You are on page 1of 403

Wave

Optics
y

z x

Natural
Polariser (unpolarised) light

Analyser

Detector

Suresh Garg
Sanjay Gupta
C.K. Ghosh
WAVE OPTICS

SURESH GARG
Professor of Physics
School of Sciences
Indira Gandhi National Open University
New Delhi

SANJAY GUPTA
Senior Lecturer in Physics
School of Sciences
Indira Gandhi National Open University
New Delhi

C.K. GHOSH
Director
National Centre for Innovations in Distance Education
Indira Gandhi National Open University
New Delhi

PHI Learning Private Limited


New Delhi-110001
2012
WAVE OPTICS
Suresh Garg, Sanjay Gupta, and C.K. Ghosh

© 2012 by PHI Learning Private Limited, New Delhi. All rights reserved. No part of this book
may be reproduced in any form, by mimeograph or any other means, without permission in
writing from the publisher.

ISBN-978-81-203-4500-3

The export rights of this book are vested solely with the publisher.

Published by Asoke K. Ghosh, PHI Learning Private Limited, M-97, Connaught Circus,
New Delhi-110001 and Printed by Baba Barkha Nath Printers, Bahadurgarh, Haryana-124507.
To
OUR PARENTS

Late Smt. Kasturi Devi and Late Sh. Bahoran Lal


—Suresh Garg

Savitri Gupta and Late Sh. R.N. Gupta


—Sanjay Gupta

Late Smt. Shefali Ghosh and Late Sh. Ajit Kumar Ghosh
—Chinmoy Ghosh
CONTENTS

Preface xi

PART I NATURE OF LIGHT


1. Nature of Light 3–42
1.1 Introduction 3
1.2 Evolution of Theories of Light 5
1.3 Electromagnetic Spectrum 9
1.3.1 Perception of Light 12
1.4 Electromagnetic Theory of Light 14
1.4.1 Propagation of Monochromatic Sinusoidal Electromagnetic Plane
Waves in Free Space 16
1.4.2 Energy Carried by Electromagnetic Waves: Poynting Vector 20
1.4.3 Energy Density and Intensity of an Electromagnetic Wave 23
1.5 Reflection and Refraction of Electromagnetic Waves 25
1.5.1 Boundary Conditions 25
1.5.2 Reflection of Electromagnetic Waves 27
1.6 Summary 40
Review Exercises 42

2. Polarisation of Light 43–83


2.1 Introduction 43
2.2 What is Polarisation? 44
2.3 States of Polarisation 48
2.3.1 Superposition of Two Linearly Polarised Waves 49

v
vi Contents

2.4 Production of Linearly Polarised Light 54


2.4.1 Polarisation by Reflection: Brewster’s Law 56
2.4.2 Polarisation by Double Refraction 59
2.4.3 Selective Absorption 67
2.4.4 Polarisation by Scattering 69
2.4.5 Wave Plates 70
2.4.6 Babinet Compensator 71
2.4.7 Optical Activity 74
2.5 Practical Applications of Polarisation 75
2.6 Summary 77
Review Exercises 78
Appendix 2A 81

PART II WAVE PHENOMENA: INTERFERENCE AND DIFFRACTION


3. Two Beam Interference by Division of Wavefront 87–111
3.1 Introduction 87
3.2 Young’s Double Slit Experiment Revisited 88
3.2.1 Interference Pattern 90
3.2.2 Displacement of Fringes 97
3.3 Production of Interference Pattern by Division of Wavefront 100
3.3.1 Fresnel’s Biprism 100
3.3.2 Fresnel’s Two-mirror Arrangement 105
3.3.3 Lloyd’s Mirror Arrangement 105
3.4 Summary 108
Review Exercises 109

4. Interference by Division of Amplitude 112–154


4.1 Introduction 112
4.2 Interference by a Plane-Parallel Thin Film 113
4.3 Newton’s Rings 125
4.4 Michelson Interferometer 133
4.4.1 Adjustment of Michelson Interferometer 136
4.4.2 Applications of Michelson Interferometer 136
4.4.3 Jamin’s Interferometer 139
4.4.4 Twyman–Green and Mach–Zehnder Interferometers 140
4.5 Multiple Beam Interferometry: Reflections from a Plane Parallel Beam 142
4.5.1 Fabry–Perot Interferometer 146
4.5.2 Width of Transmission Peaks 149
4.5.3 Sharpness of Spectral Lines: Spectral Resolution 150
4.6 Lummer–Gehrcke Interferometer 151
4.7 Summary 152
Review Exercises 153
Contents vii

5. Fresnel Diffraction 155–193


5.1 Introduction 155
5.2 Observing Diffraction 156
5.2.1 Producing Diffraction Pattern 158
5.3 Spatial Evolution of Fresnel Diffraction Pattern 159
5.4 Fresnel Construction 160
5.4.1 Half-period Zones 161
5.4.2 Fresnel Construction and Rectilinear Propagation of Light 167
5.4.3 Zone Plate 168
5.5 Fresnel Diffraction Patterns of Simple Obstacles 172
5.5.1 Diffraction by a Circular Aperture 172
5.5.2 Diffraction by a Straight Edge 176
5.6 Graphical Method: Cornu’s Spiral 180
5.6.1 Straight Edge 186
5.7 Fresnel Diffraction: A Rigorous Analysis 188
5.7.1 Circular Aperture 189
5.8 Summary 191
Review Exercises 192

6. Fraunhofer Diffraction 194–256


6.1 Introduction 194
6.2 Single Slit Diffraction Pattern 196
6.2.1 Point Source 196
6.2.2 Line Source 209
6.3 Diffraction by a Circular Aperture 210
6.4 Fraunhofer Diffraction from Two Vertical Slits 216
6.4.1 Intensity Distribution in Two Slit Diffraction Pattern 217
6.5 Fraunhofer Diffraction from N Identical Slits 224
6.5.1 Intensity Distribution in N-Slit Diffraction Pattern 225
6.5.2 Plane Diffraction Grating 231
6.5.3 Concave Reflection Grating 235
6.5.4 Echelon Gratings 239
6.6 Resolving Power of Optical Instruments 242
6.6.1 Resolving Power of an Astronomical Telescope 244
6.6.2 Resolving Power of a Microscope 247
6.6.3 Resolving Power of a Diffraction Grating 250
6.7 Summary 252
Review Exercises 255

7. Dispersion and Scattering of Light 257–294


7.1 Introduction 257
7.2 Normal Dispersion 258
7.2.1 Cauchy Equation 260
viii Contents

7.3 Anomalous Dispersion 262


7.3.1 Sellmeier Equation 263
7.4 Electromagnetic Theory of Dispersion 266
7.4.1 Undamped System 268
7.4.2 Damped System 272
7.5 Rayleigh Scattering 278
7.6 Raman Effect 284
7.6.1 Raman Spectroscopy 289
7.7 Summary 292
Review Exercises 293

PART III MODERN OPTICS


8. Lasers and Their Applications 297–338
8.1 Introduction 297
8.2 Temporal Coherence 298
8.2.1 Spectral Linewidth 301
8.3 Spatial Coherence 304
8.3.1 Visibility of Fringes 307
8.3.2 Spatial Coherence and Angular Diameter of Stars:
Michelson Stellar Interferometer 309
8.4 Spontaneous and Stimulated Emission of Radiation: Einstein’s Formulation 311
8.5 Constructing a Laser: The Prerequisites 316
8.5.1 Active Laser Medium 317
8.5.2 Pumping 320
8.5.3 Optical Resonant Cavity 324
8.6 Types of Lasers 326
8.6.1 Solid State Lasers 326
8.6.2 Helium–Neon Laser 329
8.6.3 CO2 Laser 331
8.7 Other Lasers 331
8.7.1 Semiconductor Laser 331
8.7.2 Chemical Laser 332
8.7.3 Free Electron Laser 332
8.7.4 Dye Lasers 332
8.7.5 Ion Laser 333
8.8 Applications of Lasers 333
8.9 Summary 336
Review Exercises 338
Contents ix

9. Holography 339–360
9.1 Introduction 339
9.2 Basic Principle of Holography 340
9.2.1 Transmission Hologram of a Point Source 341
9.2.2 Transmission Holograms of Extended Objects 345
9.3 Theory 346
9.3.1 Practical Requirements 350
9.4 White Light Holograms 350
9.5 Some Practical Applications of Holography 354
9.6 Summary 358
Review Exercises 360

10. Fibre Optics 361–390


10.1 Introduction 361
10.2 Optical Fibres: Working Principle 363
10.2.1 Numerical Aperture 366
10.3 Types of Optical Fibres 368
10.4 Optical Communication Through Optical Fibres 371
10.4.1 Advantages of Optical Fibre Communication 373
10.5 Attenuation and Losses in Fibres 374
10.5.1 Pulse Dispersion 376
10.6 Applications of Fibres 386
10.7 Summary 388
Review Exercises 390

INDEX 391–394
PREFACE

Light is one of the most familiar natural phenomena experienced in our lives. Our visual
contact with the world around us is facilitated by the detection of light through one of our sense
organs. We adore different hues and colours sprinkled in nature because light makes us to see.
The red colour of the rising sun and the setting sun, the green of the emerald, the blue of the
sky or the blue of the deep ocean, all involve light. Quite often, vision stimulates our thinking
and enriches our lives in many ways. Due to our ability to see, we become aware of the
existence of an infinite variety of objects around us and how they behave. It is no exaggeration
to say that light served as a catalyst in the formation of life on the earth and plays a vital role
in sustaining it on this planet.
The questions about the nature of light and how we perceive it engaged the best brains ever
since the humans moved out of the cave. The study of phenomena that arise when light
interacts with matter and the wave features of light constitute the subject of Wave Optics. Our
understanding of light has spawned several new applications and fertile channels of research
and development. For instance, the development of lasers facilitated revolutionary applications
in healthcare, industry, communication and computing, among others. New imaging technology
helped create holographic and night vision devices. Optical fibre has completely changed the
way we communicate with the world around us. For reasons such as these, optics is a vibrant
subject of research and hence a thorough grounding in its basic techniques prepares learners
to study advanced physics courses and undergo specialized training in the fields of research
and development.
A course on wave optics forms the core of the B.Sc. (honours) curriculum in physics as
well as that of the general undergraduate degree in science of every university in India. The
classroom lectures on wave optics invariably tend to become abstract, drab and uninspiring due
to lack of demonstrations. Through this book, we have made a conscious effort to bridge the
gap between the theory and its practical applications. The book assists learners in applying the
acquired theoretical knowledge to real-life applications, phenomena, and problems. To this
end, we have developed the subject matter systematically starting from simple and familiar
concepts/topics before proceeding to general cases and less familiar situations. However, no
compromise has been made to sacrifice rigorous exposition. The book is written in simple
xi
xii Preface

language, presentation is lucid and detailed mathematical steps have been worked out. The use
of access devices—expected learning outcomes, in-text questions, practice problems and
summary—should give learners adequate confidence and make learning an enjoyable experience.
The book is self-contained and should serve as an excellent reference even for those who are
engaged in self-learning.
The book can be considered to be in three parts—nature of light, wave phenomena of
interference and diffraction, and modern optics. The nature of light has been dealt with in the
first two chapters. In Chapter 1, you will navigate through an exciting and challenging journey
starting from corpuscular model to wave nature and then to wave-particle duality of light.
Thereafter, you will learn the mathematical theory concerning the nature of light worked out
by Maxwell; he established that light is a transverse electromagnetic wave made up of electric
and magnetic fields. We have also discussed reflection and refraction of electromagnetic waves,
incident on the interface separating two optically different media, in terms of reflected and
transmitted fields. In the early stages of development of wave theory, the phenomenon of
polarisation provided unflinching evidence in support of transverse nature of light. In Chapter 2,
we have discussed the orientation of electric and magnetic fields in polarised light, the different
states of polarisation, and the principles and the devices used to polarise light waves. In
particular, we have discussed the Nicol prism, wave plates, Babinet compensator, and polarimeter,
among others.
Wave phenomena such as interference, diffraction and dispersion are discussed in the next
five chapters. Interference leads to the redistribution of energy when two coherent light waves
are made to superpose, and beautiful fringes in the region of overlap are obtained either by the
division of wavefront or by the division of amplitude. The Young’s double slit experiment is
a classical example of interference by the division of wavefront and this experiment helped put
the Huygens’ wave theory on sound footing, later supported by the work of Fresnel and Lloyd.
On the other hand, interference fringes in thin films, Newton’s rings, Michelson interferometer
and Fabry–Perot interferometer are obtained by the division of amplitude. You will learn all
about these in Chapters 3 and 4.
Diffraction arises whenever a wavefront or a part of it is obstructed by an obstacle. It is
understood as mutual interference between secondary wavelets from different parts of the same
wavefront by taking phase difference into account. Diffraction is classified as Fresnel diffraction
and Fraunhofer diffraction. In Fresnel diffraction, the source or the screen or both are at a finite
distance from the diffracting object. In Chapter 5, we have discussed Fresnel construction, half
period zones, zone plate, etc. in detail and used these concepts to analyse diffraction patterns
produced by a circular aperture and a straight edge. The Cornu’s spiral helps us to gain a
clearer physical insight into the origin of the diffraction pattern. However, these methods are
approximate and a rigorous mathematical analysis for a regular-shaped obstacle is also presented
here for completeness.
In Fraunhofer diffraction, the source of light and the observation screen are effectively at
infinite distance from the diffracting aperture. A lot of good physics is involved in Fraunhofer
diffraction and in Chapter 6, we begin by discussing the diffraction pattern of a single vertical
narrow slit. These results are then extended to double and N-slits and in the limit N Æ •, we
obtain a diffraction grating, which is an extremely important optical device for spectral analysis.
Fraunhofer diffraction is of particular interest to understand the theory of optical instruments,
Preface xiii
because resolution is diffraction limited. That is, diffraction places a fundamental restriction on
optical instruments, including human eye, in respect of resolution between two nearby objects.
We have discussed Rayleigh’s criterion for resolution and obtained expressions for resolving
powers of common optical devices. This discussion will help you to appreciate why large
diameter telescopes are used to study stars and galaxies light years away in deep space and
how the electron microscope helps us to resolve and magnify microbes in medical sciences.
Chapter 7 is devoted to the discussion of dispersion and scattering of light. These phenomena
play an important role in communication of information, its quality and content. Dispersion is
intimately connected with refractive index of a material and in this chapter, you will learn
about its origin as well as the physical basis of its frequency dependence. You will learn that
refractive index changes sharply and increases with wavelength in the visible region and
beyond. This is termed anomalous dispersion. It can be understood in terms of the interplay
of absorption (damping) and restoring force (electrostatic forces between electrons and nucleus).
The blue of the sky and the red of the setting sun engaged the attention of the human mind.
Lord Rayleigh investigated this problem in detail and concluded that when monochromatic
light is scattered by a transparent substance, the scattered light has the same wavelength as the
incident light. This is known as Rayleigh scattering. Moreover, the intensity of scattered light
varies inversely as the fourth power of wavelength. However, Raman showed that when
monochromatic light is incident on a system of molecules, as in a gaseous medium such as air,
several lines appear in the scattered light on either side of, and in addition to, the incident
frequency observed by Rayleigh. The appearance of newer wavelengths is known as Raman
effect.
The portion on modern optics has three chapters and deals with the topics developed in the
second-half of the twentieth century. The development of lasers in 1960s as a highly coherent,
monochromatic, unidirectional and intense source of light has opened up immense possibilities
for new applications of tremendous technological significance in almost all branches of science
and engineering. Now we know that lasers have covered unchartered territories and led to
revolutionary developments in optical communication, healthcare, space, industry, holography,
and defence. In Chapter 8, we have discussed physics of lasers and their applications in detail.
Adequate space has been given to types of lasers, pumping mechanism, operation and
applications. In Chapter 9, we have discussed holography, which is a photographic technique
where both amplitude (intensity) and phase distributions are recorded and the pictures have
three-dimensional form. Instead of an image of the object, it carries a permanent signature of
the object in the form of an intricate interference pattern. Holography is a two-step process:
hologram recording and holographic reconstruction. We have discussed the basic concepts
involved in the holographic technique using monochromatic light from a laser and presented
analytical analysis of the process. The applications of holography in holographic interferometry,
information processing, production of optical equipment, data storage, security and other
confidential work are also discussed.
The use of internet for emailing, chatting, and looking for useful academic information are
routine activities nowadays. Have you ever thought how such communication came into the
realm of possibilities? What helped us to communicate at optical frequencies (ª1015 Hz) and
how? The first modern optical communication was based on transmission of laser beam through
open atmosphere. However, scattering and absorption of light caused severe attenuation and
xiv Preface

distortion in information and it was, therefore, considered prudent to protect the signal carrying
light beam using a guiding medium. After initial experimentation with metallic and non-
metallic waveguides, pure silica fibre has made today optical fibre communication an engineering
reality. Today, fibre optic communication is a reliable, versatile and viable proposition for
local, trunk and under the sea intercontinental applications. Finally, in Chapter 10, we have
discussed types of optical fibres, dispersion, attenuation and losses in optical fibres as well as
the advantages offered by optical fibre communication.
One of the acknowledged deficiencies in higher science education is overemphasis on rote
memorisation with little scope for nurturing independent thinking and creativity. However, the
latest researches in learning theories reinforce the concept of ‘learning by doing’. To encourage
active participation of the learner, we have structured the book in such a way that the learner
gets ample opportunities to grasp the basic concepts and develop problem solving skills. To
ensure this, solved examples have been interspersed in the text. Apart from this, nearly two
hundred graded problems—numericals and reason based questions—have been included as
in-text Practice Exercises and chapter-end Review Exercises.
Every care has been taken to produce an error-free book. Yet some omissions might have
escaped our attention. We shall be grateful to the readers for bringing such errors and omissions
to our notice. Similarly, other suggestions to improve the content and presentation of the book,
so as to make it more learner-friendly, will be welcomed by the authors.
We wish to express sincere thanks to the staff of PHI Learning for their keen interest in
this project. Part of this work was done while one of us (Suresh Garg) worked as Expert,
Commonwealth Fund for Technical Cooperation, London and Director, Centre for Learning
and Teaching, National University of Lesotho, Roma and hence would like to thank
Commonwealth Secretariat, London.

SURESH GARG
SANJAY GUPTA
C.K. GHOSH
P ART I
NATURE OF LIGHT
CH A P T E R 1
NATURE OF LIGHT

EXPECTED LEARNING OUTCOMES


After reading this chapter, the student will acquire the capability to:
• Discuss the evolution of theories of light and explain the wave–particle duality.
• List the different components of electromagnetic spectrum and their important applications.
• Derive the wave equation for electromagnetic waves and show that light waves are electromagnetic
in character.
• Derive an expression for the Poynting vector and intensity of an electromagnetic wave.
• Explain reflection and refraction of the electromagnetic waves incident on the interface separating
two optically different media.
• Solve numerical problems.

1.1 INTRODUCTION

We know that the study of wave motion is important for us because waves are present all around
us and are responsible for all phenomena associated with our communication—seeing, speaking
and hearing—with the outside world. When we speak, energy is carried by sound waves, which
enable others to hear us. This is also how music reaches us and we enjoy this creative intellect
of human mind. Sound waves find applications in SONAR (Sound Navigation and Ranging) and
prospecting for underground mineral deposits and oil. The ultrasound waves are used to obtain
the images of soft tissues in the interior of human bodies and locate the abnormalities such as
kidney and gall bladder stones, growth of fibroids in the uterus and so on.
We also know that light waves are responsible for our visual contact with the world around
us. We are able to appreciate different hues and colours sprinkled in nature because light makes
us to see. The red of the rising and setting sun, the green of the emerald and the blue of the
sky or deep ocean involve light. Quite often, vision stimulates our thinking and enriches our
lives in many ways. Due to our ability to see, we become aware of the existence of an infinite
variety of objects around us and how they behave. We all know that we do not see light;
3
4 Wave Optics

it makes us to see objects. It is no exaggeration to say that light from the sun served as a catalyst
in the formation of life on the earth and plays a vital role in sustaining it on this planet. In this
sense, light is ageless!
In this chapter, we will confine ourselves to seek answers to questions such as: What is
light? How does it travel? How is it perceived? How does it behave at an interface separating
two optically different media? And so on. These questions tested the genius of the likes of
Galileo, Newton, Huygens, Young, Fraunhoffer, Fresnel, Arago, Girimaldi, Maxwell, Michelson
and Einstein, among several others. Indian physicist CV Raman inspired several other physicists
to investigate the behaviour of light. The search for answers to these questions gave birth to a
new branch of physics: Optics finds so many and so varied applications in different branches
of medicine, technology, industry and weaponry. With the development of lasers, fibre optics,
holography, optical communication and computation, optics has emerged as a very fertile field
of research and development.
In your school physics, you have learnt about the important facts and developments made
to unfold the nature of light. From corpuscular model to wave nature and then to wave–particle
duality, this journey has been very exciting and challenging. To give a feel of how scientifically
trained minds work for further understanding of different phenomena, in Section 1.2, we will
revisit the evolution of theories of light and learn that the corpuscular model successfully
explained rectilinear propagation of light, formation of shadows, the phenomena of reflection,
refraction and (normal) dispersion, etc., and that light could propagate in vacuum, i.e. it did not
require material medium for propagation. However, it could not correctly explain the observed
fact that speed of light in material medium is less than that in free space; in fact, it predicted
the opposite. The other serious flaw of this model was its incompatibility with experimental
observations related to interference, diffraction and polarisation, which, as we now know, are
essentially wave phenomena. The idea that light propagates in a medium as wave was first
proposed by Huygens and revived by Young several years later but Maxwell put it on sound
theoretical pedestal.
When it was thought that everything is known about light, new experimental evidences were
presented which contradicted wave theory of light. Einstein proposed that rather than a wave,
light also behaved as a ‘discrete quanta’ or ‘particle’ localized in space. And the dilemma was
resolved when it was accepted that light could be treated as a particular kind of matter, possessing
both energy and momentum leading to what is now known as wave–particle duality: Light
behaves like waves when it propagates in a medium or exhibits the phenomena of interference
and diffraction. Light also behaves as particles in its interaction with matter. That is to say, in
some physical conditions, wave-like behaviour stands out and in some other situations, particle-
like behaviour dominates. And neither model singularly explains the entire range of phenomena
exhibited by light in the visible region.
In Section 1.3, we have discussed different components of electromagnetic spectrum, their
sources and applications qualitatively. The family of electromagnetic waves is quite extended;
it includes visible light, in addition to infra-red radiation, ultra-violet radiation, radio waves,
microwaves, X-rays and gamma rays. The modern communication tools such as radio, television,
telephone, mobile, fax, Internet, etc., are based on transmission and reception of electromagnetic
waves. The use of X-rays for taking images of bones to diagnose fractures and use of gamma
rays in radiation therapy for treatment of malignancy in human body is now well known.
Nature of Light 5
This is followed by perception of light. We will learn that image is formed on the retina but
information is processed/interpreted/perceived by the brain.
Section 1.4 is devoted to the basic tenets of wave model of light with particular reference
to electromagnetic (e.m.) waves. Starting from the set of four Maxwell’s equations, we derive
wave equation and show that light is indeed an electromagnetic wave. This is followed by the
calculation of the Poynting vector, which defines the direction of propagation of an electromagnetic
wave.
Reflection and refraction of light are the most fundamental optical phenomena that we learnt
earlier. The reflection of light by a silvered surface forms the basis of our daily experience while
driving a vehicle, visiting a dentist or combing the hair/getting ready for a party. Refraction
explains the working of lenses and is responsible for our visual contact with the outside world.
The formation of rainbow on a rainy day, the grand spectacles of sunrise and sunset can also
be understood in terms of refraction of light. Earlier we learnt to explain these phenomena on
the basis of Huygens’ wave model thus we will not discuss these here.
Reflection of electromagnetic waves forms the working principle of radar, which is so
important in air navigation, and reflection of radio waves by earth’s ionosphere makes signal
transmission possible, which is so crucial for information processing and long distance
communication. Total internal reflection of electromagnetic waves forms the basis of one of the
greatest technological innovations in signal transmission through optical fibres. We can now
speak to our loved ones across oceans and continents in real time as we do with next door
neighbour. Reflection and refraction of electromagnetic waves forms the subject of discussion
of Section 1.5. You will learn to derive expressions for reflected and transmitted fields when
an electromagnetic wave is incident on the interface separating two optically different media.

1.2 EVOLUTION OF THEORIES OF LIGHT

Much of what we know about light has been discovered in the past five centuries, although the
first recorded references about theories of the nature of light date back to Greek philosophers
Pythagoras, Democritus, Plato and Aristotle. The rectilinear propagation of light as well as the
law of reflection was given by Euclid (300 B.C.) in his book Catoptrics. The apparent bending
of objects partly immersed in water is referenced in Plato’s Republic. Refraction of light was
studied by Cleomedes (50 A.D.) and later by Ptolemy (130 A.D.), who tabulated fairly precise
values of the angles of incidence and refraction for several media. According to historical
records, Romans also knew about burning by magnifying glasses. Several glass and crystal
spheres were traced in Roman ruins. This suggests that some Roman artisans may have used
magnifying glasses to facilitate their very fine detailed work.
After the fall of the Roman Empire (474 A.D.), the centre of scholarship shifted to the
Arab world, where the scientific and philosophical treasures of the past were translated and
preserved. Among other scientific developments, understanding of the phenomena exhibited by
light improved through constant experimentation. Ibn-al-Haitham, also known as Alhazen
(1000 A.D.), studied reflection of light by plane, spherical and parabolic mirrors and succeeded
in putting the angles of incidence and reflection in the same plane normal to the interface.
In the 13th century, Alhazen’s work was translated in Latin and influenced the works and
6 Wave Optics

writings of later researchers. Bacon (1215–1294) initiated the idea of using lenses for correcting
vision and hinted at the possibility of combining them to form a telescope. Leonardo da Vinci
(1452–1519) described the camera obscura, which was propelled by Giovanni Battista Della
Porta (1535–1615), who discussed multiple mirrors and combinations of positive and negative
lenses in his Magia naturalis (1589).
This sequence of developments opened fertile channels for exciting accomplishments through
sustained explorations by the best researching minds. Galileo Galilei (1564–1642) is widely
credited for designing the first refracting telescope, though The Hague archives suggest that the
credit should go to a Dutch spectacle maker Hans Lippershey (1587–1619) as he reportedly
filed application for a patent on the device. The compound microscope was designed around the
same time. Minor changes in the arrangements of lenses were introduced to improve their
effectiveness. The phenomenon of total internal reflection was observed by Kepler and Snell
gave the law of refraction, which was given the formulation in terms of sines by Descartes
(1637). He viewed the light beam as pressure transmitted by an elastic medium. In essence, this
derivation assumed corpuscular model of light, which is usually attributed to Newton, probably
because of the popularity of his Opticks. Descartes theory remained undisputed until about 1657
when Pierre de Fermat (1605–1665) enunciated his principle of least time and derived Snell’s
law of refraction. He showed that the ray of light would bend towards the normal, if the velocity
of light in the second medium were lower. This contradicted corpuscular theory.
Around the middle of seventeenth century, Robert Boyle (1627–1691) and Robert Hooke
(1635–1703) observed beautiful coloured rings in thin films. (These were termed Newton’s rings
because Newton explained their formation in terms of his corpuscular model. However, his
explanation was later found to be unsatisfactory.) The phenomenon of diffraction was first
observed in 1665 by the Italian physicist Francesco Marie Grimaldi (1618–1663) in the form
of bands of light in the shadow of a rod illuminated by a small source. It was later confirmed
by Robert Hooke. He also explained the existence of Newton’s rings in terms of the ‘interaction’
between the light reflected from the front and back surfaces. He proposed the idea that light was
a rapid vibratory motion of the medium and propagated at a very high speed. He also observed
that “every pulse or vibration of the luminous body will generate a sphere”. This marked the
beginning of the wave theory.
Sir Isaac Newton (1642–1727), who regarded optics as experimental philosophy, laid greater
emphasis on direct observation than speculative hypotheses. This made him unsure about the
nature of light; initially he could not decide whether light emitted by a source consisted of a
stream of very small particles which stimulated the sense of sight when entering the eye—
corpuscular theory of light—or vibratory motion of the medium propagating at a very large
speed—wave theory of light. This became evident in his work on dispersion of light—breaking
up of light in constituent colours when light was made to undergo refraction through a triangular
glass prism. To explain his observed results, he simultaneously embraced wave as well as
corpuscular (emission) theories. He argued that the corpuscles of light associated with different
colours excited the ether—all pervading massless fluid—into characteristic vibrations. But he
became committed to corpuscular theory of light when he explained rectilinear propagation of
light, which resulted in the formation of sharp shadows. He also gave simple explanation for
the laws of reflection and refraction, formation of coloured fringes in thin films, among others.
Newton’s theory of light was therefore widely accepted by the scientific community.
Nature of Light 7
In 1678, Dutch physicist and astronomer Christian Huygens (1629–1695) put forward the
wave theory of light. He demonstrated that when two beams of light intersected, they emerged
unchanged. Huygens was also able to derive the laws of reflection and refraction. In 1669 the
Danish physicist Erasmus Bartholinus (1625–1698) observed the phenomenon of double refraction
in calcite. It was explained by Huygens, who also showed that light effectively slowed down
on entering the denser media. But this theory did not receive acceptance of most scientists, who
continued to adhere to Newton’s corpuscular theory for more than a century for several reasons.
The foremost of these were:
(i) rectilinear propagation of light, which resulted in the formation of sharp shadows, and
(ii) light (from the sun) could travel through vacuum, though the only waves known at that
time—sound and water—required medium for their propagation.
Probably Newton’s authority was too compelling and infallible!
The wave theory was reborn in 1801 by the brilliant work of Thomas Young (1773–1829).
He performed a very significant and now famous double slit interference experiment, which
could be explained only on the basis of wave model of light. In 1802, he gave an explanation
of the formation of Newton’s rings observed by Boyle. The wave theory began to receive
acceptance and more theoretical and experimental investigations were taken to explore it further.
In 1808, Malus (1775–1812) discovered polarisation of light. In 1816, French physicist Augustine
Fresnel (1788–1827) presented satisfactory explanation of the diffraction patterns resulting from
different types of obstacles, edges and apertures on the basis of wave theory. He was also able
to account for rectilinear propagation of light in homogeneous isotropic media, dispelling Newton’s
main objection. This was a big step in the right direction to understand the nature of light. In
the same year, Fresnel and Dominique Francois Jean Arago (1786–1853) performed a brilliant
experiment on interference of linearly polarised light. But they struggled to provide satisfactory
explanation of their observations because they treated light as longitudinal waves, in analogy
with sound waves in air. However, these results were explained when Young suggested that light
waves should be regarded transverse, as is a wave on a string. In 1850, Jean Foucault (1791–
1868) provided further evidence of the inadequacy of the Corpuscular theory when he showed
that light moved slower in water than in air. As a result of these and other theoretical as well
as experimental developments, the wave theory of light received general acceptance in the first-
half of the nineteenth century.
At that time, it was thought that a wave necessarily required a medium to support its
propagation. Therefore, the existence of an all pervasive substance—ether, which possessed
some rather strange properties—was postulated. Poisson (1781–1840), Cauchy (1789–1857)
and several other physicists contributed to the development of the ether theory. However,
we now know that ether does not exist—it was proved conclusively by Albert Abraham Michelson
(1852–1931) and Edward Williams Morley (1838–1923) in 1887.
The most important development concerning the theory of light was the work of Scottish
physicist James Clerk Maxwell (1831–1879). In 1873, he derived a wave equation starting from
his set of four equations, which brilliantly summarised all the then known laws of electricity and
magnetism, and predicted the existence of transverse electromagnetic waves. These details are
discussed in Section 1.3. However, it will suffice here to state that solving for the speed of
8 Wave Optics

propagation of electromagnetic waves in free space, he arrived at an expression in terms of the


electric and magnetic properties of the medium:
1
v
H 0 P0

Using the then best known values of e0 and m0, Maxwell calculated the speed of electromagnetic
waves to be 3.1074 ´ 108 m s–1. This value was amazingly close to the then most precisely
known value of 3.14858 ´ 108 m s–1 for the speed of light obtained from optical experiments
by the French scientist Armand Fizeau (1819–1896). Based on this excellent agreement and
‘faith in the rationality of nature’, Maxwell concluded that light propagated as an electromagnetic
wave.
A direct evidence for the existence of electromagnetic waves was provided by German
physicist Heinrich Hertz (1857–1894) in 1888 through a series of brilliant experiments.
He excited and detected electromagnetic waves by varying electric and magnetic fields with
time and measured their wavelength as well as frequency to obtain the value of their speed.
He also found it to be exactly equal to the speed of light and demonstrated that electromagnetic
waves exhibit reflection, refraction, interference, etc. In a way, Hertz conclusively established
electromagnetic theory of light. (In 1895, Indian scientist J.C. Bose produced electromagnetic
radio waves of wavelengths 25 mm to 5 m while working at Presidency College, Calcutta
(now Kolkata) and studied their properties.) Most of the phenomena exhibited by light were
explained by the end of 19th century and it was thought that everything had been understood
about the nature of light. But subsequent developments proved leading minds wrong!
In 1887, Hertz had made a very striking discovery—the photoelectric effect. He observed
that when a metal surface was irradiated by light of frequency above a certain critical value, it
ejected electrons. This phenomenon could not be explained on the basis of wave theory of light
because kinetic energy of an ejected electron was seen to be independent of the intensity of
light. In fact, this observation contradicted wave theory because it predicted that a more intense
beam of light should add more energy to the electron. In 1905, Albert Einstein (1879–1955),
a German-Swiss physicist, interpreted the photoelectric effect using the concept of quantum
theory developed by German theoretical physicist Max Planck (1858–1947) for blackbody
radiation. (In his theory of blackbody radiation, Planck discarded classical theories and postulated
that energy was emitted or absorbed by an individual resonator in ‘quanta’ of magnitude hn.)
Einstein proposed that rather than a wave, light also behaved as a ‘discrete quanta’ or ‘particle’
localized in space. (We may mention that the word photon was coined by Gilbert N. Lewis in
1926 to describe Einstein’s ‘discrete quanta’.)
Following Einstein, we can say that the energy of a photon is proportional to the frequency
n of the electromagnetic wave:
E = hn
where h is Planck’s constant.
Note that this equation provides a bridge, a mathematical connection, between two almost
antithetical entities—waves and particles. In Einstein’s stipulation, these concepts became almost
interchangeable and provided physical basis for the work of Danish physicist Niels Bohr
Nature of Light 9
(1885–1962) on hydrogen spectrum and American scientist Arthur Compton (1892–1962)
on scattering of X-rays from electrons as particle collisions between photons and electrons.
(This effect is known as Compton Effect.) These developments suggested that light could be
treated as a particular kind of matter, possessing both energy and momentum leading to what
is now known as wave–particle duality: Light behaves like waves when it propagates in a
medium or exhibits the phenomena of interference and diffraction. Light also behaves as particles
in its interaction with matter, as in the photoelectric effect. That is to say, in some physical
conditions, wave-like behaviour stands out and in some other situations, particle-like behaviour
dominates. And neither model singularly explains the entire range of phenomena exhibited by
light. It is instructive to mention here that wave-particle duality also holds for sub-atomic
material particles and was demonstrated in the first-half of the 20th century.

