You are on page 1of 31

SPECTROSCOPY

It is a branch of physics that deals with the study of the radiation absorbed, reflected, emitted, or
scattered by a substance.

Although the term radiation only deals with photons (electromagnetic radiation) but spectroscopy also
involves the interactions of other types of particles, such as neutrons, electrons, and protons, which are
used to investigate matter.

From a historical viewpoint, spectroscopy arose in the 17th century after a famous experiment carried
out by Isaac Newton and published in 1672. In this experiment, Newton observed that sunlight
contained all the colors of the rainbow, with wavelengths that ranged over the entire visible spectrum
(from about 390 nm to 780 nm). He actually labeled this rainbow a ‘spectrum.’ At the beginning of the
19th century, the spectral range provided by the Newton spectrum was extended with the discovery of
new types of electromagnetic radiation that are not visible; infrared (IR) radiation by Herschel (1800), at
the long-wavelength end, and ultraviolet (UV) radiation by Ritter (1801) at the short-wavelength end.
Both spectral ranges are now of great importance in different areas, such as environmental science (UV
and IR) and communications (IR).

The development of optical spectrophotometers during the first half of the 19th century allowed
numerous spectra to be registered, such as those of flame colors and the rich line spectra that originate
from electrical discharges in atomic gases. With the later development of diffraction gratings, the
complicated spectra of molecular gases were analyzed in detail, so that several spectral sharp-line series
and new fine structure details were observed. This provided a high-quality step in optical spectroscopy.

For a long time, a large number of spectra were registered but their satisfactory explanation was still
lacking. In 1913, the Danish physicist Niels Bohr elaborated a simple theory that allowed an explanation
of the hydrogen atomic spectrum previously registered by J. J. Balmer (1885). This constituted a large
impulse for the later appearance of quantum mechanics; a fundamental step in interpreting a variety of
spectra of atoms and molecules that still lacked a satisfactory explanation.

The interpretation of optical spectra of solids is even more complicated than for atomic and molecular
systems, as it requires a previous understanding of their atomic and electronic structure. Unlike liquids
and gases, the basic units of solids (atoms or ions) are periodically arranged in long (crystals) or short
(glasses) order. This aspect confers particular characteristics to the spectroscopic techniques used to
analyze solids, and gives rise to solid state spectroscopy. This new branch of the spectroscopy has led to
the appearance of new spectroscopic techniques, which are increasing day by day.

In any case, it is worthwhile to emphasize the dominant role of optical spectroscopy in the investigation
of solids. Indeed, the optical spectroscopy of solids appears as a nice ‘window’ onto the more general
field of solid state spectroscopy.

1
THE ELECTROMAGNETIC SPECTRUM:

Every day, different types of electromagnetic radiation are invading us; from thelow-frequency radiation
generated by an AC circuit (≈50 Hz) to the highest photonenergy radiation of gamma rays (with
frequencies up to 1022Hz). These types ofradiation are classified according to theelectromagnetic
spectrum (see Figure 1).

The electromagnetic spectrum is divided into seven well-known spectral regions, which are:

1. radio waves,
2. microwaves,
3. infrared,
4. visible,
5. ultraviolet light,
6. X-rays,
7. Γ-rays.

All of these radiations have in common the fact thatthey propagate through the space as transverse
electromagnetic waves and at thesame speed,c=3×108ms−1, in a vacuum.

Figure 1: The electromagnetic spectrum, showing the different microscopic excitation sources and the
spectroscopies related to the different spectral regions. XRF, X-Ray Fluorescence; AEFS, Absorption Edge
Fine Structure; EXAFS, Extended X-ray Absorption Fine Structure; NMR, Nuclear Magnetic Resonance;
EPR, Electron Paramagnetic Resonance. The shaded region indicates the optical range

2
The various spectral regions of the electromagnetic spectrum differ in wavelength and frequency, which
leads to substantial differences in their generation, detection, and interaction with matter. The limits
between the different regions are fixed by convention rather than by sharp discontinuities of the
physical phenomena involved. Each type of monochromatic electromagnetic radiation is usually labeled
by its frequency, f, wavelength, λ, photon energy, E, or wavenumber, k. These magnitudes are
interrelated by the well-known quantization equation:
ℎ𝑐𝑐
𝐸𝐸 = ℎ𝑓𝑓 = = ℎ𝑐𝑐𝑐𝑐………………………………………..(1)
𝜆𝜆

Whereh=6.62×10−34J s is Planck’s constant.

The different spectroscopic techniques operate over limited frequency ranges within the
electromagnetic spectrum, depending on the processes that are involved and on the magnitudes of the
energy changes associated with these processes.

DIFFERENT SPECTROSCOPIC TECHNIQUES:

1. Nuclear Magnetic Resonance (NMR) & Electron Paramagnetic Resonance (EPR):

In Nuclear Magnetic Resonance techniques, microwaves are used in order to induce transitions between
different nuclear spin states while in Electron Paramagnetic Resonance, microwaves are too used in
order to induce transitions between different electron spin states.

The energy separation between the different nuclear spin states or electron spin states lies in the
microwave spectral region and can be varied by applying a magnetic field. NMR transitions are excited
by frequencies of about 108 Hz, while EPR transitions are excited by frequencies of about 1010 Hz. In both
techniques, the frequency is fixed while the magnetic field is varied to find a resonant condition
between two energy levels. These techniques are of great relevance in the study of molecular structures
(NMR) and the local environments of paramagnetic dopant ions in solids (EPR).

2. Infrared absorption and Raman scattering:

Atoms in solids vibrate at frequencies of approximately 1012–1013 Hz. Thus, vibrational modes can be
excited to higher energy states by radiation in this frequency range; that is, infrared radiation. Infrared
absorption and Raman scattering are the most relevant vibrational spectroscopic techniques. Both
techniques are used to characterize the vibrational modes of molecules and solids. Thus, among other
things, vibrational techniques are very useful in identifying vibrating complexes in different materials
and in characterizing structural changes in solids.

3. Optical spectroscopy:

Electronic energy levels are separated by a wide range of energy values. Electrons located in the outer
energy levels involve transitions in a range of about 1–6 eV. These electrons are commonly called
valence electrons and can be excited with appropriate ultraviolet (UV), visible (VIS), or even near
infrared (IR) radiation in a wavelength range from about 200 nm to about 3000 nm. This wavelength

3
range is called the optical range, and it gives rise to optical spectroscopy. These electrons are
responsible for a great number of physical and chemical properties; for example, the formation of
molecules and solids.The short (UV) wavelength limit of the optical range is imposed by
instrumentalconsiderations (spectrophotometers do not usually work at wavelengths shorter thanabout
200 nm) and by the validity of the macroscopic Maxwell equations. Theseequations assume a
continuous medium; in other words, that there is a large numberof ions within a volume ofλ3. The long
(IR) wavelength limit of the optical rangeis basically imposed by experimental considerations
(spectrophotometers work up toabout 3000 nm).

