Plasmonic Optics Theory and Applications
Plasmonic Optics Theory and Applications
1
2 Chapter 1
radar, and microwaves, to the shorter wavelengths of visible light and very
short ultraviolet light.
The electromagnetic wave theory describes the fundamental concepts of
the interaction processes between matter and electromagnetic energy in
numerous scientific disciplines. Since the 20th century, it has been understood
that the theory of quantum electrodynamics can give a more accurate and
fundamental explanation of some phenomena related to photons, photon
scattering, and quantum optics. Regardless, the sophisticated electromagnetic
wave theory has been furthering our comprehension of the flourishing
scientific disciplines of plasmonic optics.
In electromagnetic wave theory, the conventional Maxwell’s equations are
formulated via vector notation in two forms of integral equations and
differential forms. The two forms are mathematically equivalent and useful in
same physical meaning. The integral equations are often used to directly
calculate fields in the conditions of symmetric distributions of electric charges
and electric currents. At the same time, the different equations are convenient
formulations in more complicated situations, such as using the finite element
analysis method.
There are also microscopic and macroscopic variants of Maxwell’s
equations. The differential forms of macroscopic Maxwell equations are as
follows:
∇E¼ B, ∇ H ¼ D þ Jext , (1.1a)
t t
∇ · B ¼ 0, ∇ · D ¼ r, (1.1b)
r ¼ rb þ rf , J ¼ J b þ Jf : (1.2)
The bound charge density rb and current density Jb are defined by introducing
the terms of polarization P and magnetization M in the form of
P
rb ¼ ∇ · P, Jb ¼ ∇ M þ : (1.3)
t
The conservation law for the electric charge density and the external current
density is described as
∇ Jext ¼ r: (1.4)
t
Equation (1.4) indicates that the divergence of the current density leaving a
volume equals the decrease rate of the charge density in the same volume. This
equation can be obtained from Eqs. (1.1) by taking the divergence of
Ampère’s circuital law and introducing Gauss’ laws for electric fields.
In the case of a nonmagnetic medium, where only the free current density
remains, the magnetization and the subsequent bound current density are
zero. The continuity equation [Eq. (1.4)] becomes the partial component
corresponding to the free charges
∇ Jf ¼ r: (1.5)
t f
In a bulk matter without electric charges or currents, the macroscopic
Maxwell’s equations are presented in the form of the microscopic variant:
B
∇ · E ¼ 0, ∇ E ¼ m0 , (1.6)
t
E
∇ · H ¼ 0, ∇ H ¼ ε0 : (1.7)
t
Equations (1.6) and (1.7) describe an electric and magnetic field in a vacuum,
where m0 and ε0 are the permittivity and permeability in the free space of
vacuum (m0 4p 10–7 henry/meter and ε0 8.85 10–12 farad/meter).
They refer only to the cases that do not account for the medium without
charges and currents; they do not mean that the space of the medium is empty
of charge or current. For an oscillating wave exp(jvt), Eqs. (1.1a) can be
rewritten in the simple form of multiplication by the time derivatives:
4 Chapter 1
X
N
Uðr,tÞ ¼ Ai expðjvi t þ jk i rÞ: (1.11)
i¼1
materials, they are position dependence within the materials. For anisotropic
materials, ε and m are in the form of vectors, while for isotropic materials, they
are denoted as scalars. Because general materials are dispersive, both ε and m
are dependent on the frequency of electromagnetic wave.
The vast majority of natural materials are electrically neutral at the
macroscopic level. This macroscopic electrical neutrality is the consequence
of the internal equilibrium of collective charge interactions. When an electro-
magnetic wave impinges on electrical neutrally media, the time-dependent
electric and magnetic field will induce separate oscillatory charge displace-
ments. Such local separation of positive and negative charges from their
original positions in opposite directions manifests the media to present in the
form of induced electric dipole momentum. These phenomena are called
polarizations. Similar effects induced by magnetic field are magnetization or
magnetic polarization. For an inhomogeneous medium with polarization and
magnetization, the continuous constitutive relations are described to be the
functions of space coordinates and time variables:2
ðt
Mðr,tÞ ¼ m0 km ðr,t tÞHðr,tÞdt, (1.14b)
`
6 Chapter 1
where the scalar functions of ke and km are the electric and magnetic
susceptibilities, both of which are time-dependent scalar. In the frequency
domain, the aforementioned induced polarizations are represented by a
multiplication operation:
pffiffiffiffiffiffiffiffiffiffi
B ¼ mH þ M0 þ ε0 m0 kme E, (1.17b)
where ε ¼ ε0(1 þ ke) and m ¼ m0(1 þ km) are the absolute permittivity and
permeability, respectively. P0 and M0 are called the static electric polarization
and static magnetization in the absence of external alternating fields. The
dimensionless tensors of kem and kme are introduced to describe the coupling
effect between the electric polarization induced by a magnetic field and the
magnetization induced by an electric field. The magnetic susceptibility
describes the linear magneto-electric effect, which is often observed in
artificial bi-anisotropic materials, such as split-ring resonators and wires.