1.3 ELECTROMAGNETIC SPECTRUM

We now know that Hertz produced electromagnetic waves by an oscillating magnetic spark.
In further experiments he showed that the spectrum of wavelengths of electromagnetic waves
ranges from 10–14 m to 104 m (3 ¥ 1022 – 3 ¥ 104 Hz). This is depicted in Figure 1.1 in order
of increasing frequency or decreasing wavelength. The visible light occupies the portion between
400 nm and 700 nm (7.5 ¥ 1014 – 4.3 ¥ 1014 Hz) and is a very small part. However, this region
is very vital for creation and sustenance of life on this planet. We are able to appreciate different
hues and colours sprinkled in nature because light makes us to see and enrich life.
Now refer to Table 1.1, where we have listed the frequency and wavelength ranges for different
types of electromagnetic waves. The lowest frequency (106–109 Hz) electromagnetic waves of
practical interest are radio and television waves. These are used for broadcasting and tele-
communication using communication satellites. Do you know that India launched its own exclusive
education satellite in 2004? The electromagnetic waves in the frequency range of 109–1011 Hz

Table 1.1 Classification of electromagnetic waves


Types of waves Frequency range Wavelength Source
(Hz) range (m)

}
AM 0.53 ¥ 106–1.70 ¥ 106 570–186
Radio waves FM 88 ¥ 106–108 ¥ 106 3.40–2.80 Electronic circuits
TV 54 ¥ 106–890 ¥ 106 5.60–0.34
Microwaves 109–1011 10–1–10–3 Special vacuum tubes
Infra-red radiation 1011–1014 10–3–10–7 Hot bodies
Visible light 4 ¥ 1014–7 ¥ 1014 10–7 Sun and lamp light
Ultraviolet radiation 1014–1017 10–7–10–10 Very hot bodies, sun and
special lamps
X-rays 1017–1019 10–10–10–12 High speed electron collisions
and atomic processes
Gamma rays >1019 <10–12 Natural radioactivity, nuclear
reactions, and particle
accelerators
10 Wave Optics

Figure 1.1 Electromagnetic spectrum.

are known as microwaves. Such waves are used in communications, radar applications, inter-
station television and microwave ovens. These also find applications in the study of the origin
of the universe, opening sensor doors, guiding airplanes and viewing the surface of the planet.
Nature of Light 11
Infrared region of electromagnetic spectrum lies in the frequency range of about 1011 Hz to
4.3 ´ 1014 Hz. Infrared waves were first detected in 1800 by Sir William Herschel (1738–1822),
a renowned astronomer. The infrared radiation is emitted in a continuous spectrum by hot bodies
such as electric heaters, glowing coal and ordinary house radiators. Almost one-half of the solar
energy is emitted as infrared radiation. An ordinary light bulb emits far more infra-red radiation
than it emits light. Even human body radiates infrared radiation starting at about 3000 nm.
Infrared radiation is trapped by greenhouse gases CO2 and water vapour and through greenhouse
effect maintains the surface of earth at a suitable temperature to sustain various life-forms.
Infrared radiation of wavelength 1000 nm is used to operate remote controls. Coupled with
a cathode ray tube, infrared radiation is used to produce thermographs. This radiation has also
been put in therapeutical use. Infrared spy satellites, which look out for rocket launchings,
resource satellites to look out for crop diseases and astronomical satellites to look deep into
space have been designed. Now we also have ‘heat seeking’ missiles guided by Infrared radiation,
in addition to IR lasers and telescopes.
The visible region of light is confined to a very small part of the electromagnetic spectrum;
less than 4% of the total range. The red of the sunrise and sunset, the green of the emerald and
the blue of the sky or deep ocean involve light in the visible region, which runs from frequency
range 4 ´ 1014 Hz to about 7 ´ 1014 Hz or 700–400 nm wavelength range. The ratio of the
longest to the shortest of these wavelengths is about 2:1. As we know, radiation in this region
activates receptors in our eyes and is responsible for our visual contact with our surroundings.
When visible light falls on an object, it may absorb some and reflect other wavelengths in the
visible range. However, a white object, such as a white sheet of paper or a white piece of cloth
reflects all incident wavelengths and our eyes see those objects as white. On the other hand, a
black object, such as the letters printed on this book, absorbs all wavelengths and no light is
reflected. We can similarly understand that a colour is the result of a filtered white light
spectrum, in cases where an object is illuminated by white light.
Ultraviolet (UV) radiations extend in the frequency range 1014 Hz–1017 Hz or wavelength
range 4–400 nm and lie between visible light and X-rays. The ratio of the longest to the
shortest of these wavelengths is 100:1. These radiations were discovered by John Wilhelm Ritter
(1776–1810). Ultraviolet radiation is produced by special lamps and very hot bodies. The sun
is a huge source of UV radiation. These radiations can cause very harmful effects to various
life-forms on the earth. Fortunately, most of the UV radiation reaching the earth is absorbed by
the ozone layer in the atmosphere at an altitude of 40–50 km. (That is why we should take
measures to prevent the ozone hole from widening in the sky.) Now sunglasses and skin creams
are also available to protect us from potentially harmful effects of UV radiations.
X-rays correspond to frequency range of 2.4 ´ 1017 Hz to 5 ´ 1019 Hz or wavelength range
0.6 nm–1.2 nm. These were discovered by Wilhelm C. Roentgen (1845–1923), a German physicist,
when he noted a glow of a piece of fluorescent paper caused by a mysterious radiation originating
in a cathode. These form a very valuable tool in medical research, diagnosis, and treatment.
Electromagnetic waves with frequencies above 1019 Hz are referred to as gamma rays.
These have such a small wavelength that it is difficult to observe any wave-like property.
However, gamma rays are highly penetrating and destructive to living cells. Nowadays, gamma
knives are used in surgical operations to remove tumours. These are also used to kill
microorganisms in food processing as well as treatment of cancerous cells.
12 Wave Optics

1.3.1 Perception of Light

Perception of light involves formation of sharp images and their interpretation. In your earlier
classes, you have learnt that image of an object is formed on the retina within the eye and can be
understood in terms of physical principles. But from image formation to its perception by the brain,
the process is physiological. So we can say that vision begins in the eye but light is sensed by the
brain. That is, vision involves a mixture of physical and physiological processes. We know that
human eyes are very versatile; they possess a great degree of adaptability as well as precision and
have the capacity for extremely fast movement. That is why from reading a book or newspaper,
we can almost instantaneously shift focus to a distant star/luminous object, adapt to dim or bright
light, distinguish colours, and estimate distance, size or direction of motion. The questions such
as how vision begins in the eye or human brain interprets images have fascinated human mind for
thousands of years. We now discuss answers to these questions qualitatively.

Human vision
We all know that human eye is an image-producing device. The light enters the pupil through
the cornea, passes through aqueous humor and reaches the biconvex eye lens, which is perfectly
transparent. (The pupil is about 2 mm in size in a normal eye and the intensity of light is
regulated by the iris. The eye lens is an elastic structure of protein fibres arranged like the layers
of an onion and has focal length of about 3 cm.) Most of the light is refracted at the air-cornea
interface and the eye lens make finer adjustments in focusing. The light coming out of the eye
lens passes through vitreous humor before an inverted image is formed on the retina.
The retina consists of five types of neuronal cells; the photoreceptors, bipolar, horizontal,
amacrine and ganglion neurons. A magnified view of the arrangement of neuronal cells is shown
in Figure 1.2. The photoreceptor or photosensitive neurons are of two types: rods and cones.

Figure 1.2 Schematic representation of the arrangement of neural cells in retina.


Nature of Light 13
(This nomenclature derives its genesis from the geometrical shapes of these neurons. On an
average, six million rods and about twenty times as many cones line up the retina in a normal
eye. These are packed together like the chambers of a honey-comb. Near the centre of the retina,
in a small region called the fovea, there are a great many cones but no rods.) Rods enable vision
in dim light, discriminate between different shades of dark and light, see shapes and movements.
On the other hand, cones help in colour vision and most precise vision with their light sensitive
pigments. That is to say, rods provide high sensitivity, whereas cones provide high visual acuity.
The rods and cones of the retina are connected to optic nerve fibres through which the light
stimulates signals to the brain.
The clarity of vision depends on the sharpness of the image. It is important to mention here
that apart from refraction of light, image formation involves accommodation of eye lens,
constriction of pupil and convergence of the eye. Accommodation is the unique ability of eye
lens to automatically change its focal control while viewing near or far objects. Constriction of
the pupil means narrowing down of the diameter of the hole through which light enters the eye.
This action occurs simultaneously with accommodation of the eye lens, prevents the entry of
light through the periphery of eye lens and helps minimise blurred vision. The pupil also
constricts in bright light to protect the retina from intense stimulation. (In dim light, the pupil
dilates to allow the retina to receive more light so that we are able to see.)
It is common experience that a normal person focuses both eyes only on one set of objects.
That is, we are empowered to single binocular vision. When we look at a distant object, the
incoming light is focused at identical spots on the retinas of our eyes. If we move close to the
object, the images are formed at the same points in both retinas; our eyes automatically make
adjustment by radial movement of two eyeballs. This is referred to as convergence.

Information processing by human brain


As mentioned earlier, the image formed on the retina is inverted laterally as well as vertically.
But we do not perceive objects around us as topsy-turvy. How does this happen? The answer
to this apparent contradiction is imbedded in the inherent capacity of our brain to automatically
process information contained in the visual images and perceive them correctly. The direct
evidence in support of this statement presents itself in the form of complete and permanent
blindness to a person who suffers severe brain injury, though her/his eyes continue to function
perfectly.
When light impulses impinge on the retina (and image is formed), these are absorbed by
rods and cones, which contain four kinds of photosensitive substances. These visual pigment
molecules undergo structural (chemical) changes. Each rod cell contains about seventy million
molecules of a purple coloured photosensitive pigment called rhodopsin and cones contain
violet coloured pigment iodopsin. (Rhodopsin consists of scotopsin protein and chromophore
retinene, which is a derivative of vitamin A in the form of cis-retinene.) These photoreceptors
help to convert light signals into electrical signals in the retina. The signals generated in the
retina are transmitted to lateral geniculate nucleus and visual cortex, which are higher centres
in the visual pathway of the brain (Figure 1.3). In this way, precise information about the image
projected on the retina is transmitted accurately to the brain.
14 Wave Optics

Figure 1.3 Retinal circuit for transfer of visual information: Electrical links between the cells of retina.

Now we have a fair idea about how light is perceived by human brain. Let us now learn
Maxwell’s electromagnetic theory of light in some detail.

1.4 ELECTROMAGNETIC THEORY OF LIGHT

Now we know that wave theory of light was proposed by Christian Huygens in 1678 and he
satisfactorily explained all phenomena associated with light then—reflection, refraction and
total internal reflection. (You have learnt about these in earlier classes and thus we will not
discuss these here.) He also provided a satisfactory explanation of the then recently discovered
phenomenon of birefringence. (The wave theory also predicted that speed of light in a material
medium will be less than the speed of light in free space.) Though Newton’s corpuscular model
failed to explain some of these phenomena, his authority and eminence was so dominant that
in spite of its success, Huygens’ proposition did not get immediate recognition and wide
acceptance. In fact, the wave model was accepted only after the classical experiments of Young
and Fresnel on interference in the early part of 19th century. However, a sound pedestal to the
wave theory was provided by Maxwell when he identified light with electromagnetic waves.
You will now learn about it.
Maxwell’s field equations are known to explain all electromagnetic phenomena: Gauss’ law,
absence of magnetic monopoles, Faraday’s law of electromagnetic induction and Ampere’s law
with the concept of displacement current. While studying electricity and magnetism, you must
have realised that Maxwell’s equations are essentially based on experimental observations.
Using vector notation, we can write these in a compact form as follows:
Ñ · D=r (1.1)
Nature of Light 15
Ñ · B=0 (1.2)

³–E ˜ B = 0 (1.3)
˜t
È ˜ DØ
Ñ ´ B = P ÉJ  Ù (1.4)
Ê ˜t Ú
where Ñ is the del operator, r signifies free charge density and J denotes conduction current.
The dot denotes divergence and cross signifies curl operation on a vector. E, D, and B respectively
represent the electric field, electric displacement, and magnetic induction. These are connected
through the following constitutive relations:
D = eE (1.5)
B = mH (1.6)
and J = sE (1.7)
where e, m and s respectively denote the dielectric permittivity, magnetic permeability and
electrical conductivity of the medium.
For free space devoid of all charges and conduction currents, r = 0 = s so that J = 0. Recall
that the permittivity and magnetic permeability for free space are e0 and m0, respectively.
Using constitutive relations for free space, Eqs. (1.1)–(1.4) take the form
Ñ · E=0 (1.8)
Ñ · B=0 (1.9)

³ – E  ˜B = 0 (1.10)
˜t
˜E
and ³ – B  H 0 P0 = 0 (1.11)
˜t

These equations can be used to derive wave equation with simple vector algebra. To this end,
we take the curl of Eq. (1.10). This gives

È ˜BØ
³ – (³ – E) ³ – É Ù
Ê ˜t Ú

Since (temporal) ¶/¶t operation is independent of (spatial) curl operation, we can interchange
these. Then using Eq. (1.11), we can decouple E and B components. This leads to the relation
2
³ – (³ – E )  ˜ (³ – B )  H 0 P0 ˜ E
2
(1.12)
˜t ˜t
To simplify the LHS of Eq. (1.12), we use the vector identity Ñ ´ Ñ ´ E = Ñ(Ñ · E) – Ñ2E
and note from Eq. (1.8) that Ñ · E = 0. Then Eq. (1.12) takes a very compact form

˜ 2E
³2E H 0 P0 2
(1.13a)
˜t
16 Wave Optics

Similarly by taking curl of Eq. (1.11) and using Eq. (1.9), we can easily show that

˜ 2B
³2B H 0 P0 2
(1.13b)
˜t
These equations are identical in form to a 3-dimensional wave equation. It means that
electromagnetic waves are made of electric field vector E and magnetic field vector B. Moreover,
each of these components satisfies a wave-like equation (Ñ2y = 1/v2 ¶ 2y/¶t 2, where y is a
physical quantity which propagates with speed v). Using this analogy, Maxwell concluded that
the speed of propagation of electromagnetic waves in free space is given by
1
v (1.14)
H 0 P0

That is, the speed of electromagnetic waves in vacuum depends only on its electric and
magnetic properties through permittivity and magnetic permeability e0 and m0, respectively.
By substituting the then known values of e 0 and m 0, Maxwell obtained the value
v = 3.1074 ´ 108 m s–1, which conformed to the then known velocity of light in free space. This
result established that light is an electromagnetic wave. It was both amazing and exciting and
brought crowning glory to Maxwell; with this he brought the entire science of optics under the
umbrella of electromagnetism—a classical example of unification of knowledge.
Before proceeding further, you should answer a Practice Exercise.

Practice Exercise 1.1 Starting from Eqs. (1.3) and (1.4), show that for a charge free material
medium electric and magnetic components of an electromagnetic wave satisfy the following
differential equations:

È 2 ˜ ˜Ø
É ³  HP 2  VP ÙE 0
Ê ˜t ˜t Ú
È 2 ˜ ˜Ø
and É ³  HP 2  VP ÙB 0
Ê ˜t ˜t Ú

1.4.1 Propagation of Monochromatic Sinusoidal


Electromagnetic Plane Waves in Free Space

We now know that electromagnetic waves can be generated by time varying electric and magnetic
fields and can be represented by the amplitudes and phases of these fields. The simplest
electromagnetic waves are monochromatic sinusoidal plane waves. We know that a monochromatic
wave has only one colour (wavelength or frequency). Moreover, in a plane wave, the phases of
all points on a plane normal to the direction of propagation are the same. Recall that the phase
of a monochromatic sinusoidal electromagnetic plane wave propagating along the +z-direction
is given by (kz – wt), where k is the wave number and w is the angular frequency. Hence, the
electric and magnetic fields for such a wave can be expressed as follows:
Nature of Light 17
E = E0 exp[j(kz – wt)] (1.15)
and B = B0 exp[j(kz – wt)] (1.16)
where E0 and B0 are space and time independent vectors. For a wave propagating along
+z-direction, the field vectors E and B will be independent of x and y. Then, from Eqs. (1.8)
and (1.9), we can write
˜ Ez
=0 (1.17)
˜z
˜ Bz
and =0 (1.18)
˜z
By writing the z-components of Eqs. (1.10) and (1.11) and using the same argument, you will
find that time variation of Ez and Bz can be expressed as

˜ Ez
=0 (1.19)
˜t
˜ Bz
and =0 (1.20)
˜t
Physically, Eqs. (1.17)–(1.20) imply that components of the electric and magnetic fields along
the direction of propagation of an electromagnetic wave do not depend on time and space
coordinates. So we choose
E z = 0 = Bz (1.21)
Note that Eq. (1.21) is not a unique choice as solution of Eqs. (1.19) and (1.20). We may as
well choose Ez = constant and Bz = constant but such fields do not correspond to a plane wave.
It means that
1. A plane electromagnetic wave has no longitudinal component. That is, electromagnetic
waves are transverse. So if we choose x-axis to be along E, then Ey = 0. Moreover, since
electromagnetic waves are transverse, the magnetic field vector will be along the y-axis
so that we can write
E = x̂E0 exp[ j(kz – wt)] (1.22)
and B = ŷ B0 exp[ j(kz – w t)] (1.23)
k 1
with B0 = E0 E0 (1.24)
Z c
Note that x̂ and ŷ are unit vectors along x-axis and y-axis, respectively.
Before proceeding further, you may like to solve a Practice Exercise.

Practice Exercise 1.2 Prove Eq. (1.24).


Hint: Express Eqs. (1.10) and (1.11) in component form and use Eq. (1.21).

2. On solving Practice Exercise 1.2, you will note that k/w is a real number. It implies that
electric and magnetic field vectors will be in phase. So if E vanishes at some instant of
18 Wave Optics

time, then B will necessarily be zero. Similarly, if E attains maximum value, B will also
be maximum. It means that electric and magnetic waves co-exist, not in isolation.
The electric and magnetic fields reinforce mutually; an electric field varying in time
gives rise to magnetic field changing in space and time. This, in turn, generates an
electric field varying in space and time. This mutual generation of electric and magnetic
fields leads to propagation of an electromagnetic wave.
Refer to Figure 1.4. It shows a plane electromagnetic wave propagating along the positive
z-direction. What can we say about the orientation of electric and magnetic fields? These are
oriented at right angles to one another as well as the direction of propagation of the wave
suggesting that electromagnetic waves are transverse in nature.

Figure 1.4 Propagation of a plane electromagnetic wave and associated electric and magnetic fields.

Before proceeding further, go through the following examples.

EXAMPLE 1.1 A monochromatic sinusoidal electromagnetic wave propagating in 3-D charge


free space can be expressed as
E(r, t) = E0 exp[ j(k · r – wt)] (i)
and B(r, t) = B0 exp[ j(k · r – wt)] (ii)
If E0(k, w) = E0 zˆ and B0(k, w) = B0 xˆ , express equivalent forms of Maxwell’s equations in terms
of wave number k.
Solution From Eq. (1.8), we note that for charge-free space
Ñ·E = 0
Using (i), we can express the divergence, curl and time derivative of E as
Ñ · E = jk · E
Ñ ´ E = jk ´ E
˜E
and = –jwE
˜t
That is, the operators Ñ and ¶/¶t are equivalent to jk and –jw, respectively. Hence, for charge-
free space, we can write Maxwell’s equations as
Nature of Light 19
k · E0 = 0
k ´ E0 = wB0
k · B0 = 0
ZE
and k ´ B0 =  0
c2
EXAMPLE 1.2 The electric field associated with a light wave propagating in vacuum is
given by
E 50 xˆ exp[  j ( az  1015 t )] Vm 1
Determine the direction of propagation, wave number, frequency and magnetic field component
of the wave.
Solution To determine the direction of propagation of the wave, we compare the argument
of the given exponential with Eq. (1.22) and note that the wave propagates along positive
z-direction.
To calculate the wave number, we use the relation

c
Z Z
k a

a
Z 10 s 15 1
3.3 – 106 m 1
Þ
c (3 – 108 m s 1 )
The frequency of the wave is given by

Q Z 1015 s 1
1.67 – 1014 Hz
2S 2S
The magnetic field associated with the light wave is given by Eq. (1.23) with

k E0
B0 E0
Z c
so that
50
B yˆ exp[  j ( az  1015 t )] Wb m 2
c

Now you should solve a Practice Exercise.

Practice Exercise 1.3 The electric field associated with a plane electromagnetic wave in free
space is given by
ˆ
Ë È (2y zˆ ) ¹ r
E 500 xˆ exp Ì j É  Z t ØÙÚ ÛÜ Vm1
Í Ê 100 Ý

Determine the associated magnetic field and calculate the frequency of the wave.
20 Wave Optics

1.4.2 Energy Carried by Electromagnetic Waves: Poynting Vector

We know that the most important characteristic of a wave is that it transports energy. Is it true
for electromagnetic waves as well? To discover the answer to this question, let us consider
electric and magnetic field vectors and calculate the divergence of their cross product.
Using the vector identity Ñ · (P ´ Q) = Q · (Ñ ´ P) – P · (Ñ ´ Q), we can write
Ñ · (E ´ B) = B · (Ñ ´ E) – E · (Ñ ´ B) (1.25)
On substituting the values of (Ñ ´ E) and (Ñ ´ B) from Eqs. (1.10) and (1.11), respectively,
we get
˜B ˜
Ñ · (E ´ B) = –B · – E · m0e0 (1.26)
˜t ˜t
The terms containing time-derivatives of B and E on the right hand side of this equation can
be expressed as
˜B 1 ˜
B¹ (B ¹ B )
˜t 2 ˜t
˜E ˜E 1 ˜
and E ¹ P0H 0 P0H 0 E ¹ H 0 P 0 ( E ¹ E)
˜t ˜t 2 ˜t
Using these results in Eq. (1.26), we get

1 ˜ 1 ˜
Ñ · (E ´ B) =  (B ¹ B )  H 0 P 0 (E ¹ E )
2 ˜t 2 ˜t
˜ Ë1 Û
=  Ì (B ¹ B + H 0 P0 E ¹ E) Ü (1.27)
˜t Í 2 Ý

Do you recognise this result? This equation is analogous to the equation of continuity in
hydrostatics. To discover its physical significance, we first integrate it over volume V bound by
surface S. This gives
˜ 1È 1 Ø
Ô ³ · (E – B) dV  P0
˜t Ô 2 ÉÊ P 0
B ¹ B  H 0 E ¹ EÙ dV
Ú
(1.28)
V V

To simplify this equation, we now use Gauss’ divergence theorem, which states that the surface
integral of a vector function taken over a closed surface bounding a given volume equals the
volume integral of the divergence of the same function over the same volume. Mathematically,
we write

vÔ M ¹ dA Ô ³¹ M dV
S V
Then Eq. (1.28) simplifies to
1 ˜ 1È 1 Ø
P0 vÔ (E – B) ¹ dA 
˜t Ô 2 ÉÊ P 0
B ¹ B  H 0 E ¹ EÙ dV
Ú
S V
Nature of Light 21
The integrand on the right-hand side of this equation denotes the time rate of flow of
electromagnetic energy associated with a plane wave in free space and electric and magnetic
field vectors contribute to it equally. It means that the quantities (1/2 m0) B · B and (1/2)e0 E · E
respectively represent the magnetic and electrical energies per unit volume and their sum gives
energy carried by the e.m. wave per unit volume:

1È 1 Ø
u= B ¹ B  H 0 E ¹ EÙ (1.29)
2 ÉÊ P0 Ú
The vector
1
S= (E – B ) (1.30)
P0
defines the Poynting Vector and we may interpret S · dA as the electromagnetic energy crossing
area dA per unit time. From Eq. (1.30), it is obvious that electric, magnetic and Poynting vectors
are mutually perpendicular. Physically it means that Poynting vector points in the direction of
propagation of an electromagnetic wave. This is illustrated in Figure 1.5.

Figure 1.5 The Poynting vector is in the direction of propagation of an electromagnetic wave.

Let us now calculate the energy carried by an electromagnetic wave per unit area. To do so,
we substitute for E and B from Eqs. (1.22) and (1.23) respectively in Eq. (1.30). This gives
k
S = ( xˆ – yˆ ) E02 cos 2 ( kz  Z t )
P0Z
k
= zˆ E02 cos 2 ( kz  Z t )
P0Z
where ẑ denotes unit vector along the z-axis. It means that the flow of energy is in the
z-direction, which is the direction of propagation of the wave. The amount of energy carried by
an electromagnetic wave across a unit area per unit time is given by
k
S E02 cos 2 ( kz  Z t )
P0Z
Since the cos2 term varies very rapidly with time for an optical beam (w = 1015 s–1), a detector
will record only an average value. You will recall from your elementary knowledge of trigonometry
22 Wave Optics

that over one cycle, the average value of cos2q = 1/2. Hence, the time-averaged energy carried
by an electromagnetic wave in the z-direction is given by

kE02
ÃS Ó (1.31)
2 P0Z
Before proceeding further, you should go through the following example carefully.

EXAMPLE 1.3 The sun radiates nearly 4 ´ 1026 J on an average per second. Calculate the
Poynting vector at its surface. How much of this energy reaches the earth? Take radius of sun,
assumed to be spherical, as 7 ´ 108 m and the distance between the sun and the earth as
1.5 ´ 1011 m.
Solution The energy radiated by the sun can be expressed as a product of Poynting vector and
area of sun’s surface:
U = áSñ ´ 4pR2
u 4 – 1026 J s 1
so that áSñ =
4S R 2 4 – 3.14 – (7 – 108 m) 2
= 6.5 ´ 107 J m–2 s–1
To calculate the amount of energy that reaches the surface of the earth, we have to calculate
the average value of Poynting vector on earth’s surface. Theoretically, we should determine the
distance between the centre of the sun and surface of the earth. But sun–earth separation (RSU)
is several orders of magnitude greater than the radius of the sun (R). If we assume that there
is no energy loss, we can write
áSEñ ´ 4pRSU2 = áSñ ´ 4pR2

2
È R Ø
Hence áS E ñ = ÉR Ù
ÃS Ó
Ê SU Ú
2
È 7 – 108 Ø
= É 11 Ù
– (6.5 – 10
7
J m 2 s 1 )
Ê 1.5 – 10 Ú
= 1.4 ´ 103 J m –2 s–1
This energy is responsible for sustaining all life-forms on this planet, changing weather patterns,
and giving beauty to nature.
Now you may like to answer a Practice Exercise.

Practice Exercise 1.4 Show that for a material medium containing free charges, Eqs. (1.29)
and (1.31) respectively can be modified as
1 1 1 2 1 2
U= B¹B  D¹E B  HE
2P 2 2P 2
kE02
and áSñ =
2PZ
Nature of Light 23

1.4.3 Energy Density and Intensity of an Electromagnetic Wave

While solving Practice Exercise 1.4, you established that energy density associated with a plane
wave propagating in a material medium characterised by dielectric permittivity e and magnetic
permeability m is given by
1 1 1 2 1 2
U B¹B  D¹E B  HE (1.32)
2P 2 2P 2
If we choose
Ex = E0 cos(kz – wt), Ey = 0 = Ez
and Bx = 0, By = B0 cos(kz – wt), Bz = 0
Equation (1.32) takes the form
1 2 2 1 2
U H E0 cos (kz  Z t )  B0 cos 2 (kz  Z t )
2 2P

For a dielectric medium, we can modify Eq. (1.24) to write B0 HP E0 . Then the expression
for energy density simplifies to
1 2 2 1 2 2
U= H E0 cos (kz  Z t )  H E0 cos (kz  Z t )
2 2
2 2
= H E0 cos ( kz  Z t )
Recall that the time-average of cos2 term is one-half. Hence, we can write
1 2
ÃU Ó H E0 (1.33)
2
If we multiply this expression for average energy density by the speed of propagation of the
wave, we will get energy crossing a unit area per unit time or power per unit area. This, by
definition, is intensity of the wave. Hence

1 2 1 H 2
I H vE0 E0 (1.34)
2 2 P
The intensity of an electromagnetic (light) wave propagating in free space is given by
1 1
I 2
H 0 cE0 (8.854 – 1012 C2 N 1m 2 ) – (3 – 108 ms 1 ) E02 1.33 – 103 E02 W V 2
2 2
Let us consider a 1 kW community radio transmitter, which emits electromagnetic waves uniformly
in all directions and calculate the amplitude of electric field at a distance of one kilometre. To
do so, we first compute the intensity of electromagnetic waves at a distance of 1000 m:
1000 W
I 7.95 – 105 W m 2
4S (1000 m) 2
24 Wave Optics

Hence, the electric field associated with the electromagnetic waves at a distance of 1 km is
given by
1/ 2
È 7.95 – 105 Wm 2 Ø
E0 É 3 Ù 2.44 – 10 1 V m 1
Ê 1.33 – 10 WV 2 Ú
We now advise you to repeat these calculations for a simpler situation.

Practice Exercise 1.5 In physics laboratory, we use a sodium lamp which gives out light
uniformly. Suppose that the lamp uses 60 W. Calculate the magnitude of electric field.
[Ans. 29.95 V m–1]

Before considering reflection and refraction of electromagnetic waves, go through the


following solved example carefully.

EXAMPLE 1.4 A laser beam is directed upward to make a tiny blackened glass sphere of
mass 10–9 kg and diameter 20 mm afloat in air. Calculate the intensity of the laser beam.
Solution Refer to the figure below. The blackened sphere of diameter D is kept afloat by a
laser beam. This can happen when the weight of the sphere acting vertically downward is
balanced by a force created by the impact of the laser beam.
The effective area of impact, A, is the equatorial circle of the
sphere of radius R = D/2. Therefore
S
A D2
4
Let the time of impact be Dt. If the intensity of the beam is I, the
energy transferred during Dt is IADt. The corresponding momentum
transfer
Figure Ex. 1.4
IA't
'p
c
where c is velocity of light.
If we assume the collision to be elastic, the change in momentum = 'p  ( 'p) 2 'p.

2 'p 2 IA't 2 IA
\ Force exerted =
't 'tc c
For equilibrium
2IA
mg
c
cmg cmg 2cmg
\ I 2
2A SD S D2

4
Nature of Light 25
On substituting the given values, we get

2 – (3 – 108 m s 1 ) – (10 9 kg) – (9.8 m s 2 )


I 4.68 – 109 J s 1m 2
3.14 – (20 – 10 6 m) 2

1.5 REFLECTION AND REFRACTION OF


ELECTROMAGNETIC WAVES

In the previous section, we learnt that light is an electromagnetic wave, which consists of
mutually reinforcing electric and magnetic fields (E and B) as these vary continuously in space
and time. Now you may logically ask the question: What happens to these fields when an
electromagnetic wave is incident (normally or obliquely) on an interface separating two optically
different media such as air and glass or air and water? To discover answers to these questions,
we need to know the boundary (continuity) conditions that these waves (or fields) must
satisfy. Moreover, we stipulate that Maxwell’s equations hold at the interface separating the two
media.

1.5.1 Boundary Conditions

Refer to Figure 1.6. It shows a thin cylinder, which encloses an area DS of the interface and
extends on either side of the boundary. If we consider Eq. (1.1) in the absence of free charges
and integrate it over the cylindrical volume, then using Gauss’ divergence theorem, we can write

H
Ô ³ ¹ E dV 0 H1
vÔ E ¹ da  H vÔ E ¹ da  vÔ D ¹ da
S1
2
S2 S3
(1.35)

Here S1 and S2 represent the flat faces of the cylinder and S3 denotes the curved surface of the
cylinder. If we let the cylinder shrink, it will take the shape of a ‘pill’ (medicine tablet), whose
thickness is extremely small in comparison to the flat surfaces and is known as ‘Gaussian pill
box’. In the limit its thickness goes to zero, the third integral will vanish. Then Eq. (1.35)
reduces to

H1

S1
E ¹ da =  H 2 vÔ
S2
E ¹ da
or e1E1 · n̂1DS = –e2 E1 · n̂ 2 DS
or e1E1n = e2E2n (1.36)
where n̂1 and n̂ 2 are unit normal vectors, as shown in Figure 1.6.
From Eq. (1.36) we note that in the absence of free charges, the normal component of
electric displacement is continuous across the interface.
Starting from Eq. (1.2) and following the same steps, we can show that the normal component
of magnetic field B is also continuous across the interface:
B1n = B2n (1.37)
26 Wave Optics

Note that only normal components of D and B are equal on both sides of the interface.
Their total magnitudes may not be equal and their directions need not be the same.

Figure 1.6 A cylinder at the interface of two optically different media.

Proceeding further, we consider Eq. (1.3) by referring to Figure 1.7, which shows a thin
Amperian (rectangular) loop ABCD of thickness L and length l. If we take surface integral over
a surface bounding this loop, we can write
˜B
Ô ( ³ – E ) ¹ da  Ô
S S ˜t
¹ da 0

Figure 1.7 An Amperian loop at the interface of two optically different media.

Using Stokes’ theorem, we can rewrite this equation as


˜B
vÔ E ¹ dl =  Ô
l S ˜t
¹ da

˜ ˜I
or Ô AB
 Ô
CD
 Ô
BC
 Ô
DA
E ¹ dl = 
˜t SÔ B ¹ da
˜t
where f denotes magnetic flux. If we let L ® 0 the integrals along BC and DA will tend to
vanish. Moreover, as width of the loop tends to zero, the magnetic flux also vanishes.

Stokes theorem states that the surface integral of the curl of a vector function taken
over a surface is equal to the closed line integral of the vector taken over the line enclosing
the surface.
Nature of Light 27
Thus, we get

ÔE ¹ dl  Ô E ¹ dl = 0
AB CD
or (E1 · p̂ )l + [E2 · (– p̂ )]l = 0
where p̂ is the unit vector in the plane of incidence along the line separating two optically
different media. Hence, we can write
E1t = E2t (1.38)
Here E1t and E2t denote the tangential components of electric field and these are continuous
across the interface.
Next by considering Eq. (1.4) and following the same steps, we will find that tangential
components of the magnetic field are also equal and continuous. Mathematically, we can write

1 1
B1t B2t (1.39)
P1 P2
To sum up, in the absence of any surface current and surface charges, when an electromagnetic
wave reaches the interface, the normal components of B and D and the tangential components
of B and E are continuous across it. We now use these boundary conditions to explain reflection
and refraction of electromagnetic waves.

1.5.2 Reflection of Electromagnetic Waves

Consider a plane electromagnetic wave that is incident on a boundary between two linear or
isotropic media for which Eqs. (1.5) and (1.6) hold and electrical permittivities (e1, e2) as well
as magnetic permeabilities (m1, m2) are independent of position and direction. (We may visualize
this as propagation of light from one medium to another such as air to glass, oil to water and
the like.) For simplicity, we assume that there are no free charges or conduction currents in
either medium. From your school physics, you may recall that a wave incident on such an interface
is partially reflected and partially transmitted (refracted). We intend to derive expressions for
(i) reflection and transmission coefficients or the fields associated with the reflected and
the refracted waves in terms of the field associated with the incident wave and
(ii) fraction of energy that is reflected and/or transmitted.
For this, we will use the boundary conditions discussed in Section 1.5.1. Let us first assume
that the waves are incident normally on the interface.

(a) Normal incidence


Refer to Figure 1.8. It shows that a plane sinusoidal electromagnetic wave of frequency w
and moving along +x-axis is incident on the yz-plane (x = 0). Let us suppose that the electric
field vector is along the y-axis. Then we can write the electric and magnetic fields associated
with the incident wave as
Ei(x, t) = E0i yˆ exp[  j (Z t  ki x )] (1.40)
28 Wave Optics

E0i
and Bi(x, t) = zˆ exp[  j (Z t  ki x )] (1.41)
v1

Figure 1.8 A sinusoidal plane electromagnetic wave incident normally at the interface separating two
optically different media.

Here j2 = –1, ŷ is the unit vector along y-axis, ẑ is the unit vector along z-axis and v1 is
the velocity of the wave in the first medium.
It is a common experience that a wave incident on an interface gives rise to reflected and
transmitted waves. In fact, the existence of both reflected and transmitted waves is necessary
to satisfy the boundary conditions. The reflected wave returns to the first medium and the fields
associated with it are respectively given by

Er(x, t) = E0 r yˆ exp[  j (Z t  ki x)] (1.42)


E0 r
and Br(x, t) =  zˆ exp[  j (Z t  ki x)] (1.43)
v1
Note that

• The electric fields associated with the incident and reflected waves lie in the same
direction but the magnetic field vectors are oppositely directed.
• The sign of the x-term in the exponent in Eq. (1.43) has changed vis-a-vis Eq. (1.41).
This is because the direction of propagation of the reflected wave is along –x-direction.
• A negative sign has been put with the amplitude of the reflected wave. This has genesis
in the transverse nature of the electromagnetic waves and the fact that electric and
magnetic field vectors must satisfy the relation

1 ˆ
Br (k i – E r ) (1.44)
v1

so that Poynting vector will be in the direction of wave propagation.


Nature of Light 29
The electric and magnetic fields associated with the transmitted wave are respectively
given by
Et(x, t) = E0t yˆ exp[  j (Z t  kt x )] (1.45)
and
1 ˆ
Bt(x, t) = (k i – Et ) (1.46)
v2
Note that the electric and magnetic fields defined by Eqs. (1.40)–(1.46) satisfy the boundary
conditions given by Eqs. (1.36)–(1.39) at each point of the interface at all times. Accordingly,
the tangential and/or normal components of these fields to the left of x = 0 will be equal to the
fields on its right.
To analyse the reflection and refraction phenomena, it is customary to express the relations
between the field vectors of reflected and transmitted waves in terms of the field vectors of the
incident wave. Since we are considering normal incidence, there are no normal field components
because neither E nor B is along the x-axis. It means that we should match only tangential
components of the electric and magnetic fields at the plane x = 0. Therefore, we can write
E0i + E0r = E0t (1.47)
and
1 1
( B0i  B0 r ) = B0t
P1 P2
or
1 È E0i E0 r Ø E0 t
 = (1.48)
P1 ÉÊ v1 v1 ÚÙ
P 2 v2
We can rewrite it as
E0i – E0r = xE0t (1.49)
where
P1v1
[ (1.50)
P2 v2
On substituting the value of E0t from Eq. (1.47) in Eq. (1.49) and rearranging the terms, we can
express E0r in terms of E0i:
È1  [ Ø
E0 r É Ù E0i (1.51)
Ê1  [ Ú

On using this result in Eq. (1.47), a little simplification leads to the required expression connecting
E0t and E0i:
È 2 Ø
E0t É Ù E0i
(1.52)
Ê1  [ Ú

For most substances, the permeability is nearly equal to the permeability for vacuum. That is,
to a very good approximation, we can take m1 = m2 = m0. Then, Eq. (1.50) simplifies to
x = v1/v2 and Eqs. (1.51) and (1.52) respectively take the form
30 Wave Optics

È v2  v1 Ø
E0r = Év  v Ù E0i (1.53)
Ê 2 1Ú
and
È 2v2 Ø
E0t = Év  v Ù E0i (1.54)
Ê 2 1Ú

These results suggest that for v2 > v1, the reflected wave will be in phase with the incident wave,
whereas for v2 < v1, the reflected wave will be out of phase with the incident wave. This is
shown in Figure 1.9.
We would now like you to answer a simple Practice Exercise.

Practice Exercise 1.6 The refractive index of a medium


is defined as n = c/v, where c is the velocity of light in
vacuum. Express Eqs. (1.53) and (1.54) in terms of
refractive indices of the two media.

On solving Practice Exercise 1.6, you will see that


n1  n2
E0r = E0i (1.55)
n1  n2
2n1
and E0t = E0i (1.56)
n1  n2
Figure 1.9 (a) In phase and (b) out
Do not confuse n1 and n2 used here with unit vectors. of phase reflected and
incident waves.
These results show that the transmitted and incident
waves are always in phase. However, the reflected and incident waves are in phase only when
n1 > n2 and out of phase by 180º when an electromagnetic wave travels from a rarer medium
to denser medium (n1 < n2). Physically, it means that the reflected wave suffers a phase change
of p when a wave is reflected from a denser medium (air-glass or air-water interface). However,
if a wave travels from a denser medium to a rarer medium, no such phase change takes place.
We can now easily derive expressions for the reflection and transmission coefficients, which
respectively measure the fraction of incident energy that is reflected or transmitted. That is,
the reflection coefficient is defined as the ratio of reflected intensity and incident intensity:
Ir
R (1.57)
Ii
Similarly, the transmission coefficient is defined as the ratio of transmitted intensity and incident
intensity:
It
T (1.58)
Ii
Here Ir, It and Ii respectively denote the intensities of reflected, transmitted and incident waves.
Nature of Light 31
To obtain expressions for reflection and transmission coefficients, we recall that intensity is
defined as the average power per unit area. Mathematically, it is given by I (1/2) H vE02 . Using
this definition in Eq. (1.57) and combining the resultant expression with Eq. (1.55) are obtain
the expression for reflection coefficient:
2
(1/ 2) H v1 E02r È n1  n2 Ø
R (1.59)
Én  n Ù
(1/ 2) H v1 E02i Ê 1 2Ú

This result shows that when n1 = n2, R = 0 implying that no reflection will occur, if the second
medium has the same refractive index as the first medium. (It essentially implies continuity of
optical media.) Thus, if we immerse a transparent solid in a liquid of the same refractive index,
the solid will not be visible!
Similarly, from Eqs. (1.58) and (1.56) we obtain the expression of transmission coefficient:
2
(1/ 2) H 2 v2 E02t v1 E02t n2 È 2n1 Ø (1.60)
T
(1/ 2) H1v1 E02i v2 E02i n1 ÉÊ n1  n2 ÙÚ

since me = n2. Equations (1.59) and (1.60) constitute Fresnel’s amplitude relations for normal
incidence.
From Fresnel’s amplitude relations, you can easily verify that the sum of reflection and
transmission coefficients is equal to one:
2
È n1  n2 Ø 4n1n2
R T Én n Ù

2
1 (1.61)
Ê 1 2Ú (n1  n2 )

Suppose that an air-glass interface is characterised by n1 = 1 and n2 = 1.5. We can easily


compute the reflection and transmission coefficients as R = 0.04 and T = 0.96. It means that
when a plane sinusoidal electromagnetic wave falls on air-glass interface normally, most of the
light will be transmitted, provided there is no loss (due to absorption, scattering, etc.) at the
boundary separating the two media.
Before we discuss the case of oblique incidence, you should solve a Practice Exercise.