4. Absorption Edge Fine Structure (AEFS) & Extended X-ray Absorption Fine Structure (EXAFS):

Inner electrons are usually excited by X-rays. Atoms give characteristic X-rayabsorption and emission
spectra, due to a variety of ionization and possible inter-shelltransitions. Two relevant refined X-ray
absorption techniques, that use synchrotronradiation, are the so-called Absorption Edge Fine Structure
(AEFS) and ExtendedX-ray Absorption Fine Structure (EXAFS). These techniques are very useful in
theinvestigation of local structures in solids. On the other hand, X-Ray Fluorescence(XRF) is an important
analytical technique.

5. Mossbauer spectroscopy:

Γ-Rays are used in Mossbauer spectroscopy. In some ways, this type of spectroscopy is similar to NMR,
as it is concerned with transitions inside atomic nuclei. Itprovides information on the oxidation state,
coordination number, and bond characterof specific radioactive ions in solids.

OPTICAL MATERIALS:

We would like to explore the optical properties of solids for the following optical materials:

i. Insulators,
ii. Semiconductors,
iii. Metals,
iv. Glasses,
v. Molecular Materials,
vi. Doped Glasses,
vii. Doped Insulators.

OPTICAL PROCESSES:

We shall now focus our attention on the optical spectroscopy of solids.

On behalf of interaction between the electromagnetic radiation and optical material, we classify the
optical process as:

a. Reflection,
b. Propagation,

4
c. Transmission.

If a solid sample is illuminated by a light beam of intensity I0, we perceive that, in general, the intensity
of this beam is attenuated after crossing the sample; that is, the intensity It of the transmitted beam is
lower than I0. The processes that contribute to this attenuation are as follows:

1. Absorption

If the beam frequency is resonant with a ground to excited state transition of the atoms in the solid, a
fraction of this intensity is generally emitted (usually at lower frequency than that of the incident beam),
giving rise to an emission of intensity Ie. The other fraction of the absorbed intensity is lost by non-
radiative processes (heat).

2. Reflection

Light reflects with an intensity IRfrom the external and internal surfaces.

3. Scattering

Light scatters with a light intensity ISspread in several directions, due to elastic (at thesame frequency as
the incident beam) or inelastic (at lower and higher frequenciesthan that of the incident beam – Raman
scattering) processes.

Figure 2: The possible emerging beams when a solid sample is illuminated with a beam of intensity I0.
The circles represent atoms or defects in the solid that are interacting with the incoming light.shows the
possible emerging beams after an incoming beam of intensity I0 reaches a solid block. These emerging
beams occur as a result of the interaction of the incoming light with atoms and/or defects in the solid:
part of the incident intensity is reflected in a backward direction as a beam of intensity IR. Emitted beams
of intensity Ie and/or scattered beams of intensity IS spread in all directions. The transmitted beam of
intensity It is also represented.

5
Optical spectroscopy (absorption, luminescence, reflection, and Raman scattering) analyzes the
frequency and intensity of these emerging beams as a function of the frequency and intensity of the
incident beam. By means of optical spectroscopy, we can understand the color of an object, as it
depends on the emission, reflection, and transmission processes of light by the object according to the
sensitivity of the human eye to the different colors. The spectral ranges (wavelength, frequency, and
photon energy ranges) corresponding to each color for an average person are given in Table 1.1.

Optical spectroscopy also provides an excellent tool with which to obtain information on the electronic
structure of absorbing/emitting centers (atoms, ions, defects, etc.), their lattice locations, and their
environments. In other words, optical spectroscopy allows us to ‘look inside’ solids by analyzing the
emerging light.

Experimental spectra are usually presented as plots of the intensity of (absorbed, emitted, reflected, or
scattered) radiation versus the photon energy (in eV), the wavelength (in nm) or the wavenumber (in
cm−1).

As optical spectroscopy studies the interaction between radiation and matter, three different
approximations can be used to account for this interaction.

classical approximation, semi-classical approximation, and quantum approximation.

classical approximation, in which the electromagnetic radiation is considered as a classical


electromagnetic wave and the solid is described as a continuous medium, characterized by its relative
dielectric constant ε or its magnetic permeability µ. The interaction will then be described by the
classical oscillator (the Lorentz oscillator).

However, a variety of spectroscopic features can only be explained by the so-called semi-classical
approximation, in which the solid material is described by its quantum response, while the propagating
radiation is still considered classically. In this case, the classical model of the oscillator must be modified
to take into account the fact that solids can only absorb or emit energy quanta according to discrete
energy levels.

Finally, in thequantum approximation the radiation is no longer treated classically (i.e., using Maxwell’s
equation), and so both radiation and matter are described by quantum methods. For most of the
features in the spectra of solids, this approach is not necessary and it will not be invoked. However, this
approximation also leads to important aspects, such as zero-point fluctuations, which are relevant in the
theory of lasers and Optical Parametic Oscillators.

6
LIGHT SOURCES:

Light has proven to be a very useful tool in measuring and scanning techniques. This is particularly true
in the field of optical spectroscopy, where the analysis of light–matter interaction phenomena provides
fundamental information about the nature of both matter and light. We will devote our attention to
describing the physical basis of different types of light sources as fundamental instruments in optical
spectroscopy.

Different types of lamps have been used in numerous applications, including spectroscopy, material
analysis, and laser pumping. We will enumerate the main characteristics of some of the most frequently
used systems.

1. Tungsten and Quartz Halogen Lamps

Tungsten lamps are used in spectrometers since they can supply a useful spectrum in the near infrared
region. In these lamps, electricity heats a coil or tungsten wire hot enough to make it glow, in a process
called incandescence. A usual temperature is 2800 K, for which we get a very bright yellow (nearly
white) color. That is the temperature of a normal light bulb filament. This filament is inside a glass
envelope that contains argon or nitrogen gas.

As a variation on these bulb lamps, we should mention halogen lamps. These also use a tungsten
filament, but it is encased inside a much smaller quartz envelope. Because the envelope is close to the
filament, it would melt if it were made from glass. The gas inside the envelope is also different. It
consists of a gas from the halogen group. These gases have a very interesting property: if the
temperature is high enough, the halogen gas will combine with tungsten atoms as they evaporate and
redeposit them on the filament. This recycling process allows the filament to last a lot longer. In
addition, it is now possible to run the filament at a higher temperature, which implies a greater light
intensity per unit of energy.

2. Spectral Lamps

Spectral lamps find a home in special laboratory applications, where they are commonly employed as
stable sources of discrete spectral lines, which correspond to the atomic spectra of specific elements
(metals, in most cases). Originally, the atomic spectra were produced either by creating an arc between
electrodes fabricated from the emitting metal within a discharge tube, or by sprinkling a powdered salt
into an ordinary gas flame. Nowadays, electrical discharge lamps have been developed with very good
performance, due to the high purity of the metals contained within the discharge tube. In the category
of spectral lamps we also include those gas discharge lamps that contain gases such as Ne, Xe, or He at a
low–medium pressure (less than 1 atmosphere). The atoms of the gas are excited to high energy levels
by an electric current flowing through the gas.