B ¼ m0 H þ M ¼ m0 ð1 þ xm ÞH, (1.18b)
where P ¼ ε0xeE and M ¼ m0xmH are the polarization density and the
induced magnetization within dielectric media, respectively. The introduced
terms of xe and xm are called the electric and magnetic susceptibility,
respectively. The relative permittivity and permeability are denoted as
εr ¼ 1 þ xe and mr ¼ 1 þ xm. Both the electric displacement D and magnetic
field B can be considered to be spatially dependent in inhomogeneous media
(and likely time-harmonic dependent, too).
Likewise, the instantaneous electromagnetic responses of a homogeneous
dielectric medium at a certain time depend on the fields at that time and the
evolutionary progress over a period of past time. Thus, the constitutive
relations involve the time evolutionary variation as:
ðt
DðtÞ ¼ ε0 EðtÞ þ ε0 xe ðt tÞEðtÞdt, (1.19a)
`
ðt
BðtÞ ¼ m0 H þ m0 xm ðt tÞHðtÞdt: (1.19b)
`
The constitutive relations are still valid as long as the susceptibilities are
independent of the strength of the electric and magnetic fields. The
evolutionary constitutive relations in the frequency domain are written as
DðvÞ ¼ ε0 ½1 þ xe ðvÞEðvÞ, (1.20a)
B0 ejðvtdÞ B0 jd
mðvÞ ¼ ¼ e : (1.22)
H0 ejvt H0
The complex permeability is a scalar for an isotropic medium and a tensor for
an anisotropic medium. The complex permeability is translated from a polar
coordinate form to a rectangular coordinate one as
mðvÞ ¼ B0 =H0 ðcos d j sin dÞ ¼ m1 ðvÞ jm2 ðvÞ, (1.23)
where m1(v) and m2(v) refer to the real and imaginary parts of the complex
permeability, respectively. Likewise, a phase delay emerges between a
displacement field and an electric field when the frequency of an
electromagnetic wave increases a certain amount. A complex permittivity is
used to describe the polarization within a medium; it is expressed in the form
of real and imaginary parts as
εðvÞ ¼ D0 =E0 ðcos d j sin dÞ ¼ ε1 ðvÞ jε2 ðvÞ: (1.24)
In Eqs. (1.23) and (1.24), the real parts of ε1(v) and m1(v) are connected with
the stored energy within a medium, while their imaginary parts are related to
the energy dissipation. The complex permittivity and permeability are usually
10 Chapter 1
ε ¼ ε0 ð1 þ xe Þ ¼ ε0 εr , (1.25a)
m ¼ m0 ð1 þ xm Þ ¼ m0 mr : (1.25b)
where n1 and n2 are the real and imaginary value of the refractive index. The
complex refractive index can then be determined by
qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 ε
n1 ¼
2
ε1 þ ε21 þ ε22 , n2 ¼ 2 : (1.26b)
2 2n1
The imaginary part n2 implies the energy absorption within a medium. Taking
into account the permeability, the refractive index is expressed as
N q2e s0
sðvÞ ¼ ¼ , (1.30)
me ð1=t jvÞ 1 jvt
where me is the effective mass of an electron, N refers to the electron number
in unit volume, and g is the damping ratio of the oscillating electrons. The
term s0 is defined as
N q2e g
s0 ¼ : (1.31)
me
In the presence of an external electric field, the loss tangent is defined by the
ratio of the loss component to the lossless component in Eq. (1.28):
vε2 þ s
tan h ¼ : (1.32)
vε1
For dielectric media with a small amount of energy loss, Eq. (1.32)
approaches a simple form by the approximation tan h h. Likewise, the
electric and magnetic loss tangent of a transparent dielectric medium are
quantified similarly in the subsequent definitions of
m2 ε2
tan hm ¼ , tan he ¼ : (1.33)
m1 ε1
The loss tangent quantifies the inherent dissipation of electromagnetic energy
due to the imaginary parts of its dielectric properties.
where z denotes the penetrating distance from the air–medium interface. The
attenuation factor a(v) depends on the frequency of the impinging
electromagnetic wave. I(z) and I0 are the penetrating intensity at depth z
and the initial intensity at the interface, respectively.