Practice Exercise 1.7 A plane electromagnetic wave is incident normally on a medium.


If the reflection amplitude coefficient and transmission amplitude coefficient of these media
are equal to 0.5, calculate the ratio of their refractive indices. [Ans. 5.8]

(b) Oblique incidence


Refer to Figure 1.10. As before, we assume that the y–z plane defines the boundary between
two optically different media. A plane monochromatic sinusoidal wave of frequency w is incident
obliquely on the interface. Its direction of propagation is given by the wave vector ki. Therefore,
it can be represented as
Ei = E0i exp[–j(wit – ki · r)] (1.62)
32 Wave Optics

Figure 1.10 A plane electromagnetic wave obliquely incident


on the interface of two optically different media.

and
1
Bi = E exp[–j(wit – ki · r)] (1.63)
v1 0i

where ki · r = kix x  kiy y  kiz z . Note that the directions of Ei and Bi are normal to ki and the
electric field associated with the incident wave is oriented arbitrarily perpendicular to the
direction of propagation.
We resolve the electric field in two components:
(i) a component in the plane of incidence, and
(ii) a component perpendicular to the plane of incidence.

The plane of incidence is defined as the plane containing the incident wave and the normal
to the surface. In Figure 1.10, y–x plane defines the plane of incidence.

The incident field vector is thus resolved in two components:


(i) the component in the y–x plane, called the parallel component, will be designated as
Eip , and
(ii) the component in the plane perpendicular to the plane of incidence, called the
perpendicular component, will be designated as Eis .
At the interface, the incident wave gives rise to reflected and transmitted waves. As
mentioned earlier, the existence of these waves is necessary to satisfy the boundary conditions
discussed in the preceding section. We can express these waves as
Nature of Light 33
Er = E0r exp[–j(wrt – kr · r)] (1.64)
1
Br = E exp[–j(wrt – kr · r)] (1.65)
v1 0r
and
Et = E0t exp[–j(wtt – kt · r)] (1.66)
1
Bt = E exp[–j(wtt – kt · r)] (1.67)
v2 0t

The boundary conditions demand that the combined fields in medium 1 must join the fields in
medium 2 seamlessly. All the boundary conditions at x = 0 are of the form

( ) exp[–j(wit – ki · r)] + ( ) exp[–j(wrt – kr · r)] = ( ) exp[–j(wtt – kt · r)] (1.68)

Recall that the boundary conditions must hold at every point on the interface at all times.
And if boundary conditions hold at one point at anytime, these will hold for all points at all
times if and only if the phases of these waves are equal. That is, the exponential functions in
Eq. (1.68) must be identically equal at the interface:
exp[–j(wit – ki · r)] = exp[–j(wrt – kr · r)] = exp[–j(wtt – kt · r)]
or
wit – ki · r = wrt – kr · r = wtt – kt · r (1.69)

So for the equality of phases at all times, we must have


wi = wr = wt (1.70)
That is, the frequency of the electromagnetic wave does not change when it undergoes reflection
or refraction and is therefore a fundamental property of a wave.
Since the reflected, refracted and transmitted waves have the same frequency, we must have
ki · r = kr · r = kt · r at x = 0
In component form, we can rewrite it as
( kiy ) y  ( kiz ) z ( kry ) y  (k rz ) z (kty ) y  ( ktz ) z for all y and z (1.71)
This relation will hold if and only if the components are separately equal. For y = 0, we have
kiz = krz = ktz (1.72a)
and for z = 0, we get
kiy = kry = kty (1.72b)
We may choose our axes so that ki lies in the y–x plane. Then kiz = 0. It means that krz = 0
and ktz = 0, implying that kr and kt also lie in the same plane. We may therefore conclude that

(i) The incident, reflected and the transmitted wave vectors lie in the plane of incidence.
The normal to the surface also lies in the same plane.
34 Wave Optics

Proceeding further, we can rewrite Eq. (1.72b) as


ki sin qi = kr sin qr = kt sin qt (1.73)
where qi, qr and qt respectively denote the angle of incidence, the angle of reflection and the
angle of refraction (transmission).
Further since the magnitudes of wave vectors ki, kr and kt are connected through the relation

v2 n1
ki kr kt kt (1.74)
v1 n2
we may conclude that
(ii) qi = qr. That is, the angle of incidence is equal to the angle of reflection (The law of
reflection) , and
(iii) n1 sin qi = n2 sin qt, which is Snell’s law.
Recall that (i), (ii), and (iii) are familiar laws of reflection and refraction obeyed by light waves.
We now impose boundary conditions on the fields in order to obtain Fresnel’s amplitude
relations for oblique incidence. As mentioned earlier, we consider the cases when E is in the
plane of incidence and in the plane perpendicular to the plane of incidence.

(i) E vectors in the plane of incidence


Refer to Figure 1.11. It shows that the electric field vectors E associated with the incident,
reflected and refracted electromagnetic waves lie in the plane of incidence and the corresponding
magnetic field vectors B point out of the page.

Figure 1.11 Reflection and transmission of an electromagnetic wave at an interface for oblique incidence.
(a) The electric field vectors E associated with the wave are in the plane of incidence.
(b) The electric field vectors are in the plane perpendicular to the plane of incidence.

We can resolve the electric field vectors associated with the incident, reflected and transmitted
waves into components as
Nature of Light 35
Eix = E0i sinq i exp(–jfi); Eiy = E0i cosq i exp(–jfi)
Erx = E0r sinq r exp(–jfr); Ery = –E0r cosq r exp(–jf r)
and Etx = E0t sinq t exp(–jft); Ety = E0t cosq t exp(–jft)
where we have abbreviated the space and time dependence as
f = wt – k · r
Using the boundary conditions that tangential components of E and B are continuous
[Eqs. (1.38) and (1.39)], we can write
(E0i – E0r) cosq i = E0t cosq t (1.75a)
and
B0i B0r B0t
 = (1.75b)
P1 P1 P2
As before, we assume that m1 = m2 = m0. Then using the relation E = vB, Eq. (1.75b) can be
rewritten as
E0i E0 r E0t

v1 v1 v2
In terms of refractive indices, we can express this result as
n1(E0i + E0r) = n2E0t (1.76)
On combining Eqs. (1.75a) and (1.76), we can write
n1
( E0i  E0r ) cos Ti ( E0i  E0 r ) cos Tt
n2
On re-arranging terms, we get
(n2 cosq i – n1 cosqt) = (n2 cosqi + n1 cosqt) E0r
so that the amplitude reflection coefficient for E vectors in the plane of incidence is given by

E0 r n2 cos Ti  n1 cos Tt
Rp (1.77)
E0i n2 cos Ti  n1 cos Tt
On using this result in Eq. (1.76) and simplifying the resultant expression, we obtain the
required expression for the amplitude transmission coefficient:
E0t 2n1 cos Ti
Tp (1.78)
E0i n2 cos Ti  n1 cos Tt
We now use Snell’s law (n1 sin qi = n2 sin qt) to eliminate the refractive indices and substitute
for n1 = n2 sin qt /sin qi in Eq. (1.77). On simplification, we can rewrite the expression for
amplitude reflection coefficient as
sin Ti cos Ti  sin Tt cos T t
Rp (1.79a)
sin Ti cos Ti  sin Tt cos T t
36 Wave Optics

Using the simple trigonometric manipulations, it is possible to recast this result in different
equivalent forms:
sin 2Ti  sin 2Tt
Rp = (1.79b)
sin 2Ti  sin 2Tt
and
tan(Ti  Tt )
Rp = (1.79c)
tan(Ti  Tt )
Similarly, using Snell’s law in Eq. (1.78), we can easily show that we can rewrite the expression
for the amplitude transmission coefficient as

2sin Tt cos T i
Tp =
cos Ti sin Ti  sin Tt cos Tt
2sin Tt cos Ti
= (1.80)
sin(Ti  Tt ) cos(Ti  Tt )
Let us now discuss a few simple but interesting phenomena exhibited by light.
(a) When n1 = n2, q1 = q2. Then, it readily follows from Eqs. (1.77) and (1.78) that
R p = 0 and T p = 1
Physically, it means that when the second medium has the same refractive index as the
first medium, there will be no reflected light wave. That is, light is completely transmitted,
as expected since it amounts to continuity of a medium.
(b) When q i + q t = p/2, Eq.(1.79c) implies that R p = 0 and the entire energy will be carried
by the transmitted beam. Thus, if natural light, which has no preferential direction of
vibration of associated electric and magnetic field vectors, is incident on an interface
at an angle such that q i + q t = p/2, the parallel component of the E-vector will not be
reflected. That is, the reflected light will have E-vector perpendicular to the plane of
incidence. (Such a light is said to be plane polarised. We will learn about it in detail
in the following chapter.) This is depicted in Figure 1.12. The corresponding angle of
incidence is known as Brewster’s angle or angle of polarisation. We usually denote it
as q B .
For this case, Snell’s law can be written in the form

n2 sin Ti sin T B
tan T B
n1 sin Tt ÈS Ø
sin É  T B Ù
Ê2 Ú
or
È n2 Ø
TB tan 1 É Ù (1.81)
Ê n1 Ú
Nature of Light 37

Figure 1.12 When unpolarised light is incident at the interface of two optically different media at the
polarising angle, the reflected beam is plane polarised.

This result shows that tangent of the Brewster angle gives the refractive index of the
second medium with respect to the first. This gives a convenient method to measure the
refractive index of a dielectric medium.
(c) When light is incident on a denser medium, q t < q i and for q i + q t > p /2,
(i.e., q i > q p), R p will be negative implying a phase change of p.
(d) Suppose that light is incident from a rarer medium to a denser medium at grazing
incidence (qi » p/2). If we put n2/n1 = n, z1 = p/2 – qi and z2 = p/2 – qt, Eq. (1.77)
takes the form
n sin ]1  sin ] 2
Rp (1.82)
n sin ]1  sin ] 2
Note that
sin Ti cos ] i
n
sin Tt cos ] t
so that
1/ 2
2 1/2
È cos 2 ] i Ø
sin ] t (1  cos ] t ) É1  Ù
Ê n2 Ú

Using these results in Eq. (1.82), we can rewrite the expression for the amplitude
reflection coefficient as
1/ 2
È cos2 ]i Ø
n sin ]i  É1  Ù
Ê n2 Ú
Rp 1/ 2
È cos2 ]i Ø
n sin ]i  É1  Ù
Ê n2 Ú
Since zi and zt are small for grazing incidence, it is a good approximation to take
sin zi » zi and coszi » 1. Then the expression for R p takes the form
38 Wave Optics

1/ 2
È 1Ø
n] i  É1  Ù
Rp »
Ê n2 Ú (1.83)
1/ 2
È 1Ø
n] i  É1  Ù
Ê n2 Ú

ËÈ 1/ 2 Û

Ì É1  Ù  n]i Ü
ÍÊ n2 Ú Ý
= 
1/ 2
ËÈ 1Ø Û
Ì É1  Ù  n]i Ü
ÍÊ n2 Ú Ý

1/ 2
È 1Ø Ë n] i Û
É1  Ù Ì1  Ü
Ê n2 Ú È 1Ø
1/ 2
Ì 1  Ü
É Ù
ÌÍ Ê n2 Ú ÜÝ
= 
1/ 2
È 1Ø Ë n] i Û
É1  Ù Ì1  Ü
Ê n2 Ú Ì È 1Ø
1/ 2
Ü
É1  Ù
ÌÍ Ê n2 Ú ÜÝ

1
Ë n] i ÛË n] i Û
=  Ì1  Ü Ì1  Ü
Ì (n 2  1) ÜÌ ( n2  1) Ü
ÌÍ n2 ÜÝ Ì
Í n2 ÜÝ

If we retain terms only up to first order in zi in the binomial expansion of the second bracketed
term and simplify, we get
2
Ë n]i Û
Rp  Ì1  Ü
Ì ( n 2  1) Ü
Ì
Í n2 Ü
Ý

Hence the expression for the amplitude reflection coefficient can be written in a compact
form as

Rp
Ë
 Ì1 
2n2]i
 " Û
Ü (1.84)
Ì
Í (n 2  1) Ü
Ý

Note that as z i ® 0, R p ® –1. That is, at grazing incidence, reflection is complete and the
amplitude transmission coefficient tends to zero. So if you hold a glass plate horizontally at the
eye level and allow light to reach it, the plate will reflect the entire incident light and act as a
mirror. Even a rather poor surface, such as the cover of this book, will be mirror-like at glancing
incidence. You should convince yourself by doing this activity.
Nature of Light 39
You may now like to answer a Practice Exercise.

Practice Exercise 1.8 An electromagnetic wave travels from a denser medium to a rarer
medium. Starting from Eq. (1.80), show that the entire energy will be reflected back into the
first medium, if the angle of incidence is greater than the critical angle.

(ii) E vectors in the plane perpendicular to the plane of incidence


Let us now suppose that E-vectors lie in the plane perpendicular to the plane of incidence and
point upward. Then, B-vectors will lie in the plane of incidence, as shown in Figure 1.11(b).
As before, we resolve B-vectors into components as follows.
Bix = –B0i sinqi exp[–j(wit – k · r)]; Biy = –B0i cosqi exp[–j(wit – k · r)]
Brx = –B0r sin qr exp[–j(wrt – k · r)]; Bry = B0r cosqr exp[–j(wr t – k · r)]
and Btx = –B0t sinqt exp[–j(wtt – k · r)]; Bty = –B0t cosqt exp[–j(wtt – k · r)]
We now apply the boundary conditions that tangential components of E and B are continuous
at the interface. This gives
E0i + E0r = E0t (1.85)
and
–(B0i – B0r) cosqi = –B0t cosqt (1.86a)
Since B = (1/v )E and n = c/v, Eq. (1.86a) can be rewritten as
n1(E0i – E0r) cosq i = n2E0t cosqt (1.86b)
On combining this result with Eq. (1.85), we get
E0 r n1 cos Ti  n2 cos Tt
Rs (1.87a)
E0i n1 cos Ti  n2 cos Tt
If we use Snell’s law (n1 sinq i = n2 sinq t), we can rewrite the expression for amplitude
reflection coefficient when E-vectors are perpendicular to the plane of incidence as
sin Tt cos Ti  sin Ti cos Tt
Rs = (1.87b)
sin Tt cos Ti  sin Ti cos Tt
sin(Ti  Tt )
=  (1.87c)
sin(Ti  Tt )
From this expression, we note that for grazing incidence (qi = p/2), Rs = –1 indicates that
reflection is complete. It means that if angle of incidence exceeds the critical angle, total internal
reflection will occur.
If we eliminate E0r from Eq. (1.86b) using Eq. (1.85), we get the expression for amplitude
transmission coefficient when E-vectors are perpendicular to the plane of incidence:
E0t 2 n1 cos Ti
Ts (1.88a)
E0i n1 cos Ti  n2 cos Tt
40 Wave Optics

As before, we can rewrite it in a slightly different form using Snell’s law as

2 sin Tt cos Ti
Ts (1.88b)
sin(Ti  Tt )
Equations (1.79), (1.80), (1.87) and (1.88) are known as Fresnel’s formulae.
Let us now sum up what you have learnt in this chapter.

1.6 SUMMARY

• Light exhibits wave–particle duality. Light behaves like waves when it propagates in a
medium or exhibits the phenomena of interference and diffraction. Light also behaves as
particles in its interaction with matter, as in the photoelectric effect.
• Visible light occupies the portion between 400 nm and 700 nm (7.5 ´ 1014 to
4.3 ´ 1014 Hz) and is a very small part (» 4%) of the entire electromagnetic spectrum.
• Human vision is a mix of physical and physiological processes; it begins in the eye but
light is sensed by the brain.
• Maxwell identified light with electromagnetic waves. The electric and magnetic fields
satisfy the following equations for free space:
˜ 2E
³2E H 0 P0
˜t 2
and
˜ 2B
³2B H 0 P0
˜t 2
• The speed of propagation of electromagnetic waves in free space is given by
1
v
H 0 P0
where e0 and m0 are permittivity and magnetic permeability respectively for free space.
• The electromagnetic waves are transverse. If we choose x-axis to be along the electric field
vector E, then the magnetic field vector will be along the y-axis and we can write
E = x̂ E0 exp[j(kz – wt)]
and
B = ŷ B0 exp[j(kz – wt)]
where
k E0
B0 = E0
Z c
• The Poynting vector is defined as
1
S (E – B )
P0
and we may interpret S · dA as the electromagnetic energy crossing area dA per unit time.
Nature of Light 41
• The amount of energy carried by an electromagnetic wave across a unit area per unit time
is given by
k
S E02 cos 2 ( kz  Z t )
P0Z
• The intensity of an electromagnetic (light) wave propagating in free space is given by

1 2 1 H0 2
I H 0 c E0 E0
2 2 P0

• When an electromagnetic wave is incident normally on the interface between two optically
different media, the amplitude reflection and transmission coefficients are given by
2
È n1  n2 Ø
R= Én  n Ù
Ê 1 2Ú
2
n2 È 2n1 Ø
and T=
n1 ÉÊ n1  n2 ÙÚ

• When an electromagnetic wave is incident obliquely on the interface between two optically
different media and electric vectors are in the plane of incidence, the amplitude reflection
and transmission coefficients are given by

E0 r n2 cos Ti  n1 cos Tt
Rp =
E0i n2 cos Ti  n1 cos Tt
E0t 2 n1 cos Ti
Tp =
E0i n2 cos Ti  n1 cos Tt

• When an electromagnetic wave is incident obliquely on the interface between two optically
different media and electric vectors are normal to the plane of incidence, the amplitude
reflection and transmission coefficients are given by

E0 r n1 cos Ti  n2 cos T t
Rs = =
E0i n1 cos Ti  n2 cos T t
sin Tt cos Ti  sin Ti cos Tt
=
sin Tt cos Ti  sin Ti cos Tt
E0 t 2n1 cos Ti
Ts = =
E0i n1 cos Ti  n2 cos T t
2sin Tt cos Ti
=
sin(Ti  Tt )
42 Wave Optics

REVIEW EXERCISES

1. The electric field associated with an electromagnetic wave is given by

È 14 S Ø
Ex 0, E y 50cos É 2S – 10 t xÙ , and Ez 0
Ê 3 Ú

Calculate the wavelength and the associated magnetic field.


Ë zˆ È S ØÛ
Ì Ans. O 6 m; B 50cos É 2S – 1014 t  xÙ Ü
Í c Ê 3 ÚÝ
2. The direction of motion of a plane sinusoidal light wave is represented as –xˆ. Its frequency
is 1014 Hz. The electric field is normal to the z-axis. Derive the expressions for electric
and magnetic fields associated with this wave.
Ë ÈS 14 Ø 1 ÈS ØÛ
Ì Ans. E yˆ E0 cos 2S É  10 t Ù ; B zˆ E0 cos 2S É  1014 t Ù Ü
Í Ê3 Ú c Ê 3 ÚÝ
3. –1
The electrical breakdown in air occurs at electric field strength of 3.5 MV m . Calculate
the corresponding intensity of a plane wave. [Ans. 1.626 ´ 1010 J s–1 m –2]
4. The laser beam described in Solved Example 1.4 is made to fall on 5 ´ 10–4 m2 area of
a perfectly reflecting plane mirror for half an hour. Calculate the average force exerted on
the target area. [Ans. 3.33 ´ 10–7 N]
5. Light from the sun is incident on the Earth at the rate of about 1.4 kW m–2 on an area
perpendicular to the direction of light. If the sun is assumed to emit light only of wavelength
600 nm, calculate the number of photons per second on one square metre of the Earth’s
surface directly from the Sun. [Ans. 4.2 ´ 1021]
6. Using the data given in Problem 5, calculate the strength of the magnetic and electric
fields of sunlight at Earth’s surface.
[Ans. B = 1.93 ´ 10–6 T; E = 7.26 ´ 102 V m–1]
7. Show that the time average of the energy density in a monochromatic plane wave
propagating in an isotropic non-conducting medium is distributed equally between the
electric and magnetic fields.
8. The root mean square of the displacement current density in a monochromatic plane wave
in free space is 10 mA m–2. Calculate the amplitudes of the electric and magnetic fields
associated with the wave.
[Ans. E0 = 2.54 ´ 10–3 V m–1; B0 = 8.47 ´ 10–12 T]
9. In a TV picture tube, electrons are accelerated through a potential difference of 10 kV.
Calculate the frequency of the X-rays that are emitted when these electrons strike the
screen. [Ans. 2.4 ´ 1018 Hz]
10. The average wavelength of light emitted by a 100 W lamp is 550 nm. Calculate the
number of photons emitted per second. [Ans. 2.8 ´ 1020]
CH A P T E R 2
POLARISATION OF LIGHT

EXPECTED LEARNING OUTCOMES


After reading this chapter, the student will acquire the capability to:
• Discuss how polarisation of light provided the surest test of transverse nature of light.
• Distinguish between linearly polarised, circularly polarised and elliptically polarised lights.
• Describe polarisation of light by reflection, double refraction, scattering and dichroism.
• Explain the working of Nicol prism and wave plates to produce polarised light.
• Discuss practical applications of polarisation.
• Solve numerical problems.

2.1 INTRODUCTION

In Chapter 1, you learnt that in some physical conditions, wave-like behaviour of light stands
out while in some other situations, particle-like behaviour dominates. According to classical
electromagnetic wave theory, light consists of electric and magnetic fields oriented perpendicular
to one another. Moreover, the direction of propagation of the wave is perpendicular to the plane
containing electric and magnetic fields (Figure 1.4). Recall that if the electric field is oriented
along the x-axis, and the magnetic field along the y-axis, the wave propagates along the z-axis.
This means that vibrations in the electric field are parallel to one another at all points in the
wave, so that the electric field forms a plane called the plane of vibration. Similarly, all points
in the magnetic field component of the wave lie in a plane that is normal to the plane containing
electric field. You also learnt that electric and magnetic fields support mutually as they oscillate
(in phase) and the rate of oscillation determines the frequency of the wave.
In Section 1.5.2, you learnt that when an electromagnetic wave falls obliquely on the
interface of two optically different media, and E-vector is in the plane of incidence, the wave
is plane polarised. As such, polarisation is a characteristic of transverse waves only. (You can
convince yourself about the latter part of this statement by taking a string and a narrow slit.)
In fact, in the early stages of development of wave theory, the phenomenon of polarisation
43
44 Wave Optics

provided unflinching evidence in support of wave theory of light. Therefore, we begin our study
by revisiting the property of polarisation of light. In particular, we seek answers to questions
such as what is polarisation and how it is visualized in terms of the orientation of electric and
magnetic fields? What are the different states of polarisation of light? How can we polarise a
light wave? Where does polarisation find applications? Why do we observe patterns of coloured
light on the windshield of a car? In this Chapter, we propose to build upon qualitative knowledge
of polarisation.
In Section 2.2, we have discussed a simple experiment to explain what we mean by
polarisation and the simple states of polarised light. Section 2.3 is devoted to the discussion of
different states of polarisation. You will appreciate the subtle differences between plane polarised,
circularly polarised and elliptically polarised lights. In fact, you will see that circular polarisation
is a special case of elliptical polarisation. Section 2.4 is devoted to principles of producing
linearly polarised light by reflection, double refraction and selective absorption. Simple instruments
used in the study of polarisation such as Nicol prism, Wave plates, Babinet compensator form
the subject matter of discussion of this section. You will observe that natural as well as artificial
devices are fabricated for selective absorption. The practical and modern applications of
polarisation are discussed in Section 2.5. Some of these include stress analysis, optical activity
and liquid crystal displays (LCDs), which can be seen on calculators and thin flat computer/TV
screens.

2.2 WHAT IS POLARISATION?

To understand what is meant by polarisation of light, we revisit a mechanical analogy. Refer to


Figure 2.1, which shows a transverse wave on a string. Note that in Figure 2.1(a), the string is
always in the same vertical plane, whereas in Figure 2.1(b), the string is always in the horizontal
plane. Such waves are said to be plane polarised because in each case vibrations of the string
are always confined to just one (fixed) plane. If a screen with a vertical slit is placed in the path
of the vertically polarised wave (Figure 2.1(c)), the wave will pass through the slit unchanged/
unimpeded. But if a screen with a horizontal slit is placed in the path of the wave emerging from
the vertical slit as shown in Figure 2.1(d), it will be blocked. That is, no wave will be transmitted
beyond the second obstacle.
Now refer to Figure 2.2. The light from an ordinary source is made to pass through
a Polaroid (P), which is just like the vertical slit in Figure 2.1(c). You will see that the
intensity of light drops to nearly one-half of the original value. If we rotate the polariser in its
own plane, no further change in the intensity of light is observed. Next, we allow the light
coming out of P to pass through an identical Polaroid (A) arranged parallel to P. The intensity
of light is not affected if the axes of A and P are parallel. But on rotating A in its own plane,
keeping Polaroid P fixed, a remarkable effect is observed! When A is rotated by 90o with respect
to P, the light is almost completely cut-off. However, if we rotate A further, the intensity of light
coming out of it begins to increase gradually, becoming maximum when P and A are parallel
again. How do we interpret this observation? This result shows that light propagates as a
transverse wave.
Polarisation of Light 45

Figure 2.1 (a) A vertically plane polarised wave on a string, (b) A horizontally plane polarised wave
on a string, (c) A vertically plane polarised wave passes through a vertical slit unimpeded,
and (d) A vertically plane polarised wave is blocked by a horizontal slit.

Figure 2.2 Schematics of the experimental arrangement used to polarise light.


46 Wave Optics

Representation of Field Vibrations in Light


In Chapter 1, we learnt that according to electromagnetic theory of light, light consists of electric and
magnetic field vectors, which are perpendicular to each other as well as to the direction of propagation
of the wave. For simplicity, we concentrate only on the electric field and the way it oscillates with
respect to the direction of propagation of the wave. Let us assume that instantaneous electric vector
in the light beam travelling along the positive z-direction, say executes a linear vibration with the
direction and amplitude as shown in Figure 2.3(a). If this vibration continues unchanged, we say that
the light is plane polarised, since its vibrations are confined to the plane containing the direction of
propagation of the wave and oriented at an angle q. For unpolarised light, q changes randomly and
very rapidly. And every orientation is equally probable so that the average effect is completely
symmetrical about the direction of propagation of the wave.

Figure 2.3 Vibrations in unpolarised light: (a) Equally probable planes, (b) Mutually
orthogonal components of electric field, and (c) pictorial representation of
plane polarised and unpolarised light. The double-pointed arrows denote the
vibrations confined to the plane of the paper and dots represent the vibrations
normal to the plane of the paper.

The electric vector in unpolarised light can be resolved in two components. If its amplitude is E,
the mutually orthogonal components of electric field vector oriented at an angle q with the horizontal
axis will be E cosq and E sinq in magnitude. In general, the magnitudes of these components will be
different. But due to random orientation of the electric field, q takes all possible values and the net result
is equivalent to having two incoherent vibrations of equal amplitude at right angles to one another. The
most common method of depicting plane polarised components is shown in Figure 2.3(c).

To visualise this result in terms of electromagnetic theory of light, by convention,


we concentrate only on the electric field vector and the way it oscillates with respect to the
direction of propagation of the wave. When electric field vectors of a light source are randomly
oriented, the light is said to be unpolarised. Most of the light around us, say from the sun, an
electric lamp or a candle, is unpolarised. If the field vectors exhibit partial preferential orientation,
light is said to be partially polarised. A wave is said to be plane polarised (or linearly polarised),
if the associated electric field vector always lies in the same plane as the propagating wave.
The plane of polarisation is that plane in which no vibrations occur. It could be vertical
or horizontal. In Figure 2.4, we have depicted the plane of polarisation by ABCD. The plane
in which vibrations take place is known as the plane of vibration and it is shown as EFGH in
Figure 2.4.
Polarisation of Light 47

Figure 2.4 Cross-sectional view of planes of polarisation and plane of vibration. A polariser allows
components of electric fields parallel to the vertical transmission axis to pass unchanged.

Again refer to the arrangement shown in Figure 2.2. We see that all orientations of the
electric vector in the unpolarised light from the source are in the yz-plane. It propagates as such
till it reaches the polariser P, which has a vertical transmission axis and allows vertical electric
fields to go through it unhindered. If the transmission axis is along y-axis, the electric field
along y-direction (Ey) passes through it unaffected. Thus, after passing through the Polaroid P,
the electric vectors oriented only along y-axis will be present. That is, the light coming out of
P will be vertically polarised, as shown in Figure 2.4. So we can say that when electric vector
oscillates along a straight line in a plane perpendicular to the direction of propagation,
the light is said to be plane polarised. The electric field vectors of an unpolarised, partially
polarised and plane polarised light are shown in Figure 2.5. Note that in all cases, the direction
of propagation of the wave is into the plane of the paper.

Figure 2.5 Electric field vectors of (a) unpolarised, (b) partially polarised, and (c) plane polarised light.
48 Wave Optics

Unpolarised light can be polarised by passing it through a polariser, which is a sheet of material
with a molecular structure that only allows a specific orientation of the electric field to go through
it unchanged. The most common polariser is a plastic called Polaroid. It was invented by Edwin
Land in 1928 when he was 19 years old and an undergraduate student at Harvard.

When the plane polarised wave reaches Polaroid A, which is identical to P and has been
rotated at 90º with respect to it, it can allow only the z-component of E to pass. Since the wave
coming out of P and incident on A has only y-component, light will be completely blocked by
A. From this discussion, we may conclude that
• transverse waves can only be polarised; longitudinal waves do not exhibit polarisation
as vibrations are along the line of transmission.
• polarisation is the surest test of transverse nature of light and its electromagnetic origin.
In our country, the antenna transmitting TV signals is horizontally oriented. Before the
advent of cable TV supported by a set-top box, a TV antenna was mounted on the roof-top in
horizontal position. (You may still get to see this in the countryside.) This was done to ensure
that the receiving antenna got maximum pick up when it was oriented parallel to the transmitting
antenna. (In a set-top box, the reception is controlled through the disc mounted by the service
provider. In recent years, technology has undergone tremendous change and people even use
Internet Protocol TV.)

2.3 STATES OF POLARISATION

We now know that for plane polarised light, the electric vector oscillates up and down or left
and right in a straight line in the same plane and is perpendicular to the direction of propagation
of the wave. The spatial variation of electric field vector E for a linearly polarised light wave
is shown in Figure 2.6. Note that the envelope of the tips of the electric vector is a sinusoidal curve.

Figure 2.6 Spatial variation of electric field vector for a linearly polarised light.
The envelope of the tip of the electric vector is sinusoidal in time.
Polarisation of Light 49
It means that the tip of E executes one full cycle as one full wavelength passes through a
reference plane. Note that we have used a right-handed coordinate system and the
yz-plane (or x = 0 plane) is the plane of polarisation of the wave. (In a right handed coordinate
system, a right-handed screw, when turned, rotates the x-axis towards the y-axis, the direction
of advancing screw represents z-axis.)
Mathematically, a linearly polarised wave moving along z-axis can be expressed as
E( z , t ) zˆ E0 cos(kz  Z t ) (2.1)
Here ẑ is a unit vector along z -axis.
Two other states of polarisation are: circular polarisation and elliptical polarisation. The
paths followed by the tip of E for a right circular wave is shown in Figure 2.7. (In the following
section, we will learn that circular polarisation is a special case of elliptical polarisation.) Note
that in case of circular polarisation, both electric and magnetic field vectors have equal amplitude,
whereas in elliptical polarisation, these fields have unequal amplitudes. (In linearly polarised
light, the field vectors vibrate along one axis.)

Figure 2.7 Path traced by the tip of the electric vector for a right circular wave.

Let us now consider superposition of two linearly polarised waves of unequal amplitudes
and having a constant phase difference f. Suppose that these waves have the same frequency
and are moving in the same direction. Physically, we expect that the resultant electric field
vector will change in magnitude as well as orientation and the path traced out should generate
an elliptical, circular or linear figure in the reference plane as the wave propagates.

2.3.1 Superposition of Two Linearly Polarised Waves

Suppose that two light waves are moving along the z-direction. We choose their electric field
50 Wave Optics

vectors along the x- and y-axes. That is, the electric field vectors characterizing these waves are
mutually perpendicular. We can write these as
E1(z, t) = x̂E01 cos(kz – wt) (2.2)
and
E2(z, t) = ŷ E02 cos(kz – wt + f) (2.3)
where x̂ and ŷ are unit vectors along the x- and y-axes, respectively. (These are also called
polarization vectors.) Here f denotes the phase difference between the waves.
Proceeding further, we consider the scalar part of Eq. (2.3) and write its expanded form as

E2
= cos(kz – wt) cosf – sin(kz – wt) sin f
E02

On substituting for cos(kz – wt) from Eq. (2.2), we can write

E2 E1
= cos I  sin(kz  Z t ) sin I
E02 E01
so that
E2 E
 1 cos I = –sin(kz – wt) sin f (2.4)
E02 E01

Using the elementary trigonometry, Eq. (2.2) can be used to express the sine function in terms
of electric field vectors characterizing the first wave:

2 1/2
Ë È E1 Ø Û
sin( kz  Z t ) Ì1  É Ù Ü
Ì
Í Ê E01 Ú ÜÝ
Using this result in Eq. (2.4), we get

2 1/ 2
E2 E1 Ë È E1 Ø Û
 cos I  Ì1 
ÉE Ù Ü
sin I
E02 E01 Ì
Í Ê 01 Ú ÜÝ

On squaring both sides, we get


2 2 Ë 2
È E2 Ø È E1 Ø È E2 Ø È E1 Ø È E1 Ø Û
ÉE Ù É Ù cos 2 I  2 É ÙÉ Ù cos I Ì1  É Ù
2
Ü sin I
Ê 02 Ú Ê E01 Ú Ê E02 Ú Ê E01 Ú ÌÍ Ê E01 Ú Ü Ý

On rearranging terms and using the trigonometric relation sin2f + cos2f = 1, we get
2 2
È E2 Ø È E1 Ø ÈE2 Ø È E1 Ø
ÉE Ù

ÉE Ù
 2É Ù É Ù
cos I sin 2 I (2.5)
Ê 02 Ú Ê 01 Ú Ê 02 Ú Ê E01 Ú
E

Let us pause for a while and interpret Eq. (2.5). We come across a similar equation while
discussing the formation of Lissajous figures due to superposition of two oscillations of unequal
Polarisation of Light 51
frequencies amplitudes. Equation (2.5) represents an ellipse whose principal axis is inclined to
the (E1, E2) coordinate system. If a is the angle of inclination between the coordinate axes and
the axes of the ellipse, then

2 E01 E02 cos I


tan 2D 2 2
(2.6)
E01  E02
We may now conclude that superposition of two linearly
polarised light waves having different amplitudes and a
finite phase difference results in the production of an
elliptically plane polarised wave. That is, the orientation
of the resultant electric vector changes continuously and
its tip moves along an ellipse as the wave propagates.
This is shown in Figure 2.8.
For f = ±(2n + 1)p/2; n = 0, 1, 2, ..., Eq. (2.5) Figure 2.8 Schematic representation of
simplifies to elliptically polarised light.
2 2
È E2 Ø È E1 Ø
ÉE Ù

ÉE Ù
1 (2.7)
Ê 02 Ú Ê 01 Ú

This equation represents an ellipse with its minor and major axes along the coordinate axes.
If we further have E01 = E02, Eq. (2.7) represents a circularly polarised light wave. It means that
when two orthogonal linearly polarised light waves differing in phase by a multiple of p/2 are
made to superpose, the resultant wave is circularly polarised. That is, the orientation of the
resultant electric vector changes continuously and its tip moves along a circle as the wave
propagates with time. This means that E is not restricted to a single plane. You may now like
to know the direction of rotation of E. Obviously, it can either be clockwise or anticlockwise.
To discover the answer to this question, you should answer Practice Exercise 2.1.

Practice Exercise 2.1 Consider two orthogonal plane polarised light waves having equal
amplitude but their relative phase difference is 2np – p/2, n = 0, ±1, ±2, ... . That is to say,
these differ in phase by p/2. Obtain the expression for resultant electric field vector.
[Ans. E( z , t ) E0 [xˆ cos(kz  Z t )  yˆ sin(kz  Z t )]

If you tabulate the resultant electric vector obtained in Practice Exercise 2.1 at t = 0 for
space points at intervals of l/8 starting from z = 0, you will get the following result:

O O 3O O 5O 3O 7O
z 0 z z z z z z z z=l
8 4 8 2 8 4 8
eˆ x  eˆ y eˆ y  eˆ x eˆ x  eˆ y eˆ x  eˆ y
E eˆ x E0 E0 eˆ y E0 E0 eˆ x E0  E0 eˆ y E0 E0 eˆ x E0
2 2 2 2

If we position ourselves in the reference plane and note the evolution of electric vector from
z = l to z = 0 (backward towards the source), we will find that tip of E rotates clockwise,
52 Wave Optics

as shown in Figure 2.7. The electric field makes one complete rotation as the wave advances
through one wavelength (Figure 2.9).

Figure 2.9 Rotation of electric vector in a right circular wave.

To check your understanding of the basic concepts developed so far, we would like you to
answer the following Practice Exercise.

Practice Exercise 2.2 For the waves considered in Practice Exercise 2.1, we wish to observe
evolution of resultant electric field E with time rather than in space. For this, fix an arbitrary
point z = z0 and tabulate the values of electric vector. Depict your result graphically.

