 226–227, 244–247,
Slab, 257, 266, 321, 329–330, 339
plane wave incident upon, 16–18 line, 26–29, 45–50, 73–74, 121–123, 150, 205–209
reflection by, 16 left-handed, 209–210
silver, 101, 177, 181–182, 188–189 spectra, 128–131
surface plasmon–polaritons on, 92–102, 116–118 Transverse
transfer function of, 16, 70 electric mode (TE), 7–10, 17, 21, 79, 107–117, 138–139,
Slotted-tube resonator, 121–122 156, 160–163, 168, 174, 312
Skin depth, 86–87 magnetic mode (TM), 7–10, 15–16, 67–70, 79–117,
Slow wave structures, 320–322 138–139, 156–160, 167–169, 176, 300, 313, 339
Small resonators, 119–154, 199 permeability, 203–204
Snell’s law, 8–12, 59, 70, 306, 316, 333 Triangle particle, 143–144, 303
Spatial frequency, 31–33, 66–68, 158–159, 163, 174–179, Tungsten mask, 185
188, 197, 297–298 Tunnelling, 191, 206–207
Spatial resonance, 251, 259–265 perfect, 206–207, 296
Spherical scatterer, 309–311
Varactor diodes, 134, 250, 278
Split pipe, 235–241, 249, 265, 308
Veselago’s lens, 70–73, 155–156, 298
Split-ring resonator (SRR)
Via, 133
broadside coupled, 132, 265
Voltage vector, 218–219, 258, 271
complementary, 45, 127, 205
edge coupled, 209 Wave,
mathematical model of, 144–152 Brewster, 80, 333–334
Square waveguide, 200, 206 equation, 2–7, 21, 27–28, 261, 297, 336
Staple structure, 138–139 four-pole, 28–30, 38, 213
Stop band, 37, 47–50, 63, 73, 78–79, 110, 132, 200–208, idler, 278, 288
223–224, 232, 239–241, 247–249, 277, 334 leaky, 99, 269
Stored energy, 18, 225, 253 nanoparticle, 87, 268, 289, 299–306, 321–322
Stratified media, 190 number (see also propagation constant), 1, 5, 21–22,
Stripe, 103–106, 139–140 31–33, 37, 57, 76–81, 89, 203, 269–270, 274–276
Subwavelength imaging, 68–72, 79–81, 87, 117, 155–199, pump, 227–231, 250, 278–280, 284–288
290, 294, 321 signal, 227–231, 283–287
Superdirective, superdirectivity, 308–310, 321 vector diagram, 10, 251
Superlens (see also perfect lens), 69, 168, 175, 184–197 Wavefront conversion, 307–308
Superresolution, 321 Waveguide,
Surface plasmon–polariton (SPP), 76–118, 156–162, components, 244–248
168–169, 174–189, 195–197, 290, 294, 334, coplanar, 204–208, 320
339–340 couplers, 209–210, 247
Surface waves, 14–16, 33, 69, 83, 98, 103, 107, 333–336 hollow, 199
Swiss roll, 126, 243, 264–265 microstrip, 133, 199, 204–205, 209–210, 233
rectangular, 199
Tensor, 2, 297 square, 200, 206
permeability, 19–20, 52, 123, 200–201, 298–300, 312, 318 Wire medium, 42–44
permittivity, 19–20, 123, 193, 312
polarizability, 20–21, 51, 124, 343–345 ZnS, 138
This page intentionally left blank
Authors