A penetration depth is defined as the penetrating distance at which the wave
intensity decays to the value 1/e (0.37) of its initial intensity. Conventionally,
the penetration depth is defined in terms of the attenuation factor
In the condition of a static electric field, the current density within a conductor
wire distributes uniformly throughout the cross-section of the wire. On the other
hand, when the frequencies of an electromagnetic wave increase, the current
flow tends to approach the surface of a conductor medium. The penetration
depth decreases sharply with the frequency increasing. A similar concept of skin
depth is used to describe the current flow in a conductor medium.
The energy of a wave propagating through a particular medium is
proportional to the square of its intensity. From the energy point of view,
when the amplitude of the energy attenuates to 1/e of its initial energy, the
depth below the air–medium interface is defined to be the skin depth of a
medium, which relates with the penetration depth as
1 1
ds ðvÞ ¼ ¼ dp ðvÞ: (1.36)
2aðvÞ 2
Figure 1.1 Schematic diagram of current distribution within conductor media. (a) Current
flows through a cylindrical conductor wire and (b) in metal strips of a composite structure.
The dark color indicates regions with a higher electromagnetic amplitude. The regions of
metal and dielectric materials are labeled by their permittivity.
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
dS ðvÞ ¼ 2rc =vm 1 þ ðrc εvÞ2 þ rc εv, (1.37)
Equation (1.38) applies to the materials with a larger skin depth at a much
lower frequency. For instance, it would hold true for metal copper at
frequencies below 1018 Hz. Nevertheless, for materials with a rather low
conductivity corresponding to a shallow skin depth or materials that work at
frequencies much higher than the value of 1/rcε, the skin depth approaches an
asymptotic value rather than monotonously decreasing:
pffiffiffiffiffiffiffiffi
dS ðvÞ 2rc ε=m: (1.39)
Figure 1.2 Spectral dependence of skin depth for frequently used metals when a plane
wave impinges on the air–metal interface at normal incidence. The parameter values for the
estimation of skin depth are adopted from Rodrigo et al.7
Table 1.1 Bulk resistivity and skin depths of frequently used noble metals.
Skin Depth (mm)
Bulk Resistivity
Material (Ω 10 8 m) 1 MHz 10 MHz 100 MHz 1 GHz 10 GHz 100 GHz
Aluminum 2.65 81.9 25.9 8.19 2.59 0.819 0.259
Copper 1.69 65.4 20.7 6.54 2.07 0.654 0.207
Gold 2.2 74.7 23.6 7.47 2.36 0.747 0.236
Graphite 783.7 1409 446 141 44.6 14.1 4.46
Lead 20.6 228 72.2 22.8 7.22 2.28 0.722
Nickel 6.9 9.3 3.0 .093 0.30 0.093 0.030
Silver 1.63 64.3 20.3 6.43 2.03 0.643 0.203
Titanium 54 370 117 37.0 11.7 3.70 1.17
Optical Properties of Plasmonic Materials 15
where n2(v) denotes the imaginary part of the complex refractive index. So
both the penetration depth and the skin depth can be estimated directly from
the imaginary part of the refractive index.
E Empty
hw
Filled
k
Band Band
gap gap
Valence
Valence band
Valence band
band
insulator semiconductor metal
Momentum k
Figure 1.4 Typical band structures of insulator, semiconductor, and metal materials.
bound electrons from lower energy levels into the empty conduction band.
Such interband transitions often occur in the light–matter interaction at the
visible or infrared wavelength. The process of interband transitions produces a
pair carrying an electron and a hole. In practice, only a portion of interband
transitions contribute to the conduction mechanism, which corresponds to a
few specific energy bands.
Solid materials have their own energy band structures, as shown in Fig. 1.4.
The various characteristics of energy band structures determine the multitudi-
nous electrical properties of solid materials. For example, the bandgap of the
undoped silicon is 1.1 eV, whereas diamond has a 5-eV bandgap.8
The electron transitions in a medium between the energy levels determine its
dielectric function and then the optical properties. For an insulator, a large gap
exists between the valence and conduction band, as shown in Fig. 1.4(a). No
electron jumps into the conduction band without enough energy. The electrons
in insulators are rare compared with semiconductors. As shown in Fig. 1.4(b),
semiconductor materials have basically the similar band structure as insulators
have; however, the bandgap Eg of semiconductors is much smaller than that of
insulators. The number of electrons within semiconductors would be increased
greatly after the photons are excited to achieve electron conduction. A deep
doping will greatly reduce their semiconductor band gaps, which enables more
available electrons to contribute to the interband transition.