In case the relative phase difference between the waves considered in Practice Exercise 2.1
is 2np + p/2, n = 0, ±1, ±2, ..., the resultant electric field vector is given by

E( z , t ) E0 [xˆ cos( kz  Z t )  yˆ sin( kz  Z t )] (2.8)


It shows that electric field vector rotates anticlockwise in the reference frame. Such a wave is
referred to as left-circular wave. You should check this fact and convince youself by tabulating
the values of E at t = 0 for different space points.
Now suppose that two oppositely polarised circular waves of equal amplitude are superposed.
Their resultant is given by
E 2 E0 xˆ cos( kz  Z t ) (2.9)
Note that this equation is similar to Eq. (2.1) which represents a linearly polarised plane wave.
It means that superposition of two oppositely polarised circular light waves (of equal amplitudes)
gives rise to a linearly polarised light wave.
Proceeding further, we note that for f = np; n = 0, 1, 2, ..., Eq. (2.5) takes the form
2
Ë E1 n E2 Û
Ì E  ( 1) E02 ÜÝ
=0
Í 01
Polarisation of Light 53
or
E2 n E02
= ( 1) (2.10)
E1 E01
This equation represents a straight line, implying that superposition of two linearly polarised
light waves differing in phase by zero or an integral multiple of 2p give rise to a linearly
polarised wave, which can be specified as

E( z , t ) ( xˆ E01  yˆ E02 ) cos( kz  Z t ) (2.11)


2 2
The amplitude of the resultant wave is given by E01  E02 and the electric field oscillates in
–1
the reference frame with an angle q = tan (E02/E01) with the x-axis.
For the special case of in-phase waves of equal amplitude (E01 = E02 = E0), the resultant
wave will have an amplitude 2E0 and the associated electric vector will be oriented at 45º.
It means that when two in-phase linearly polarised light waves superpose, the resultant wave
is also linearly polarised and has fixed amplitude as well as orientation, as shown in
Figure 2.10(a). In the reference (observation) plane, we see a single electric vector oscillating
continuously along an inclined line Figure 2.10(b). The E-field progresses through one complete
cycle as the wave advances along the z-axis through one wavelength.

Figure 2.10 (a) Schematic representation of a plane polarised light wave arising from superposition of two
in-phase linearly polarised light waves and (b) Oscillations of electric vector in reference plane.

If we reverse this process, we can say that any linearly polarised light can be seen as made
up of two linearly polarised lights with planes of polarization parallel to x = 0 and y = 0
planes. Can you give an analogy for this from vector algebra? It is as if we are resolving a
vector along two mutually perpendicular directions. Is this result not expected?
Having discussed different states of polarisation, we now discuss different methods used to
produce linearly polarised light.
54 Wave Optics

2.4 PRODUCTION OF LINEARLY POLARISED LIGHT

Now we know that an ordinary light is an unpolarised wave and can be considered to be made
up of two orthogonal linearly polarised waves. We can produce polarised light by reflection of
light from a surface, by refraction and birefringence or double refraction through a crystal, by
selective absorption in a crystal and scattering by medium particles. The most common technique
for obtaining polarised light is to use a polarising material. These materials transmit when the
electric field components vibrate in a plane parallel to a specific direction. The material absorbs
waves whose electric field components vibrate in the other direction. Before we discuss
these in some detail, refer to Figure 2.11. The polariser P changes unpolarised light (from a
natural source, a sodium/mercury lamp or an ordinary lamp) into some form of polarised light.
(In this sense, polariser is the most important optical device/component in any arrangement/
equipment used to produce polarised light.) The light coming out of a polariser is in a definite
polarisation state; the electric vectors oriented parallel to the transmission axis of the polariser
are allowed to emerge and all others are discarded. This can be understood in terms of Malus Law.

Figure 2.11 Experimental arrangement for obtaining linearly polarised light.

Now refer to Figure 2.12. Natural light is incident on a polariser, whose transmission axis
makes an angle q with y-axis. For this arrangement, a polarisation state parallel to the transmission
axis of the polariser will be transmitted. The light emerging from the polariser is made to fall
on an analyser, whose transmission axis is parallel to y-axis. If there is no loss of light due to
any absorptive mechanism and E is the electric field transmitted by the polariser, its component
E cos q parallel to the transmission axis of the analyser will pass through. The intensity of the
transmitted polarised light reaching the detector varies as
I(q) = I(0) cos2q (2.12)
where I(0) is the intensity of the polarised light incident on the polariser. It is measured
in W m–2. q is the angle between the transmission axes of the polariser and the analyser.
For q = 0, i.e. when the transmission axes of the polariser and analyser are parallel, the
transmitted intensity is maximum. On the other hand, no light is transmitted for q = 90º.
It means that light of variable intensity and polarisation can be obtained by adjusting the
orientation of the polariser.
Equation (2.12) is known as Law of Malus, after the French Engineer Etienne-Louis Malus
(1775–1812). To illustrate this, let us consider sunlight that falls on two ideal polarisers placed
Polarisation of Light 55

Figure 2.12 (a) Transmission of unpolarised light through a polariser; (b) Transmission of unpolarised
light through a polariser and an analyser.

one after the other. The intensity of the unpolarised light will reduce to one-half after passing
through the first polariser. The intensity of transmitted light is obtained by averaging cos2q over
all angles since there is no fixed value of q. And the average value of cos2q is one-half so that
I = I(0)/2.
If the angle between the transmission axes of the polariser and analyzer is 30o, the intensity
of light transmitted finally will be

1 2ÈSØ
I(30º) = – cos É Ù
2 Ê 6Ú
2
1 È 3Ø 3
= –É Ù 0.375
2 Ê 2 Ú 8

It means that from a combination of two ideal polarisers whose transmission axes are inclined
at 30o, only 37.5% of the incident light will be transmitted finally. However, when transmission
axes of the polarisers are parallel (q = 0), nearly 50% of the incident light will be transmitted.
56 Wave Optics

We now advise you to go through the following example carefully.

EXAMPLE 2.1 Consider unpolarised light of intensity 500 W m–2. It is made to pass through
a polariser and an analyser. The axis of the polariser is vertical and the axis of the analyser is
at an angle of 45º with respect to the polariser axis. Calculate the intensity of the transmitted
light.
Solution The intensity of polarised light transmitted through the polariser will be half of the
intensity of unpolarised light. Hence
500
I1 250 W m 2
2
The intensity of polarised light transmitted through the analyser can be calculated using the Law
of Malus:
250
I out I1 cos 2 (45º ) 125 W m 2
2

You may now like to answer a Practice Exercise.

Practice Exercise 2.3 Unpolarised light is incident on two polarisers whose transmission
axes are parallel to each other. Calculate the angle at which one of them must be rotated so
that the transmitted intensity is half of the intensity incident on the second polariser.
[Ans. 45º]

So far we have considered an ideal polariser to produce linearly polarised light. We now
discuss polarisation by reflection.

2.4.1 Polarisation by Reflection: Brewster’s Law

One evening Malus was examining a calcite crystal while standing at the window of his house.
The image of the sun was reflected towards him from the windows of Luxembourg Palace.
When he looked at it through the calcite crystal, two images of the sun appeared due to
birefringence or double refraction. (We will discuss it in the next section.) But as he rotated the
crystal, he was amazed to observe that one of the double images disappeared. This effect was
studied by Malus in detail and he concluded that it occurred due to reflection of light at a
particular angle of incidence. You may have also experienced reflection of light from the jacket
of a book or ‘glare’ from reflections off the sea. Such a light is partially polarised. (The glare
can be reduced by wearing Polaroid sunglasses.) In fact, the reflection of light from glass or a
plastic sheet is one of the most convenient methods for producing polarised light. Now we
discuss it in some detail.
Suppose that an unpolarised light wave is incident at an angle q on a smooth interface of
two optically different media such as glass and air. We expect that light will be partially
reflected and partially refracted according to the laws of light, as shown in Figure 2.13(a).
The reflected light is completely polarised, partially polarised or unpolarised, depending on the
Polarisation of Light 57
angle of incidence. If the angle of incidence is 0º or 90º, the reflected light is unpolarised.
That is, when incident beam is normal to or grazing along the interface, the reflected beam is
unpolarised. For angles in-between 0º or 90º, the reflected light is partially polarised. However,
for one particular angle of incidence, the reflected light is completely polarised and the transmitted
light is partially polarised, as shown in Figure 2.13(b). This angle of incidence is called polarizing
angle or the Brewster’s angle after Sir David Brewster (1781–1868), a Scottish physicist. We
denote it as qB. Brewster also discovered experimentally that when the angle of incidence is
equal to the polarising angle, i.e. when reflected light is 100% polarised, the reflected light and
the refracted light are perpendicular to each other. In other words, Brewster’s angle is that angle
of incidence for which the angle between the reflected and the refracted rays is 90º.

Figure 2.13 Polarisation by reflection. (a) Reflection and refraction of light at an interface of two
optically different media, and (b) At Brewster’s angle, the reflected light is plane polarised.

To visualize this phenomenon in terms of the vibrations of electric field vector, we recall
that unpolarised light can be thought of as made up of two linearly polarised waves with electric
field vector components parallel and perpendicular to the plane of incidence. Note that in
58 Wave Optics

Figure 2.13(b), unpolarised light is represented as to indicate two components of electric


field; dots (‘•’) indicate electric field component perpendicular to the plane of the paper and
double pointed arrows (‘«’) indicate electric field vibration in the plane of the paper. When the
angle of incidence is equal to the polarising angle, the reflected beam has no field vibrations
parallel to the plane of incidence, i.e., vibrations of the electric field in the reflected beam are
normal to the plane of the paper. But the refracted beam has both field components.
From this result we can conclude that if unpolarised light wave is incident on a reflecting
surface at an angle equal to Brewster’s angle, the parallel component of E-vector will not be
reflected and the reflected light will be polarised with its E-vector perpendicular to the plane
of incidence. This is known as Brewster’s Law and is illustrated in Figure 2.13(b). Note that the
angle of refraction will be (p/2) – qB. Proceeding further, we recall from Snell’s law that
m 1 sin qB = m2 sin qr
where m 1 and m 2 are the refractive indices of the media at whose interface light waves are
incident. Since qr = (p/2) – q B, it readily follows that
m 1 sinqB = m 2 cosqB
or
P2
tanqB = (2.13)
P1
That is, for an air-crown glass interface, Brewster’s angle qB = tan–1(1.52) » 56.7º. For an
air–water interface, qB = tan–1(1.33) » 53º. It means that when the sun is 37o above the horizontal,
the light reflected by a calm pond or lake will be linearly polarised. However, at all angles other
than the Brewster’s angle, the reflected light will be partially polarised and the electric field
vector has a strong horizontal component. To absorb this and minimise its undesirable effects
on our eyes, we should wear polarising sunglasses with vertical optical axis. Note that the
transmitted beam will be partially polarised and sunglasses can reduce the overall intensity of
light transmitted through the lenses, depending on the quality of the material of the lenses.
If a glass slab (m = 1.5) is immersed in water (m = 1.33), the Brewster’s angle will be
qB = tan–1(1.5/1.33) = tan–1(1.13) » 48.4º. It means that polarising angle or Brewster’s angle is
a function of the index of refraction. The Brewster’s angle also varies to some degree with the
wavelength of the incident light.
Though polarisation by reflection is a very convenient and the most common method,
designing an effective polariser poses some problems on account of two reasons:
• The reflected beam, although completely polarised, is weak.
• The transmitted beam, though strong, is partially polarised.
To overcome these difficulties, we use a pile of plate Polarisers. It is fabricated with glass
plates for light in the visible region, silver chloride (AgCl) plates for light in the infrared region
and quartz for light in the ultraviolet region. Using a dozen or so microscope slides, you can
fabricate a pile and convince yourself about the authenticity of the preceding statement. In fact,
you may observe beautiful colours when the slides are in contact. These arise due to the
interference phenomenon, which forms the subject matter of discussion of Chapter 3.
Polarisation of Light 59
We now advise you to solve a Practice Exercise before proceeding further.

Practice Exercise 2.4 A plate of flint glass (m = 1.67) is immersed in a liquid of refractive
index 1.47. Calculate the angle at which a completely polarised reflected wave is obtained
[Ans. 48º]

Let us pause for a while and reflect as to what have we learnt and what more we wish to
achieve. We now know that light can be polarised by reflection. It is therefore natural to ask:
Can we use the phenomenon of refraction to polarise light. The answer is yes as well as no.
Let us understand it now.

2.4.2 Polarisation by Double Refraction

We know that refraction of light in isotropic crystals like NaCl or non-crystalline substances like
glass or water does not produce polarised light. However, when an unpolarised light beam is
made to fall on an anisotropic substance such as calcite, it splits into two beams. This remarkable
phenomenon of splitting of a light beam into two beams is known as double refraction or
birefringence. If we eliminate one of these through selective absorption, total internal reflection
or otherwise, we can obtain linearly polarised light. Before discussing how double refraction
can be used to produce polarised light, it is important to know how it arises and familiarise
ourselves with some basic concepts.

An isotropic substance exhibits the same behaviour in all directions, whereas an anisotropic substance
exhibits different properties in different directions.

(a) What and how of double refraction?


When we look at the book with unaided eye, we see only one image of every letter/word.
But if we tell you that it is possible to see more than one image of every letter/word, you may
not believe easily. However, if we give you a calcite crystal and advise you to place it on the
book to see through it, you will observe two images of every letter. Similarly, if we mark a dot
on a piece of paper and see it through this (calcite) crystal, we will observe two grey dots. And
if we rotate the crystal, we will note that one of the dots remains stationary while the other one
appears to move in a circle. As of now, you do not know anything whatsoever about the basic
phenomenon and the underlying physical principle. This phenomenon is known as double
refraction or birefringence. The phenomenon of double refraction was discovered by Bartholinus,
a Scottish Physician, in 1669 and later studied by Huygens and Newton. Bartholinus was so
excited by his investigations that he termed the calcite crystal one of the grandest creations of
nature. To quote him,
“Greatly prized by all men is the diamond, precious stones and
pearls but he, who prefers the knowledge of unusual phenomena
to these delights, will, I hope, have no less joy in a transparent
crystal of calcite (Author’s emphasis), , which perhaps is one of the
60 Wave Optics

greatest wonders that nature has produced. As my investigation of this


crystal proceeded, there showed itself a wonderful and extraordinary
phenomenon: objects which are looked at through the crystal do not
show, as a single refracted image, but they appear double.”

When a beam of unpolarised light enters a calcite crystal, it splits into two refracted beams
having different angles of refraction. One of these refracted beams obeys the Snell’s law of
refraction. Bartholinus termed it the ordinary ray (o-ray). The beam which does not obey Snell’s
law was termed the extraordinary ray (e-ray). Moreover, the velocity of ordinary ray is the same
in all directions, whereas the velocity of extraordinary ray is different in different directions.
The double refraction depends on the purity of the crystal and the wavelength of light. In other
words, the genesis of this spectacular effect is in two distinct principal indices of refraction in
crystalline substances such as calcite (calcium carbonate CaCO3), quartz (silicon dioxide SiO2),
mica, sugar, topaz, aragonite and ice. These indices correspond to the oscillations of E-vector
parallel and perpendicular to the optic axis. We can therefore visualise each beam as being
characterized by a certain state of polarisation; the ordinary beam is linearly polarised with its
vibrations in one plane and the extraordinary beam is linearly polarised with its vibrations in
a plane perpendicular to the first plane.
A birefringent crystal has a special direction, called optic axis, along which ordinary as well
as extraordinary rays travel with the same velocity. It means that the refractive indices for
ordinary and extraordinary rays are equal along the optic axis. For a calcite crystal, the direction
of the optic axis is determined by joining the two blunt corners of the crystal. In Figure 2.14,
the line AA¢ shows the direction of optic axis. It may be mentioned here that the birefringent
crystal is cut, grounded and polished to exact angles with respect to the optic axis. (In a calcite
crystal, each face is a parallelogram, whose angles are 78º and 102º. The two opposite surfaces
of the calcite crystal are parallel to one another. As a result, the refracted beams always emerge
parallel to the incident beam and are parallel to each other.)

Figure 2.14 The optic axis in a calcite crystal is obtained by joining


the blunt corners of the crystal. It is shown by line AA¢.
Polarisation of Light 61
Birefringent crystals having only one optic axis are called uniaxial crystals and crystals
having two optic axes are called biaxial crystals. Calcite, tourmaline, quartz, sodium nitrate and
ice are uniaxial crystals, whereas mica is a biaxial crystal. If refractive index for o-ray in a
uniaxial crystal is greater than the refractive index for e-ray (mo > me), it is said to be a negative
uniaxial crystal. On the other hand, if the refractive index for e-ray in a uniaxial crystal is
greater than the refractive index for o-ray (me > mo), it is said to be a positive uniaxial crystal.
The values of refractive indices for ordinary and extraordinary rays for a few typical birefringent
materials are given in Table 2.1. Note that calcite, sodium nitrate and tourmaline are negative
uniaxial crystals, whereas ice, quartz, titanium oxide and zinc oxide, etc., are positive uniaxial
crystals. The difference Dm = me – mo gives a measure of birefringence.

Table 2.1 Refractive indices for a few typical uniaxial


birefringent crystals for l = 589.3 nm
Crystal mo me Dm = me – mo
Ice 1.309 1.313 0.004
Quartz 1.5443 1.5534 0.0091
Calcite 1.6584 1.4864 –0.1920
Sodium nitrate 1.5854 1.3368 –0.2486
Tourmaline 1.669 1.638 –0.31

(b) Polarisation by a uniaxial crystal


Refer to Figure 2.15(a). It shows one of the principal sections ABCD of a calcite crystal. It is
perpendicular to the cleavage faces BECF and AHDG and contains the optic axis AC. Now we
know that when unpolarised light is made to pass through a calcite crystal, it splits into o- and
e-rays. The electric field vector of e-ray vibrates in the plane containing the optic axis while the
electric field vector of o-ray vibrates perpendicular to the plane containing the optic axis, as
shown in Figure 2.15(b). It means that due to birefringence, an unpolarised light wave splits into
two plane polarised waves.

Figure 2.15 (a) The optic axis and principal sections of a calcite crystal, and (b) When unpolarised
light passes through a calcite crystal, it splits into two components.
62 Wave Optics

We may now conclude that in double refraction


• Both o- and e-waves are plane polarised and their E-vibrations are perpendicular to each
other.
• The vibrations of the electric vector corresponding to the ordinary wave are always at
right angle to the direction of the wave as well as the optic axis.
• The direction of the vibrations of the electric vector corresponding to the extraordinary
wave is always at right angle to the propagation vector k and lies in the plane containing
the optic axis and the direction of the incident wave.
Huygens explained the phenomenon of birefringence in terms of wave surfaces using the
principle of secondary wavelets. (A wave surface is a wavefront or a pair of wavefronts completely
surrounding a point source of monochromatic light.) To appreciate the developments in historical
perspective, it would be worthwhile to spend some time here and learn about it now.

Huygens explanation of birefringence


According to Huygens, a point source of light in a birefringent medium is the origin of two sets
of different wavefronts. Refer to Figure 2.16. It shows a circle and an ellipse which represent
the traces (in 2-D) of the wavefronts emitted by a monochromatic source at P. (Recall that a
wavefront is foci of points of equal phase in the waves emitted by a source. In 3-D, the rotation
of circle about any axis generates a sphere and rotation of an ellipse about either the major or
the minor axis generates an ellipsoid of revolution.)

Figure 2.16 Wave surfaces in (a) a negative, and (b) a positive crystal.

We can now say that, the wave surface will be in the form of a sphere in 3-D for the
ordinary ray since it propagates with uniform velocity in all directions in the crystal. However,
for the e-ray, which propagates with different velocities in different directions in the crystal, the
wave surface will be an ellipsoid of revolution. Note that the axis of rotation is along the optic
axis. Further, to account for the equal velocities of o- and e-rays along the optic axis, Huygens
assumed that both these surfaces touch each other at the extremities on the optic axis. For a
negative crystal, the ellipsoid of revolution lies completely outside the sphere because the
velocity of the extraordinary wave is greater than that of the ordinary wave everywhere, except
along the optic axis. This is illustrated in Figure 2.16(a). On the other hand, in a positive crystal,
the ellipsoid of revolution lies completely inside the sphere because the velocity of extraordinary
Polarisation of Light 63
wave is less than that of the ordinary wave everywhere, except along the optic axis, as shown
in Figure 2.16(b).
It may be pointed out here that
• The wave surfaces shown in Figure 2.16 correspond to one wavelength only. For a
polychromatic source and/or dispersive media, larger or smaller surfaces will be obtained.
• Radii of wave surfaces are proportional to phase velocity and therefore do not measure
the rate of propagation of energy.
We have discussed propagation of plane waves in uniaxial crystals based on Huygens
construction in detail in Appendix–2A.
Proceeding further, we note that if we can somehow remove either o- or e-ray produced in
birefringence, it is possible to produce linearly polarised light. In practice, this is readily done
using a Nicol prism, which removes the ordinary ray through total internal reflection. It was
designed by William Nicol in 1828. We will discuss it now.

Nicol prism
Refer to Figure 2.17(a). It shows the principal section PQRS of a Nicol prism, which is made
from a naturally occurring crystal of calcite. There are many different forms of Nicol prism but
the most common prism is made from a crystal whose length to width ratio is 3:1. The smaller
faces (PQ and RS) are grounded from 71º to 68º. The crystal is then cut along P¢R¢ by a plane
passing through P¢ and R¢ and perpendicular to both the principal section PQRS and end-faces.
The two cut surfaces are grounded and polished to optical flatness and then cemented together
with a layer of Canada balsam, which is a non-refringent material and transparent to visible
spectrum.
The natural question that arises here is: Why do we use Canada balsam as the cementing
material. To discover answer to this question, we note that for sodium light, the refractive index
of Canada balsam (mCB) is 1.552. On the other hand, the refractive indices of Canada balsam
for o- and e-rays are mo = 1.658 and me = 1.486 respectively. That is, mCB lies in-between mo
and me. It means that Canada balsam is optically rarer with respect to o-ray and optically denser
with respect to e-ray. Hence, the critical angle for total internal reflection of o-rays is
sin–1(1.552/1.658) = 69º. So if natural light is incident on the Canada balsam surface at an angle
of 69º, the o-ray will be totally internally reflected and the emergent light will be made up only
of e-component. That is, the output of Nicol prism is a linearly polarised light beam. (It is for
this reason that end faces are grounded from 71º to 68º.)
Note that if o-ray strikes at an angle less than the critical angle, it may also be transmitted.
This sets a limit on the angular field on one side of the axis. Moreover, the refractive index for
e-ray depends on the direction of propagation; mo = 1.658 when propagating along the optic axis
and 1.486 when travelling perpendicular to it. Therefore, for oblique incidence, e-rays may also
be lost due to total internal reflection. As a result, the field of view of Nicol prism is limited
by total internal reflection of the e-ray on one side and partial transmission of the o-ray at the
other side. Nowadays Nicol prisms with the useful field of ±14º are available. Another very
serious drawback of Nicol prism is that it can be used only for visible light; the Canada balsam
layer absorbs ultraviolet light.
64 Wave Optics

There are a few other disadvantages. For instance, the transmitted beam may be laterally
shifted, the region of incidence of linearly polarised light is not fixed and a lot of light is lost
due to reflection. Some of these difficulties are overcome in Glan–Thompson prism. The entrance
face of Glan–Thompson prism is normal to the axis, and the optic axis of the crystal lies in the
entrance face, as shown in Figure 2.17(b). Though Glan–Thompson prism is slightly longer than
Nicol prism for the same aperture; its field is about 30º. Moreover, the state of polarisation of
the transmitted beam is uniform.

Figure 2.17 (a) Nicol prism. The dashed outline corresponds to the natural crystal. (b) Glan–Thompson prism.

For use in the ultra-violet region of the spectrum, we use Glan–Foucault prism in which an
air gap is kept between the two segments. However, even in this case, and in fact in all crystal-
based polarisers, the aperture is small.

Rochon and Wollaston prisms


Sometimes, it is desirable to retain both o- and e-rays for later comparison of their intensities,
and maximise separation between the beams.
For this purpose, different types of beam splitters have been designed but the Rochon prism
and Wollaston prism are the most widely used optical devices. These devices are also referred
to as double-image prisms. As such, these are also made of quartz or calcite but at certain
definite angles and cemented together with glycerine or castor oil.
Rochon prism consists of two quartz prisms ABC and BCD joined together as shown in
Figure 2.18(a). The optic axis of the prism ABC is in the plane of the paper (indicated by the
broken line). The optic axis of the prism BCD is perpendicular to the plane of the paper
(indicated by the dots). Light is made to enter the prism ABC normal to the surface. It travels
along its optic axis undeviated up to the boundary BC and undergoes double refraction at the
boundary BC of the second prism. In the prism BCD, the o-vibrations pass without any deviation.
In quartz prisms, the e-vibrations are deviated downwards, as shown in Figure 2.18(a).
(In a calcite prism, the e-vibrations are deviated upwards.)
Polarisation of Light 65
In the Wollaston prism [Figure 2.18(b)], the optic axis of the prism ABC is in the plane of
the paper but normal to the horizontal (indicated by the broken line). The optic axis of the prism
BCD is perpendicular to the plane of the paper (indicated by the dots). The light falls on the
face AC and travels perpendicular to the optic axis without deviation as in case of Rochon
prism. However, unlike the case of Rochon prism, here both types of vibrations are deviated
leading to greater separation between them. In quartz, me > mo suggesting that the o-beam will
be refracted towards the normal, whereas the e-vibrations will be refracted away from the
normal. (For calcite prisms, the directions of o- and e-vibrations will be interchanged.)

Figure 2.18 Typical quartz (a) Rochon and (b) Wollaston prisms.

It is worth mentioning here that in Rochon prism, light is made to enter from the left to
enable it to travel along the optic axis. If it enters from the other direction, we will observe
rotatory dispersion. (This is discussed later in this chapter.) You may now like to go through
the following Solved Example.

EXAMPLE 2.2 A birefringent prism comprises a glass prism of refractive index m = 1.66 and
a calcite prism with mo = 1.658 and me = 1.486. The angle of the prism is 30º. Calculate the
angle between the rays emerging from the prism when the ray enters the glass prism normally.
Solution Refer to Figure 2.19. The incident light splits into o- and e-vibrations at the
glass–calcite interface. The o-ray propagates without any deviation and emerges normally.

Figure 2.19 A glass-calcite birefringent prism.


66 Wave Optics

For the vibrations in the plane of the paper, we can write


mo sin30º = me sin r
Po
\ sin r = sin 30º
Pe
or
P
r = sin 1 È o sin 30º Ø sin 1 ÈÉ
1.658 Ø
– 0.5Ù
ÉP Ù Ê 1.486 Ú
Ê e Ú
= sin–1(0.5579) = 33.91º
Hence, the angle of incidence at the second surface will be 33.91º – 30º = 3.91º. And the
emerging angle will be given by
sinq = me sin (3.91º) = 1.486 ´ 0.0682 = 0.1013
Hence
q = 5.81º

(c) Polarisation by a biaxial crystal


Birefringence in biaxial crystals, just like in uniaxial calcite and quartz crystals, is most easily
described in terms of wave-surface diagrams and Huygens’ Principle. In Figure 2.20, we have
depicted three cross-sectional views of the wave surfaces which started from O and spread out
inside the biaxial crystal. As before, the directions of vibrations are shown by dots and arrows.
Note that each section cuts the two surfaces in one circle and one ellipse. Moreover, these are
different in three sections depicted here.

Figure 2.20 Cross-sections of wave surfaces for a biaxial crystal. The figures are drawn
for semi-axes a = 3, b = 2 and c = 1.

Of the three cross-sections, the section of the wave surface in the x–z plane is of special
interest. Note that there are four singular points (R1, R2, R3, R4), where circle (outer wave
surface) and the ellipse (inner wave surface) cut each other. We have enlarged this diagram in
Figure 2.21.
By referring to this figure, you can convince yourself that OR1 and OR2 signify directions
in which there is one wave velocity. These do not correspond to optic axes. (The optic axes are
located by drawing the tangent planes, A1M1 and A2M2. It is not possible to show in
2-D that these tangent planes touch the 3-D outer surface in circles whose diameters are A1M1
Polarisation of Light 67

Figure 2.21 Enlarged view of wave surfaces shown in Figure 2.20(b).

and A2M2.) Since the cross-section of one surface is a circle, the lines OA1 and OA2 will be
normal to the tangent planes. These therefore give the same normal velocity for both the ellipse
and the circle, so that OA1 and OA2 are the optic axis for the point O. The angle between the
two axes is denoted by 2a. A biaxial crystal is said to be positive, if 2a < 90º and negative if
2a > 90º.
In terms of geometry of the wave surfaces, the angle a is given by

È b2 c Ø
2
D cos 1 É Ù
Ê a 2  c2 Ú

When a ® b, a ® 0. Physically it means that OA1 and OA2 coincide and we get a positive
uniaxial crystal. It means that a uniaxial crystal is a special case of biaxial crystal whose two
optic axes are coincident. When b ® c, a ® 90º and we get a negative crystal.

2.4.3 Selective Absorption

We know that unpolarised light can be considered as made up of two linearly polarised light
waves differing in phase by p/2. Many naturally occurring as well as artificial materials exhibit
the characteristic property of selective absorption of one of these waves and allow the other one
without any significant attenuation. For instance, when unpolarised light is made to pass through
a crystal of tourmaline, it absorbs the o-rays while allowing e-rays to pass through unaffected.
(Another dichroic crystal is quinine sulphide periodide, called herapathite, after Herapath,
an English physician who discovered its polarising properties in 1852. In 1928, Edwin Land,
an American scientist discovered a dichroic material, which is now known as a Polaroid.)
The property of selective absorption is known as dichroism. And the net effect of allowing an
unpolarised light to pass through a dichroic material is to produce linearly polarised light.
A particularly simple dichroic device is the Wire-Grid Polariser. A discussion of this device and
how it polarises light forms the subject matter of discussion of the following few paragraphs.
68 Wave Optics

The Wire-Grid polariser


The working of the wire-grid polariser is very simple. It consists of a grid of a large number
of parallel copper wires, as shown in Figure 2.22. Suppose that unpolarised light is incident
on the grid from the right. The component of the electric vector along the length of the wire
(y-component in Figure 2.22) drives the electrons of each wire and generates electric current,
which appears as Joule heat. The net effect is that energy is transferred from the field to the
wire grid and transmission of y-component is almost blocked. However, since the wires are
assumed to be very thin, the component of the electric field perpendicular to the wires
(x-component) passes through without being attenuated much. Thus, the light emerging from
the wire-grid polariser is linearly polarised with the electric vector along the x-axis.

Figure 2.22 The wire-grid polariser.

For the system to be effective, i.e. for the y-component of the electric vector to be
completely blocked, the spacing between the wires should be less than or equal to the
wavelength of the incident wave. Obviously, fabrication of such a polariser for visible light
(4.0 ´ 10–7 < l < 7.2 ´ 10–7 m) is extremely difficult since the wires have to be placed at
distances of (5 – 7) ´ 10–7 m. Nevertheless, in 1950, Bird and Parrish reportedly put about
30,000 wires in 2.54 cm.
A possible way out of this difficulty lies in the use of a long chain of polymer molecules—
the H-sheet—invented by Land in 1938. It is fabricated in thin sheets of long-chain hydrocarbons
such as polyvinyl chloride. The sheets are stretched while these are being manufactured so that
the long chain molecules align. After the sheet has been impregnated with iodine (by dipping
it in a solution containing iodine), the molecules begin to conduct. (The X-ray diffraction
studies of these dichroic films show that iodine is present as independent long strings.)
The conduction takes place along the hydrocarbon chain because the valence electrons (of the
molecules) can easily move along the chain. These long chain molecules are aligned so that they
are almost parallel to each other and behave like the wires in the wire-grid polariser. Because
of high electrical conductivity, the electrical vector of unpolarised light parallel to the chain is
absorbed while the perpendicular component passes almost unaffected. Since the spacing between
two adjacent molecular chains in a Polaroid is small compared to the optical wavelength,
the Polaroid is usually extremely effective in producing linearly polarised light.
Polarisation of Light 69
Dichroic crystals
Some naturally occurring crystalline materials have anisotropic
structure and exhibit dichroism inherently. The best known
dichroic material is tourmaline, a precious stone used in
jewellery. Tourmalines are boron silicate crystals of differing
chemical composition. The component of electrical field
normal to the principal axis is strongly absorbed by the crystal.
It is observed that a thicker crystal absorbs the electrical field
more completely. A plate cut from a tourmaline crystal parallel
to its optic axis acts as a linear polariser (Figure 2.23).

2.4.4 Polarisation by Scattering

When light is incident in the atmosphere, some of it is Figure 2.23 A tourmaline crystal
polariser.
absorbed by the air molecules (nitrogen and oxygen) and the
dust particles. The absorbed light may be reradiated. This process is called scattering of light.
However, the size of molecules present in the atmosphere must be smaller than the wavelengths
of visible light. According to Lord Rayleigh (1842–1919), a British physicist, the light scattered
by a molecule is inversely proportional to the fourth power of wavelength (1/l4). Since the
wavelength of blue light is less than the wavelength of red light, blue light is scattered towards
the ground more than the red. This explains why the sky appears blue. But the sunset or sunrise
appear reddish because the sunlight passes through a maximum length of atmosphere. In this
process, much of the blue colour is scattered out or removed by the air, dust and smoke
molecules and the light that reaches the earth has blue-deficit.
We can polarise light by filtering it through a polariser. When ordinary light falls on a
polariser, a part of it is absorbed and a part is transmitted, as shown in Figure 2.24. The light
emerging from this polariser is completely polarised along the direction of propagation.

Figure 2.24 Polarisation of light by scattering.


70 Wave Optics

2.4.5 Wave Plates

In the preceding subsection we learnt how in double refraction an incident unpolarised wave
splits into two-rays, each characterized by a certain state of polarisation with their E-vibrations
perpendicular to each other. The vibrations of the electric vector corresponding to the o-ray are
always at right angle to the direction of propagation of the wave as well as the optic axis,
whereas the direction of the vibrations of the electric vector corresponding to the e-ray is always
at right angle to the propagation vector k and lies in the plane containing the optic axis and the
direction of the incident wave. We now consider a class of optical equipment—wave plates—
which can be used to introduce a phase change and hence change in the state of polarisation
of the incident wave. (For this reason, wave plates are also known as phase plates or retarders.)
A wave plate introduces a fixed phase difference of p/2 or p between the polarisation states.
This concept is also used to change the nature of polarisation, say from plane polarisation to
circular or elliptical polarisation. Now we discuss it in some detail.
Refer to Figure 2.25. It shows a calcite wave
plate of thickness d with the optic axis parallel to
the entrance face. Consider a plane wave incident
on the crystal. It splits into o- and e-rays. These
waves propagate with different velocities in the
crystal resulting in a phase difference between
waves emerging from the plate. Since calcite is a
negative uniaxial crystal, mo > me and ve > vo. It
means that after travelling a calcite crystal of
thickness d, the path difference between the o- and
e-rays is given by
D = d(mo – me)
Figure 2.25 A calcite wave plate.
In case of a positive uniaxial crystal like quartz,
mo < me so that the path difference will be D = d(me – mo). Therefore, we can write the general
expression for path difference as d(|mo – me|). Hence, the phase difference between o- and
e-rays is given by
2S 2S
G ' Po  Pe d (2.14)
O O
Note that the o- and e-rays originated in phase when incident light split into components on
entering the crystal.
The state of polarisation of emerging light depends on the phase difference and the amplitudes
of incoming field components. We now consider a few special cases:
(i) When the phase difference d = 2mp, where m is an integer, the relative path difference
will be ml. A birefringent device which introduces a path difference between the two
orthogonal field vibrations in integral multiples of l is called a full-wave plate. It will
not introduce any observable effect on the polarisation of the incident beam and is
therefore not of much practical interest.
Polarisation of Light 71
(ii) When d = (2m + 1)p, the relative path difference will be (2m + 1)l/2. A device which
introduces a path difference between the two orthogonal field vibrations in odd integral
multiples of p/2 is called a half-wave plate. The thickness of a half-wave plate is
given by
(2 m  1)O
d (2.15a)
2 ( Po  Pe )
(iii) When d = (2m + 1)p/2, the path difference will be (2m + 1)l/4. A device which
introduces a path difference between the two orthogonal field vibrations in odd integral
multiples of l/4 is called a quarter-wave plate. The thickness of a quarter-wave plate
is given by
(2 m  1)O
d (2.15b)
4 ( Po  Pe )
So when a linearly polarised wave propagates through a quarter wave plate, the emergent
light will, in general, be elliptical. However, if the plane of vibration of plane polarised
light is inclined at an angle of 45º to the optic axis, the emergent light will be circularly
polarised.
Thus we may conclude that thickness of a birefringent device determines the state of the
emergent light beam.
For a calcite crystal, mo = 1.65836 and me = 1.48641. Therefore, the thickness of the crystal
required to introduce a phase difference of p/2 for sodium light of wavelength 589.3 nm will
be given by
1 O 589.3
d 856.8 nm
4 Po  Pe 4 – 0.17195
Thus a calcite crystal of thickness 856.8 nm will introduce a phase difference of p/2 between
the o- and e-rays.
To get a feel of the numbers, you should answer the following Practice Exercise.

Practice Exercise 2.5 A half-wave plate is rotated between two crossed Polaroids. Light
from a lamp is incident on the first Polaroid. Discuss how the intensity of emergent beam will
vary as the half-wave plate is rotated. What will happen if the half-wave plate is replaced by
a quarter-wave plate?

2.4.6 Babinet Compensator

We now know that a wave plate introduces a fixed path difference between the ordinary and
extraordinary rays. Moreover, these can be used for light of fixed wavelength only. It means that
for different wavelengths, we will have to use different wave plates and such an arrangement
can be quite inconvenient. You may logically ask: Is there any way out of this situation?
A possible solution was provided by Babinet, who designed a crystal plate of variable thickness.
The faces of such a plate are cut parallel to the optic axis. It is called Babinet compensator and
can produce a desired path difference.
72 Wave Optics

A Babinet compensator consists of two wedge-shaped right-angled quartz prisms ABD and
A¢CD¢ which are cut at a very small angle and one of them can be translated precisely relative
to the other by a micrometer. These are placed together with their hypotenuse AD in contact with
each other so that they together form a rectangular slab, as shown in Figure 2.26(a).
The optic axes of ABD is on the face and parallel to the refracting edge, whereas the optic axis
of A¢CD¢ is on the face but perpendicular to the refracting edge. It means that the axes of the two
prisms are at right angles to each other so that the ordinary ray in one will behave as extraordinary
ray while propagating in the other and vice versa. The thicknesses of the prisms are very small
and the separation between the ordinary and extraordinary rays is usually negligibly small.