7
3. Fluorescent Lamps

Fluorescent lamps are based on low-pressure gas (mercury, in most cases) dischargelamps. The central
element in a fluorescent lamp is a sealed glass tube, as shown in Figure 3. The tube contains a small
amount of mercury and an inert gas, which arekept under very low pressure (a few hundredths of an
atmosphere). The tube alsocontains a phosphor powder, which coats the inside of the glass. The tube
has twoelectrodes, one at each end, connected to an electrical circuit. When the lamp is turnedon,
electrons from the electrodes migrate through the gas from one end of the tube tothe other. As
electrons and charged atoms move through the tube, some of them will

Figure 3:The schematic design of a fluorescent lamp

collide with the gaseous mercury atoms. These collisions excite the atoms, pumpingelectrons up to
higher energy levels. When the electrons return to their original energylevel, they release light photons
in the ultraviolet (UV) region (185 and 254 nm.This UV radiation is converted into visible radiation by
means of the fluorescenceof the phosphor powder coating. The phosphor material uses the UV
radiation as anexcitation source and produces fluorescent emission in the visible region, with a
broadspectrum to give off the white light that we can see. A good variety of combinationsof phosphors
are used (Shionoya and Yen, 1999). The principal field of applicationof fluorescent lamps is general
lighting, for which they constitute efficient devices.

4. High-Pressure Discharge Vapor Lamps

The fundamental principle behind these types of lamp is that the monochromaticresonance lines from
the gas can be considerably broadened when the gas pressure isincreased. A collision-induced additional
broadening (often called pressure broadening) affects the energy levels of the atomic gas. This effect is
due to the different shiftsof the energy levels produced by the mutual interactions of atoms at short
distances.Instead of the the typical lines of a diluted gas, the discharge through the high-pressuregas
produces emission bands. Figures 2.2(c) and 2.2(d) show the spectra of a highpressure discharge sodium
and a mercury lamp, compared to those of low-pressurelamps. Typical pressures in these lamps are
higher than 200 atmospheres.These lamps show moderately good color rendering properties, the very

8
broadspectrum being one of their main utilities. For instance, high-pressure Xe lamps ordeuterium
lamps are commonly used in optical spectroscopy techniques.

5. Solid State Lamps

Finally, we should mention the so-called solid state lamps, which are based on semiconductor
technology, namely thelight emitting diodes(LED). Here, the origin of ther adiation is the radiative
recombination of electrons from the conduction band with holes in the valence band in a p–n junction
(see Section 2.4). One of the most interesting characteristics is that, due to the facilities provided by
semiconductor technology, they can be custom-designed to produce radiation over a wide spectral
range, from the UV to the infrared region. An additional selection of colors is also made possible by
controlling specific combinations of red, blue, and green LEDs. The applications mainly include displays
and outdoor lamps. White sources that could replace tungsten lamps are available nowadays using
source color devices based on GaN and InGaN diodes. The flexibility of the arrangements allows for the
thermal radiation spectra to be artificially reproduced by nonthermal radiation sources.

Maxwell Equations

Maxwell equations describe the coupling between the electric field and magnetic field and their
interaction with the material, resulting in all electromagnetic phenomena.

In the stduy of optics, we are concerned with the following four vector quantities called electromagnetic
fields:

1. The electric field strenth E (V/m)


2. The elelctric flux density D (C/m2)
3. The magnetic field strength H (A/m)
4. The magnetic flux density B (Wb/m2)

The fundamental theory of these electromagnetic fields is based on Maxwell’s equations. The
differential form of these Maxwell’s equations are:

∇. 𝑫𝑫 = 𝜌𝜌
∇. 𝑩𝑩 = 0
𝜕𝜕𝑩𝑩
∇ × 𝑬𝑬 = −
𝜕𝜕𝜕𝜕
𝜕𝜕𝑫𝑫
∇ × 𝑯𝑯 = 𝑱𝑱 = 𝑱𝑱𝒄𝒄 +
𝜕𝜕𝜕𝜕

Where 𝑱𝑱 is the current density (A/m2) and 𝜌𝜌 denotes the electric charge density (C/m3).

9
∇. 𝑫𝑫 = 𝜌𝜌 (1)
It is the differential form of the Gauss’s Law.

Gauss's Law
The total of the electric flux out of a closed surface is equal to
the charge enclosed divided by the permittivity.

The electric flux through an area is defined as the electric field multiplied by the area
of the surface projected in a plane perpendicular to the field. Gauss's Law is a general
law applying to any closed surface. It is an important tool since it permits the
assessment of the amount of enclosed charge by mapping the field on a surface
outside the charge distribution. For geometries of sufficient symmetry, it simplifies
the calculation of the electric field.

Another way of visualizing this is to consider a probe of area A which can measure
the electric field perpendicular to that area. If it picks any closed surface and steps
over that surface, measuring the perpendicular field times its area, it will obtain a
measure of the net electric charge within the surface, no matter how that internal
charge is configured.

To explain the above differetial form let we convert it into an integral form as:

Let we integrate above eqaution with respect to a volume V bounded by the surface S

� ∇. 𝑫𝑫 𝑑𝑑𝑑𝑑 = � 𝜌𝜌 𝑑𝑑𝑑𝑑

Use divergence theorem or Gauss’s theorem that is, ∫ ∇. 𝑫𝑫 𝑑𝑑𝑑𝑑 = ∮ 𝑫𝑫. 𝒅𝒅𝒅𝒅. Therefore above eqaution can
be written as:

∮ 𝑫𝑫. 𝒅𝒅𝒅𝒅 = ∫ 𝜌𝜌 𝑑𝑑𝑑𝑑

10
This shows that the electric flux flowing out of the closed surface S equal the total charge enclosed in
the volume V.

∇. 𝑩𝑩 = 0 (2)
It is magnetic analog of Eq. (1) that is the magnetic flux flowing out of the closed surface S is equalt to
zero. To make it more physica, we convert it into an integral form as:

Integrate above Eq. (2) over a volume V bounded by the closed surface S.

� ∇. 𝑩𝑩 𝑑𝑑𝑑𝑑 = 0

Using divergence theorem or Gauss’s theorem that is ∫ ∇. 𝑩𝑩 𝑑𝑑𝑑𝑑 = ∮ 𝐵𝐵. 𝑑𝑑𝑑𝑑. therefore above equation
becomes:

� 𝐵𝐵. 𝑑𝑑𝑑𝑑 = 0

It explains that the magnetic flux flowing out of the closed surface S is equalt to zero. This is only
possible that there exist two magnetic poles inside the closed surface S. In other words we can say that
magnetic monpole does not exist.