In the case of metal materials, however, the energy bands overlap each
other or, in certain conditions, are partially filled, as shown in Fig. 1.4(c). The
overlap of valence bands and free conduction bands enables electrons to move
freely across the bands. Interband transitions within metals can easily occur
under the excitation by photons with low frequencies. The free carriers
contribute largely to the conduction. Although the densities of free carriers are
low, the high mobility of free carriers contributes to the conductivity of metals.
For semiconductors, the energy gaps are often temperature dependent.
For example, the energy gap of germanium depends on the temperature as
Optical Properties of Plasmonic Materials 17
C, γ e¯ m
·
Nucleus r
P = -er
Figure 1.5 Schematic of electron motion for a harmonic oscillator model in an external
electric field.
18 Chapter 1
2 r r
me þ me g þ me v20 r ¼ qe Eo ejvt : (1.42)
t 2 t
In general, the condition g . 0 leads to the damping attenuation.
The instaneous distance solution of Eq. (1.42) for a monochromatic
electric field is solved to be
qe
rðvÞ ¼ EðvÞ: (1.43)
me ðv20 v2 jgvÞ
The polarization of a dipole moment causes the electric-charge timing
distance to deviate from its balanced position. The gross polarization of the
whole electrons in unit volume is
N e q2e
P ¼ N e qe rðvÞ ¼ EðvÞ, (1.44)
me ðv20 v2 jgvÞ
where Ne is the electron density per unit volume. The induced dielectric
polarization density is proportional to an electric field by the constant electric
susceptibility as
P ¼ ε0 xe EðvÞ ¼ ε0 ðεr 1ÞEðvÞ: (1.45)
The susceptibility is xe ¼ εr – 1. The permittivity function is thus obtained:11
v2p
εðvÞ ¼ 1 þ , (1.46)
v20 v2 jgv
where vp is the plasma frequency of the bulk metal, which is calculated by5
N e q2e
v2P ¼ : (1.47)
me ε0
The dielectric function [Eq. (1.46)] presents explicitly the real and imaginary
parts in the form of
Optical Properties of Plasmonic Materials 19
v2p ðv20 v2 Þ
ε1 ðvÞ ¼ 1 þ , (1.48a)
ðv20 v2 Þ2 þ g2 v2
gv2p v
ε2 ðvÞ ¼ : (1.48b)
ðv20 v2 Þ2 þ g2 v2
d 2r dr
me þ me g ¼ qe Eo ejvt , (1.49)
dt2 dt
where qe is the electric charge of a free electron. the damping effect g is
proportional to the Fermi velocity g ¼ n/l ¼ 1/t, where n denotes the Fermi
velocity, and l is the mean free path of an electron between successive collision
events. The relaxation time t is the averaged interval time between subsequent
collisions of an electron. In general, the relaxation time is about 10–14 s. The
solution of instaneous distances in Eq. (1.49) for a monochromatic electric
field is solved to be
qe
rðvÞ ¼ EðvÞ: (1.50)
me ðv þ jvgÞ
2
The polarization density is the total dipole moment per unit volume. The
gross polarization of all of the electrons in the unit volume is
N q2e
P ¼ N qe rðvÞ ¼ EðvÞ, (1.51)
me ðv20 þ jgvÞ
where N is the electron density per unit volume. A comparison of Eqs. (1.45)
and (1.51) leads to the frequency-dependent permittivity of metals as
v2P
εðvÞ ¼ 1 , (1.52)
v2 þ jvg
where vp is the plasma frequency of the bulk media, which has a similar
physical mechanism as the definition in Eq. (1.47). At the plasma frequency,
the electromagnetic response of a material changes from the metallic
behaviors to those of dielectric materials. At the frequency below the plasma
20 Chapter 1
v2P v2P
εðvÞ 1 þ j g: (1.54)
v2 v3
The permittivity of metal gold is calculated by Eq. (1.53), ranging from
visible to infrared wavelengths. As shown in Fig. 1.6, the real part of the
permittivity for metal gold remains negative throughout the wavelengths in
which the frequencies are less than the plasma frequency. The negative
permittivity results in an imaginary part of the refractive index. For metals,
the imaginary part of the permittivity implies an energy dissipation that is
relevant with the thermal motion of electrons.