Figure 2.26 A Babinet compensator (a) Schematic representation, and


(b) Field of view between two adjacent fringes.

When plane polarised light is made to fall normally on the face AB of the compensator so
that its plane of vibration is at an arbitrary angle q to the optic axis, it splits into o- and e-rays.
The e-component, parallel to the optic axis in the first crystal, travels slower than the
o-component, which is perpendicular to the axis of ABD. (This is because the prism is made
of quartz.) The path difference between o- and e-rays when they have traversed thickness t1 of
the prism ABD is given by
d1 = (me – mo)t1
When e-component enters the prism A¢CD¢, it begins to behave as o-component and vice versa.
That is, these components exchange velocities in passing from one prism to the other.
This change occurs because e-component is normal to the optic axis in the second prism.
The path difference introduced by the prism A¢CD¢ as these rays traverse can be expressed as
d2 = (mo – me)t2
Hence, the total path difference between the o- and e-rays as these propagate through the
compensator is obtained by adding the expressions for d1 and d2:
d = (me – mo) (t1 – t2) (2.16a)
That is, these prisms tend to mutually annihilate the effect of each other. Therefore, to achieve
desired result, these are mounted such that prism ABD is fixed and prism A¢CD¢ can slide along
its surface with the help of a rack and pinion arrangement. In this way, Babinet compensator
Polarisation of Light 73
facilitates introduction of desired value of (t1 – t2) and hence a given path difference can be
introduced for light of any wavelength.
The phase difference between the o- and e-rays as these propagate through the compensator
can be written as
È 2S Ø
I É Ù ( Pe  Po ) (t1  t2 ) (2.16b)
Ê O Ú

Note that the difference in the thicknesses (t1 – t2) is variable along the length of the compensator,
and it helps to generate variable phase difference. At the centre of the compensator, t1 = t2 so
that the resultant phase difference will be zero and the light emerging from the centre will
remain plane polarised and the direction of vibration will be parallel to that of the incident light.
When we move away from the centre through a distance l/2, the phase difference will be p.
It means that if light emerging from the centre of the compensator and from a distance of
l/2 from the centre are superposed, the resultant will be a linear vibration whose direction is
rotated at an angle 2q from the incident light.
If a beam of linearly polarised light is incident on the compensator such that the direction
of vibration makes an angle of 45º with the axes, then after traversing a distance a from the
centre of the compensator (along its length), we will obtain a linear vibration whose direction
makes an angle of 90º with the incident plane polarised light. The path difference between the
two emergent rectangular vibrations after a distance x from the centre is given by lx/2a. Hence
Ox
( Pe  Po ) (t1  t2 ) (2.17)
2a
Suppose that light emerging from the central region of the compensator is fed into an analysing
Nicol prism fitted with an eyepiece and a fine cross-wire. This result shows that if the central
region is dark, the field would again be dark when the Nicol has shifted by multiple of 2a from
the centre, provided q = 45º. It means that we will get alternate dark and bright lines as the
eyepiece with the Nicol is made to move parallel to CD in the field of view. Various stages of
polarisation between two adjacent dark fringes are shown in Figure 2.26(b).
The major disadvantage of a Babinet compensator is that changeable thickness is confined
to a very small field of view. This was overcome by Jamin when he envisaged a more accurate
arrangement, wherein one of the prisms was made to shift by a micrometer screw, instead of
the eyepiece with the Nicol. As you will realise, this arrangement is more sensitive because to
make two consecutive dark fringes coincide with cross-wire, the shift of the prism would be
double the shift of the eyepiece with the Nicol. If the prism is shifted though 2b to make two
consecutive dark bands coincide with cross-wire of fixed eyepiece then the path difference
between the two emergent rectangular components would be l. Hence, the path difference
introduced for a shift of the prism would be lx/2b. Hence
Ox (2.18)
( Pe  Po ) (t1  t2 )
2b
A Babinet compensator can be used to determine the wavelength of monochromatic light,
produce elliptically and circularly polarised light and analyse an elliptically polarised light.
We will not go into these details here. However, we wish you to solve a Practice Exercise.
74 Wave Optics

Practice Exercise 2.6 The prism ACD in Figure 2.26 is shifted through a distance b/2 from
its position of coincidence with the prism ABD. Discuss the nature of emerging light when the
incident plane polarised light makes an angle of
(i) 45º and
(ii) less than 45º.

2.4.7 Optical Activity

We know that no light is transmitted through an analyser whose transmission axis is at right
angles to a polariser and the field of view of the analyser would be black. You may recall that
this happens because the light emerging from the polariser has vibrations parallel to the principal
plane of the polariser but the principal plane of the analyser is perpendicular to it. Suppose we
place a sugar solution or a quartz crystal cut perpendicular to its axis between the polariser and
the analyzer. Will the field of view of the analyser continue to be dark? No, some light does
get transmitted and the field of the analyser becomes bright. It means that the direction of
vibration of the emergent light has changed with respect to the incident light. This change arises
because the sugar solution/quartz rotates the plane of polarization of light entering it. If the
analyser is now rotated by a certain angle either towards the left or the right, the field of view
will again become dark. This angle of rotation of the analyser gives a measure of the rotation
of the plane of polarisation by the sugar solution/quartz crystal. The phenomenon of rotation of
the plane of polarization is called optical activity and materials showing this phenomenon are
said to be optically active.
This phenomenon was first studied by Argo and Biot. It is exhibited by organic compounds
like sugar solution, tartaric acid, turpentine, and cinnabar as well as crystals such as quartz and
sodium chloride. A fascinating characteristic with interesting applications in biology and
biochemistry is that some optically active substances rotate the plane of polarization clockwise
and some anticlockwise. These are respectively referred to as dextrorotatory or right-handed
and laevorotatory or left-handed. (It means that when an observer looks towards the light
travelling towards him, the plane of vibration is rotated to the right or the left.) It is now well
established that some crystals of quartz are dextrorotatory while others are laevorotatory.
The angle of rotation of the plane of polarisation (q) produced by an optically active
substance is directly proportional to the distance (l) travelled by light within the material, the
wavelength of the incident light and the temperature of the substance. For example, the angle
of rotation is about 21º for every millimetre travelled by yellow light (l = 600 nm) in quartz.
In a solution, the rotation of the plane of polarisation is proportional to the concentration
of the optically active substance such as sugar in water, camphor in alcohol, etc., thickness of
the medium and wavelength of light used. Mathematically, we can write
q = S (t, l)lC (2.19)
where constant of proportionality S (t, l) is called the specific rotation at temperature T
and wavelength l, q is the angle of rotation, l is the length of the column of solution (in cm)
through which the plane polarised light passes and C is concentration of the active substance
Polarisation of Light 75
(in gcm–3) in the solution. In your physics laboratory, you will get an opportunity to study the
concentration variation of specific rotation. (The product of specific rotation and the molecular
weight of the substance is called molecular rotation. So molecular rotation is the rotation
produced by a solution of one decimetre in length containing 1 g-molecule of the active substance
per cubic centimetre of the solution.
Biot’s relation between the rotation of the plane of polarisation and wavelength of light used
resembles Cauchy’s formula for normal dispersion:
B (2.20)
T A
O2
This relation implies that violet light will be rotated almost four times more than the red light.
It means that when white light is made to pass through a polariser-analyser arrangement interposed
with an optically active substance such as quartz, we will obtain a coloured image on the screen.
This phenomenon is known as rotatory dispersion.

2.5 PRACTICAL APPLICATIONS OF POLARISATION

The phenomenon of polarisation finds several practical applications. These include stress analysis,
measuring solution concentrations and liquid crystal displays (LCDs). We will now discuss
these in brief.

(a) Stress analysis


In 1816, Brewster discovered that certain materials that were normally not optically active
become so if subjected to stress. The degree to which the substance becomes optically active
is directly proportional to stress. When a piece of plastic under stress is placed in between two
polarisers at right angles, a complicated pattern is seen and examination of the pattern reveals
information about stress variation in the material. The coloured patterns on the windshield of
a car arise if the glass has not been fitted properly, i.e., it is under stress.

(b) Measuring solution concentrations


The angle through which plane of polarisation is rotated by an optically active substance is
determined with the help of Laurent’s half-shade polarimeter. The experimental arrangement is
shown in Figure 2.27. It consists of two Nicol prisms N1 and N2. Of these, N1 is a polariser and

Figure 2.27 Laurent’s half-shade polarimeter used to study concentration variation of the angle of
rotation of plane of polarisation.
76 Wave Optics

N2 is an analyser. A half-shade device comprising quartz half-wave plate Q and glass plate G
is placed behind N1. The half-wave plate Q covers one-half of the field of view. The glass plate
G absorbs the same amount of light as the quartz half-wave plate Q. The solution containing
an optically active substance is filled in a glass tube T, which bulges at the middle. While
performing the measurement, you must make sure that there is no air bubble inside the tube.
Light from a monochromatic source such as sodium
lamp is made to fall on the Nicol prism N1 by a converging
lens L. After passing through N1, the light is plane polarised.
One-half of this beam passes through the quartz plate Q
and the other half passes through the glass plate G. Suppose
that the plane of vibration of the plane polarised light
emerging from Nicol prism N1 and incident on the half-
shade plate is along AB, as shown in Figure 2.28, and
makes an angle q with the vertical axis y1 y2. The beam
passing through the quartz plate Q breaks up into o- and
e-components. Recall that these components will emerge
from the half-wave plate with a phase difference of p or a
Figure 2.28 Change in orientation
path difference of l/2. The vibrations of the beam emerging of vibrations emerging
from the quartz plate will be along CD, whereas the from the quartz plate.
vibrations of the beam traversing the glass plate will be
along AB. If the principal plane of analyser N2 is along y1 y2, i.e. along the direction which
bisects ÐAOC, the amplitudes of light incident on the analyser from both halves will be equal.
As a result, the entire field of view will be uniformly illuminated.
If the analyser N2 is rotated to the right of y1 y2, the right half will be brighter than the left
half and vice versa. Therefore, to study concentration variation of specific rotation of an optically
active substance, we first set the analyser N2 in the position which corresponds to uniform
illumination of the field of view without any solution in the glass tube T. Next, we fill the tube
with a solution of known concentration. We observe that the vibrations of the beam emerging
from the quartz and glass plates will be rotated and the field of view will not be uniformly
illuminated. (For sugar solution, the rotation will be in the clockwise direction.) However,
by rotating the analyser in the clockwise direction, we can again make the field of view position
equally bright. The difference between these two positions of the analyser defines the angle
through which the plane of vibration of the incident beam rotates for a given concentration.
We can repeat this process for different concentrations and plot a graph between angle of
rotation and concentration. The graph should be a straight line, whose slope gives Dq/DC.
By using this in Eq. (2.19), we can easily calculate the value of specific rotation. For sugar
solution, the value of specific rotation is about 66º.
(c) Liquid crystal displays
A more modern application of polarisation is in liquid crystal displays (LCDs). These can be
seen on calculators, watches and thin flat computer and TV screens.
An LCD consists of a surface of tiny rectangles called pixels. Each pixel has liquid crystals
in between two glass plates. The liquid crystals are relatively long, thin molecules that attract
each other rather weakly. The first glass plate has very thin (of the order of 1 nm) slits so
Polarisation of Light 77
that the long, rod-like molecules align themselves with the slits. The other glass plate with
similar slits is rotated by 90º with respect to the first. Thus, if molecules next to the first glass
plate are vertical, those in contact with the other plate will be horizontal. The in-between molecules
slowly change their orientation from vertical to horizontal due to intermolecular forces.
When a polariser with its axis vertical is placed in front of the top glass plate, the transmitted
light is vertically polarised. As light propagates from molecule to molecule, its plane of polarisation
changes so as to align with the orientation of the molecules. By the time the light reaches the
back plate, the plane of polarisation changes by 90º. When a second polariser is placed behind
the back plate with an axis of transmission at 90º with respect to the first polariser, the light
will pass as such and the pixel will be bright. But if a potential difference is applied between
the two glass plates, the molecules tend to align their long axes along the electric field.
Then light reaching the back polariser will be vertically polarised and fail to pass through.
The pixel will then be dark.
In actual practice, voltage is applied to certain pixels so that these appear black against the
bright background of those pixels where no voltage has been applied. The background is made
brighter by placing a mirror there to reflect the light that passed through the lower polariser.
Colour is introduced in the LCDs by using green, blue and red filters on sub-pixels.
Let us now summarise what you have studied in this chapter.

2.6 SUMMARY

• When electric field vectors of a light source are randomly oriented, the light is said to be
unpolarised.
• If the field vectors exhibit partial preferential orientation, light is said to be partially polarised.
• A wave is said to be plane or linearly polarised if the associated electric field vector always
lies in the same plane as the propagating wave. In fact, in plane polarised light, the electric
vector oscillates along a straight line in a plane perpendicular to the direction of propagation.
• We can polarise transverse waves only; longitudinal waves do not exhibit polarisation as
vibrations are along the line of transmission. That is why polarisation is the surest test
of transverse nature of light and its electromagnetic origin.
• A linearly polarised wave moving along z-axis is represented as
E(z, t) = ẑ E0 cos(kz – wt)
Here ẑ is unit vector along z-axis.
• Superposition of two linearly polarised light waves having different amplitudes and a
finite phase difference results in the production of an elliptically plane polarised wave.
• Superposition of two oppositely polarised circular light waves (of equal amplitudes) gives
rise to a linearly polarised light wave.
• Light can be polarised by reflection from a surface, by refraction by birefringence or
double refraction through a crystal, by selective absorption in a crystal and scattering by
medium particles.
• When angle of incidence is equal to the Brewster’s angle, the reflected light is completely
polarised and the transmitted light is partially polarised. When reflected light is 100%
polarised, the angle between the reflected and the refracted rays is 90º. The tangent of
78 Wave Optics

Brewster’s angle is equal to the ratio of the refractive indices of the media at whose
interface incident light is reflected.
• In terms of the electric field vibrations, we can say that the reflected beam has no
field vibration parallel to the plane of incidence, i.e., vibrations of the electric field in
the reflected beam are normal to the plane of the paper. But the refracted beam has both
field components.
• When an unpolarised light beam is made to fall on an anisotropic substance like calcite,
it splits into two beams. This phenomenon is known as double refraction or birefringence.
• In double refraction, the ordinary and extraordinary waves are plane polarised with their
E-vibrations perpendicular to each other. The vibrations of the electric vector corresponding
to the ordinary wave are always at right angle to the direction of propagation of the wave
as well as the optic axis.
• The direction of the vibrations of the electric vector corresponding to the extraordinary
wave is always at right angle to the propagation vector k and lies in the plane containing
the optic axis and the direction of the incident wave.
• When unpolarised light is made to pass through a tourmaline crystal, it absorbs the o-rays
and allows e-rays to pass through unaffected. This property of selective absorption is
known as dichroism. The net effect of allowing an unpolarised light to pass through a
dichroic material is to produce linearly polarised light.
• A device which introduces a path difference between two orthogonal field vibrations in
odd integral multiples of l/4 is called a quarter-wave plate. So when a linearly polarised
wave propagates through a quarter wave plate, the emergent light will, in general, be
elliptical.

REVIEW EXERCISES

1. The x- and y-components of electric field are given by


1
(a) Ex = E0 cosq; Ey E0 cos(S  T )
2
(b) Ex = E0 sin q; Ey = E0 cos q
ÈS Ø È SØ
(c) Ex E0 sin É  IÙ , Ey E0 sin É I  Ù
Ê2 Ú Ê 2Ú
ÈS Ø 1
(d) Ex E0 sin É  IÙ , Ey E0 sin I
Ê4 Ú 2
where q = kz + w t and f = kz – wt. Determine the state of polarisation in each case.
[Ans. (a) Linearly polarised, (b) circularly polarised,
(c) left circularly polarised, (d) left circularly polarised]
2. The electric field at z = 0 associated with right circularly polarised and left circularly
polarised lights propagating in the +z-direction is respectively given by
Exr = E0 coswt; Eyr = E0 sinwt
Polarisation of Light 79
and
Exl = E0 cos(wt – a); Eyl = –E0 sin(wt – a)
The subscripts r and l respectively signify right circularly polarised and left circularly
polarised lights. If these are made to superpose, determine the state of polarisation of the
beam. [Ans. Elliptical]
3. An extraordinary wave propagates through a LiNbO3 crystal. If the wave vector k makes
an angle of 45º with the optic axis, calculate the angle between the polarised vector and
k. Take mo = 2.2967 and me = 2.2082. [Ans. 2.25º]
4. Two crossed Polaroids are placed in the path of an unpolarised beam of intensity I0.
A third Polaroid, whose optic axis is at 45º to either of the Polaroids, is placed in between
the two. Calculate the intensity of the transmitted beam. [Ans. I0/8]
5. A left circularly polarised beam of wavelength 706.5 nm is incident on a calcite quarter
wave plate corresponding to 404.6 nm. Describe the state of polarisation of the emergent
beam. Use the following values of mo and me for these wavelengths:

l (nm) mo me
404.6 1.68134 1.49694
706.5 1.65207 1.48359

[Ans. Left circularly polarised]


6. A half-wave plate (HWP) in introduced between two crossed Polaroids P1 and P2.
The optic axis of the plate makes an angle of 15º with the optic axis of Polaroid P1.
If natural light is made to fall on P1, calculate the intensities after the stages P1, P1 plus
HWP, and P1 plus HWP plus P2. [Ans. I0/2; I0/2; I0/32]
7. Refer to Figure 2.29. It shows two calcite prisms connected together. A beam of unpolarised
light is incident normally from Region 1. Determine the path of the rays in regions
2, 3, and 4. Indicate the direction of vibrations. Given that mo = 1.658 and me = 1.486.
The angle of the prism is 15º.

Figure 2.29 The lines and dots represent the directions of


optic axis in two calcite prisms put together.

[Ans. Polarisation vector normal to the plane of paper will pass


through undeviated; Polarisation vector in the plane
of paper: Ði on second surface = 1.8º, Ðr » 2.7º]
80 Wave Optics

8. A quarter-wave plate (QWP) and a half-wave plate are made of calcite (mo = 1.66584
and me = 1.4864). Calculate (i) their thicknesses for an operating wavelength of 589 nm.
Repeat the calculations for quartz (mo = 1.5442 and me = 1.5574). How will the output of
the state of polarisation change if linearly polarised wave at 45º to the axes of the wave
plate is incident on the QWP? [Ans. (i) 0.8206 µm, 1.6412 µm;
(ii) 11.155 µm, 22.311 µm]
9. A circularly polarised light of wavelength 600 nm propagates along the x-axis of a biaxial
crystal. Calculate the distance after which it will become linearly polarised. Given mx =
1.619, my = 1.620 and m2 = 1.626. [Ans. 25 µm]
10. For a biaxial crystal, mx = 1.56, my = 1.59 and mz = 1.60. Determine the direction along
which a circularly polarised light will propagate so that it does not change its state of
polarisation as it propagates. [Ans. 45º]
Polarisation of Light 81

APPENDIX 2A

Propagation of Plane Waves in Uniaxial Crystals


Huygens’ Construction

In Section 2.3, you learnt that the phenomenon of birefringence of light at the surface of a
crystal can be readily understood in terms of the wave surfaces using Huygens’ construction.
Let us begin by considering a relatively simple situation where a beam of parallel light is
incident normally on the surface of a negative crystal such as calcite.

Normal Incidence

Refer to Figure 2A.1. ABC is an incident plane wavefront. For simplicity, we assume that the
optic axis (shown by broken line) is in the plane of incidence and inclined at an angle to the
crystal surface SS¢. According to Huygens’ Principle, we may choose points anywhere along the
wavefront as new point sources of light. Here A, B and C signify the points at which the wave
is incident on the surface of a calcite crystal. After a short while, the Huygens wavelets entering
the crystal from these points will generate spherical and ellipsoidal surfaces corresponding to
o- and e-waves. (In 2-D, we depict these as a circle and an ellipse, respectively.) The common
tangents to these surfaces are shown as OO¢ and EE¢. These signify secondary wavefronts
corresponding to ordinary and extraordinary waves, respectively.

Figure 2A.1 Huygens’ construction for a plane wave incident normally on a calcite crystal.

Note that
• The wavefronts OO¢ and EE¢ are parallel to the surface of the crystal, i.e. the wavefront
AC remains parallel even after refraction; and
• o- and e-waves travel along different directions.
Recall that vibrations of the o-wave are normal to the principal section, whereas vibrations
of the e-wave occur in the principal section. It means that the e-rays, which connect the origins
of secondary wavefronts with the points of tangency diverge from the o-wave and are not
82 Wave Optics

perpendicular to the wavefront EE¢. Moreover, the velocity of e-wave is greater than the velocity
of o-wave. Note that in constructing Figure 2A.1, we have assumed that the optic axis is in the
plane of the paper. A logical question to ask now is: How will the surfaces shown in Figure 2A.1
get modified if optic axis is not in the plane of the paper? In that case, the tangent to the
ellipsoidal wavelets will make contact at points in front or in back of the plane of the paper.
Before proceeding further, you should solve a Practice Exercise.

Practice Exercise 2A.1 Depict Huygens’ construction when the optic axis is (a) parallel, and
(b) perpendicular to the crystal surface for normal incidence.
So far we have confined our discussion to the case where a plane wave was incident
normally on the surface of a calcite crystal, whose optic axis was inclined at an arbitrary angle
to crystal surface in the plane of the paper. Let us now consider a slightly complex situation of
oblique incidence of a light wave.

Oblique Incidence

Refer to Figure 2A.2. It shows Huygens’ construction for a light beam obliquely incident on the
surface of a calcite crystal, whose optic axis is in the plane of the paper (incidence) and inclined
to the crystal surface, as before. The incident wavefront AB is inclined to the crystal surface SS¢,
as shown. You should expect that the o-wave will generate a spherical surface and the e-wave
will generate an ellipsoidal surface. The radius of the spherical surface is determined by the
condition that the ratio BC/AD equals the refractive index mo of the o-wave.

Figure 2A.2 Huygens’ construction when a plane wave is incident obliquely and the optic axis of a
uniaxial calcite crystal lies in the plane of incidence

The ellipsoidal wavesurface is then drawn tangent to the circle at the point where it intersects
the optic axis. The points D and F as well as refracted wavefronts DC and FC are located by
drawing tangents from C to the circle and the ellipse, respectively.
Note that during the time incident light travels from B to C in air, the o-wave travels
from A to D in the crystal and e-wave travels from A to F. If we denote the velocity of light
in air as c and the velocities of o- and e-waves along AD and AF by vo and ve, respectively, we
can write
Polarisation of Light 83

BC AD AF
c vo ve
so that
vo BC
AD BC
c Po
and
ve BC
AF BC
c Pe
Here mo and me denote the refractive indices for the ordinary and extraordinary waves, respectively.
P ART II
WAVE PHENOMENA
INTERFERENCE AND DIFFRACTION
CH A P T E R 3
TWO BEAM INTERFERENCE BY
DIVISION OF WAVEFRONT

EXPECTED LEARNING OUTCOMES


After reading this chapter, the student will acquire the capability to:
• Discuss how energy is redistributed due to superposition of two light waves.
• Explain the term coherence and its relevance in obtaining interference pattern by division of
wavefront.
• Derive expressions for fringe width and intensity distribution in Young's double slit experiment.
• Describe how interference pattern changes when a thin sheet of mica is introduced in the path of
one of the interfering waves.
• Highlight the advantages offered by Fresnel two-mirror arrangement.
• Differentiate between Biprism and Lloyd's mirror fringes.
• Solve numerical problems.

3.1 INTRODUCTION

We now know that superposition of waves in air (sound waves) or on strings leads to redistribution
of energy, which reappears in the form of beats and stationary waves, respectively. The latter
of these phenomena is responsible for adding quality to life through music. We hope you
will agree that life would have been very stressed and devoid of fun without music.
This may tempt you to ask: Do light waves, which are transverse in nature, also exhibit some
similar phenomenon on superposition? To discover answer to this question, you may like to
recall the brilliant colours that we observed in our childhood while playing with soap bubbles.
Similarly, sitting in a vehicle or while walking on foot on a rainy day, you must have observed
shining colours which arise due to oil film spread on water surface on the road. These arise
due to interference of light. But if we illuminate the same part of a screen with two torches,
the region of overlap is uniformly lit without any evidence of redistribution of energy.

87
88 Wave Optics

It suggests that for interference to occur, some specific conditions must be satisfied. We shall
discuss these for completeness at appropriate place.
When we look for the possible conditions for interference to take place, we recall that the
visible range of light extends from 400 nm to 700 nm. It means that interference can take place
only if the dimensions of the equipment used to observe this effect are of the order of the
wavelength of light. You may also recall that a natural source emits light in short trains of
random pulses lasting for about 10–9 s. It means that interference pattern changes every billionth
of a second. But human eye can notice changes which last at least for one-sixteenth of a second.
So we can say that during the period of persistence of vision, interference pattern averages out
leading to a uniform intensity on the screen. That is why some special conditions have to be
created for observing stable interference pattern.
In your elementary physics classes, you have learnt that when two light waves derived
from a point source are made to superpose, the intensity of light is seen to vary on a screen
(i.e. in space). It is characterised by regions of brightness and darkness. This is referred to as
the interference pattern. The methods used to obtain interference pattern are classified as
(i) division of wavefront wherein waves having constant phase difference are derived from
two closely spaced regions of the parent wave and made to superpose,
(ii) division of amplitude wherein a beam is divided at two or more reflecting/transmitting
surfaces and such beams are made to interfere, and
(iii) multiple beam interferometry where multiple beams are superposed.
We will discuss the relative advantages offered by each method as we progress. However,
the emergence of laser sources almost completely changed the situation for better. The observation
and display of interference phenomenon has become particularly simple, notwithstanding some
other complexities.
To given you a feel for the development of the subject of physical optics, we have opted
for classical approach. For this reason, in Section 3.2, we have revisited Young’s double slit
experiment and obtained expressions for fringe-width, intensity distribution in the fringe system
and shape of interference fringes. You will also learn how interference fringe pattern is influenced
when a thin sheet of transparent material such as glass or mica is introduced in the path of one
of the interfering beams. In Section 3.3, we have discussed different methods for obtaining
interference pattern by division of wavefront. In particular, we have discussed the use of Fresnel’s
biprism to study interference in Section 3.3.1. With Fresnel biprism, we obtain symmetrical
interference fringe pattern and the central fringe is bright. Sections 3.3.2 and 3.3.3 are devoted
to other simple devices such as Fresnel’s two-mirror arrangement and Lloyd’s single mirror
experiment used to observe interference fringes. With Lloyd’s mirror, only a few fringes are
visible on one side of the central dark fringe. You will learn about the unique advantages offered
by each of these methods.

3.2 YOUNG’S DOUBLE SLIT EXPERIMENT REVISITED

You may recall that to observe stationary interference pattern due to light waves, we need a pair
of coherent sources, which give out waves having the same frequency and bear a fixed phase
Two Beam Interference by Division of Wavefront 89
relationship in time between them. Moreover, the waves must travel in the same direction,
have equal amplitude and be in the same state of polarisation for maximum contrast. It is quite
easy to have sources emitting waves of the same frequency. For example, we may use two
sodium lamps or mercury lamps. But if we closely examine the mechanism of light emission
by a source of light, we realise that the phase relationship between the waves may vary with
time. This is because the basic mechanism for emission of light waves involves a large number
of independent atoms and each atom emits light for about 10–9 s. It means that waves are
emitted randomly by different atoms and they bear no definite relationship with each other.
Hence, coherence persists only over one-billionth of a second, implying that interference
pattern will vary every billionth of a second. Since persistence of vision of human eye is (1/
16)th of a second, an unaided eye cannot take note of these changes and we fail to observe
a stationary interference pattern. The screen exhibits uniform intensity. (However, we can
record an interference pattern on the film of a camera whose shutter opening time is less than
10–9 s.) From this discussion, it is clear that light waves from two independent sources do not
have a fixed relationship and ordinarily we fail to observe a stationary interference pattern.
In 1801, Thomas Young devised an ingenious but simple experiment to lock the phase
relationship between two independent sources. He divided a single wavefront in two wavefronts
and these behaved as if they emanated from two sources having a fixed phase relationship.
When these waves were allowed to superpose, a stationary interference pattern was obtained.
In the actual experiment, Young allowed sunlight to fall on a pinhole S punched in a screen (A),
as shown in Figure 3.1. The light emerging from this pinhole spread out and was allowed to
fall on another screen (B), which had two pinholes S1 and S2 punched close to one another.
Moreover, these pinholes were equidistant from S so that waves reached there in the same
phase. As a result, these pinholes behaved as coherent sources. The spherical waves emanating
from these pinholes overlapped as these propagated in space and generated a beautiful symmetrical
interference pattern. The intensity of the pattern varied and consisted of alternate bright and dark
fringes on the observation screen C. When either S1 or S2 was covered, the interference fringes
disappeared.

Figure 3.1 Young’s double slit arrangement to observe interference pattern.

Usually, when this experiment is performed in a physics laboratory, we replace the pinholes
by narrow slits and a monochromatic source such as sodium lamp is used. Does the nature of
waves coming out of the slits differ from that emanating from the pinholes? If you think that
90 Wave Optics

the waves superposing in this case are cylindrical, you are on the right track. We can also
perform this experiment in physics laboratory by using small photographic plates to create the
slits. The spacing between double slits should be kept about 0.2 mm.
Now refer to Figure 3.2. It shows the sections of the wavefronts on the plane containing
S, S1 and S2. Young explained the appearance of interference pattern by considering the principle
of superposition of waves. This simple experiment provided much needed boost to wave theory
of light. We will now discuss the interference pattern and relate distance between the fringes
to wavelength of light. In fact, to support his experiment, Young calculated the wavelength of
light on the basis of the measurement of fringe-width, distance between the slits and distance
of the observation screen C from the source screen B.

Figure 3.2 Sections of the spherical wavefronts emanating from S, S1 and S2.

3.2.1 Interference Pattern

To analyse the interference pattern and investigate the spacing of the interference fringes, refer
to Figure 3.3. S is a narrow slit illuminated by monochromatic light. Let S1 and S2 be two
extremely narrow parallel slits placed close to each other (and represent the two pinholes of

Figure 3.3 Schematics of Young’s interference experiment.


Two Beam Interference by Division of Wavefront 91
Young’s interference experiment). The waves emanating from S1 and S2 propagate as if they
start from there in the same phase because these slits are equidistant from S. Moreover, since
waves emanating from S1 and S2 essentially arise by division of (same) wavefront from the same
source, we can safely assume that their amplitudes are practically equal and they are in same
state of polarisation.
To calculate the intensity at a point P on the screen, we join S1P and S2P. Suppose that the
distance between S1 and S2 is d and the distance between the slits in the source plane and
observation plane is D. Similarly, we denote the distance between P0, the foot of the perpendicular
from the point S on the observation screen, and the observation point P as y. Note that waves
originating from S1 and S2 reach P with path difference D = S2P – S1P = S2A. It means that
A and S1 are equidistant from the point of observation P.
We now express D in terms of physically measurable quantities like d, D and y. Note that
we can write
(S2P)2 – (S1P)2 = [(S2B)2 + (PB)2] – [(S1C)2 + (PC)2]
Ë 2Û Ë 2Û
È dØ È dØ
= ÌD2 Éy  Ù
2
Ü  ÌD  É y  Ù Ü
Í Ê 2Ú Ý Í Ê 2Ú Ý
2 2
È dØ È dØ
= Éy  Ù Éy Ù 2 yd [Œ (a + b)2 – (a – b)2 = 4ab]
Ê 2Ú Ê 2Ú
On simplification, we get
2yd
S2 P  S1 P (3.1)
S2 P  S1 P
In a typical interference experiment, y and d are several orders of magnitude less than D. For
example, if d = 0.02 cm, D = 100 cm and PP0 = 0.5 cm, we get
S2P + S1P = [1002 + (0.51)2]1/2 + [1002 + (0.49)2]1/2
= 200.005 cm
It shows that if we replace S2P + S1P by 2D, no significant error will be introduced. Therefore,
we can rewrite Eq. (3.1) as
yd
S2P – S1P =
D
or
D'
y= (3.2)
d
The intensity at point P will be maximum or minimum depending on whether the path difference
D is an integral multiple of l or an odd multiple of l/2. Hence for maximum, which corresponds
to bright fringes, we have
nO
yn D n = 0, 1, 2, ... (3.3)
d
The number n defines the order of the fringe. The zeroth order fringe (n = 0) corresponds to
92 Wave Optics

central maximum, which is always bright and occurs around point P0. The first order fringe
corresponds to the first bright fringe (n = 1) on either side of the central maximum and so on.
For minimum, which corresponds to dark fringes, we have

È 1Ø OD
yn Én  Ù n = 0, 1, 2, ... (3.4)
Ê 2Ú d
Equations (3.3) and (3.4) define the positions of bright and dark fringes. It means that when two
waves of the same amplitude and frequency arrive at a point in different phases, they either
reinforce or annihilate the effect of one another depending on whether the path difference is an
integral multiple of wavelength or odd multiple of half-wavelength. Note that the redistribution
of energy arises essentially due to superposition of (two or more) waves. It is worthwhile to
point out here that whenever we wish to count the number of fringes we have to use a magnifier
or eyepiece because they are so fine.
Before proceeding further, study the following example carefully.

EXAMPLE 3.1 In a typical interference experiment, light from a monochromatic lamp is


incident on two narrow slits, which are 2 ´ 10–4 m apart. The fourth bright fringe of the
interference pattern is observed at a distance of 5 ´ 10–3 m from the central maximum on a
screen placed at a distance of 0.50 m from the plane containing the slits. Calculate the wavelength
of light used.
Solution From Eq. (3.3) we recall that the distance of a bright fringe from the central maximum
is given by
nO
yn D n = 0, 1, 2, ...
d
On rearrangement, we can write
yn d
O
nD
For n = 4, we are told that y4 = 5 ´ 10 m, d = 2 ´ 10–4 m and D = 0.50 m. Hence
–3

(5 – 10 3 m) – (2 – 104 m)
l=
4 – (0.50 m)
= 5 ´ 10–7 m = 500 nm

Fringe-width
If we denote the distances of the nth and (n + 1)th bright fringes by yn and yn+1, respectively,
the spacing between these fringes defines the fringe-width. Then, we can write
( n  1) O D nO D OD
yn 1  yn œ E  (3.5)
d d d
Note that b is independent of n. It means that spacing between any two consecutive bright
fringes is constant.
Two Beam Interference by Division of Wavefront 93
Starting from Eq. (3.4), you can easily convince yourself that, as in the case of bright
fringes, any two consecutive dark fringes are also equidistant. So we can say that bright and
dark fringes are equally spaced.
Proceeding further, we note that fringe-width is directly proportional to the wavelength of
light used, l as well as the distance between the plane containing the coherent sources and the
observation screen and inversely proportional to the distance between the slits, d.
You may now like to solve a Practice Exercise.

Practice Exercise 3.1 In the experiment discussed in Example 3.1, the lamp is changed to
emit light of wavelength 600 nm. The central maximum and twentieth order maximum occur
at 2 ´ 10–3 m and 6 ´ 10–3 m.
(i) Calculate the fringe-width.
(ii) If the source of light is again changed to 400 nm, determine the new positions of these
maxima. [Ans. (i) 2 ´ 10–4 m;
(ii) 2 ´ 10–3 m, 5.334 ´ 10–3 m]

Fringe shape
Refer to Figure 3.3 again. Here S1 and S2 behave as coherent sources at point P, which may be
anywhere along the observation plane, determines the position of maximum (bright fringe) or
minimum (dark fringe) depending on the path difference between the interfering waves reaching
there. We are now interested in determining the shape of the interference fringes.
So, for a given value of n, we have to determine the locus of the point P, subject to the condition
D = S2P – S1P
The point P may correspond to a minimum for D = (n + 1/2)l and maximum for D = nl.
We choose the mid-point of S1S2 as the origin. Since the distance between S1 and S2 is d,
the coordinates of the source points are respectively (–d/2, 0) and (d/2, 0). If the coordinates
of point P are (x, y), we can write
2 1/ 2
ËÈ dØ Û
2
S2 P ÌÉ y  Ù  x Ü
ÍÊ 2Ú Ý
and
2 1/2
ËÈ dØ Û
2
S1 P ÌÉ y  Ù  x Ü
ÍÊ 2Ú Ý
Hence,
2 1/2 2 1/ 2
ËÈ dØ Û ËÈ dØ Û
2 2
' S 2 P  S1 P ÌÉ y  Ù  x Ü  ÌÉ y  Ù  x Ü
ÍÊ 2Ú Ý ÍÊ 2Ú Ý
To simplify this expression, we rearrange terms as

2 1/ 2 2 1/ 2
ËÈ dØ Û ËÈ dØ Û
2 2
ÌÉ y  Ù  x Ü '  ÌÉ y  Ù  x Ü
ÍÊ 2Ú Ý ÍÊ 2Ú Ý
94 Wave Optics

On squaring both sides, we get


2 Ë 2Û Ë 2 Û1/ 2
2 È dØ 2 2 È dØ 2 È dØ
x Éy  Ù '  Ìx  É y  Ù Ü  2' Ì x  É y  Ù Ü
Ê 2Ú Í Ê 2Ú Ý Í Ê 2Ú Ý
We can rewrite it as
2 2 Ë 2 Û1/2
È dØ È dØ 2 2 È dØ
Éy  Ù Éy  Ù '  2' Ì x  É y  Ù Ü
Ê 2Ú Ê 2Ú Í Ê 2Ú Ý
Using the algebraic relation (a + b)2 – (a – b)2 = 4ab and rearranging terms, we can rewrite
this expression as
2 1/ 2
dØ Û ËÈ
2 2
2 yd  ' Ù x Ü 2' Ì É y 
2Ú Ý ÍÊ
On squaring both sides again and simplifying the resultant expression, we get
d2 2Ø È
4 y2d 2  4 y '2 d '
4
4' 2 É y 2
x Ù  yd 
Ê 4 Ú
We now collect coefficients of x and y and divide the resultant expression by D2(d 2 – D2). This
2 2

gives
y2 x2
4 2 4 2 1
' (d  '2 )
We can rewrite it as
y2 x2
 1 (3.6)
1 2 1 2
' (d  ' 2 )
4 4
Do you recognise this equation? It represents a
hyperbola with both S1 and S2 as foci. This is shown in
Figure 3.4. So we may now conclude that the inter-
ference pattern in space will be a hyperboloid of
revolution, which is obtained by revolving the hyperbola
about the line S1S2. However, note that interference
fringes are observed on a screen in a plane perpendicular
to the plane of the paper (figure) and parallel to the
line joining S1S2. It means that interference fringes on
the observation screen are hyperbolae. But on a screen
we observe fringes more or less as straight lines. This
is because wavelength of light is of the order of 600 Figure 3.4 Interference fringes are
nm and the path difference D is also to be of this order. hyperboloids in space.