𝜕𝜕𝑩𝑩
∇ × 𝑬𝑬 = − 𝜕𝜕𝜕𝜕 (3)

It is the differential form of the Faraday’s Law of electromagnetic induction.

Faraday's Law
Any change in the magnetic environment of a coil of wire will cause a voltage (emf)
to be "induced" in the coil. No matter how the change is produced, the voltage will be
generated. The change could be produced by changing the magnetic field strength,
moving a magnet toward or away from the coil, moving the coil into or out of the
magnetic field, rotating the coil relative to the magnet, etc.

11
Faraday's law is a fundamental relationship which comes from Maxwell's equations. It
serves as a succinct summary of the ways a voltage (or emf) may be generated by a
changing magnetic environment. The induced emf in a coil is equal to the negative of
the rate of change of magnetic flux times the number of turns in the coil. It involves
the interaction of charge with magnetic field.

12
To make it more physically transparent, we integrate Eq. (3) over an open surface S bounded by
a curve line C,

𝜕𝜕𝑩𝑩
�(∇ × 𝑬𝑬). 𝒅𝒅𝒅𝒅 = − � . 𝒅𝒅𝒅𝒅
𝜕𝜕𝜕𝜕

Using the Stoke’s law that is ∫(∇ × 𝑬𝑬). 𝒅𝒅𝒅𝒅 = ∮ 𝐄𝐄. 𝐝𝐝𝐝𝐝. Therefore above equation can be written as:

𝜕𝜕(𝑩𝑩. 𝒅𝒅𝒅𝒅) 𝜕𝜕𝜑𝜑𝑚𝑚


� 𝐄𝐄. 𝐝𝐝𝐝𝐝 == − � == − �
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕

This states that electromotive force (∮ 𝐄𝐄. 𝐝𝐝𝐝𝐝) induces in a loop is equal to the time rate of change of the
magnetic flux passing through the area of the loop. The emf is induced in a sense such that it opposes
the variation of the magnetic field, as indicated by the negative sign, known as Lenz’s law.

13
THE LORENTZ OSCILLATOR MODEL (1878)

The interaction of light with matter has been a burning issue since before the Quantum
Mechanics born both for experimentalists and theorists. Before the advent of Quantum
Mechanics, the most well-known attempt to describe the interaction of light with matter in terms
of Maxwell’s equations was carried out by a Hendrik Antoon Lorentz, a Dutch Physicist in the
late 19th century.

He proposed a model that the electron is bound to the nucleus of the atom by a force that
behaves according to the Hook’s Law and he named it as an atomic oscillator. In this model,
nucleus of an atom is treated as stationary because of its large mass as compared with electron.
The attractive force between an electron and nucleus is represented by a spring with some
particular spring constant. An applied electric field would then interact with the charge of the
electron, causing stretching or compression of the spring, which would set the electron into
oscillating motion. This is so called Lorentz Oscillator Model. Despite being a purely classical
description, the Lorentz oscillator model was adopted to Quantum Mechanics in the 1900s and
still of the considerable use today.

Assumptions:

1. The nucleus of an atom is much more massive than the electron, so it is supposed at rest.
2. The mass of an electron is 𝑚𝑚 = 9.1 × 10−31 kg. It may be reduced mass or effective mass
depending upon the case.
3. The displacement of the electron is small enough to obey the Hook’s Law.
4. Each atomic oscillator oscillates with a characteristic natural frequency, called resonance
frequency 𝜔𝜔0 .

14
Consider an electromagnetic light of frequency 𝜔𝜔 which polarized along x-axis, interacts with
an atomic oscillator whose natural frequency ω0 as shown in the figure below. The electron will
oscillate along the polarized direction of light with the same frequency 𝜔𝜔 and its amplitude of
oscillation becomes large when, 𝜔𝜔 = 𝜔𝜔0 . The following possible forces act on this electron:

a) The external force due to the oscillating electric field of the light, 𝐹𝐹𝑒𝑒𝑒𝑒𝑒𝑒 = −𝑒𝑒𝐸𝐸𝑙𝑙𝑙𝑙𝑙𝑙 ,
where 𝐸𝐸𝑙𝑙𝑙𝑙𝑙𝑙 , is the local oscillating electric field.
b) The restoring force due to the Coulomb attraction in between electron and nucleus,
𝐹𝐹𝑟𝑟𝑟𝑟𝑟𝑟 = −𝑚𝑚𝜔𝜔0 2 𝑥𝑥.
c) The viscous force due to the effect of the solid on the motion of the electrons, 𝐹𝐹𝑣𝑣 =
𝑑𝑑𝑑𝑑
−𝑚𝑚𝛾𝛾 𝑑𝑑𝑑𝑑 , where 𝛾𝛾 is the damping rate.

In this case, the general motion of a valence electron bound to a nucleus is a damped oscillator,
which is forced by the oscillating electric field of the light wave. This atomic oscillator is called
a Lorentz oscillator. The motion of such a valence electron is then described by the following
differential equation:

𝑑𝑑 2 𝑥𝑥 𝑑𝑑𝑑𝑑
𝑚𝑚 𝑑𝑑𝑑𝑑 2 + 𝑚𝑚𝛾𝛾 𝑑𝑑𝑑𝑑 + 𝑚𝑚𝜔𝜔0 2 𝑥𝑥 = −𝑒𝑒𝐸𝐸𝑙𝑙𝑙𝑙𝑙𝑙 ……………………………(1)

The above equation can be obtained by using the Newton’s second law. Where 𝑚𝑚 and e are the
electronic mass and charge respectively, and x is the electron’s position with respect to
equilibrium in the x-direction.

Assuming that time-varying electric this field is

𝐸𝐸𝑙𝑙𝑙𝑙𝑙𝑙 = 𝐸𝐸0 𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖 ……………………………………………(2)

Where 𝐸𝐸0 is the amplitude of the light. This electric field will drive oscillations n an electron at
its own frequency ω. Therefore, the time-varying displacement of the electron from its
equilibrium position is given by

𝑥𝑥(𝑡𝑡) = 𝑥𝑥0 𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖 ……………………………………………(3)

Where, 𝑥𝑥0 is the amplitude and keep in mind both 𝐸𝐸0 and 𝑥𝑥0 are complex quantities. Therefore,
Eq. (1) becomes

15
−𝑚𝑚𝜔𝜔2 𝑥𝑥0 𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖 − 𝑖𝑖𝑚𝑚𝛾𝛾𝛾𝛾𝑥𝑥0 𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖 + 𝑚𝑚𝜔𝜔0 2 𝑥𝑥0 𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖 = −𝑒𝑒𝐸𝐸0 𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖

Or

−𝑚𝑚𝜔𝜔2 𝑥𝑥0 − 𝑖𝑖𝑚𝑚𝛾𝛾𝛾𝛾𝑥𝑥0 + 𝑚𝑚𝜔𝜔0 2 𝑥𝑥0 = −𝑒𝑒𝐸𝐸0

−𝑒𝑒𝐸𝐸0 /𝑚𝑚
⇒ 𝑥𝑥0 =
𝜔𝜔0 − 𝜔𝜔 2 − 𝑖𝑖𝑖𝑖𝑖𝑖
2

Or

𝑒𝑒𝐸𝐸0
𝑥𝑥0 = 2 2 ………………………………(4)
𝑚𝑚 �𝜔𝜔 −𝜔𝜔 0 +𝑖𝑖𝑖𝑖𝑖𝑖 �

It is the amplitude of the oscillator that is the maximum displacement from its equilibrium
position and it is a complex number because of there is a phase shift between incoming light and
the oscillator. The time-varying displace of the oscillator in Eq. (3) is then given by

𝑒𝑒𝐸𝐸0
𝑥𝑥(𝑡𝑡) = 2 𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖
𝑚𝑚�𝜔𝜔 2 − 𝜔𝜔0 + 𝑖𝑖𝑖𝑖𝑖𝑖�

𝑒𝑒
= 2 𝐸𝐸𝑙𝑙𝑙𝑙𝑙𝑙 …………………………………(5)
𝑚𝑚 �𝜔𝜔 2 −𝜔𝜔 0 +𝑖𝑖𝑖𝑖𝑖𝑖 �

The time-varying displacement in the oscillator is because of the oscillating electric field 𝐸𝐸𝑙𝑙𝑙𝑙𝑙𝑙
that induces varying dipole moment in the oscillator and is given by:

𝑝𝑝(𝑡𝑡) = −𝑒𝑒 𝑥𝑥(𝑡𝑡)

𝑒𝑒 2
= 𝑚𝑚 (𝜔𝜔 2 −𝜔𝜔 2 −𝑖𝑖𝑖𝑖𝑖𝑖 )
𝐸𝐸𝑙𝑙𝑙𝑙𝑙𝑙 …………………… (6)
0

If we have N number of such oscillators per unit volume, then the microscopic resonant
polarization that is total dipole moment per unit volume is given by

16
𝑃𝑃𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟 = 𝑁𝑁𝑁𝑁(𝑡𝑡)

𝑁𝑁𝑒𝑒 2
= 𝑚𝑚 (𝜔𝜔 2 −𝜔𝜔 2 −𝑖𝑖𝑖𝑖𝑖𝑖 )
𝐸𝐸𝑙𝑙𝑙𝑙𝑙𝑙 ………………… (7)
0

Eq. (7) shows that resonant polarization will be maximum, when 𝜔𝜔 = 𝜔𝜔0 that is when the
incident light frequency becomes exactly equal to the natural frequency of the oscillator. Hence
we can say that the response of the material media to the light is strong close to the natural
frequency of the oscillator.

The electric displacement D of the media is defined as

𝐷𝐷 =∈0 𝐸𝐸 + 𝑃𝑃

=∈0 𝐸𝐸 + 𝑃𝑃𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟 + 𝑃𝑃𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏𝑏

𝑁𝑁𝑒𝑒 2
=∈0 𝐸𝐸 + 𝑚𝑚 (𝜔𝜔 2 −𝜔𝜔 2 −𝑖𝑖𝑖𝑖𝑖𝑖 )
𝐸𝐸 +∈0 𝜒𝜒𝑒𝑒 𝐸𝐸
0

𝑁𝑁𝑒𝑒 2
=∈0 𝐸𝐸 �1 + 𝑚𝑚 (𝜔𝜔 2 −𝜔𝜔 2 −𝑖𝑖𝑖𝑖𝑖𝑖 )
+ 𝜒𝜒𝑒𝑒 �…………..(8)
0

For isotropic medium, we have

𝐷𝐷 =∈0 ∈𝑟𝑟 𝐸𝐸……………………………..(9)

By comparing Eq. (8) and Eq. (9), we get

𝜔𝜔 𝑝𝑝 2
∈𝑟𝑟 (𝜔𝜔) = 1 + 𝜒𝜒𝑒𝑒 + (𝜔𝜔 2 −𝜔𝜔 2 −𝑖𝑖𝑖𝑖𝑖𝑖 )
……………………………..(10)
0

It is the frequency dependent dielectric constant of the material. It shows that permittivity not
𝑁𝑁𝑒𝑒 2
only depends upon the frequency of the electric field but also on plasma frequency, �𝑚𝑚 ∈ = 𝜔𝜔𝑝𝑝
0

and damping as well. In this equation, 𝜒𝜒𝑒𝑒 is the electric susceptibility of a material which
represents the response of the material to polarize in the presence of an external field. The
plasma frequency is a frequency at which the collective oscillations of electrons are taking into
consideration. Note that the plasma frequency is proportional to the electron density, and can be
calculated for any material for which such density is known. The plasma frequency for non-

17
plasma materials stands for the natural collective oscillation frequency of the sea of electrons in
the material, not of individual dipoles.

In the ionosphere, 𝑁𝑁 = 1012 electrons/volume, so 𝜔𝜔𝑝𝑝 = 5.64 × 107 rad./sec and 𝑓𝑓𝑝𝑝 = 9𝑀𝑀𝑀𝑀𝑀𝑀.

This natural resonance of plasma has some interesting effects. For example, if one tries to
propagate a radio wave through the ionosphere, one finds that it can penetrate only if its
frequency is higher than the plasma frequency. Otherwise the signal is reflected back. We must
use high frequencies if we wish to communicate with a satellite in the space. On the other hand,
if we wish to communicate with a radio station beyond the horizon, we must use frequencies
lower than the plasma frequency, so that the signal will be reflected back to the earth.

When, 𝜔𝜔 = 0, Eq. (I0) reduces to

𝜔𝜔 𝑝𝑝 2
∈𝑟𝑟 (0) = 1 + 𝜒𝜒𝑒𝑒 + 𝜔𝜔 2
≡∈𝑠𝑠𝑠𝑠 …………………………………(11)
0

It represents the dielectric response to static electric field.

When, = ∞ , Eq. (I0) reduces to

∈𝑟𝑟 (∞) = 1 + 𝜒𝜒𝑒𝑒 ≡∈∞ …………………………… (12)

It represents the dielectric response to high electric field.

From Eq. (11) and Eq. (12), we get

𝜔𝜔 𝑝𝑝 2
∈𝑠𝑠𝑠𝑠 −∈∞ = 𝜔𝜔 2 ………………………………………(13)
0

Therefore, Eq. (10) in terms of low and high frequency limits can be written as

(∈ −∈ )𝜔𝜔 2
∈𝑟𝑟 (𝜔𝜔) =∈∞ + (𝜔𝜔 𝑠𝑠𝑠𝑠2 −𝜔𝜔∞2 −𝑖𝑖𝑖𝑖𝑖𝑖
0
)
…………………………(14)
0

18
It is another form of the Eq. (10). A medium whose permittivity depends on the frequency of the
wave is called dispersive. Dispersion occurs in a dispersive media e.g. a phenomenon exhibited
in a prism and raindrop that causes white light to be spread out into a rainbow of colors. That is
the white light is the mixture of many different colors-all travelling at the same speed, but having
different frequencies and wavelengths.