Figure 1.6 indicates that, in the infrared wavelengths longer than
650 nm, the Drude model gives an accurate permittivity of gold, and the
experimental data clearly agree well with the Drude theory. However, there is
an obvious deviation of the imaginary part in the visible wavelengths shorter
than 650 nm. The measured imaginary part increases much more sharply than
that predicated by the Drude theory. This difference is attributed to the fact
that the interband transitions of the bound electrons excited by the photons
Figure 1.6 Permittivity of metal gold estimated by the Drude model (dotted line) and the
measurement (solid line). The measured data come from Ref. [10]. The parameter values in
the Drude model are adopted from Table 1.2.
Optical Properties of Plasmonic Materials 21
with higher energy have not been taken into account in the Drude model,
where, however, the interband transitions become significant.
The contribution from the interband transitions of bound electrons to the
dielectric permittivity looks like the corresponding resonance in dielectric
media. These transitions are obtained from the governing equations without
consideration of the free electrons as
v21
εðvÞ ¼ 1 þ , (1.55)
v20 v2 jgv
where v0 is the oscillation frequency of a bound electron under an electric
field, and v1 and g are related to the corresponding frequency and damping of
a bound electron, respectively.
Figure 1.7 shows the dielectric permittivity of metal gold considering
only the interband contribution of bound electrons. The real part shows a
dispersion behavior while the imaginary part displays a clear resonant behavior.
The vertical dotted line represents the resonant frequency v0 that is consistent
with the zero value of the second part on the right side of Eq. (1.52). The
resonant frequency v0 also agrees well with the peak value of the imaginary
part, so it is sometimes called the absorption frequency. For the real parts, there
is a peak value in the lower frequency and a valley value in the higher frequency
that is greater than the resonant frequency. The frequency region between the
two extreme values is called the abnormal dispersion region, where the
permittivity values decrease with the increasing frequency.
For a Lorentz oscillator, the real part of permittivity is a positive value in
the range less than the resonance frequency, while there is a negative valley at
the slightly higher frequency than the resonance frequency. Another feature in
Fig. 1.7 is that the dielectric permittivity has a nonzero asymptotic value when
Figure 1.7 Contribution of the bound electrons in metal gold to the permittivity. The values
of the parameters used are v1 ¼ 4.55 1015 s1, g ¼ 9.10 1016 s1. The central
resonance wavelength is 450 nm (v0 ¼ 4.19 1017 s1).
22 Chapter 1
the wavelength increases much longer than the resonance wavelength. After
all, if the interband electron transitions within media have been taken into
account, such an asymptotic constant should approach one summary value.11
If the higher order of interband electrons transition is taken into account,
the dielectric permittivity is expressed as the model superposition of the
Lorentz model as
X
K
f i v2i
εðvÞ ¼ 1 þ , (1.56)
i v2i v2 jgi v
Table 1.2 Electric properties of noble metals and bulk plasma frequencies.
Plasma Frequency Plasma Frequency Damping Constant
Metal (eV) (1015 s–1) (1015 s–1)
Au 9.1 13.8 0.011
Ag 9.2 14.0 0.032
Al 15.1 22.9 0.92
Cu 8.8 13.4 0.14
(a) (b)
Dielectric function
ε2 ε2
0 ε1
ε1 0
ωp ωp ωr
Figure 1.8 Components of the complex dielectric functions for (a) metals described using
the Drude model and (b) dielectrics using the Lorentz equations.
Optical Properties of Plasmonic Materials 23
v2p X
K
f i v2i
εðvÞ ¼ 1 þ , (1.57)
v2 þ jvgd i¼1 v2i v2 jvgi
v2p v2p g
εðvÞ ¼ ε` þ j : (1.58)
v2 þ g2 vðv2 þ g2 Þ
For example, the interband offset for gold is about ε` ¼ 1.2 from Eq. (1.58) by
setting v ! 0.
Optimal values of typical metals for the Drude–Lorentz model are listed
in Table 1.3. These include noble metal materials (Au, Ag, and Cu),
aluminum, and transition metals (Cr, Ni, Pd, Pt, and Ti). These metals have
been frequently used in optoelectronic devices. The measurements of these
optical data are strongly dependent on the experimental conditions (in terms
of geometry morphology and the fabrication process), so the relevant optical
data must be carefully chosen.12
A correction may be made to the model data shown in Fig. 1.5 obtained
by the Drude–Lorentz model. Taking into account the contribution from both
free electrons and a single interband transition, the dielectric function of gold
progresses toward the measured results. Only one interband transition has
24 Chapter 1
Table 1.3 Parameter values for the Drude–Lorentz model. Adapted from Ref. 12 with
permission from the American Physical Society, © 1998.