Intensity distribution
To determine the intensity distribution in the fringe system, let us assume that E1 and E2 are the
electric fields produced at point P by the coherent sources S1 and S2, respectively. In general,
Two Beam Interference by Division of Wavefront 95
these fields will have different directions and magnitudes. However, if S1P and S2P are very
large in comparison to d(=S1S2), the directions of these fields will be almost the same. So, we
can write
È 2S Ø
E1 = ˆi E01 cos É S1 P  Z t Ù
Ê O Ú
and
È 2S Ø
E2 = ˆi E02 cos É S2 P  Z t Ù (3.7)
Ê O Ú

where î is the unit vector along the direction of either of the electric fields. The resultant field
at P is obtained by their superposition:
E = E1 + E2
Ë È 2S Ø È 2S ØÛ
= ˆi Ì E01 cos É S1 P  Z t Ù  E02 cos É S2 P  Zt Ù Ü (3.8)
Í Ê O Ú Ê O ÚÝ

Since intensity is proportional to the square of the electric field, we can write

Î 2 2 È 2S Ø 2 2 È 2S Ø Þ
ÑÑ E01 cos ÉÊ O S1 P  Z t ÙÚ  E02 cos ÉÊ O S2 P  Z t ÙÚ  ÑÑ
I KÏ ß
Ñ 2 E01 E02 ËÌ cos ÈÉ 2S S1 P  Z t ØÙ cos ÈÉ 2S S2 P  Z t ØÙ ÛÜ Ñ
ÑÐ Í Ê O Ú Ê O Ú Ý Ñà

Here K is constant of proportionality. In free space, it is equal to e 0c2, where e 0 is permittivity


of space and c is speed of light in vacuum.
Using the trigonometric identity cos(A + B) + cos(A – B) = 2 cosA cosB, we can rewrite
the expression for intensity as
Î 2 È 2S Ø È 2S Ø Þ
Ñ E01 cos 2 É S1 P  Z t Ù  E02 2
cos 2 É S2 P  Z t Ù  Ñ
Ñ Ê O Ú Ê O Ú Ñ
I KÏ ß (3.9)
Î
Ñ E01 E02 Ïcos Ë 2S Û Ë 2S Û ÞÑ
( S P  S P )  cos 2Z t  ( S P  S P ) ß
Ð ÌÍ O
2 1 Ü ÌÍ 1 Ü
Ñ
Ð Ý O 2 Ñ
Ý àà
15 –1
For light, the frequency is very large (w » 10 s ) and all terms containing wt will vary very
rapidly. Therefore, it is customary to take the average value over one cycle. From the basic
knowledge of trigonometry, we know that the average value of cos2q is one-half. Moreover,
the last term in Eq. (3.9) will vary between +1 and –1. Therefore, its average value will be zero.
Hence, we can write
È1 2 1 2 Ø
I = K É E01  E02  E01 E02 cos G Ù
Ê2 2 Ú

= I1  I 2  2 I1 I 2 cos G (3.10)
2
where I1 = (1/2) K E01 is the intensity produced by the source S1 if no light from S2 is allowed
2
to fall on the screen. Similarly, I2 = (1/2) K E02 is the intensity produced by the source S2 if no
96 Wave Optics

light from S1 is allowed to fall on the screen. Further, d = 2p (S2P – S1P)/l represents the phase
difference between the displacements reaching the point P from S1 and S2.
Do you recognise Eq. (3.10)? It is identical to the one that is obtained while studying
superposition of oscillations.
From Eq. (3.10) we conclude that
• The maximum value of intensity is
2
I max I1  I 2 (3.11a)

The maximum value occurs at d = ±2np; n = 0, 1, 2, ... or S2P – S1P = nl. In this case,
the waves differ in phase by an integral multiple of 2p and are said to be in-phase.
It is referred to as constructive interference.
• The minimum value of intensity is
2
I min I1  I 2 (3.11b)

and it occurs at d = ±(2n + 1)p ; n = 0, 1, 2, ... or S2P – S1P = (n + 1/2)l. In this case,
the waves differ in phase by odd multiples of 180º and are said to be out of phase.
Physically it means that troughs of one wave overlap crests of another. It is referred to
as destructive interference.
• If I1 = I2 = I0, say, then Eq. (3.10) takes the form
G (3.12)
I 2 I 0  2 I 0 cos G 4 I 0 cos2
2
For d = 0, 2p, 4p, ..., the intensity at a given point will be four times the intensity of
either beam. But if d = p, 3p, 5p, ..., the intensity distribution at corresponding points
will be zero. At in-between points, the intensity distribution exhibits cos2f pattern,
as shown in Figure 3.5.

Figure 3.5 Intensity distribution as a function of phase difference d between


the displacements reaching a point from two coherent sources.

From this discussion we can say that when two waves reach a point on an observation
screen exactly in opposite phases, the resultant intensity is zero and they are said to interfere
destructively. Now the question arises what happens to the energy of the interfering beams and
Two Beam Interference by Division of Wavefront 97
is the law of conservation of energy violated. The answer to this question is that energy is not
destroyed, though it appears to disappear. But the fact is that it reappears at the maximum,
where intensity is four times the intensity due to individual beams. That is, energy is redistributed
in the interference pattern and there is no violation of the principle of conservation of energy
in the interference phenomenon.
Before proceeding further, you are advised to solve a Practice Exercise.

Practice Exercise 3.2 Starting from Eq. (3.12) show that the average intensity in an
interference pattern is equal to the sum of individual intensities.

In a typical interference experiment with a monochromatic source, the number of interference


fringes obtained is large and it is fairly difficult to locate the position of the central fringe.
So let us see how to locate the position of the central fringe. To discover the answer to this
question, we replace the monochromatic source by a source of white light. It is observed that
the central white fringe is surrounded by a few coloured fringes on both sides. The genesis of
such a pattern can be understood by recalling that a pair of white light coherent sources is
equivalent to a number of pairs of monochromatic sources. Each monochromatic pair produces
its own system of fringes with a fringe width characteristic of the wavelength corresponding to
a particular colour. At the centre of the pattern, the interfering waves arrive in phase and there
is no path difference between them. This is valid for all colours. Hence, the waves corresponding
to different colours combine to produce a bright fringe at the centre and their superposition
leads to a white central fringe.
As we move on either side of the centre, the path difference gradually increases and may
be equal to l/2 first for the violet colour. This defines the position of the first dark fringe of
violet. As we move further, we will observe first minimum for blue, green, yellow and red in
the order. But the inner edge of the first dark fringe receives sufficient intensity due to red
colour. Hence it appears reddish. Similarly, the outer edge of the first dark fringe, which is
minimum for red, receives sufficient intensity due to violet and is therefore violet. The same
applies to all other dark fringes resulting in a few coloured fringes on both sides of the central
white fringe, which can be easily identified.

3.2.2 Displacement of Fringes

Refer to Figure 3.6. In this arrangement for observing interference pattern, a thin transparent
plate of glass or a sheet of mica has been introduced in the path of one of the two interfering
beams. What change we can expect in the interference pattern? You will note that the entire
fringe pattern is displaced through a finite distance determined by the refractive index and
thickness of the transparent sheet. The displacement occurs towards the side on which the
transparent sheet has been introduced. You will discover that this result offers us an opportunity
to determine either the wavelength of light or refractive index of the material of the transparent
sheet, provided we know one of these. This experiment also conclusively established that light
travels faster in air than a material medium and proved an important milestone in the developments
which led to the acceptance of the wave theory of light.
98 Wave Optics

Figure 3.6 Observing the effect of introducing a thin transparent sheet in


the path of one of the interfering waves.

Suppose that thickness of the transparent sheet is t and it has been introduced in the path
S1P. It means that light from S1 travels partly in air and partly in the transparent sheet before
it reaches the observation point P. The light travels a distance (S1P – t) in air and equal to t
inside the sheet. If we denote the speed of light in air by c and in the sheet by v(=c/m), the time
taken (T) by light waves in reaching P from S1 is given by
S1 P  t t S1 P  t P t S1 P  ( P  1)t (3.13)
T  
c v c c c
This result shows that the effect of introducing a sheet of refractive index m is to increase the
path traversed by light waves in air from S1P to S1P + (m – 1)t, i.e. by (m – 1)t.
Suppose O denotes the point where optical distances S1O and S2O are equal. It will correspond
to the position of central bright fringe when the transparent plate has not been introduced. As
soon as the transparent plate is introduced, a path difference equal to (m – 1)t arises. So we
should expect a shift in the position of the central fringe, say to Q. It means that at Q, these
optical paths become equal. A similar argument applies to other fringes as well.
Now at the observation point P, the effective path difference is given by
D = S2P – [S1P + (m – 1)t]
= (S2P – S1P) – (m – 1)t
On combining this result with Eq. (3.2), we get
d
' y  (P  1)t (3.14)
D
For point P to be the centre of the nth bright fringe, the effective path difference D should be
equal to nl, i.e.
d
yn  ( P  1)t nO
D
so that
D
yn [ nO  ( P  1)t ] (3.15)
d
Note that when there is no transparent plate (t = 0), Eq. (3.15) reduces to Eq. (3.3). That is,
the expression for the distance of the nth bright fringe from O reduces to (D/d) nl.
Two Beam Interference by Division of Wavefront 99
On combining Eqs. (3.3) and (3.15), we can determine the distance through which the nth bright
fringe is displaced on introduction of a sheet of glass in the path of one of the interfering waves:
D D
z= [ nO  ( P  1)t ]  O
d d
D
= ( P  1)t (3.16)
d
From this result we note that the distance through which a fringe shifts is independent of its
order implying that the shift is the same for all bright fringes. Similarly, you can convince yourself
that this result holds for dark fringes as well. It means that introduction of a transparent sheet of
optically active medium in the path of one of the superposing waves displaces the entire fringe
system through a distance z = D(m – 1)t/d towards the side on which the plate has been introduced.
To give you an idea of the numbers involved, we recall that in a typical experiment,
D = 0.5 m and d = 0.1 cm. If the central fringe shifts by 0.2 cm when a thin mica sheet of
refractive index 1.6 is introduced, the thickness of the sheet is
d] (0.1 cm – 0.2 cm)
t 6.67 – 10 4 cm
D ( P  1) (50 cm) – 0.6
This result also suggests that light travels faster in air than in a material medium of refractive
index greater than one.
It may be mentioned here that Eq. (3.16) is used to determine the thickness of extremely
thin transparent sheets by measuring the shift of the fringe system. As explained earlier, if a
monochromatic source of light is used, it will be difficult to locate the position where the central
fringe shifts after introduction of the transparent sheet. Therefore, we invariably use white light
source which gives white central fringe and coloured fringes around it.
Now you may ask: Does introduction of the sheet also influence the fringe-width in any
way? To seek answer to this question, let us calculate the fringe-width. By considering the nth
and (n + 1)th fringes, we can write from Eq. (3.15)
b = yn+1 – yn
D D
= [(n  1) O  ( P  1)t ]  [ nO  ( P  1)t ]
d d
D
= O (3.17)
d
Note that this expression is identical to Eq. (3.5) indicating that introduction of the plate does
not affect the fringe-width in any way whatsoever.
Now study the following example carefully.

EXAMPLE 3.2 The central fringe in the fringe pattern obtained using a monochromatic
source, which emits light of wavelength 589 nm, is shifted to a position occupied by the third
bright fringe when a thin glass plate of refractive index 1.5 is introduced in the path of one of
the interfering waves. Calculate the thickness of the glass plate.
100 Wave Optics

Solution When n fringes cross the field of view, we rewrite Eq. (3.16) as
d
( P  1)t ] nO
D
Here, m = 1.5, n = 3 and l = 589 nm. Hence, the thickness of the glass plate is given by
nO 3 – (589 nm) 3 – (589.0 nm)
t 3534 nm
P 1 1.5  1 0.5

Now you should answer a Practice Exercise.

Practice Exercise 3.3 A monochromatic source, which emits light of wavelength 589.0 nm,
is used in Young’s double slit experiment. When a 2945 nm thick sheet of glass is introduced
in the path of one of the interfering beams, the central bright fringe shifts through a distance
equal to the width of four fringes. Calculate the refractive index of the sheet of glass.
[Ans. 1.8]

3.3 PRODUCTION OF INTERFERENCE PATTERN


BY DIVISION OF WAVEFRONT

Soon after Young produced interference pattern in his now famous double slit experiment,
objections were raised about the authenticity/origin of the interference pattern. In particular, it
was argued that bright fringes observed by Young were in all probability due to some complicated
modification of light by the edges of the slits rather than interference phenomenon. This attracted
the attention and interest of Fresnel in the subject. On the basis of a series of new experimental
arrangements, Fresnel rediscovered Young’s principle of interference and succeeded in producing
interference pattern by division of wavefront. It may be mentioned here that these investigations
of Fresnel put wave theory of light on a very sound pedestal, beyond any doubt, criticism
or suspicion. In fact, Huygens’ brilliant ideas began to receive due approval and acceptance
of physics fraternity only after the work of Fresnel. One of his arrangements used a biprism.
We have discussed it in some detail in the following section.

3.3.1 Fresnel’s Biprism

Refer to Figure 3.7, which depicts schematics of Fresnel biprism experiment. abc is a biprism,
which was actually a single prism with extremely small base angles (»20¢). The prism is
assumed to stand perpendicular to the plane of the paper. S is a narrow vertical slit, which is
also perpendicular to the plane of the paper. It is illuminated by a monochromatic source of
light. Light from the slit S is made to fall symmetrically on the biprism, placed at a small
distance away from S and having its refracting edges parallel to slit S, i.e. perpendicular to the
plane of the paper. Thus light is refracted by the upper as well as lower halves of the prism.
The light falling on the lower portion of the biprism is bent upward and appears to come from
the virtual source S2. Similarly, light falling on the upper portion of the biprism is bent downwards
Two Beam Interference by Division of Wavefront 101

Figure 3.7 Ray diagram of Fresnel’s biprism experiment.

and appears to come from virtual source S1. It means that biprism produces two virtual images
of S : S1 and S2. These images act as coherent sources.
To give you an idea of the separation between these virtual images produced by the biprism,
we note that if the refractive index of the material of the biprism is m and its base angle is a,
the angular deviation S1a or S2a produced by the biprism will be (m – 1)a. Hence, the total
angular separation between the virtual images is 2(m – 1)a, and separation between the virtual
sources is equal to 2(m – 1)y0a, where y0 denotes the distance between S and the base of the
prism. So, if m = 1.5, a » 20¢ = 5.8 ´ 10–3 rad and y0 = 2 cm, the separation between virtual
sources d = 0.012 cm.
If screens A1 and A2 are placed as shown in Figure 3.7, the superposition of light apparently
emanating from S1 and S2 in the cones qS1s and pS2r produces interference fringes of equal width
in the region of overlap qr on the observation screen A3A4 placed to the right of the biprism.
To observe the fringe pattern, we can replace the observation screen A3A4 by a photographic
plate, an eyepiece or a low power microscope. When a photographic plate is used, we observe
closely spaced straight line fringes of equal width in the centre of the photograph, as shown in
Figure 3.8(a). These arise due to interference of light. Note that there are wide fringes at the
edges of the pattern and arise due to diffraction of light by the vertices and of the biprism and
each of these acts as a straight edge. (We will discuss more about it in Chapter 5.)
What will happen if screens A1 and A2 are removed? Obviously, the light from S1 and S2
will overlap over the entire region ps on the observation screen and you will observe equally
spaced interference fringes superimposed over the diffraction pattern, as shown in Figure 3.8(b).
In this way, Fresnel conclusively demonstrated interference without diffraction beams through
the two slits used in Young’s double slit experiment.
In physics laboratory, we work with Fresnel biprism to determine the wavelength
of light emitted by a monochromatic source such as sodium lamp. A fine vertical slit is
adjusted close to the source of light and the refracting edge of the biprism is set parallel to the
slit. These are mounted on an optical bench. The interference fringes are observed in the focal
plane (cross-wires) of an eyepiece mounted on the optical bench at some distance from the
biprism. The fringe-width b is determined by means of a micrometer attached to it.
Suppose that the distance between the source and eyepiece is D and the distance between
the (virtual) coherent sources is d. The eyepiece is moved horizontally (perpendicular to the
102 Wave Optics

Figure 3.8 Interference and diffraction fringes produced in Fresnel biprism experiment.

length of the optical bench) to determine fringe-width. If the eyepiece moves through a distance
x in crossing n bright fringes, the fringe-width b = x/n. But b = lD/d or l = bd/D = xd/nD.
From this result, it is clear that once we know the distance between the virtual sources, we can
easily determine the wavelength of light using Fresnel biprism arrangement.
We determine d in two ways: either by measuring the refracting angle a or by using the
displacement method, which is widely used to determine the focal length of a convex lens. But
the latter method is more convenient and we will discuss it here in some detail.
Refer to Figure 3.9. It shows a convex lens between the biprism and the eyepiece. It is
placed in such a way that the images of the virtual sources S1 and S2 are clearly seen in the field
of view of the eyepiece. Let us denote this position of the lens as L1. Suppose that the distance
between the images of S1 and S2 as seen in the eyepiece is d1. Then, we can write*
d1 v1
d u1 (3.18)
*We know that
1 1 1

v u f
Since v + u = D, we can write
1 1 1
 =
v Dv f
Þ v2 – Dv + fD = 0
This is quadratic in v and has roots
D “ D 2  4 fD
\ v=
2
Hence sum of the roots v1 + v2 = D. Using this result for adjoint positions, we can write
D = u1 + v1 = u2 + v2
\ u1 = v2 and u2 = v1 (i)
d1 v1 u 2
Hence m1 = (ii)
d u1 u1
d 2 v2 u1
and m2 = (iii)
d u 2 u2
so that m1m2 = 1 Þ d2 = d1d2
Two Beam Interference by Division of Wavefront 103

Figure 3.9 Displacement method to determine the distance between the coherent sources in Fresnel
biprism experiment. L1 and L2 denote the adjoint positions of the lens where the images
of S1 and S2 are seen clearly in the eyepiece.

Next, the lens is moved towards the eyepiece and adjusted so that the images of S1 and S2
can be again seen clearly in the field of view of the eyepiece. Let the distance between the
images of S1 and S2 as seen in the eyepiece in the new position of the lens be d2. Then we can
write
d 2 v2
(3.19)
d u2
On combining Eqs. (3.18) and (3.19), we get

d1d 2
1
d2
or
d d1d 2 (3.20)

Note that d is geometrical mean of d1 and d2. In a typical experiment, d = 0.01 cm,
l = 600 nm, D = 0.50 m and b = 3 ´ 10–3 m.
Now study the following examples carefully.

EXAMPLE 3.3 The inclined faces of a glass biprism (m = 1.5) make angles of 1º with its
base. If the source slit is 10 cm away from the biprism, calculate the separation between the
coherent sources formed by it.
Solution The separation between the coherent sources formed by a biprism is given by

d = 2(m – 1)ay0

Here m = 1.5, y0 = 0.1 m and a = 1º = p/180 rad. Hence

2 – (1.5  1) – (0.1 m) – S
d 1.74 – 103 m
180
104 Wave Optics

EXAMPLE 3.4 In Example 3.3, suppose that the biprism is illuminated by a sodium lamp,
which emits light of wavelength l = 589 nm. Calculate the fringe-width observed at a distance
of 1 m from the biprism.
Solution From Eq. (3.17), we recall that fringe-width is given by

OD
E
d
where d = 2(m – 1)ay0. .
Here m = 1.5, y0 = 0.1 m and a = 1º = p/180 rad, D = 1.1 m and l = 589 nm. Hence

(589 – 10 9 m) – (1.1 m) – 180 – 7


b=
2(1.5  1) – 22 – (0.1 m)
= 0.37 ´ 10–3 m
EXAMPLE 3.5 In Example 3.3, the biprism is illuminated by a sodium lamp (l = 589 nm)
and the eyepiece is at a distance of one metre from the slit. A convex lens inserted between the
biprism and the eyepiece gives clear images of coherent sources in the focal plane of the
eyepiece. If these images are 0.4 cm apart in one case and 0.16 cm apart in the second case,
calculate the width of interference fringes observed on the screen.
Solution From Eq. (3.17), we recall that fringe-width is given by
OD
E
d
where d d1d 2
Here d1 = 0.004 m, d2 = 0.0016 m, D = 1.0 m and l = 589 nm. Hence

d (0.004 m) – (0.0016 m) 2.53 – 10 3 m


and
(589 – 10 9 m) – (1.0 m)
E 2.33 – 104 m
2.53 – 103 m

You should now solve a Practice Exercise.

Practice Exercise 3.4 In a Fresnel biprism experiment, the fringes of width 2 ´ 10–4 m
are observed with a light of wavelength 500 nm at a distance of 0.9 m from the biprism.
The refractive index of the material of the prism is 1.5 and it is placed at a distance of 0.1
m from the slit illuminating the prism. Calculate
(i) the acute angle and
(ii) the angle of the vertex of the biprism. [Ans. 0.01 rad; (ii) (p – 0.02) rad]
Two Beam Interference by Division of Wavefront 105

3.3.2 Fresnel’s Two-mirror Arrangement

Another simple arrangement devised by Fresnel for production of interference pattern comprised
two plane glass mirrors M1 and M2, which were silvered on the front side and inclined to each
other at a small angle q such as at a point M, as shown in Figure 3.10. S is a narrow slit placed
perpendicular to the plane of the paper. A portion of the wavefront from S is reflected from M1M
and another portion is reflected from MM2. These appear to diverge from S1 and S2, respectively.
Thus we can say that Fresnel’s mirrors produce two virtual images, which act as coherent
sources as in the case of biprism. These coherent sources illuminate the regions ps and qr of
the screen. Since the reflected wavefronts have been derived from the same source, these give
rise to interference fringes in the region qr of overlap. If you use this arrangement to study
interference phenomenon in your physics laboratory, you must make sure that no reflection
takes place from the back of the mirrors.

Figure 3.10 Fresnel’s two-mirror arrangement for production of interference pattern.

From simple geometrical considerations, we can show that the points S1, S2 and S lie on a
circle whose centre is at the point M. Further, if the angle between the mirrors is q, the angle
subtended by the coherent sources at the point of intersection is twice the angle between the
mirrors and S1S2 is 2Rq, where R is the radius of the circle.
When Fresnel two-mirror experiment is performed in physics laboratory using an optical
bench, light is reflected from the mirrors at nearly grazing angles. We can demonstrate the
interference pattern by using ordinary glass plates. One of these is fitted with an adjusting screw
for changing angle q and the other with a screw for making the edges of the two mirrors
parallel.

3.3.3 Lloyd’s Mirror Arrangement

Lloyd’s mirror is a very simple device to produce an interference pattern. This experiment was
performed by Lloyd in 1834. He used a 30 cm ´ 8 cm plane mirror. In this arrangement,
monochromatic light from a narrow slit S1 is made to fall on a plane mirror, which has been
polished on the front surface and blackened at the back (to avoid multiple reflections),
106 Wave Optics

at grazing incidence (Figure 3.11). The reflected beam appears to diverge from S2, which is a
virtual image of S1. Thus S1 and S2 act as coherent sources. It means that light directly coming
from the slit S1 superposes over the reflected light and forms an interference pattern in the
region of overlap. By referring to Figure 3.11, we can say that superposition of the direct cone
of light pS1s and the reflected cone of light qS2r results in production of interference pattern
in the region qr.

Figure 3.11 The Lloyd’s mirror arrangement for production of interference pattern.

Recall that in Young’s double slit, Fresnel biprism as well as Fresnel two-mirror arrangement,
the central fringe is bright. But in case of Lloyd’s mirror, the central fringe is dark and
cannot be observed. To discover the reason for such a significant difference, we note that if the
screen is brought in contact with the mirror such that the point O of the screen MN touches the
end of the mirror, it comes at the centre of a dark fringe, though it is equidistant from S1 and
S2 and under such circumstances, we expect it to lie at the centre of a bright fringe. Such a
physical situation can be explained by assuming that one of the interfering beams undergoes a
phase change of p. Since the direct beam from S1 cannot undergo a phase change, the only
possibility is that the reflected wave undergoes a phase change of p. This is an important
result: On reflection from an optically denser medium, light undergoes a phase change of
p. (However, no such abrupt phase change occurs when reflection takes place at the interface
with a rarer medium.)
Note that in case of Fresnel’s two-mirror arrangement, the central fringe is bright because
both interfering waves undergo a phase change of p and there is no change in path difference,
as happens in case of Lloyd’s mirror.
You should now go through the following example carefully.

EXAMPLE 3.6 In Lloyd’s mirror experiment, a mirror of length 0.05 m is illuminated with
monochromatic light of wavelength 589 nm. If the slit opening is 0.001 m and is at a distance
of 0.05 m from the nearer edge, calculate the fringe-width and the total width of the observed
pattern on a screen 1.0 m from the slit.
Solution Refer to Figure 3.12. MM¢ is Lloyd’s mirror of length 0.05 m. The interference
pattern is observed in the region AB. From Eq. (3.17), we recall that fringe-width is given by
OD
E
d
Two Beam Interference by Division of Wavefront 107

Figure 3.12 Interference pattern produced by Lloyd’s mirror.

Here, l = 589 ´ 10–9 m, D = 1.0 m and d = 2 ´ 0.001 = 0.002 m = 2 ´ 10–3 m. Hence

(589 – 10 9 m) – (1.0 m)
E 294.5 – 10 6 m
2 – 10 3 m
To determine the width of the interference region, we note from Figure 3.12 that
0.001 m 0.001 m
tan T1 0.02 and tan T 2 0.01
0.05 m 0.1 m
Also from the right angled triangle DAMC, we can write
AC = MC tanq1 = (0.95 m) ´ 0.02 = 0.019 m
Similarly, from the right angled triangle DBM¢C we can write
BC = M¢C tanq2 = (0.90 m) ´ 0.01 = 0.009 m
Hence, the width of the interference region is given by
AB = AC – BC = (0.019 – 0.009) m = 0.01 m

You may now like to solve a Practice Exercise.

Practice Exercise 3.5 Interference fringes are obtained in Lloyd’s single mirror arrangement
with a monochromatic source emitting waves of wavelength 589 nm. A thin transparent sheet
of refractive index 1.6 is then introduced normally in the path of one of the interfering waves.
As a result, the central dark fringe moves into the position earlier occupied by the third dark
fringe from the centre. Calculate the thickness of the sheet. [Ans. 1.96 ´ 10–6 m]

In the above discussion, you must have noted the following important differences in the
fringes produced by Fresnel Biprism and Lloyd’s single mirror:
108 Wave Optics

• With Biprism, we obtain symmetrical fringe pattern, whereas with Lloyd’s mirror,
usually only a few fringes are visible on one side of the central fringe.
• The central fringe in the interference pattern obtained with a Biprism is bright, whereas
the central fringe in the interference pattern obtained with Lloyd’s mirror is dark.
Let us now sum up all that you have learnt in this chapter.

3.4 SUMMARY

• Two sources are said to be coherent if they give out light waves of the same frequency,
which bear a fixed phase relationship in time between them, travel in the same direction
and have equal amplitude.
• In Young’s double slit experiment, interference pattern is obtained by division of wavefront
wherein two coherent sources are obtained from a single source and waves having constant
phase difference are made to superpose.
• The condition for a bright fringe at a point on the screen is
nO
yn D; n = 0, 1, 2, ...
d
• The condition for a dark fringe at a point on the screen is

È 1Ø O
yn Én  Ù D; n = 0, 1, 2, ...
Ê 2Ú d
• The fringe-width in an interference pattern is given by
OD
E
d
• The interference pattern in space is a hyperboloid of revolution, which is obtained by
revolving the hyperbola about the line joining the coherent sources. However, the
interference fringes on the observation screen are more or less straight lines.
• When a glass sheet is introduced in the path of one of the interfering waves, the distance
through which the nth bright fringe is displaced is given by:
D
] ( P  1)t
d
• In a Fresnel biprism experiment, we obtain closely spaced straight line interference
fringes of equal width.
• With biprism, we obtain symmetrical fringe pattern, whereas with Lloyd’s mirror, usually
only a few fringes are visible on one side of the central fringe.
• The central fringe in the interference pattern obtained with a biprism is bright, whereas
the central fringe in the interference pattern obtained with Lloyd’s mirror is dark.
Two Beam Interference by Division of Wavefront 109

REVIEW EXERCISES

1. A monochromatic beam of wavelength 585 nm illuminates two narrow slits 2.5 ´ 10–4 m
apart. The observation screen is located at a distance of 0.3 m from the source plane.
Calculate fringe-width. [Ans. 7.0 ´ 10–4 cm]
2. Interference fringes are produced by a Fresnel biprism in the focal plane of a microscope,
which is located at a distance of 1 m from the slit. Two images of the slits are produced
by a convex lens placed between the microscope and the biprism. At the positions of
the images, the distance between the slits are 4.10 mm and b. Calculate the value of b,
if fringe-width 0.175 mm is obtained with wavelength 589 nm.
[Ans. 2.76 ´ 10–3 m]
3. The obtuse angle of a prism is 176º. A slit illuminated by a monochromatic light of
wavelength 571.5 nm is placed at a distance of 0.1 m from the biprism. The distance
between the biprism and the screen is 1 m. Calculate the distance between two consecutive
dark fringes. [Ans. 1.8 ´ 10–4 m]
4. In a Fresnel’s biprism experiment, interference fringes of width 0.0195 cm are observed
at a distance of 1 m from the slit. A convex lens is then put between the observer and the
prism so as to give an image of the source at a distance of 1 m from the slit. The distance
between the slits and their two image positions are 0.838 cm and 0.585 cm. Calculate the
distance of the slit from the lens, if a monochromatic source of wavelength 588 nm
is used. [Ans. 0.30 m]
5. A thin glass sheet (m = 1.5) is placed normally in the path of one of the interfering beams
originating from a source emitting light of 540 nm wavelength. It is observed that the
central bright band of the fringe system moves to the position previously occupied by the
third bright band from the centre. Calculate the thickness of the glass sheet.
[Ans. 3.24 ´ 10–6 m]
6. In Young’s double slit experiment performed with monochromatic light of wavelength
500 nm, a thin film of transparent material of thickness 10 µm is put behind one of the
slits. The zero order fringe moves to the position previously occupied by the nth order
bright fringe. Calculate n, if refractive index of the film is 1.2. [Ans. 4]
7. The width of a spectral line of wavelength 500 nm is 0.02 nm. Calculate the approximation
to the largest path difference for which the interference fringes produced will be clearly
visible. [Ans. 3.125 ´ 10–3 m]
8. Refer to Figure 3.13. A point source of light placed at the origin of a rectangular coordinate
system emits a spherical sinusoidal wave. The electric field associated with the wave is
given by

È DØ Ë r ØÛ
E1 (Z , r ) cos Ì 2S É Q t 
È

Ê r ÚÙ Í Ê O ÙÚ ÜÝ
where r denotes the distance of the point of observation from the origin.
110 Wave Optics

Figure 3.13 Cross-sectional view of the experimental arrangement


used to observe interference fringes.

The electric field associated with another wave propagating along the x-axis is given by
Ë xØÛ
E2 (Z , r ) a cos Ì 2S É Q t
È

Í Ê O ÙÚ ÜÝ
Both these waves reach the flat screen placed perpendicular to the x-axis as shown in
Figure 3.13. Calculate the resultant intensity I on the screen as a function of y, D, and l,
if y = D, and lk = D.
Ë È S y2 Ø Û
Ì Ans. I I 0 cos2 É ÙÜ
Í Ê D2 Ú Ý

9. Refer to the arrangement of three slits as shown in Figure 3.14. Suppose that the width
of each slit is approximately equal to half the wavelength and these are illuminated by a
monochromatic source of light.
(a) Determine the value of q at which the first principal maximum is observed.
(b) If we denote the angle at which first principal maximum is observed as q0 and the
flux in the direction of zeroth order maximum as F0, calculate the flux in the direction
q0/2.

Figure 3.14 Cross-sectional view of three slits illuminated by a monochromatic source.


2O
Ë 5 Û
ÌÍ Ans. (a) T0
, (b) I (0) Ü 
d 9 Ý
10. A double slit interference pattern is produced by a parallel beam of light passing through
two slits of unequal separations 20l and 40l, respectively, where l is the wavelength of
the light used (see Figure 3.15). The distance between the centres of the slits is 1,000 l.
Two Beam Interference by Division of Wavefront 111

Figure 3.15 Cross-sectional view of two slits of unequal widths


illuminated by a monochromatic source.

If the fringes are observed on a screen far off from the slits (D ? 1,000 l), calculate the
separation of two adjacent maxima.
Ë L Û
Ì Ans. 1,000 ÜÝ
Í
CH A P T E R 4
INTERFERENCE BY
DIVISION OF AMPLITUDE

EXPECTED LEARNING OUTCOMES


After reading this chapter, the student will acquire the capability to:
• Discuss the formation of interference fringes in a thin film and a wedge-shaped film.
• Explain how Newton's rings are used to determine the wavelength of light.
• Differentiate between fringes of equal inclination and fringes of equal thickness.
• Describe the working of Michelson interferometer and how it produces different types of fringes
and used to measure the wavelength of light.
• Describe the working of Fabry–Perot interferometer and explain the salient points which differentiate
it from Michelson interferometer.
• Discuss other types of interferometers based on two-beam as well as multiple beam interferometry.
• Solve numerical problems.

4.1 INTRODUCTION

In Chapter 3, we have discussed two-beam interference produced by division of wavefront, i.e.


different portions of the wavefront are made to propagate in different directions and then
recombined. In other words, the same source is used to generate two coherent sources and when
spherical or cylindrical wavefronts emanating from these sources are allowed to superpose,
beautiful fringes are seen in the region of overlap. As mentioned in the preceding chapter, some
of the most beautiful colours of interference arise by division of amplitude where two light
waves are derived from a single incident wave. When incident wave is partially reflected from
the upper surface of a thin transparent plane-parallel film and partly reflected from its lower
surface, their amplitudes are reduced. However, their superposition leads to interference pattern.
The brilliant colours in soap bubbles, thin oil films on water or by cracks in a piece of glass
appear due to interference by division of amplitude when white light undergoes multiple

112
Interference by Division of Amplitude 113
reflections. Hooke observed such colours in thin films of mica and other transparent plates. Newton
was able to show interference rings when a convex lens was placed on a plane glass plate.
In Section 4.2, we have discussed the theory of interference by transparent plane-parallel as
well as wedge-shaped thin films when illuminated by a point as well as an extended source.
The applications of interference in thin films are very important for production of coated
surfaces, increase in reflectivity, test deviation of optical surfaces from parallelism and grinding
of a lens. These are also discussed in this section at appropriate places. In Section 4.3, we have
discussed formation of Newton’s rings and how this arrangement can be used to determine the
wavelength of light.
In Chapter 3, we learnt that Young’s double slit arrangement, Fresnel’s biprism, Lloyd’s mirror
and Fresnel’s double mirrors, which are used to observe interference phenomenon based on division
of wavefront, can be categorized as interferometers. We can similarly design an interferometer
which divides a wave by partial reflection and partial transmission, as in thin or wedge-shaped
films. Such an interferometer belongs to the category of division of amplitude. (The resulting
waves will have reduced amplitude or intensity.) The Michelson interferometer is an important
example of this class. In this interferometer, the original wave is split by a semi-reflecting
metallic film and the two resulting beams are made to interfere subsequently. In Section 4.4, we
will learn about the working of Michelson interferometer and how it is used to observe different
types of fringes or measure the wavelength of light. This discussion is then extended to multiple-
beam interferometry. In particular, we have discussed Fabry–Perot interferometer. The relative
advantages offered by both these interferometers are also discussed in some detail.

4.2 INTERFERENCE BY A PLANE-PARALLEL THIN FILM

Refer to Figure 4.1(a). Light from a source S is incident obliquely on a thin transparent film of
uniform thickness h. A part of this wave is reflected from the upper surface at A along AR1 and
a part is refracted along AB. On reaching B, part of the latter is reflected along BC and a part
is transmitted along BT1. At C, the wave will again get partly reflected along CD and partly
refracted along CR2, which is paralled to AR1. A continuation of this process gives a set of
waves on either side of the film. These move parallel to each other but the amplitudes or
intensity decreases gradually from one wave to the next. (For films of low reflectivity, the drop
in intensity of successive waves is very rapid and it is possible that we are left with only a
couple of waves in reflection as well as transmission.) If these parallel sets are made to focus
at a point, say P, each wave would have travelled a different distance and the phase relations
may be such as to produce a constructive or destructive interference at P. In case of unaided
eye, P conforms to the retina and in case of an eye-piece, it corresponds to the cross-wires.
The basic condition for interference is that reinforcing waves must originate from two
coherent sources. Therefore, you may logically ask: Are coherent sources produced in this case
also? If so, how? To discover answer to this question, let us consider the waves propagating
along AR1 and CR2. As such, these waves arise from the point source S due to reflection at the
top and bottom surfaces of the thin film which acts as amplitude-splitting device. But these
waves may respectively be considered as arising from two point coherent virtual sources S¢ and
S¢¢, which are obtained by producing AR1 and AR2 behind the film, as shown in Figure 4.1(b).
114 Wave Optics

Thus if we put a photographic plate parallel to the surface of the film, we will, in general, obtain
interference fringes.

Figure 4.1 (a) Multiple reflections in a thin plane-parallel film of transparent material, and (b) the
interference pattern due to the first set of rays is the same as would have been produced
by two coherent point sources.