The dielectric constant is a complex quantity, so we can break it into real and imaginary parts as:

Rationalizing Eq. (14), we get

(∈ −∈ )𝜔𝜔 2 �𝜔𝜔 2 −𝜔𝜔 2 +𝑖𝑖𝑖𝑖𝑖𝑖 �


∈𝑟𝑟 (𝜔𝜔) =∈∞ + (𝜔𝜔 𝑠𝑠𝑠𝑠2 −𝜔𝜔∞2 −𝑖𝑖𝑖𝑖𝑖𝑖
0
)
× (𝜔𝜔 0 2 −𝜔𝜔 2 +𝑖𝑖𝑖𝑖𝑖𝑖 )
0 0

(∈𝑠𝑠𝑠𝑠 −∈∞ )�𝜔𝜔 0 2 −𝜔𝜔 2 +𝑖𝑖𝑖𝑖𝑖𝑖 �𝜔𝜔 0 2


=∈∞ + (𝜔𝜔 0 2 −𝜔𝜔 2 )2 +(𝛾𝛾𝛾𝛾 )2

(∈𝑠𝑠𝑠𝑠 −∈∞ )�𝜔𝜔 0 2 −𝜔𝜔 2 �𝜔𝜔 0 2 (∈ −∈ )(𝛾𝛾𝛾𝛾 )𝜔𝜔 2


=∈∞ + (𝜔𝜔 0 2 −𝜔𝜔 2 )2 −(𝛾𝛾𝛾𝛾 )2
+ 𝑖𝑖 (𝜔𝜔𝑠𝑠𝑠𝑠2 −𝜔𝜔∞2 )2 +(𝛾𝛾𝛾𝛾0)2
0

∈𝑟𝑟 (𝜔𝜔) =∈1 + 𝑖𝑖 ∈2

Finally, We get the real and imaginary parts of the dielectric constant as:

(∈𝑠𝑠𝑠𝑠 −∈∞ )�𝜔𝜔 0 2 −𝜔𝜔 2 �𝜔𝜔 0 2


∈1 (𝜔𝜔) =∈∞ + (𝜔𝜔 0 2 −𝜔𝜔 2 )2 +(𝛾𝛾𝛾𝛾 )2
(∈𝑠𝑠𝑠𝑠 −∈∞ )(𝛾𝛾𝛾𝛾 )𝜔𝜔 0 2
� ……………………………(16)
∈2 (𝜔𝜔) = (𝜔𝜔 0 2 −𝜔𝜔 2 )2 +(𝛾𝛾𝛾𝛾 )2

Eq. (15) and Eq. (16) represent real and imaginary parts of the dielectric constant of a material.
These depend on frequency of incoming light, damping constant and natural frequency of the
material used. In other words, we can say that dielectric constant of a material depends upon the
frequency of light used and the nature of the material.

The frequency dependent real and imaginary parts of the refractive index and the extinction
coefficient are calculated simply by using Eq. (16) in the following mathematical relations:

19
1
1
𝑛𝑛(𝜔𝜔) = [∈1 + √∈1 +∈2 ]2
√2
1 � …………………………(17)
1
𝜅𝜅(𝜔𝜔) = [−∈1 + √∈1 +∈2 ]2
√2

The absorption coefficient 𝛼𝛼(𝜔𝜔) and the reflection coefficients 𝑅𝑅(𝜔𝜔) can then be calculated by
using the following relations:

4𝜋𝜋𝜋𝜋(𝜔𝜔)
𝛼𝛼(𝜔𝜔) =
𝜆𝜆(𝜔𝜔)
2𝜔𝜔𝜔𝜔 (𝜔𝜔 )
= ………………………………..(18)
𝑐𝑐

(𝑛𝑛−1)2 +𝜅𝜅 2
𝑅𝑅(𝜔𝜔) = (𝑛𝑛+1)2 +𝜅𝜅 2 …………………………….(19)

Ionic Crystal:

The crystals formed by ionic bond between cation and anion like NaCl, NaBr etc. So the main
absorption line occurs in them in the infrared (IR) spectral region. Let us we describe a particular
ionic crystal for which, 𝝎𝝎𝟎𝟎 = 𝟏𝟏𝟏𝟏𝟏𝟏𝟏𝟏 𝑯𝑯𝑯𝑯 = 𝟏𝟏𝟏𝟏𝟏𝟏 𝑻𝑻𝑻𝑻𝑻𝑻, 𝜸𝜸 = 𝟓𝟓 × 𝟏𝟏𝟏𝟏𝟏𝟏𝟏𝟏 𝑯𝑯𝑯𝑯 = 𝟓𝟓 𝑻𝑻𝑻𝑻𝑻𝑻, 𝜺𝜺𝒔𝒔𝒔𝒔 = 𝟏𝟏𝟏𝟏. 𝟏𝟏
and 𝜺𝜺∞ = 𝟏𝟏𝟏𝟏.

Let we use 0-140 THz a range of incident light which belongs to infrared spectrum of light
(300GHz-430THz). This light is absorbed by phonon in the lattice and hence creating the ionic
polarization that is, it is utilized in displacing positive and negative ions of the ionic crystal.

MATLAB-CODE:

1. Graph between ∈1 (𝜔𝜔) and 𝜔𝜔.

omega=0:0.2:140; % range of incident light

omega_0=100; % in the units of THz

gamma=5; % in the units of THz

eps_st=12.1;

eps_inf=10;

20
eps_1=eps_inf+(eps_st-eps_inf).*((omega_0).^2-
(omega).^2).*(omega_0).^2./(((omega_0).^2-(omega).^2).^2+(gamma.*omega).^2);

plot(omega,eps_1)

xlabel('\omega (THz)')

ylabel('\epsilon_1(\omega)')

35

30

25

20

15
ε1(ω)

10

-5

-10

-15
0 20 40 60 80 100 120 140
ω (THz)

21
2. Graph between ∈2 (𝜔𝜔) and 𝜔𝜔.
omega=0:0.2:140; % range of incident light
omega_0=100; % in the units of THz
gamma=5; % in the units of THz
eps_st=12.1;
eps_inf=10;
eps_2=(eps_st-eps_inf).*(omega_0).^2.*(gamma.*omega)./(((omega_0).^2-
(omega).^2).^2+(gamma.*omega).^2);
plot(omega,eps_2)
xlabel('\omega (THz)')
ylabel('\epsilon_2(\omega)')

45

40

35

30

25
ε2(ω)

20

15

10

0
0 20 40 60 80 100 120 140
ω (THz)