Parameters Ag Au Cu Al Be Cr Ni Pt Ti W
f0 0.845 0.760 0.575 0.523 0.084 0.168 0.096 0.333 0.148 0.206
gd 0.048 0.053 0.030 0.047 0.035 0.047 0.048 0.080 0.082 0.064
f1 0.065 0.024 0.061 0.227 0.031 0.151 0.100 0.191 0.899 0.054
g1 (eV) 3.886 0.241 0.378 0.333 1.664 3.175 4.511 0.517 2.276 0.530
v1 (eV) 0.816 0.415 0.291 0.162 0.100 0.121 0.174 0.780 0.777 1.004
f2 0.124 0.010 0.104 0.050 0.140 0.150 0.135 0.659 0.393 0.166
g2 0.452 0.345 1.056 0.312 3.395 1.305 1.334 1.838 2.518 1.281
v2 4.481 0.830 2.957 1.544 1.032 0.543 0.582 1.314 1.545 1.917
f3 0.011 0.071 0.723 0.166 0.530 1.149 0.106 0.547 0.187 0.706
g3 0.065 0.870 3.213 1.351 4.454 2.676 2.178 3.668 1.663 3.332
v3 8.185 2.969 5.300 1.808 3.183 1.970 1.597 3.141 2.509 2.580
f4 0.840 0.601 0.638 0.030 0.130 0.825 0.729 3.576 0.001 2.590
g4 0.916 2.494 4.305 3.382 1.802 1.335 6.292 8.517 1.762 5.836
v4 9.083 4.304 11.18 3.473 4.604 8.775 6.089 9.249 19.43 7.498
f5 5.646 4.384 — — — — — — — —
Г5 2.419 2.214 — — — — — — — —
v5 20.29 13.32 — — — — — — — —
been considered in Fig. 1.9, so the mode curves still fail to reproduce the
experimental results at wavelengths below 500 nm. For a practical metal,
there are usually multiple interband transitions in the short-wavelength
spectrum because the higher energy of photons promotes interband transitions
of bound electrons.
Figure 1.9 Permittivity of gold estimated by the Drude–Lorentz model (dotted line)
considering the contribution from both the free electron and a single interband transition. The
experimental values (solid line) from Ref. [10] are compared. The model parameters for Au
are taken from Table 1.3. Note the different scales for the real part in the lower half and the
imaginary part in the upper half.
Optical Properties of Plasmonic Materials 25
Figure 1.10 (a) Real and (b) imaginary parts of the permittivity for metal aluminum at
wavelengths ranging from 200 nm to 1200 nm. The first, second, and third oscillators of
interband transitions are integrated into the effect. The experimental values are taken from
Ref. [10], and the parameter values for the model are from Table 1.3.
When more interband transitions are integrated into Eq. (1.57), the
calculated date would progress closely to the experimental data. For
example, the dielectric functions of aluminum are shown in Fig. 1.10. Both
the real part and the imaginary part of the permittivity approach the
experimental results when the higher-energy interband transitions are
integrated into the effect. Some rapid variations occur around 800 nm,
which corresponds to the lowest-lying interband transition of the bound
electrons in the outer shells.
The dielectric permittivity of metal silver and the complex refractive index
are plotted in Fig. 1.11. The contribution of both free and bound electrons is
considered. The measured data are consistent with the results estimated by the
Drude–Lorentz model. At optical frequencies, the imaginary part of the
dielectric permittivity is positive, whereas the real part has negative values at
wavelengths longer than the resonant wavelength. The dielectric function of
Im(ε) ,, Re(|ε|) holds true. However, the imaginary part of the refractive
index is incrementally greater than the corresponding real part, which is
slightly greater than zero.
26 Chapter 1
Figure 1.11 (a) Permittivity and (b) corresponding refractive index of metal silver. The
Drude–Lorentz model in (a) considers the contribution of both free and bound electrons. The
experimental data come from Ref. [10].
Figure 1.12 (a) Real part and (b) imaginary part of the dielectric functions for metals Au,
Cu, Ni, and Cr at wavelengths ranging from 200 nm to 900 nm. Fitting lines are shown for
clarity. The experimental data come from Ref. [10].
v2pm
mr ðvÞ ¼ 1 þ , (1.59)
v2om v2 igm v
where vpm, vom, and gm are the plasma frequency, resonant frequency, and the
damping coefficient in the magnetization of a magnetic dipole. For multiple
orders of magnetic resonance, Eq. (1.59) is extended into a multiple form as
X
M v2pmk
mr ðvÞ ¼ 1 þ : (1.60)
k¼1
v2omk v2 igmk v
Figure 1.13 Dielectric functions for semiconductor materials of silicon and germanium.
The experimental data comes from Ref. [10].