Interference due to reflected light


Refer to Figure 4.2. It shows that light wave is incident at an angle a at A on a film of thickness
h and refractive index m (>1). At A, the wave is refracted along AB and makes an angle f with
the normal. We draw CN ^ AR1 and BM ^ AC, so that the distance travelled by the waves along
AR1 and CR2 beyond CN is the same. Hence, the path difference between these waves is given by
D = ABC – AN
= m(AB + BC) – AN (4.1)

Figure 4.2 Optical path difference between two consecutive waves reflected by a thin plane-parallel
film of thickness h and refractive index m.
Interference by Division of Amplitude 115
From this equation, we note that to calculate the path difference D, we have to express optical
paths AB, BC and AN in terms of the physical properties of the thin film. To do so, we note
that in Ds ABM and CBM
BM h
AB BC (4.2)
cos I cos I
and in DCNA, CN ^ AN so that
AN = AC sin a
Proceeding further, we note that
AC = AM + MC
and from Ds ABM and CBM, we can write
AC = BM tanf + BM tan f
= 2h tan f
Hence,
AN = 2h tan f sin a
sin I
= 2h sin D
cos I

= 2h
sin I
( P sin I )
È
' sin D P ØÙ
cos I É
Ê sin I Ú
2
sin I
= 2P h (4.3)
cos I
On using these results for AB, BC and AN in Eq. (4.1), we get

È h h Ø sin 2 I
D = PÉ   2P h
Ê cos I cos I ÙÚ cos I
2P h
= (1  sin 2 I )
cos I
= 2mh cosf (4.4)
Note that at A, we have rarer–denser interface, i.e. the wave is reflected while propagating from
a rarer medium to a denser medium and undergoes a phase change of p. But at B, the wave is
reflected while travelling in a denser medium and there is no additional phase change. We
further note that a phase change of p is equivalent to a path difference of l/2. Hence, the
effective path difference between waves travelling along AR1 and CR2 is given by
O
' eff 2 P h cos I “ (4.5)
2
Note that we can choose either ‘+’ sign or ‘–’ sign. But we will choose the negative sign for
simplicity.
In Chapter 3, we learnt that if Deff = nl, n = 1, 2, 3, ..., i.e. the effective path difference
is an integral multiple of l, constructive interference will take place and the film will appear
bright:
116 Wave Optics

O
' eff 2P h cos I  nO
2
O
or 2P h cos I (2n  1) n = 0, 1, 2, 3, ..., maxima (4.6)
2
Similarly, if Deff = (2n – 1) l/2, where n = 1, 2, 3, ..., i.e. the effective path difference is an
odd multiple of l/2, the reflected waves propagating along AR1 and CR2 will be out of phase
and interfere destructively. Hence the condition of minimum or destructive interference in this
case becomes
O O
' eff 2 P h cos I  (2n  1)
2 2
or 2mh cosf = nl n = 1, 2, 3, ..., minima (4.7)
We have obtained the conditions of constructive and destructive interferences produced by a plane-
parallel transparent thin film by division of amplitude. Note that these conditions are opposite to
the conditions for two wave interference by division of wavefront. However, symmetry considerations
require the fringes to be circular in a plane parallel to the plane of the film. The perpendicular
from the source on the plane of the film forms the axis of symmetry. Further, if the intensities of
the waves AR1 and CR2 decrease gradually, the interference pattern will not be perfect. In fact, this
difference in intensities may not allow formation of completely dark fringes if we had waves
propagating along AR1 and CR2 only. (This will happen only for films of low reflectivity.) But it
is not so when multiple reflections take place giving rise to other sets of reflected waves.
To understand the effect of multiple reflections, refer to Figure 4.1 again and examine the
phases of the waves propagating along ER3, GR4, IR5, ... . Note that the path difference between
any pair of these waves will be the same, equal to D, since the geometry is the same. Moreover,
these waves undergo only internal reflections so that Deff = D and is given by Eq. (4.4). It means
that if Eq. (4.7) is satisfied, the wave along ER3 will be in phase with the wave along CR2.
The same holds for all other successive pairs. It implies that under the condition specified by
Eq. (4.7), waves along AR1 and CR2 will be out of phase but those travelling along CR2, ER3,
GR4, IR5, ... will be in phase with each other. Moreover, the addition of these waves (ER3, GR4,
IR5, ...) gives a net amplitude just sufficient to make up for the difference to produce complete
darkness.
Refer to reflected waves depicted as 1, 2, 3, in Figure 4.3. Suppose that amplitude of the
incident wave is a. Let r be the reflection coefficient, t be the transmission coefficient from
rarer to denser medium and t¢ be the transmission coefficient from denser medium to rarer
medium. (For normal incidence, a = f = 0). Then the amplitudes of the reflected rays are
ar, atrt¢, atr3t¢, atr5t¢, ... and so on. The resultant amplitude of 2, 3, 4, waves is given by
E = atrt¢ + atr3t¢ + atr 5t¢ + L
= att¢r(1 + r2 + r4 + L)
Since r is necessarily less than one, the terms within the brackets form an infinite geometric
series with common ratio r2. So its sum is finite and we can write
È 1 Ø
E artt „ É
Ê1  r 2 ÙÚ
Interference by Division of Amplitude 117

Figure 4.3 Amplitudes of successive waves undergoing multiple reflections.

According to the principle of reversibility (Stokes’ law), tt¢ = 1 – r2, so that the expression
for resultant amplitude takes a very compact form:
E = ar (4.8)
This result shows that the resultant amplitude of 2, 3, 4, waves is equal in magnitude to
the amplitude of wave 1 but out of phase with it. It means that there will be complete destructive
interference and the minima of the reflected system will have zero intensity.
Note that in Figure 4.1, one of the interfering waves is reflected from the upper surface of
the film and the second wave is reflected from the back surface of the transparent film. Moreover,
these are derived from the same incident wave. The interference fringes in this case are located
in the focal plane of a lens or at infinity. Such fringes are said to be localized fringes. On the
other hand, if two different waves are incident at slightly different angles and interfere at a finite
distance from the film after reflection from the upper and lower surfaces, the point of interference
changes with the angle of incidence. Such fringes are said to be non-localized.
Before proceeding further, you should go through the following examples carefully. These
will help you to appreciate the formation of colours by reflected light.

EXAMPLE 4.1 A typical thin film of refractive index 1.33 and thickness 1.5 ´ 10–6 m is
illuminated by a point source of white light at an angle of 60º. When reflected light was
examined, a dark band was found to occur corresponding to a wavelength of 500 nm. Determine
the order of interference of the dark band.
Solution From Eq. (4.7), we recall that position of minima in the interference pattern is
given by
2mh cosf = nl n = 1, 2, 3, ...
sin D
Here, m = 1.33, l = 500 nm, h = 1.5 ´ 10–6 m and a = 60º. Since m = , we can write
sin I
sin 60’ 3 1.732
sin I 0.6495
1.33 2 – 1.33 2.66
so that
f = 40.5º
118 Wave Optics

and
cos 40.5º = 0.7604
Hence
2 P h cos I 2 – 1.33 – (1.5 – 106 m) – 0.7604
n 6
O (500 – 10 9 m)
EXAMPLE 4.2 A parallel beam of light from a sodium lamp (l = 589 nm) is incident on a
thin glass plate of refractive index 1.5 so that the angle of refraction is 60º. For what minimum
thickness of the glass plate will it appear completely dark in reflected light?
Solution From Eq. (4.7), we recall that
2mh cosf = nl n = 1, 2, 3, ...
Here, m = 1.5, f = 60º, l = 589 nm and for minimum thickness n = 1. Hence,
nO
h=
2 P cos I
589 – 109 m
=
2 – 1.5 – 0.5
= 392.7 ´ 10–9 m
EXAMPLE 4.3 In Example 4.1, light is made to fall
(i) normally (f = 0º), and
(ii) at an angle so that f = 60º.
Which colours will appear in the reflected light?
Solution From Eq. (4.6), we know that the condition for constructive interference of reflected
light is
O
2 P h cos I (2n  1) n 0, 1, 2, 3, ...
2
Here, m = 1.33 and h = 1.5 ´ 10–6 m.
(i) When f = 0, cosf = 1. Hence,
O
2 ´ 1.33 ´ (1.5 ´ 10–6 m) ´ 1 = (2n  1)
2
6 9
2 – 2 – 1.33 – (1.5 – 10 m) 798 – 10
or O m
(2n  1) (2n  1)
For n = 0, 1, 2, ... the wavelengths which will be reflected most strongly are
l = 798 nm, 266 nm, 159.6 nm, ... . Of these, the wavelength corresponding to
798 nm lies in the visible range. It means that the film will appear reddish when
viewed in the reflected light. All other waves lie in the ultraviolet region and will be
invisible.
Interference by Division of Amplitude 119
(ii) For f = 60º, cos 60º = 0.5. Hence,
O
2 – 1.33 – (1.5 – 10 6 m) – 0.5 (2n  1)
2
6
2 – 2 – 1.33 – (1.5 – 10 m) – 0.5 399 – 109
or O m
(2n  1) (2n  1)
For n = 0, 1, 2, ... the wavelengths which will be reflected most strongly are
l = 399 nm, 133 nm, 79.8 nm, . Of these, only 399 nm corresponds to violet colour,
which lies in the visible range. It means that the film will appear violet when viewed
in the reflected light. All other waves lie in the invisible ultraviolet region.

On going through these examples, you must have realised when light falls on a thin film,
the colour(s) seen in reflected light depends on its thickness, refractive index of the material and
angle of inclination (or position of the eye relative to the region of the film under focus).
It means that if f and h are constant, the colour will be uniform. But in the case of oil in water,
different colours are visible because f and h vary.
You may now like to answer a Practice Exercise.

Practice Exercise 4.1 Refer to Figures 4.1 and 4.3 and consider the light transmitted by a
thin film. Show that the conditions for maxima and minima are given by 2mh cosf = nl;
n = 1, 2, 3, ... and 2mh cosf = (2n + 1)l/2; n = 0, 1, 2, 3, ..., respectively.

If we compare the conditions of maxima and minima in the reflected and transmitted lights,
we note that these are exactly opposite. It means that the colours which are found missing in
reflected light will be visible in transmitted light. In other words, the appearance of colours in
the transmitted and reflected lights is complementary. However, since the amplitudes of reflected
and transmitted waves differ, in general, the interference fringes in reflected light have higher
visibility than those in transmitted light.
A simple but important application of the principle of interference in thin films is the
production of coated surfaces. To accomplish this, a film of a transparent substance of refractive
index m¢ is deposited on a glass plate or lens having refractive index m(> m¢). The thickness of
O
the film is kept one-quarter of the wavelength of light in the film, so that h = . For normal
4P „
incidence, the path difference between light waves reflected from the upper and lower surfaces
of the film will be l/2. But waves undergo a phase change of p since reflections occurring at
both surfaces are from ‘rarer-to-denser’. It means that the reflected waves will remain out of
phase by p and will interfere destructively. Such a film is known as non-reflecting film.
You should not think that a non-reflecting film destroys light; it only redistributes energy—
decrease of intensity in reflected light is compensated by a corresponding increase of intensity
of transmitted light.
Obviously, a non-reflecting film can be used to reduce loss of light due to reflection by
lenses and prisms used in binoculars, cameras, etc., and improve contrast. Usually, glass
is coated with a very thin layer of magnesium fluoride, whose refractive index (=1.38) is
in-between those of glass and air.
120 Wave Optics

Can you use thin film interference phenomenon to increase reflectivity? If so, how? Thinking
logically, we may conclude that we can do so by using a film of refractive index greater than
that of glass and suitable thickness.
Interference fringes produced in thin films are classified as fringes of equal inclination and
fringes of equal thickness. We now discuss these in brief.

Fringes of equal inclination: Need for a broad source


We know that the entire pattern of interference fringes formed by division of wavefront, as in
the case of Fresnel’s biprism, Fresnel’s two-mirror arrangement and Lloyd’s single mirror, can
be obtained on the screen or viewed with an eye-piece. Moreover, the source used is narrow.
But in case of interference by division of amplitude, as in case of thin films, the narrow source
limits the visibility of the film; i.e. only that portion becomes visible from where waves are
directly reflected into the eye/eye-piece/focusing lens, as shown in Figure 4.4(a). As may be
noted, the interference fringes produced only by the waves reflected from region A along 3 and
4 (and generated from wave incident along (a) become visible because these waves reach the
eye. On the other hand, the wave along (b) is reflected along 5 and 6 (region B), and these do
not reach the eye. Thus, region B does not become visible.
Similarly, other waves incident on the film at different angles may not become visually
observable as these fail to reach the eye/eye-piece/focusing lens. It means that only portion A
of the film becomes visible.
Now refer to Figure 4.4(b). It shows an extended monochromatic source that illuminates
a thin film. Such an extended source may be produced by illuminating a ground glass plate by
a sodium lamp. An extended source can be assumed to consist of a large number of independent
point sources. As a result, the field of view gets extended significantly since waves from
different points on the source are incident at various angles on the thin plane-parallel film.
It means that the fringe pattern will spread out over a larger area of the film. That is why a broad
source of light is preferred while observing interference phenomenon in thin films.

Figure 4.4 Fringes formed by (a) a thin film illuminated by a narrow source, and (b) an extended source.
Interference by Division of Amplitude 121
If we use a photographic plate to record the fringe pattern, no definite pattern will appear
because each point on the extended source gives rise to its own interference pattern. Moreover,
these patterns will be displaced with respect to one another. However, if we view the film with
unaided eye, waves from all parts of the film will reach the eye after reflection and focus on
the retina. For any fringe, the value of a is fixed so that the fringe will have the form of the
arc of a circle whose centre is at the foot of the perpendicular drawn from the eye to the plane
of the film. So, these fringes are of equal inclination.
As the film becomes thicker, the reflected waves may get separated so much that both waves
may fail to reach the eye simultaneously. In such conditions, no interference can occur. However,
we can use a telescope of large aperture to focus these waves and fringes may become visible.
Alternatively, we can reduce the separation between reflected waves by viewing the film at
nearly normal incidence, i.e. by reducing angle of incidence and therefore the angle of refraction.
The fringes seen with thick plates near normal incidence are of equal inclination and are often
called Haidinger fringes (Figure 4.5).

Figure 4.5 Schematics of the arrangement used to observe Haidinger fringes.

Fringes of equal thickness: Interference by a wedge-shaped film


We know that fringes of equal inclination are formed when a thin plane-parallel film of uniform
thickness is illuminated by an extended source. You may now like to know as to what will
happen if the film is not plane-parallel, i.e. the film has continuously varying thickness.
Such a film can be realized by putting together two pieces of ordinary glass plates having plane
surfaces and with a thin sheet of paper spacer along one edge, as shown in Figure 4.6(a).
As may be noted, we obtain a wedge-shaped film of air between the plates. When illuminated
by a monochromatic light source, say sodium lamp, equidistant, practically straight line
122 Wave Optics

interference fringes parallel to the thin edge of the wedge are formed. These are referred to as
fringes of equal thickness. Each fringe is the locus of all points in the film for which thickness
is a constant. Such fringes are localized on the film itself and can be observed using a microscope.
To derive the conditions of maxima and minima, refer to Figure 4.6(b). AB and CD are two
plane surfaces inclined at angle q and enclosing a film of refractive index m. When produced
backward, the lines corresponding to these surfaces will meet at O. The thickness of the film
increases from A to B. Suppose that the film is illuminated by a monochromatic source of light
from a slit held parallel to the edge of the wedge so that light falls on the upper surface of the
film practically perpendicularly. (The edge is the line passing through O and normal to the plane
of the paper.) When the film is viewed in the reflected light, interference fringes parallel to the
line of intersection of these surfaces are seen.

Figure 4.6 Fringes of equal thickness. (a) Method of observation and (b) Schematic representation.

When the wedge angle q is small, the optical path travelled by refracted and reflected light
waves is essentially equal to twice the thickness of the film at the point of incidence. Hence,
the path difference for a given pair of waves for a wedge with small angle is essentially given
by Eq. (4.4). Since the light is made to fall on the film near normal, the cosf factor in this
expression can be taken to be equal to unity. Thus, the expression for the path difference
between the waves reflected at the upper and lower surfaces simplifies to 2mh, where h is the
thickness of the film at P.
As before, an additional path difference of l/2 arises due to change of phase by p at P.
O
Hence the effective path difference can be expressed as Deff = 2mh – and the condition for
2
bright fringes becomes
O
' eff 2P h  nO
2
O
or 2P h (2n  1) n = 0, 1, 2, ... (4.9)
2
Similarly, the condition for the dark fringe is given by
2mh = nl n = 1, 2, 3, ... (4.10)
Note that Eqs. (4.9) and (4.10) are analogous to Eqs. (4.6) and (4.7) respectively. Further, for
near normal incidence on a wedge shaped film with small wedge angle, a bright or a dark fringe
of a particular order is obtained, if h is constant. But in a wedge-shaped film, this condition is
Interference by Division of Amplitude 123
satisfied only along lines parallel to the thin edge of the wedge. That is why the interference
pattern in this case comprises straight lines parallel to the thin edge of the wedge. Moreover,
at the thin edge, h = 0 so that the path difference is l/2. This signifies the general condition
of minimum intensity. It means that the edge of the film will be dark.
Proceeding further, we derive the expression for fringe width, i.e. the spacing between two
consecutive bright or dark fringes. Suppose that the nth bright fringe occurs at a distance xn
from the thin edge. Then h = xn tan q = xnq, if q is small. (It is measured in radians.) Hence,
Eq. (4.9) can be rewritten as
O
2P xnT (2n  1) (4.11)
2
Similarly, if the (n + 1)th bright fringe is obtained at a distance xn + 1 from the thin edge, then
we can write
O
2P xn 1T (2n  3) (4.12)
2
On combining Eqs. (4.11) and (4.12), we get
2m(xn+1 – xn)q = l
Hence, the expression for fringe-width b is given by
O
E xn 1  xn (4.13)
2PT
Such fringes are commonly referred to as fringes of equal thickness. We can similarly show that
the spacing between two consecutive dark fringes will also be given by Eq. (4.13).
The phenomenon of interference in wedge-shaped films finds important practical application
in the testing of optical surfaces for flatness, i.e. deviation from parallelism. If an air film is
formed between two surfaces, only one of which is perfectly plane, we expect the fringes to be
irregular in shape. That is, if both surfaces are not parallel, fringes of equal thickness (straight
and parallel) will not be formed. So the standard method to test the flatness of an optical surface
is to take an optically plane surface and form its wedge shaped thin film with the test surface.
If fringes of equal thickness are seen in the field of view, we can say that the test surface is
plane. If the fringes are irregular, the test surface is not plane. Then it is polished and the above-
stated process is repeated till fringes of equal thickness are obtained.
Before proceeding further, you are advised to study the following examples carefully.

EXAMPLE 4.4 A vertical rectangular soap film of total length 0.12 m is illuminated with
light of wavelength 600 nm. Just before the film breaks, there are 12 dark and 11 bright
interference fringes between the upper and lower ends. Calculate the angle of the wedge so
formed and the thickness of the film at the base just before the film breaks. Take the refractive
index of the soap solution as 1.33.
Solution From Eq. (4.10), we can write
nO
T
2P xn
where we have used the result h = xnq.
124 Wave Optics

Here, x = 0.12 m, n = 11, l = 600 nm = 600 ´ 10–9 m and m = 1.33. Hence

11 – (600 – 109 m)
T 2.07 – 105 rad
2 – 1.33 – (0.12 m)
When the wedge angle is small, we can write
h = xq
= (0.12 m) ´ 2.07 ´ 10–5 = 2.48 ´ 10–6 m
EXAMPLE 4.5 Interference fringes are produced by a source of monochromatic light.
The light is incident normally on a wedge-shaped film of refractive index 1.40, and wedge angle
20¢¢. If fringe width is 2.5 ´ 10–3 m, calculate the wavelength of light.
Solution We rewrite Eq. (4.13) as
l = 2bmq
Here b = 2.5 ´ 10–3 m, m =1.4 and q = 20¢¢. We have to express q in radian. To do so,
we note that
20 – S
T rad
60 – 60 – 180
Hence,
2 – 20 – 22 – 1.4 – (2.5 – 103 m)
O 679 nm
60 – 60 – 180 – 7
EXAMPLE 4.6 Two 0.15 m long glass plates are made to touch at one end. A sheet 6 ´ 10–5 m
thick separates the other end. How many bright fringes will be observed over the entire plate,
if light of wavelength 720 nm is reflected normally from it?
Solution From Eq. (4.13), we know that
O
E
2PT
Here, m = 1 and q = h/l, where h is the thickness of the wedge at the open end and l is the
length of the plate. Hence we can rewrite the expression for fringe width as
OA
E
2h
Suppose that the number of fringes observed over the entire plate is N. Then l = Nb so that
the expression for the number of fringes observed over the entire plate is given by
2h
N
O
On inserting the given data, we get
2 – (6 – 10 5 m)
N  167
(720 – 10 9 m)
Interference by Division of Amplitude 125
EXAMPLE 4.7 Calculate the fringe-width for the arrangement described in Example 4.6.
Solution The expression for fringe-width derived in Example 4.6 is
OA
E
2h
On inserting the given data, we get
(720 – 10 9 m) – (0.15 m)
E 9 – 10 4 m
2 – (6 – 10 5 m)

We would now like you to answer a Practice Exercise.

Practice Exercise 4.2 The thicknesses of a wedge-shaped film of refractive index 1.5
at its two ends are h1 and h2. If 10 fringes are observed in the film with a light of wavelength
600 nm, calculate the difference h2 – h1. [Ans. 2 ´ 10–6 m]

We know that when light from an extended monochromatic source like sodium lamp is
incident normally on a wedge-shaped thin transparent plane film, equally spaced dark and bright
fringes are observed. The distance between any two successive bright or dark fringes is determined
by the wavelength of incident light, angle of the wedge and the refractive index of the film. If
a polychromatic source such as incandescent lamp is used, we observe coloured fringes. Moreover,
if thickness of the film varies arbitrarily, each fringe represents the locus of constant film
thickness. This is what we see when sunlight falls on a soap bubble. But when optical path
difference between waves reflected from the upper and lower surfaces of the film exceeds a few
wavelengths, the interference pattern disappears due to overlapping of many colours and no
fringes are seen. It means that to observe interference fringes with white light, the thickness of
the film should not be more than a few wavelengths.
You may now like to know as to how is the fringe shape influenced when an air film is
enclosed between a convex surface, say of a lens, and a plane glass surface. We expect that
fringes of equal thickness will be produced with circular contour lines because thickness of the
air film remains constant on the circumference of a circle, whose centre is at the point of
contact. The ring-shaped fringes so produced were first studied in detail by Newton. He was
the first to measure their radii. But he failed to explain the basic physics adequately. Boyle and
Hooke had independently observed the fringes and the original discovery is attributed to Hooke.
The proper explanation of formation of Newton’s rings was given by Thomas Young.
The discussion of Newton’s rings forms the subject matter of discussion of the following section.

4.3 NEWTON’S RINGS

Refer to Figure 4.7. It shows the schematics of the experimental arrangement used to observe
Newton’s rings. S is an extended source of monochromatic light such as a sodium lamp. It is
placed at the first principal focus of a double-convex lens so that parallel beam of light is
incident on the glass plate G, which is held at 45º. The glass plate partially reflects the light
incident on it as a parallel beam towards the air film enclosed by the curved surface AOB of
126 Wave Optics

Figure 4.7 Schematic representation of the arrangement used for observing Newton’s rings.

a plano-convex lens of long focal length and the upper surface of the plane glass plate POQ.
The light waves reflected from the upper and lower surfaces of the air film interfere and the
Newton’s rings can be viewed directly or though a travelling microscope M.
Note that the thickness of the air film is zero at the point of contact O and increases as we
move away from it. Therefore, the pattern of dark and bright rings will consist of concentric
circles. We will observe dark fringe wherever thickness of air film satisfies the condition for
minima:
2h = nl n = 1, 2, ... (4.14)
Similarly, we will observe bright fringes wherever thickness of air film satisfies the condition
for maxima:
O (4.15)
2h (2 n  1) n 0, 1, 2, ...
2
Since the convex side of the lens is a spherical surface, the thickness of the air film will be
constant over a circle whose centre is at O. That is why we obtain a pattern of concentric bright
and dark circular rings. To observe these, the microscope (or the eye) is focused on the upper
surface of the film. (In case the film is made of any material other than air, the left hand sides
of Eqs. (4.14) and (4.15) will be multiplied by its refractive index, m.)
Let us now derive the expression for the radii of various rings and relate these to the
wavelength of light used. Since thickness of the air film is constant over a circle whose centre
is at the point of contact, let us denote the radius of nth dark ring by rn. Suppose that the
thickness of the air film where nth dark ring is formed is h and corresponds to point M in
Figure 4.8, which shows the plano-convex lens AOB and the glass plate POQ. Suppose the
radius of curvature of the curved surface of the lens is R. Then, from the property of a circle,
we can say that ÐOME = 90°. Hence N is the foot of the perpendicular drawn on the hypotenues
of a right angled triangle from the vertex containing the right angle. From similarity of Ds EMN
and EMO, we can write
(MN)2 = EN ´ ON = ON ´ (2R – ON) (4.16)
Interference by Division of Amplitude 127

Figure 4.8 Cross-sectional view of Newton’s rings rn is radius of the nth Newton ring.

But MN = rn and ON = h. On using these results in Eq. (4.16), we get

rn2 h(2 R  h) (4.17)


In a typical arrangement used for observing Newton’s rings, R = 1.0 m and h £ 10–5 m, so that
we can ignore h in comparison to 2R in Eq. (4.17) without introducing any significant error. In
this approximation, Eq. (4.17) reduces to

rn2 2 Rh
so that
rn2
2h (4.18)
R
On using this result in Eq. (4.14), we find that the radius of a dark fringe is given by

rn2 nRO
or
rn nRO where n = 1, 2, 3, ... (4.19)
Let us pause for a while and reflect as to what we have achieved. We note that when an air film
is enclosed between a convex surface and a plane glass surface, we obtain a pattern of concentric
dark and bright circular rings. The radius of a dark fringe is directly proportional to the square
root of
(i) the radius of curvature of the lens used,
(ii) the order of the fringe, and
(iii) wavelength of light used.
Between any two dark rings, there will be a bright ring. The radius of the bright ring is
obtained by combining Eqs. (4.15) and (4.18):

(2n  1) O R
rn n = 0, 1, 2, 3, ... (4.20)
2
128 Wave Optics

Proceeding further, we note that for l = 600 nm and R = 1.0 m, the radii of dark fringes are
given by
rn 7.74 – 10 4 n m (4.21)
It means that the radii of the first, second and third dark rings are 7.74 ´ 10–4 m,
1.1 ´ 10–3 m and 1.34 ´ 10–3 m, respectively. Note that the spacing between the second and
third rings is less than the spacing between the first and second dark rings. We can generalize
this result as: Newton’s rings get closer and closer to each other as their order increases.
Equation (4.19) implies that the central spot of Newton’s rings will be dark. Therefore,
while counting the order of dark fringes 1, 2, 3, , the central fringe is not counted. That is why
it is always desirable to measure the radii of nth and (n + p)th ring and determine the difference
in the squares of the radii. This gives rn2 p  rn2 pO R, which is independent of n.
Note that when the contact between the lens and the glass plate is perfect, the central spot
will be completely dark. This is a direct evidence of the relative phase change of p between air-
to-glass and glass-to-air reflections. If there was no such phase change, we should have obtained
a bright spot at the centre. Now we would like to know if it is possible to somehow obtain a
bright central spot in reflected light. If we argue logically, we can envisage an interesting
modification of the conventional Newton’s rings arrangement. It would be in the form of a glass
plate of a material of refractive index higher than that of the lens and the space between the
two surfaces is filled by a liquid of intermediate refractive index. Then both reflections will be
at rarer-to-denser surfaces.
Usually, it is more convenient to measure the diameter of a ring. In terms of diameters, the
wavelength is given by
Dn2 p  Dn2
O (4.22)
4 pR
You should now answer a Practice Exercise.

Practice Exercise 4.3


(i) Suppose that a liquid of refractive index m is introduced between the lens and the
glass plate. Show that the diameter of the dark rings seen in the reflected light will
nO R
be given by Dn 2 .
P
(ii) Show that the central ring in Newton’s rings seen in transmitted light is bright and the
diameter of bright rings is given by Dn2 4nO R.

On working out Practice Exercise 4.3, you will note that


• The radius of a dark ring seen in the reflected light is inversely proportional to the square
root of the refractive index of the material of the film. It means that if air is replaced by
water, the radius of the dark fringes will diminish by a factor of 0.0867.
• The rings seen in transmitted light are complementary to those seen in the reflected light.
• The ring-contrast is comparatively poor in transmitted light.
Interference by Division of Amplitude 129
Now go through the following examples carefully.

EXAMPLE 4.8 A plano-convex lens of radius 1.0 m is placed on an optically flat glass plate
and is illuminated by an extended monochromatic source. Assume that the point of contact is
perfect. The diameters of the 15th and 5th dark rings in the reflected light are 5.90 ´ 10–3 m
and 3.36 ´ 10–3 m, respectively. Calculate the wavelength of the light used.
Solution From Eq. (4.22), we can write
2
D15  D52
O
4 pR
On substituting the given values, we get

(5.90) 2 – 106 m 2  (3.36) 2 – 106 m 2


O 588 nm
4 – 10 – (1.0 m)
EXAMPLE 4.9 In the arrangement described in Example 4.8, the space between the lens and
the glass plate is filled with a liquid. The diameter of the 5th ring changes to 3.0 ´ 10–3 m.
Calculate the refractive index of the liquid when the ring is (i) dark, and (ii) bright. The
wavelength of light used is 589 nm.
Solution
(i) While working out Practice Exercise 4.2, you discovered that when a liquid of refractive
index m is introduced between the lens and the glass plate, the radii of the dark rings
seen in the reflected light is given by
nO R
rn =
P
In terms of diameter, we can rewrite it as
4nO R
Dn2 =
P
so that
4 nO R
m=
Dn2
Here, n = 5, l = 589 nm, R = 1.0 m and D5 = 3.0 ´ 10–3 m. On inserting these values
in the expression for m, we get
4 – 5 – (589 – 10 9 m) – (1.0 m)
m=
9 – 10 6 m 2
= 1.31
(ii) From Eq. (4.20), we recall that radius of the bright ring in an air film is given by
(2n  1) O R
rn n = 0, 1, 2, 3, ...
2
130 Wave Optics

Since rn is inversely proportional to refractive index, we can write


2(2n  1) O R
Dn2 =
P
so that
2(2n  1) O R 2 – 11 – (589 – 10 9 m) – (1.0 m)
m= =
Dn2 9 – 10 6 m 2
= 1.44
If we use the expression
2(2n  1) O R
Dn2 =
P
we will get m = 1.18.
EXAMPLE 4.10 Again refer to the arrangement described in Example 4.8 to observe Newton’s
rings. The lens is raised slowly vertically upward above the plate. Discuss the changes in the
interference pattern as seen through the travelling microscope, if light of wavelength 490 nm is
used.
Solution Refer to Figure 4.9. Suppose that the first dark ring is formed at N1, where
N1H1 = l/2 and the radius of this ring OH1 O R 7.0 – 104 m, as shown in Figure 4.9(a).
Similarly, the radius of the second ring will be OH 2 2O R 9.89 – 104 m and so on. If the
lens is raised vertically through l/4 = 122.5 nm, 2h corresponding to the central spot
will be l/2 and instead of the dark spot at the centre, we will obtain a bright spot
[Figure 4.9(b)]. The radii of the first couple of rings will be OH1„ (O R/2)1/ 2 4.95 – 104 m
and OH 2„ (3lR/2)1/2 = 8.57 ´ 10–4 m. Note that OH1 ! OH1„ and OH 2 ! OH 2„ implying that
the fringes move towards the centre as the lens is raised.

Figure 4.9 The rings collapse to the centre as the lens is moved vertically upward.
Interference by Division of Amplitude 131
On raising the lens through a further distance of l/4, the first dark ring collapses to the
centre and the central spot will be dark again. The ring which was at N2 initially now shifts to
N 2„„ and so on, as shown in Figure 4.9(c). From this we may conclude that as the lens moves
upward, the rings collapse to the centre. It means that if we can measure the distance by which
the lens is moved upward and also count the number of dark spots collapsing to the centre,
we can determine the wavelength of light. For example, if the lens moves through a distance
of 4.9 ´ 10–5 m, 200 rings will collapse to the centre.

We now advise you to answer a Practice Exercise.

Practice Exercise 4.4 In a Newton’s rings experiment, the light of wavelength 600 nm is
used. If a drop of water is placed between the lens and the plate, and the diameter of the 10th
ring is 6 ´ 10–3 m, calculate the radius of curvature of the lens. [Ans. 2.0 m]

So far we have considered formation of Newton’s rings using a monochromatic source.


Now we will investigate as to what changes will occur when a source such as sodium lamp,
which gives out two closely spaced wavelengths (589 nm and 589.6 nm), is used. Since the two
fringes are very close, the low order bright and dark rings due to l1 will superpose on the bright
and dark rings due to l2, respectively. We can easily check this statement by calculating the radii
of 10th or 20th bright and dark rings formed by l1 and l2. However, for large value of n, the
two ring patterns may produce uniform illumination. To be precise, suppose that the air film
thickness h is such that
O2
2h = nO1 (2n  1) (4.23a)
2
or
2h 1 2h 1
= n 
O2 2 O1 2
We can rewrite it as
2h 2h 1
 (4.23b)
O2 O1 2
This result shows that rings will completely disappear around the point of contact of the curved
surface with the glass plate. That is, bright ring due to l1 will superpose on the dark ring due
to l2 and vice versa. Thus, the contrast will be zero and no interference rings will be visible.
To get an idea about the order of rings which will produce uniform illumination, we rewrite
Eq. (4.23b) as
È O2  O1 Ø 1
2h É
Ê O1O2 ÙÚ 2
or
1 È O1O2 Ø 1 È 589 – 589.6 – 1018 m 2 Ø
2h É Ù 3 – 104 m
2 Ê 'O Ú 2 ÉÊ 0.6 – 109 m
Ù
Ú

Using this value of 2h in Eq. (4.23a), we find that its will correspond to n » 500.
132 Wave Optics

We can also observe this phenomenon if we slowly raise the lens upward as discussed in
Example 4.10. Suppose that the lens is moved through a distance y0 such that it satisfies
Eq. (4.23b). Then we can write
2 y0 2 y0 1

O2 O1 2
or
O1O2
y0
4(O1  O2 )
Thus, if the point of contact corresponds to a dark spot for l1, it will correspond to a bright
spot for l2 and vice versa. Similarly, the dark ring for l1 will coincide with the bright ring for
l2 so that the interference pattern will be washed out.
Next, the lens is moved further upwards to a distance y1, so that
2 y1 2 y1
 1
O2 O1
So, if the new position corresponds to a dark spot for l1, then it will also correspond to a dark
spot for l2. As a result, the fringe pattern will reappear but the contrast will be somewhat
È 1 O1O 2 Ø
weaker. If the lens is moved upward, the rings will reappear for a distance 2 y1 É Ù . This
Ê 2 'O Ú
forms the working principle of Michelson interferometer to measure the difference in wavelength
of the sodium doublet.
Before we discuss Michelson interferometer, you should solve a Practice Exercise.

Practice Exercise 4.5 Newton’s rings experiment is used with a source emitting colours of
wavelengths l1 = 800 nm and l2 = 600 nm. It is observed that the nth dark ring corresponding
to l1 coincides with the (n + 1)th dark ring due to l2. If the radius of curvature of the curved
surface is 2.0 m, calculate
(i) the value of n and
(ii) the diameter of the nth dark ring due to l1. [Ans. (i) 3; (ii) 43.82 ´ 10–4 m]

We now know that interference between two coherent beams can be obtained by division
of wavefront as well as by division of amplitude. Accordingly, the apparatus used to observe
the phenomenon can be divided into two main classes. The Young’s double slit arrangement,
Fresnel’s biprism, Lloyd’s mirror, etc,. belong to the former category. Broadly speaking, we can
categorize these instruments as interferometers. An interferometer which divides a wave by
partial reflection and partial transmission such that the resulting waves have reduced intensity
(or amplitude) belongs to the latter category. Michelson interferometer is an important example
of the second class. In this interferometer, the original wave is split by a semi-reflecting metallic
film and the two resulting beams are made to interfere subsequently. However, the path difference
between them can be varied at will by moving one of the mirrors or by introducing a refracting
material in one of the beams. A few variants of this interferometer find applications in metrology,
plasma diagnostics and the other allied fields, apart from optics. Now we discuss Michelson
interferometer in some detail.
Interference by Division of Amplitude 133

4.4 MICHELSON INTERFEROMETER


Refer to Figure 4.10. It shows basic configuration of a Michelson Interferometer. Its principal
optical components are two high quality plane mirrors M1 and M2, which are silvered on their
front surfaces and are mounted on two arms. These have very high reflectivity and can be
adjusted in any position by means of fine screws attached to them. The mirror M1 is fixed but
M2 is mounted on a carriage, which can be moved forward as well as backward. P1 and P2
are two identical glass plates, which are parallel to one another and inclined to M1 and M2.
The surface of P1 towards P2 is partially silvered so that it acts as a beam splitter, i.e. a beam
incident on it is divided into reflected and refracted beams of equal intensity. In the normal
adjustment of the interferometer, the mirrors M1 and M2 are perpendicular to one another and
the plates P1 and P2 are at 45º to M1 and M2.
Light from an extended source S is made to fall as a parallel beam on plate P1, which
splits it into two parts of equal intensity through partial reflection (wave 1) and partial transmission
(wave 2). These beams travel along two mutually perpendicular paths. The wave reflected by
P1 and shown as 1 in Figure 4.10, undergoes second reflection at M2 (wave 4). This wave is
partially transmitted through P1 (and is shown as 5). The transmitted wave, shown as 2,
undergoes reflection at M1 (wave 3) and is partially reflected by P1 towards the eye (telescope).
This results in the wave shown as 6 in Figure 4.10. Since waves 5 and 6 arise from the same
incident wave, these are coherent and in a position to interfere.

Figure 4.10 A schematic diagram of Michelson interferometer. S is an extended source, mirror M1 is


fixed and mirror M2 is moveable, P1 is a beam splitter and P2 is the compensating plate.