22
3. Graph between (𝜔𝜔) and 𝜔𝜔.
omega=0:0.2:140; % range of incident light
omega_0=100; % in the units of THz
gamma=5; % in the units of THz
eps_st=12.1;
eps_inf=10;
eps_1=eps_inf+(eps_st-eps_inf).*((omega_0).^2-
(omega).^2).*(omega_0).^2./(((omega_0).^2-(omega).^2).^2+(gamma.*omega).^2);
eps_2=(eps_st-eps_inf).*(omega_0).^2.*(gamma.*omega)./(((omega_0).^2-
(omega).^2).^2+(gamma.*omega).^2);
n=1./sqrt(2).*sqrt(eps_1+sqrt((eps_1).^2+(eps_2).^2));
plot(omega,n)
xlabel('\omega (THz)')
ylabel('n(\omega)')

5.5

4.5

4
n(ω)

3.5

2.5

1.5

1
0 20 40 60 80 100 120 140
ω (THz)

23
4. Graph between 𝑘𝑘(𝜔𝜔) and 𝜔𝜔.

omega=0:0.2:140; % range of incident light


omega_0=100; % in the units of THz
gamma=5; % in the units of THz
eps_st=12.1;
eps_inf=10;
eps_1=eps_inf+(eps_st-eps_inf).*((omega_0).^2-
(omega).^2).*(omega_0).^2./(((omega_0).^2-(omega).^2).^2+(gamma.*omega).^2);
eps_2=(eps_st-eps_inf).*(gamma.*omega).*(omega_0).^2./(((omega_0).^2-
(omega).^2).^2+(gamma.*omega).^2);
k=sqrt(1./2).*sqrt(-eps_1+sqrt((eps_1).^2+(eps_2).^2));
plot(omega,k)
xlabel('\omega(THz)')
ylabel('\kappa(\omega)')

4.5

3.5

3
κ(ω)

2.5

1.5

0.5

0
0 20 40 60 80 100 120 140
ω(THz)

24
5. Graph between 𝛼𝛼(𝜔𝜔) and 𝜔𝜔.

omega=0:0.2:140; % range of incident light


omega_0=100; % in the units of THz
gamma=5; % in the units of THz
eps_st=12.1;
eps_inf=10;
c=3.*10.^8;
eps_1=eps_inf+(eps_st-eps_inf).*((omega_0).^2-
(omega).^2).*(omega_0).^2./(((omega_0).^2-(omega).^2).^2+(gamma.*omega).^2);
eps_2=(eps_st-eps_inf).*(gamma.*omega).*(omega_0).^2./(((omega_0).^2-
(omega).^2).^2+(gamma.*omega).^2);
k=1./sqrt(2).*sqrt(-eps_1+sqrt((eps_1).^2+(eps_2).^2));
a=2.*omega.*k./c;
plot(omega,a)
xlabel('\omega (THz)')
ylabel('\alpha(\omega)')
-6
x 10
3.5

2.5

2
α (ω)

1.5

0.5

0
0 20 40 60 80 100 120 140
ω (THz)

25
6. Graph between 𝑅𝑅(𝜔𝜔) and 𝜔𝜔.
omega=0:0.2:140; % range of incident light
omega_0=100; % in the units of THz
gamma=5; % in the units of THz
eps_st=12.1;
eps_inf=10;
c=3.*10.^8;
eps_1=eps_inf+(eps_st-eps_inf).*((omega_0).^2-
(omega).^2).*(omega_0).^2./(((omega_0).^2-(omega).^2).^2+(gamma.*omega).^2);
eps_2=(eps_st-eps_inf).*(gamma.*omega).*(omega_0).^2./(((omega_0).^2-
(omega).^2).^2+(gamma.*omega).^2);
n=1./sqrt(2).*sqrt(eps_1+sqrt((eps_1).^2+(eps_2).^2));
k=1./sqrt(2).*sqrt(-eps_1+sqrt((eps_1).^2+(eps_2).^2));
R=((n-1).^2+k.^2)./((n+1).^2+k.^2);
plot(omega,R)
xlabel('\omega (THz)')
ylabel('R(\omega)')

0.7

0.6

0.5

0.4
R(ω)

0.3

0.2

0.1

0
0 20 40 60 80 100 120 140
ω (THz)

26
7. Graph between 𝑇𝑇(𝜔𝜔) and 𝜔𝜔.
omega=0:0.2:140; % range of incident light
omega_0=100; % in the units of THz
gamma=5; % in the units of THz
eps_st=12.1;
eps_inf=10;
c=3.*10.^8;
eps_1=eps_inf+(eps_st-eps_inf).*((omega_0).^2-
(omega).^2).*(omega_0).^2./(((omega_0).^2-(omega).^2).^2+(gamma.*omega).^2);
eps_2=(eps_st-eps_inf).*(gamma.*omega).*(omega_0).^2./(((omega_0).^2-
(omega).^2).^2+(gamma.*omega).^2);
n=1./sqrt(2).*sqrt(eps_1+sqrt((eps_1).^2+(eps_2).^2));
k=1./sqrt(2).*sqrt(-eps_1+sqrt((eps_1).^2+(eps_2).^2));
R=((n-1).^2+k.^2)./((n+1).^2+k.^2);
plot(omega,R)
xlabel('\omega (THz)')
ylabel('R(\omega)')

0.9

0.8

0.7
T(ω)

0.6

0.5

0.4

0 20 40 60 80 100 120 140


ω (THz)

27
Drude Model (1900)
It explains a number of important optical properties, such as the fact that metals are excellent
reflectors in the visible while they become transparent in the ultraviolet.

Consider a free electron of mass m that can make oscillations due to an external electric field
E(t) of an electromagnetic wave. A metal contains such free electrons called delocalized
electrons they can move from one atom to another atom and form a sea of electrons. For
simplicity, we assume that this electron is oscillating along x-axis. Therefore equation of motion
would be

𝑑𝑑 2 𝑥𝑥 𝑑𝑑𝑑𝑑
𝑚𝑚 𝑑𝑑𝑑𝑑 2 + 𝑚𝑚𝛾𝛾 𝑑𝑑𝑑𝑑 = −𝑒𝑒𝐸𝐸(𝑡𝑡)…………………(1)

The above equation can be obtained by using the Newton’s second law. Where 𝑚𝑚 and e are the
electronic mass and charge respectively, and x is the electron’s position with respect to
equilibrium in the x-direction.