Optical Properties of Plasmonic Materials 29
Table 1.4 Bandgap and optical properties of frequently used dielectric materials.
Material InAs Ge Si InP GaAs TiO2 ZnO SiO2 Al2O3
Bandgap (eV) 0.35 0.66 1.12 1.34 1.424 3.0 3.44 8.5 8.9
l (mm) 3.50 1.87 1.10 0.92 0.87 0.41 0.36 0.15 0.14
Carrier density (cm–3) 1015 2 1013 1010 1.3 107 2.1 106 — — — —
ðv
2v h ε1 ðṽÞ
ε2 ðvÞ ¼ P d ṽ, (1.61b)
p c vl ṽ2 v2
where Pc is the Cauchy principal value. For the refractive index, the real part
and its imaginary part are related with the Kramers–Kronig relations18
30 Chapter 1
ðv
2 h ṽ n2 ðṽÞ vn2 ðvÞ
n1 ðvÞ ¼ 1 þ Pc d ṽ, (1.62a)
p vl ṽ2 v2
ðv
2v h n1 ðṽÞ n1 ðvÞ
n2 ðvÞ ¼ P d ṽ: (1.62b)
p c vl ṽ2 v2
The relation between the reflectivity R(v) and its phase difference Dw(v) are
also defined by the Kramers–Kronig form
ðf
2 h ṽ DðṽÞ v DwðvÞ
ln RðvÞ ¼ Pc d ṽ, (1.63a)
p fl ṽ2 v2
ðv
2v h ln RðṽÞ ln RðvÞ
DwðvÞ ¼ P d ṽ: (1.63b)
p c vl ṽ2 v2
ðn0 n1 Þ2 þ n22
R¼ , (1.64a)
ðn0 þ n1 Þ2 þ n22
2n0 n2
Dw ¼ arctan , (1.64b)
ðn20 n21 n22 Þ
where n0 refers to the refractive index of the incident medium, and n1 þ jn2 is
the complex refractive index of the substrate material. In general, at the
boundary of free space and a nonmagnetic substrate material, the reflection
coefficient is presented through Snell’s law:3
n1 1 þ jn2
R ejvu ¼ , (1.65)
n1 þ 1 þ jn2
where u refers to a series of incident angles that depend on the wave frequency.
The optical functions are then related to the physical observables by
Optical Properties of Plasmonic Materials 31
1 R2
n1 ¼ , (1.66a)
1 þ R2 2R cos u
2R sin u
n2 ¼ : (1.66b)
1 þ R2 2R cos u
In Eqs. (1.64)–(1.66), optical parameters are the function of optical frequency.
The optical refractive index can be determined in Eq. (1.66) by the reflectivity
over a frequency region.
The reflection coefficient and the corresponding phase difference can be
calculated from the frequency-dependent reflectivity. In order to calculate
the complex refractive index n1 þ jn2, two independent measurements could
be made to relate the complex dielectric function to physical observables.
One convenient method measures the reflectivity in different conditions, for
example, different incident angles of light with the same wavelength or
various wavelengths with the same incident angles. Both the real and
imaginary parts of a complex dielectric function can be related with the
Kramers–Kronig relations.
EO
Ed
EI
A
P B
Figure 1.15 Lorentz cavity model to calculate the local field of the central particle.
EL ¼ E þ Ed þ EI þ Ep : (1.67)
It is notable that the term Ep denotes the field induced by other particles
dispersed throughout the whole system. When the particles are randomly
distributed, such as the conditions of particles dispersing in a liquid or gas
environment, the term Ep would vanish due to the counterbalance of the
symmetric lattice. In the familiar cubic crystal lattice, the Ep term is also zero
due to the lattice symmetry.
The depolarized local field Ed relates the macroscopic polarization P by the
relation Ep ¼ –P/ε0, where ε0 is the permittivity of the free space in the
aforementioned cavity. It is assumed that the particles are immersed in the host
medium of vacuum. Both E and Ed are field values averaged within the entire
composite medium. For the electric fields inside the particle, the total charge is
integrated over its normal surface. Because of the structural lattice symmetry,
the total field EI, ascribed to all of surface charges accumulated on the particles,
is parallel with the external field. The total polarization field is estimated with a
magnitude of11,24
EI ¼ P=3ε0 : (1.68)
Therefore, the local field upon the particle is
3ε0 εr 1
a¼ , (1.71b)
f εr þ 2
where ε ¼ ε0εr is the permittivity of the individual particle. This notion is called
the Clausius–Mossotti theory, which calculates the microscopic polarizability of
an individual particle from the observable dielectric function of bulk materials.