Assuming that time-varying electric this field is

𝐸𝐸(𝑡𝑡) = 𝐸𝐸0 𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖 ……………………………………………(2)

Where 𝐸𝐸0 is the amplitude of the light. This electric field will drive oscillations n an electron at
its own frequency ω. Therefore, the time-varying displacement of the electron from its
equilibrium position is given by

𝑥𝑥(𝑡𝑡) = 𝑥𝑥0 𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖 ……………………………………………(3)

Where, 𝑥𝑥0 is the amplitude and keep in mind both 𝐸𝐸0 and 𝑥𝑥0 are complex quantities. Therefore,
Eq. (1) becomes

−𝑚𝑚𝜔𝜔2 𝑥𝑥0 𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖 − 𝑖𝑖𝑚𝑚𝛾𝛾𝛾𝛾𝑥𝑥0 𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖 = −𝑒𝑒𝐸𝐸0 𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖

Or

−𝑚𝑚𝜔𝜔2 𝑥𝑥0 − 𝑖𝑖𝑚𝑚𝛾𝛾𝛾𝛾𝑥𝑥0 = −𝑒𝑒𝐸𝐸0

𝑒𝑒𝐸𝐸0 /𝑚𝑚
⇒ 𝑥𝑥0 =
𝜔𝜔 2 + 𝑖𝑖𝑖𝑖𝑖𝑖

28
Or

0𝑒𝑒𝐸𝐸
𝑥𝑥0 = 𝑚𝑚 (𝜔𝜔 2 +𝑖𝑖𝑖𝑖𝑖𝑖 )
………………………………(4)

It is the amplitude of the oscillator that is the maximum displacement from its equilibrium
position and it is a complex number because of there is a phase shift between incoming light and
the oscillator. The time-varying displace of the oscillator in Eq. (3) is then given by

𝑒𝑒𝐸𝐸0
𝑥𝑥(𝑡𝑡) = 𝑒𝑒 −𝑖𝑖𝑖𝑖𝑖𝑖
𝑚𝑚(𝜔𝜔 2 + 𝑖𝑖𝑖𝑖𝑖𝑖)

𝑒𝑒
= 𝑚𝑚 (𝜔𝜔 +𝑖𝑖𝑖𝑖𝑖𝑖 ) 𝐸𝐸(𝑡𝑡)…………………………………(5)

The time-varying displacement in the oscillator is because of the oscillating electric field 𝐸𝐸𝑙𝑙𝑙𝑙𝑙𝑙
that induces varying dipole moment in the oscillator and is given by:

𝑝𝑝(𝑡𝑡) = −𝑒𝑒 𝑥𝑥(𝑡𝑡)

−𝑒𝑒 2
= 𝑚𝑚 (𝜔𝜔 2 +𝑖𝑖𝑖𝑖𝑖𝑖 ) 𝐸𝐸(𝑡𝑡)…………………… (6)

If we have N number of such oscillators per unit volume, then the microscopic resonant
polarization that is total dipole moment per unit volume is given by

𝑃𝑃𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟𝑟 = 𝑁𝑁𝑁𝑁(𝑡𝑡)

−𝑁𝑁𝑁𝑁 2
= 𝑚𝑚 (𝜔𝜔 2 +𝑖𝑖𝑖𝑖𝑖𝑖 ) 𝐸𝐸(𝑡𝑡)………………… (7)

The electric displacement D of the media is defined as

𝐷𝐷 =∈0 𝐸𝐸 + 𝑃𝑃

−𝑁𝑁𝑁𝑁 2
=∈0 𝐸𝐸(𝑡𝑡) + 𝑚𝑚 (𝜔𝜔 2 +𝑖𝑖𝑖𝑖𝑖𝑖 ) 𝐸𝐸(𝑡𝑡)

29
𝑁𝑁𝑁𝑁 2
=∈0 𝐸𝐸 �1 − 𝑚𝑚 (𝜔𝜔 2 +𝑖𝑖𝑖𝑖𝑖𝑖 )�…………..(8)

For isotropic medium, we have

𝐷𝐷 =∈0 ∈𝑟𝑟 𝐸𝐸……………………………..(9)

By comparing Eq. (8) and Eq. (9), we get

𝜔𝜔 2
𝑝𝑝
∈𝑟𝑟 (𝜔𝜔) = 1 − (𝜔𝜔 2 +𝑖𝑖𝑖𝑖𝑖𝑖 )……………………………..(10)

It is the frequency dependent dielectric constant of the material. It shows that permittivity not
𝑁𝑁𝑒𝑒 2
only depends upon the frequency of the electric field but also on plasma frequency, � = 𝜔𝜔𝑝𝑝
𝑚𝑚 ∈0

and damping as well. Note that the plasma frequency is proportional to the electron density, and
can be calculated for any material for which such density is known. The plasma frequency for
non-plasma materials stands for the natural collective oscillation frequency of the sea of
electrons in the material, not of individual dipoles.

The same result in Eq. (10) can also be deduced from the Lorentz Oscillator Model as:

The Lorentz Oscillator Model is

𝜔𝜔𝑝𝑝 2
∈𝑟𝑟 (𝜔𝜔) = 1 + 𝜒𝜒𝑒𝑒 +
(𝜔𝜔0 2 − 𝜔𝜔 2 − 𝑖𝑖𝑖𝑖𝑖𝑖)

This is mostly used for insulators where the elastic force and background polarization are
necessary i.e., 𝜔𝜔0 and 𝜒𝜒𝑒𝑒 both are non-zero for insulators. In metals the electrons are already
present in the conduction band and act as free electrons without the bondage with any specific
atom. So for metals elastic force and background polarization both will be zero. This implies that
𝜔𝜔0 = 0 and 𝜒𝜒𝑒𝑒 = 0 for metals. Put these values in the above Lorentz Oscillator Model and the
same Eq. (10) will obtain which is the Drude Model.

To find the real and imaginary parts of the Eq. (10), we proceed as:

𝜔𝜔𝑝𝑝 2 (𝜔𝜔2 − 𝑖𝑖𝑖𝑖𝑖𝑖)


∈1 + 𝑖𝑖 ∈2 = 1 − 2 ×
(𝜔𝜔 + 𝑖𝑖𝑖𝑖𝑖𝑖) (𝜔𝜔 2 − 𝑖𝑖𝑖𝑖𝑖𝑖)
𝜔𝜔𝑝𝑝 2 𝜔𝜔2 𝜔𝜔𝑝𝑝 2 𝛾𝛾𝛾𝛾
=1− 4 + 𝑖𝑖
𝜔𝜔 + 𝛾𝛾 2 𝜔𝜔 2 𝜔𝜔 4 + 𝛾𝛾 2 𝜔𝜔 2

30
𝜔𝜔𝑝𝑝 2
∈1 = 1 − 2 �
𝜔𝜔 + 𝛾𝛾 2

𝜔𝜔𝑝𝑝 2 𝛾𝛾
∈2 =
𝜔𝜔(𝜔𝜔 2 + 𝛾𝛾 2 )

Thus, the Drude model predicts that ideal metals are 100 % reflectors for frequencies up to 𝜔𝜔𝑝𝑝
and highly transparent for higher frequencies. This result is in rather good agreement with the
experimental spectra observed for several metals. In fact, the plasma frequency 𝜔𝜔𝑝𝑝 defines the
region of transparency of a metal. It is important to realize that this frequency only depends on
the density of the conduction electrons N, which is equal to the density of the metal atoms
multiplied by their valency. This allows us to determine the region of transparency of a metal
provided that N is known.

31

You might also like