It provides a distinct connection between the electric response of a microscopic
particle and the macroscopic behavior of a bulk material.
If the particle polarizability volume a0 ¼ 4pε0a(m3) is used in Eq. (1.71),
the Clausius–Mossotti formula is rewritten in the form of
ε ε0 4p 0
¼ Va : (1.72)
ε þ 2ε0 3
The following section adopts the particle polarizability a, as in Eqs. (1.71).
(a) (b)
2.6 80
2.2 60
Re (ε)/εh
Im (ε)/εh
1.8 40
1.4 20
1.0 0
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
f f
Figure 1.16 Effective dielectric functions as a function of the volume fraction for composite
materials as predicted by the Maxwell–Garnett and Bruggeman formula: (a) dielectric
function of a composite for permittivity ratio εi/εh ¼ 3, and (b) imaginary part of the permittivity
for εi/εh ¼ 5 j100. Both feature predictions using the Maxwell–Garnett (solid) and the
Bruggeman (dashed) formula.
1 þ 2Gi εi εh
ε ¼ εh , Gi ¼ f i : (1.75)
1 Gi εi þ 2εh
Equation (1.75) formulates the equivalent dielectric function of a composite
medium by the individual dielectric functions of the inclusions and the host
material, i.e., the Maxwell–Garnett theory. In the MGT, the embedded
particles and the host solid material are treated in an unsymmetrical manner.
When the difference of the dielectric functions between the two composite
materials are obvious, this asymmetry becomes particularly sharp. For a
dilute composite material with the volume-filling factor less than a unit away
fi ,, 1, the equivalent dielectric function of the composite materials appears
in the function expansion of the Taylor series as
3ðε1 f 1 þ ε2 f 2 Þ ðε1 þ ε2 Þ
G2 ¼ : (1.80b)
4
Optical Properties of Plasmonic Materials 37
The þ/ sign in Eq. (1.80a) is determined to ensure that the imaginary part of
the obtained dielectric function is positive.11
Based on Eq. (1.79) for an effective medium expression, it is reasonable to
get the general theory for a composite material with multiple components in a
symmetrical manner
X εi ε X
fi ¼ 0, f i ¼ 1: (1.81)
i
εi þ 2ε i
Equation (1.81) is the popular Bruggeman form of the EMT. This formula
can be expanded to any number of components in a variety of composite
materials because each component is treated equally in the mixture. For two-
phase composite materials, the Taylor series expansion of the Bruggeman
approach in the case of low-volume fractions (fi ,, 1) becomes
References
1. J. Aukong, Electromagnetic Wave Theory, John Wiley & Sons, New York
(1986).
2. P. Šolín, Partial Differential Equations and the Finite Element Method,
John Wiley & Sons, New York (2006).
3. N. Johnson et al, Electromagnetic Characterization of Complex Nano-
structured Materials: An Overview, ECONAM CSA project, http://
econam.metamorphose-vi.org/index.php?option=com_docman&task=
doc_download&gid=263&Itemid=24 (2011).
4. J. M. Lourtioz, Photonic Crystals: Towards Nanoscale Photonic Devices,
p. 122, Springer, Berlin (2005).
5. J. E. Conrad, Electromagnetic Waves and Radiating Systems, 2nd ed.,
Prentice Hall, Englewood Cliffs, NJ (1968).
6. A. vander Vorst, A. Rosen, and Y. Kotsuka, RF/Microwave Interaction
with Biological Tissues, John Wiley & Sons, New York (2006)
7. S. G. Rodrigo, F. J. García-Vidal, and L. Martín-Moreno, “Influence of
material properties on extraordinary optical transmission through holes
arrays,” Physical Review B 77(7), 075401 (2008).
8. M. S. Dresselhaus, “Solid State Physics Part II: Optical Properties of
Solids,” lecture notes, Massachusetts Institute of Technology, Cambridge,
MA (1999).
9. R. J. Van Overstraeten and R. P. Mertens, “Heavy doping effects in
silicon,” Solid State Electron. 30(11), 1077–1087 (1987).
10. E. D. Palik, Handbook of Optical Constants of Solids, Academic Press
Limited, New York (1998).
11. W. Cai and V. M. Shalaev, Optical Metamaterials: Fundamentals and
Applications, pp. 26–37, Springer, New York (2010).
12. D. Rakić, A. B. Djurišić, J. M. Elazar, and M. L. Majewski, “Optical
properties of metallic films for vertical-cavity optoelectronic devices,”
Appl. Optics 37(22), 5271–5283 (1998).
Optical Properties of Plasmonic Materials 39