Note that the wave reflected at the rear surface of P1 propagates towards mirror M2. It means
that the wave reflected at M2 passes through P1 three times. But the light reflected at M1 will
pass through it only once. It means that waves 1 and 2 may not travel equal distance in glass.
To compensate this difference and make sure that both waves travel equal distance in glass, a
compensator plate P2 is placed in the path of the second wave parallel to the beam splitter P1.
Note that P2 is identical to plate P1, but without the reflector coating on it. Usually, the
compensating plate has to be used with white light source (mercury lamp) but it can be dispensed
134 Wave Optics

with while using a monochromatic source such as laser light. The fringes produced by interference
between the two beams can be observed with an unaided eye or with a telescope.
Formation of Fringes: The nature of fringes formed in a Michelson Interferometer
depends on the inclination of mirrors M1 and M2 shown in Figure 4.10. To visualize this, refer to
Figure 4.11, where M 1„ is the virtual image of stationary mirror M1 formed by the beam splitter.
(Depending on the position of mirrors, image M 1„ may be in front of, behind or exactly coincident
with mirror M2.) Similarly, the extended source S has also been brought in line with the direction
of observation. The observed fringes can be interpreted as two-wave interference fringes formed
by an air film bounded between mirror M2 and M 1„ . In replacing mirror M1 by its virtual image,
we have to take note of phase changes, if any, produced by reflections from the beam splitter.

Figure 4.11 Air film equivalence of Michelson interferometer: Formation of fringes and their nature.

When mirrors M1 and M2 are perfectly perpendicular to each other, the bounding surfaces
M 1„ and M2 of the air film will be exactly parallel and thickness of the film so formed equals
the difference in axial distance (d = l1 – l2) of the mirrors from the rear face of plate P1.
The distance between coherent sources from which the light waves appear to originate will be
2d. Can this arrangement be identified with any familiar situation? It is exactly equivalent to the
one used to observe Haidinger fringes from a thick film. So in Michelson interferometer, we
obtain localized fringes at infinity, which can be seen with a telescope. Alternatively, these fringes
can be observed in the focal plane of a converging lens. The symmetry of the optical arrangement
leads to the formation of circular fringes with the centre of the fringe pattern on the optic axis.
For a perfectly collimated beam incident normally on the mirrors, the entire field of view
will be uniformly illuminated. The nature of illumination—brightness, darkness or any value in
between—depends on the thickness d of the air film. For d = 0, i.e. when two paths are equal
and M 1„ coincides with M2, we expect the waves to reinforce each other and form a maximum.
But this is not so because phase change of p takes place on external (air-to-glass) reflection.
It means that when M 1„ coincides with M2, the centre of the field will be dark.
When one of the mirrors is moved through a quarter of a wavelength (d = l/4), the distance
between coherent sources will change by l/2 and the interfering waves go out of phase by p.
Interference by Division of Amplitude 135
But the phase change of p on external (air-to-glass) reflection gives a maximum. Moving the
mirror further by l/4 leads to minimum and so on. The relation
2d = ml m = 0, 1, 2, ... (4.24)
denotes Michelson interferometer equation.
If we look obliquely into the interferometer so that our line of sight makes an angle q with
the axis (Figure 4.12), Eq. (4.24) modifies to
2d cosq = ml m = 0, 1, 2, ... (4.25)

Figure 4.12 Looking obliquely in a Michelson interferometer.

Note that for a given separation between the mirrors and order m, wavelength l and angle
q remain constant. The maxima will be in the form of circles about the foot of the perpendicular
from the eye to the mirrors.
When d is a few centimetre, the rings are very closely spaced. If mirror M2 is moved away or
towards the glass plate P1, the fringes cross the centre of the field of view of the observer.
If the path difference d decreases, Eq. (4.25) implies that radius of a ring characterized by a given
value of m decreases. (This is because the product 2d cosq must remain constant.) Then, fringes
appear to move in and become broader. For a decrease in d by one-half of the wavelength, one
fringe crosses the field of view. When M2 coincides with M 1„, the central fringe covers the whole
field of view. This process reverses if d increases. That is, if either M2 overtakes M 1„ or the mirror
separation increases, fringes become sharp and move out of the centre. A new fringe appears at
the centre for every increase of l/2 in mirror separation. These are shown in Figure 4.13.

Figure 4.13 Fringes observed in Michelson interferometer under different conditions.


136 Wave Optics

When the mirrors are not perfectly perpendicular, the air film between M 1„ and M2 will be
wedge-shaped and result in the formation of straight line fringes of equal thickness for small
values of path difference. Note that in this case, the path difference between the interfering
beams varies primarily due to changes in the thickness of the wedge. As the wedge angle
increases, fringes begin to show curvature.

4.4.1 Adjustment of Michelson Interferometer

To adjust the interferometer, the distances of the mirrors M1 and M2 from the silvered surface
of P1 are made nearly as equal as possible by moving mirror M2. A pinhole is then placed
between the lens and the plate P1. If mirrors M1 and M2 are not mutually perpendicular, four
images of the pinhole will be seen; two by reflection at the semi-silvered surface of plate P1
and the other two by reflection at its other surface. The pair of images obtained by reflection
from the silvered surface of plate P1 will be brighter. Using small screws fixed at the back of
fixed mirror M1, the bright images are made to coincide. The mirrors M1 and M2 are now nearly
perpendicular to each other. The pinhole is then removed.
Since the mirrors are not exactly perpendicular to each other, localized fringes will appear.
At this stage, the eyes should be focused in the neighbourhood of mirror M1. The fine tilting
screws on the fixed mirror are now adjusted to obtain circular fringes with the centre of the
fringe pattern in the middle of the field of view. When this happens, we can be sure that mirrors
M 1 and M 2 are mutually perpendicular. The fringes will be quite thin and sharp.
The position of the movable mirror is now adjusted with the fine pitched screw to equalize the
lengths of the two arms of the interferometer. In this adjustment, the fringes in the field of view
should become broad and fewer in number. If only a few fringes cover the entire field of view,
we can be certain that the arms are nearly balanced.
To produce interference with white light, the path difference between the interfering
waves should not exceed a few wavelengths of light. This condition is first achieved with the
quasi-monochromatic source in place as explained above. The mirror M2 is then moved until
the fringes become straight. At this stage, monochromatic source is replaced by white light
and mirror M2 is further moved in the same direction until the central fringe is achromatic,
i.e. all wavelengths combine to show no colour. This observation is taken as an indication
of the two arms of the Michelson interferometer being exactly balanced. This observation has
been used with great success for calibration of the standard metre with a Michelson
interferometer.

4.4.2 Applications of Michelson Interferometer

We can perform three types of measurements with a Michelson interferometer:


(i) wavelength of light,
(ii) width and fine structure of spectrum lines, and
(iii) refractive indices.
We discuss these now.
Interference by Division of Amplitude 137
(a) Determination of wavelength of monochromatic light
The interferometer is first adjusted for circular fringes. Thereafter, mirror M2 is adjusted so as
to obtain a bright spot at the centre of the field of view. If thickness of the air film is d and
the order of the fringe is n, then we have
2d cosq = nl
At the centre, q = 0 so that the above relation reduces to
2d = nl
If we move M2 away from M1 by l/2, 2d increases by l and n is replaced by (n + 1),
i.e. the centre is now occupied by (n + 1)th bright spot. In fact, each time, M2 moves through
l/2, the next bright spot appears at the centre. If p new fringes appear at the centre of the field
when M2 moves through a distance x, we can write
O
x p
2
so that
2x
O (4.26)
p
Thus we can easily measure the wavelength of light emitted by a monochromatic source if we
can count the number of fringes that appear in moving the mirror M2 through a distance x,
which can be easily measured. The value of l measured by a Michelson interferometer is very
accurate since x can be measured to an accuracy of 10–7 m. For example, if a shift of 1000
fringes is obtained when the movable mirror of Michelson interferometer is moved through a
distance of 295 ´ 10–6 m, the wavelength of light is 590 nm.
This method was used by Michelson for standardization of the metre since split-beams can
be widely separated and any desired path difference can be introduced between them. In fact,
this makes the Michelson interferometer an extremely versatile tool in optical research and
testing. (It played an important role in the development of electromagnetic theory of light
following the negative results of Michelson-Morley experiment.) However, coherence length of
the source and the ability to count a large number of fringes crossing the field of view pose
practical constraints. Michelson used the red cadmium line of wavelength 643.84696 nm as a
reference and defined the metre as
1m = 1553164.13 red cadmium wavelength
In 1960, the metre was expressed in terms of the orange-red line of krypton (86
36Kr) of wavelength
605.78021 nm. The standard metre was defined as equivalent to
1m = 1650763.73 orange-red line of krypton
The precision of measurement allowed detection of a displacement of less than 1/100 of a
fringe, which is less than the width of the lines engraved in the platinum-iridium alloy bar kept
at 0ºC in Paris as International Prototype Metre.
138 Wave Optics

(b) Determination of fine structure of spectral lines


When a source of light emits closely spaced spectral lines, such as sodium doublet, having
wavelengths l1 and l2, each wavelength produces its own system of rings. Suppose that l1 is
only slightly greater than l2. Then, for a small thickness film, the fringes corresponding to these
wavelengths will almost coincide in the entire field of view. But if mirror M2 is moved away
from splitter plate P1, the fringes due to l1 and l2 begin to separate out. For a particular
thickness of the air film, the dark fringes due to l1 will coincide with the bright fringes due to
l2 and become indistinguishable again. Moving mirror M2 farther away will, however, make
them distinct. Suppose that movement of mirror M2 through a distance x makes n dark fringes
due to ll and (n + 1) bright fringes due to l2 to appear at the centre. Then, we can write
O1 O2
x= n ( n  1)
2 2
or
2x
n=
O1
and
2x
n+ 1=
O2
so that
2x 2x
 1
O2 O1
Hence, the difference in wavelengths is given by
O1O2
O1  O2 (4.27)
2x
If l1 » l2, we can replace the numerator on RHS of Eq. (4.27) by l2, where l is the mean of
l1 and l2. Therefore,
O2
O1  O 2 (4.28)
2x
This result shows that once we measure the distance moved by the movable mirror between two
consecutive positions of disappearance of the fringe pattern in the field of view and know the
mean wavelength l, we can easily determine the difference between the two wavelengths.

(c) Determination of refractive index of a thin film


The refractive index of a thin transparent plate of known thickness h can be measured by
introducing it in the fixed arm of the interferometer. This will increase the optical path of this
beam by (m – 1)h. Moreover, insertion of the plate produces a discontinuous shift in the fringe
pattern and the number of fringes that cross the field of view cannot be counted. In fact, if we
use a monochromatic light source, it is impossible to identify as to which fringe in the displaced
set corresponds to one in the original set. For this reason, the interferometer is set to see straight
Interference by Division of Amplitude 139
fringes with monochromatic light and white light simultaneously before the plate is inserted.
We first focus the cross-wire on the achromatic fringe. Then, the given plate is inserted in the
path of one of the interfering waves. Since the wave traverses the plate twice, an extra path
difference of 2(m – 1)h is introduced between the two interfering beams. As a result, the fringe
pattern gets shifted. Therefore, the movable mirror M2 is moved till the fringes are brought back
to their initial positions and the achromatic fringe is made to coincide with the cross-wires. If
the distance moved by the mirror M2 is x, we can write
2x = 2(m – 1)h
or
x
m = 1 (4.29)
h
Alternatively, if p fringes cross the field of view, we can write
2(m – 1)h = pl
or
pO
m = 1 (4.90)
2h
This result shows that once we know p, h and l, we can easily determine the refractive index
of the material of the plate. Alternatively, if we know m, we can determine the thickness of the
plate very precisely.

4.4.3 Jamin’s Interferometer

One of the many variants of Michelson interferometer is Jamin’s interferometer. It is particularly


useful in determining the refractive index of a gas at different pressures. Refer to Figure 4.14,
which depicts Jamin’s interferometer schematically. It consists of two thick glass plates G1 and
G2, which are identical and silvered at their back surfaces. T1 and T2 are evacuated glass tubes

Figure 4.14 Schematics of Jamin’s interferometer.


140 Wave Optics

of equal length l. Monochromatic light from a broad source S placed at the focal plane of a lens
L is split into two parallel beams 1 and 2 by reflection at the parallel faces of glass plate G1.
When these beams reach plate G2, they recombine after reflection and form interference fringes
known as Brewster fringes, which can be seen in the field of view of the telescope. When the
glass plates G1 and G2 are parallel, the path difference between them will be zero.
Next, we allow the gas to fill one of the tubes and note the number of fringes that cross
the centre of the field of view. If n fringes cross the centre of the field of view while the gas
reaches the desired pressure and temperature, we can determine the refractive index of the gas
using the relation
(m – 1)l = nl
The value of (m – 1) is seen to be directly proportional to pressure at a given temperature.
Many a time, it may not be very convenient to count the number of fringes. To avoid this,
two identical compensating plates (of equal thickness) made of the same material are introduced
in the path of beams reflected by glass plate G1. These are shown as C1 and C2 in Figure 4.14.
These plates can be rotated about a common horizontal axis with the help of a calibrated
circular disc, D. When this disc is rotated, the path traversed by one of the waves increases and
for the other wave, it decreases. The circular disc, D. is calibrated by counting the number of
fringes directly and is marked in terms of the refractive index and the number of wavelengths.
In the experimental situation, the tubes T1 and T2 are initially evacuated and using white light,
the telescope is focused such that the central white fringe is in the field of view. Next, the gas
is introduced gradually at the desired pressure and temperature in tube T1. You may now like
to know: What happens to the interference pattern? We expect that the central fringe will shift.
It is brought back to the original position by rotating the compensating plate placed in the path
of wave 2 using the circular disc. Since the circular disc had been calibrated, we obtain the
value of refractive index directly.

4.4.4 Twyman–Green and Mach–Zehnder Interferometers

Twyman–Green and Mach–Zehnder interferometers are some other variants of Michelson


interferometer. While the former is particularly useful for testing optical elements such as lenses
and prisms, the latter is widely used for plasma diagnostics and gas flow, say in a wind tunnel.
You will recall that the field of view of a Michelson interferometer gives uniform illumination
when perfectly collimated light beams fall on the beam-splitter held at 45º and the mirrors M1
and M2 are mutually perpendicular to each other. Twyman–Green interferometer makes use of
this property. The optical element to be tested (lens or prism) is put in one of the arms of the
interferometer. Instead of the plane silvered mirror M1, this arm carries a distortion free spherical
convex mirror M when lenses are tested. (When flat surfaces of prisms and cubes are tested,
the convex mirror is replaced by a flat mirror.) If the lens has no aberrations, the plane wavefront
returning from it is exactly perpendicular to the plane wavefront reflected by the movable mirror
(Figure 4.15). As in a Michelson interferometer, we get uniformly illuminated field of view. But
a distorted fringe pattern characterizes lens aberration.
Interference by Division of Amplitude 141

Figure 4.15 Schematic representation of Twyman–Green interferometer.

Mach–Zehnder interferometer differs from Twyman–Green and Michelson interferometers


in that instead of one beam-splitter, it uses two beam-splitters P1 and P2. While P1 is used for
splitting the incident beam, P2 is used for combining the split beams, as shown in Figure 4.16.
Plane silvered mirrors M1 and M2 reflect the light beams towards second beam-splitter, which
combines the split beams. The centres of the beam-splitters and mirrors lie on the corners of
a parallelogram. As a result, the split-beams travel widely separated paths before these are made
to combine by the second beam-splitter. The test chamber is put in one arm and the compensating
elements in the other so as to equalize optical path lengths. The contours of the fringes determine
changes in local density that take place in the test chamber.

Figure 4.16 Schematics of Mach–Zehnder interferometer.

So far we have discussed interference of two beams which were derived from a single beam
either by division of wavefront or by division of amplitude. The interference pattern was
characterized by sinusoidal variation of light intensity with phase difference between interfering
waves. We now consider interference involving many beams which are derived from a single
142 Wave Optics

beam by multiple reflections. Note that the interference fringes so formed are much sharper than
those by two beam interference and the Fabry–Perot or other interferometers based on multiple
beam interference exploit the high contrast of interference fringes in light transmitted by high
reflectivity films and find wide applications in high resolution spectroscopy.

4.5 MULTIPLE BEAM INTERFEROMETRY: REFLECTIONS


FROM A PLANE PARALLEL BEAM

As before, consider a plane wave incident on a plate of thickness h and refractive index m,
as shown in Figure 4.17(a). Suppose that the (complex) amplitude of the incident wave is A0.
The wave undergoes multiple reflections at both interfaces.

Figure 4.17 Multiple internal reflections in a high reflectivity film.

Let r1 and t1 represent the amplitude reflection and transmission coefficients when the wave
is incident from rarer medium on an interface with denser medium and let r2 and t2 represent
the corresponding coefficients when the wave is incident from denser medium to rarer medium.
Thus the amplitude of successive reflected waves will be

A0 r1 , A0t1r2 t2 eiG , A0t1r23t2 e2iG , ...


where
2S 4SP h cos T 2
G k' ' (4.31)
O0 O0
denotes the phase difference (between two successive waves emanating from the plate) due to
the additional path traversed by the beam in the film. q2 is the angle of refraction inside the film,
h is the thickness of film and l0 is the free space wavelength. If we assume that there are
infinitely many interfering waves in reflected light (no attenuation), the resultant amplitude
(complex) of the reflected wave is given by

Ar A0 [ r1  t1t2 r2 eiG (1  r22 eiG  r24 e2iG  ")]


Interference by Division of Amplitude 143

Note that the infinite geometric series in the parentheses has common ratio r22 eiG and its sum
is finite, since r22  1. Summing up the series, we obtain

È t1t2 r2 eiG Ø
Ar A0 É r1  Ù (4.32)
Ê 1  r22 eiG Ú

For non-absorbing media, i.e. if we assume that there is no loss in intensity of light due to
reflection (r1 = –r2), using the principle of reversibility, we can write

R = r12 r22
and
T = t1t2 = 1 – R
In terms of R and T, we can rewrite Eq. (4.32) as

Ë (1  R)eiG Û 1  eiG
Ar A0 r1 Ì1  iG Ü
A0 r1
Í 1  ( R )e Ý 1  ( R)eiG

Hence, the reflectivity of the film is given by

2 2
Ar 1  eiG
Ir = r12
A0 1  ( R)eiG
2
(1  cos G )  i sin G
= R
(1  R cos G )  iR sin G

Using the relation z2 = zz*, where z* is complex conjugate of z, we can write

(1  cos G ) 2  sin 2 G
Ir = R
(1  R cos G ) 2  R 2 sin 2 G
2 R (1  cos G )
=
1  R 2  2 R cos G
G
4 R sin 2
= 2
G
(1  R) 2  4 R sin 2
2
2G
4 R sin
= 2
2Ë 4R GÛ
(1  R ) Ì1  2
sin 2 Ü
Í (1  R ) 2Ý
144 Wave Optics

G
F (r ) sin 2
= 2 (4.33)
G2
1  F ( r ) sin
2
where we have defined the Coefficient of Finesse as
4R (4.34)
F (r )
(1  R )2
For R = 1, the coefficient of Finesse is small and the reflectivity will be proportional to
G
sin2 . Recall that in case of two-beam interference pattern, the beam intensity is proportional to
2
2 G
cos . This difference arises due to the additional phase change of p in one of the reflected beams.
2
The maximum value of reflectance
F (r )
( I r ) max for d = (2m + 1)p (4.35)
1  F (r )
Note that the maximum value of reflectance remains less than one, except for F(r) ® ¥ or
r ® 1. (This is in contrast to the transmission maxima with unit transmittance, irrespective of
the value of reflection coefficient.) The amplitudes of multiply reflected waves from a high
reflectivity film decrease slowly and lead to an intensity distribution which is much different
from the sinusoidal distribution observed in two-wave interference.
The reflectance of a film with a large value of F(r) is not very sensitive to changes in phase
difference, except when d = 2mp. This condition signifies the minima of the intensity distribution.
So we should expect narrow dark fringes among broad regions of brightness in reflected light
for films with high interface reflection coefficient, as shown in Figure 4.18.

Figure 4.18 Plot of reflectance of a non-absorbing plane parallel film as a function of phase difference
between successive interfering waves for different values of reflection coefficient.

Let us now consider transmitted waves. We can obtain an expression for transmittance, by using
the relation
It = 1 – Ir
Interference by Division of Amplitude 145
Alternatively, we can start from the first principle. For the sake of completeness, we follow the
latter approach and assume that the first transmitted wave has zero phase. Then, we can write
the amplitudes of successive transmitted waves as

A0t1t2 , A0t1t2 r22 eiG , A0 t1t2 r24 e 2iG , ...


Hence the resultant amplitude of the transmitted wave is given by

At A0 t1t2 (1  r22 eiG  r24 e2iG  ")


As before, we note that the infinite series in the brackets is geometric with common ratio r22 eiG .
We can easily write its sum as
t1t2 1 R
At A0 A0
1 r22 eiG 1  ( R )eiG
As before, we can derive expression for transmittivity of the film as
2
At (1  R)2
It =
A0 [1  ( R)eiG ] [1  ( R)e iG ]
(1  R ) 2
=
[(1  R cos G )  iR sin G ] [(1  R cos G )  iR sin G ]
(1  R ) 2
=
(1  R cos G ) 2  R 2 sin 2 G
In terms of the Coefficient of Finesse, we can express film transmittance function as
1
It (4.36)
G
1  F ( r ) sin 2
2
It is known as Airy function or the Airy formula.
From Eqs. (4.33) and (4.36), we observe that when a beam of light undergoes multiple
reflections and transmissions from a plane parallel film, its reflectivity and transmittivity add up
to unity. Further,
(It)max = 1 for d = 2mp m = 1, 2, 3, ... (4.37)
In Figure 4.19, we have plotted transmittivity It as a function of d for a few representative
values of the reflection coefficient. As may be noted, the transmittance peaks are characterized
by unit transmittance, irrespective of the reflection coefficient. However, the minimum film
transmittance depends on the reflection coefficient of the film:
1
( I t )min for d = (2m + 1)p (4.38)
1  F (r )
Note that for films of low reflection coefficient, transmittance does not drop much below one
with the result that visibility of the interference fringes is low. However, for the large reflection
146 Wave Optics

Figure 4.19 Plot of transmittivity of a plane parallel film as a function of phase difference between
successive interfering waves for different values of reflection coefficient.

coefficient, the denominator of the Airy function becomes quite large and transmittance drops
rapidly as we move away on either side of the transmittance peak. It means that for films of
higher reflectivity, the bright interference fringes are sharper. This gives rise to extremely sharp
and bright interference fringes separated by broad regions of almost complete darkness in the
transmitted light.

Practice Exercise 4.6 Consider an absorbing film defined by r2 + tt¢ + A = 1, where A


denotes the absorptance of the film. Show that transmittance of the film is given by

( I At )max
I At
G
1  F (r ) sin 2
2
where peak transmittance (IAt)max is given by
2
È A Ø
I At É1  1  R Ù
Ê Ú

Discuss physical implications of the result.

We now discuss interferometers based on multiple beam interference. We first discuss


Fabry–Perot interferometer.

4.5.1 Fabry–Perot Interferometer

Refer to Figure 4.20, which shows schematic representation of a Fabry–Perot interferometer.


In its simplest form, a Fabry–Perot interferometer consists of two identical, optically plane glass
or quartz plates with highly reflecting silver (or aluminium) coatings on inner sides facing each
other. These plates are arranged so that these enclose a plane parallel slab of air between the
coated surfaces. Screws are provided for finer adjustments. To prevent formation of undesirable
fringes due to multiple reflections in the plates, the outer uncoated surface of each plate is
slightly wedged. One of the two plates is fixed and the other one can be moved to vary the
Interference by Division of Amplitude 147

separation between the plates. Sometimes a fixed distance is maintained between these plates
with the help of spacers. (If the reflecting glass plates are parallel to each other and have fixed
spacing between them, we have what is known as Fabry–Perot etalon.)
S1 is a broad source of monochromatic light and convex lens L1 helps to collimate the beam
emanating from it. A large number of plane waves enter the interferometer at different angles
with the optical axis of the interferometer. Each incident wave undergoes multiple internal
reflections successively at the silvered surfaces. One such wave is shown in Figure 4.20.
(At each reflection, a small fraction of light may be transmitted.) So, we find that each incident
wave gives rise to a large number of coherent and parallel waves in transmitted light with a
constant path difference between any two successive waves. These waves are brought to focus
by lens L2 and produce interference fringes, which are obtained on screen S2.
If the separation between the silvered surfaces is d and the angle of incidence of a wave
is q, the path difference between any two successive transmitted waves corresponding to the
incident wave will be 2d cosq. Since the space between two mirrors is filled with air, phase
changes of p occur at both surfaces. Hence, the condition for maximum intensity can be
expressed as
2d cos q = ml (4.39)
where m is an integer. This condition is the same as that for two-wave interference.

Figure 4.20 Schematics of Fabry–Perot interferometer.

Note that the condition for maximum intensity depends only on angle of incidence rather
than the exact location of the point on the extended source. Therefore, in the absence of
collimating lens L1, all waves emanating from different points of the extended source but
travelling in the same direction will reinforce each other. However, since different points on an
extended source are mutually incoherent, only those emergent waves which undergo multiple
reflections within the interferometer but originate from the same point on the source can interfere.
However, note that for waves which originate from different points of the source but propagate
along parallel directions and converge on the same point in the focal plane of the converging
lens L2, we have to add intensities rather than amplitudes.
Proceeding further, we recall that locus of points in the source, which give waves of
constant inclination, is a circle. Hence, with an extended source, the interference pattern obtained
with Fabry–Perot interferometer consists of a system of bright concentric rings (circular fringes)
against a wide dark background and each ring corresponds to a particular q. That is, these are
148 Wave Optics

fringes of equal inclination and have an axis of symmetry (Figure 4.21). The fringe at the centre
of the pattern has the highest order (Haidinger fringes). Now compare Figures 4.13 and 4.21,
which show fringe patterns obtained with Fabry–Perot interferometer and Michelson
interferometer. Is there any difference between them? Note that although qualitative nature in
both cases is the same, there are differences in detail. In particular, the fringes in the light
transmitted by the Fabry–Perot interferometer are extremely sharp. That is why Fabry–Perot
interferometer is more suited to study the hyperfine structure of spectral lines.

Figure 4.21 Fringe pattern obtained in Fabry–Perot interferometer.

Before proceeding further, read the following example carefully.

EXAMPLE 4.11 Consider a Fabry–Perot etalon whose mirrors are at a fixed distance of
1.2 ´ 10–2 m with air between them. Calculate the angle of incidence at which bright fringes
will be obtained for spectral lines of wavelengths l1 = 600 nm and l2 = 599.998 nm.
Solution For spectral line of wavelength l1 = 600 nm, Eq. (4.39) gives

È m – 600 – 109 m Ø
cosq = É 2 Ù
Ê 2 – 1.2 – 10 m Ú
so that
1 È m Ø
q = cos É 4Ù
Ê 4 – 10 Ú

This result shows that the first bright fringe will form at q = 0º corresponding to m = 40000.
Other bright fringes will form at q = 0.41º, 0.57º, ... for m = 39999 and m = 39998, respectively.
These are shown as solid curves in Figure 4.22.
For l2 = 599.998 nm, you will find that bright fringes will occur at q = 0.162º, 0.436º,
0.595º at aforementioned values of m. These spectral lines are shown as dashed curves in Figure
4.22. Note that even though these spectral lines differ in wavelength by 0.002 nm, these are
quite well resolved. This characteristic of Fabry–Perot interferometer is used to test flatness of
coated surfaces with very high degree of precision.
Interference by Division of Amplitude 149

Figure 4.22 Variation of intensity of spectral lines corresponding to l = 600 nm and 599.998 nm
with angle of incidence in a Fabry–Perot interferometer whose mirrors are separated by
1.2 ´ 10–2 m.

4.5.2 Width of Transmission Peaks

Refer to Figure 4.20 again. If separation between plates of a Fabry–Perot interferometer is d and
the waves are incident normally (q = 0), the phase difference between successive waves d = kD
= 2nkd. Let us now calculate the wave number and frequency spreads of the light coming out
of the interferometer when wave number of the light entering the interferometer is changed.
Figure 4.23 shows a typical transmission profile of a Fabry-Perot interferometer with peak
transmission at nkd = mp. Suppose that the wave numbers corresponding to the points on the
transmission curve at Full Width at Half Maximum (FWHM) are (k – dk) and (k + d k). Hence,
Eq. (4.36) gives
1 1
2
2 1  F (r ) sin ( mS “ nd G k )

Figure 4.23 Transmission profile of a Fabry–Perot interferometer.

This will be satisfied if


F (r ) sin 2 ( “ ndG k ) 1
150 Wave Optics

For large values of F(r), sinq » q and we can write

1
Gk
nd F ( r )

Hence, FWHM for a spectral line subject to above conditions is given by

2 1 (1  R)
FWHM 2G k (4.40)
nd F (r ) nd R

In many situations, it is more convenient to work in terms of frequency or wavelength. You can
readily express this result in units of frequency and wavelength as

c c 1 R
Gf (4.41a)
O 2nS d R
and
O2 1  R
GO (4.41b)
2nS d R
Before proceeding further, you may like to convince yourself about the truthfulness of Eq. (4.41).
For this, you may like to solve a Practice Exercise.

Practice Exercise 4.7 Starting from Eq. (4.40), derive Eqs. (4.41a) and (4.41b).

It may be pointed out here that widths of spectral lines observed in transmission light can
be reduced by increasing film reflectivity and/or plate separation. A few typical values of
FWHM for Fabry–Perot interferometer peaks are given below.
d f = 10 MHz for r = 99% and d = 10 cm
= 10 MHz for r = 99.99% and d = 1 mm
= 1 MHz for r = 99.99% and d = 1 cm

4.5.3 Sharpness of Spectral Lines: Spectral Resolution

The minimum wavelength separation between two spectral lines that an instrument can distinguish
distinctly defines the limit of its spectral resolution. However, the arbitrariness as to when two
spectral lines can be considered to be distinct or resolved has led to several resolution criteria.
Of these, the most widely used resolution criterion is due to Rayleigh. It makes use of the
minimum intensity point. (We will discuss about it in detail in Chapter 5 on diffraction.)
For a Fabry–Perot interferometer, two spectral lines are said to be just resolved if their half
intensity points coincide. This is shown in Figure 4.24. Alternatively, we can say that two
spectral lines with identical sharpness are said to be just resolved by a Fabry–Perot interferometer
if their peaks are separated by the FWHM of each spectral line. When this happens,
the minimum of the resultant intensity distribution is about 74% of the corresponding maximum
value. We will not go into mathematical details here.
Interference by Division of Amplitude 151

Figure 4.24 Resolution of two spectral lines by a Fabry–Perot interferometer: Spatial variation of
intensities of individual as well as total intensity when two lines are just resolved.

Now we know that sharpness of fringes of a Fabry–Perot interferometer depends on the


reflectivity of the plates. It means that we will require highly reflecting coatings for greater
sharpness. But we cannot use very thick coating of metals as intensity of the beam decreases due
to absorption and scattering losses in optical coatings. To overcome this difficulty, we use the
phenomena of total internal reflection at an interface; light is made to fall at angles very close to
but below the critical angle. The Lummer–Gehrcke (LG) interferometer makes use of this
phenomenon for high reflection coefficient. We discuss it now.

4.6 LUMMER–GEHRCKE INTERFEROMETER

A Lummer–Gehrcke interferometer, also known as LG plate, consists of a glass or quartz plane


parallel plate. The plate is a few mm thick, several mm in width and a few tens of cm in length.
A right angled prism of the same material is optically bonded to one end of the plate, as shown
in Figure 4.25. The angle of the prism is chosen so as to obtain the desired angle of incidence
(less than the critical angle) for the internal reflections. Since the surfaces of the
LG-plate are parallel, all successive (multiple) internal reflections within it give rise to waves
propagating in parallel directions at near grazing angles from its upper as well as lower surfaces.
These waves can be focused with a lens just as in a Fabry–Perot interferometer and produce
nearly straight line interference fringes parallel to the surfaces of the plate on its either side.
We will not go into the details of the underlying theory. However, two points need special mention:

Figure 4.25 A Lummer–Gehrcke plate.


152 Wave Optics

• Unlike in the case of Fabry–Perot interferometer, the space between the reflecting surfaces
is a dispersive medium.
• The number of reflections and resolving power depend on the length of the plate.
The LG-plates have been quite useful in the study of fine structure of spectral lines in
the ultraviolet region. But it has been replaced by Fabry–Perot interferometer because of
flexibility and higher resolution capability.
We now summarise what you have studied in this chapter

4.7 SUMMARY

• For a thin film in reflected light, the conditions for constructive and destructive interference
are expressed as
O
2mh cosf = (2n  1) n = 0, 1, 2, 3, ... maxima
2
2mh cosf = nl n = 1, 2, 3, ... minima
where m is the refractive index of the film of thickness h, f is the angle of refraction in
the film and l is the wavelength of the light used to illuminate the film.
• For a thin film in transmitted light, the conditions for constructive and destructive
interference are expressed as
2mh cosf = nl n = 1, 2, 3, ... maxima
O
2mh cosf = (2n  1) n = 0, 1, 2, 3, ... minima
2
• Fringes of equal inclination are obtained in thin films of uniform thickness depending on
the angle of inclination f inside the film. Fringes of equal inclination are also called
Haidinger fringes.
• Fringes of equal thickness are obtained in films of variable thickness mainly due to
changes in m.
• The fringe width, which denotes the spacing between two consecutive bright (or dark)
fringes produced by a wedge-shaped film, is given by
O
E
2PT
where q (measured in radian) is the wedge angle.
• In Newton’s ring experiment, the radii of dark rings are proportional to the square roots
of natural numbers, the radius of curvature of the plano-convex lens and the wavelength
of light used:
rn nRO n = 1, 2, 3, ...
• The radii of bright Newton’s rings are proportional to the square roots of odd natural
numbers, the radius of curvature of the plano-convex lens and the wavelength of light used:
Interference by Division of Amplitude 153

(2n  1) O R
rn n = 1, 2, 3, ...
2
• In Newton’s ring experiment, the wavelength of light can be expressed in terms of the
diameter of a ring and radius of curvature as

Dn2 p  Dn2
O
4 pR
• The Michelson interferometer, which uses an extended monochromatic source, gives circular
fringes when the mirrors M1 and M2 are mutually perpendicular. At any other angle,
a pattern of straight parallel fringes are obtained.
• Michelson interferometer finds applications in the measurement of refractive index of a
thin film and wavelengths of two closely spaced spectral lines.
• Jamin’s interferometer, Twyman–Green and Mach–Zehnder interferometers are variants of
Michelson interferometer.
• Fabry–Perot interferometer, which is based on the principle of multiple beam interference,
produces very sharp fringes.
• In Fabry–Perot interferometer, we observe fringe pattern formed by transmitted light.
The intensity of the pattern is given by

1
It
G
1  F ( r )sin 2
2
where F(r) = 4R/(1 – R)2 is known as coefficient of Finesse.
• The Lummer–Gehrcke interferometer is based on the phenomenon of total internal reflection.
It produces straight line fringes. It is particularly useful in the study of fine structure of
spectral lines in the ultraviolet region.

REVIEW EXERCISES

1. The radii of the nth and (n + 5)th dark rings in a Newton’s ring experiment are 4 mm and
6 mm respectively. The radius of curvature of the lower surface of the lens is 10 m.
Calculate the wavelength of light used as well as n. [Ans. 400 nm; 4]
2. Newton’s rings are formed by reflected light of wavelength 600 nm using a combination
of a plane plate of glass and a plane-convex lens. On introducing a liquid of refractive
index 1.326 between the lens and the plate, the diameter of the 7th dark ring decreases
by 0.054 cm. Calculate the radius of curvature of the plane-convex lens. [Ans. 1 m]
3. White light is used for producing fringes by an air film of thickness 10 mm. It is incident
at an angle q and the fringes are examined by a spectroscope. 20 dark bands are seen in
the spectrum in the wavelength range from 400 nm to 700 nm. Calculate q.
[Ans. cos–1(0.933)]
154 Wave Optics

4. Calculate the thickness of a soap film (m = 1.33) that gives constructive second order
interference of reflected red colour (l = 700 nm) incident in the form of a parallel beam
at 30º with the normal. [Ans. 425.8 nm]
5. A wedge-shaped film of air is illuminated by a monochromatic beam of light (l =
465.5 nm). The wedge angle is 40¢¢. Calculate the distance between the two consecutive
fringes. [Ans. 0.12 ´ 10–2 m]
6. A parallel beam of sodium light (l = 589 nm) is incident on a thin glass plate of refractive
index 1.5 such that the angle of refraction into the plate is 60º. Calculate the minimum
thickness of the plate for which we will record complete darkness by reflection.
[Ans. 392.7 nm]
7. The diameter of the 10th dark ring in a Newton’s ring experiment is 0.5 cm in the reflected
system. Calculate the thickness of air film at the position and the radius of curvature of
the lens. Given l = 590 nm. [Ans. 2.95 ´ 10–6 m; 1.06 m]
8. Newton’s rings are formed with reflected light of wavelength 589 nm using a plano-
convex lens and a plane glass plate with a liquid between them. The diameter of the third
bright ring is 2 mm. If the radius of curvature of the plano-convex lens is 0.9 m, calculate
the refractive index of the liquid. [Ans. 1.86]
9. In a Michelson interferometer, when the movable mirror moves through a distance of
0.233 mm, the number of fringes counted are 790. Calculate the wavelength of light used.
[Ans. 589.9 nm]
10. A Michelson interferometer is adjusted to give a fringe pattern of concentric circles when
illuminated by an extended source of light (l = 500 nm). How far must the movable arm
be displaced for 1,000 fringes to emerge from the centre? If the centre is bright, calculate
the angular radius of the first dark ring in terms of the path difference between the two
arms and l. [Ans. 0.025 cm, 1.81º]
CH A P T E R 5
FRESNEL DIFFRACTION

EXPECTED LEARNING OUTCOMES


After reading this chapter, the student will acquire the capability to:
• Discuss simple situations/experiments to illustrate diffraction phenomenon.
• Differentiate between Fresnel and Fraunhofer class of diffraction patterns.
• Explain the concept of Fresnel half-period zones and apply it to explain diffraction patterns of
some common apertures.
• Draw Cornu's spiral and use it to explain diffraction pattern of a few typical obstacles.
• Carry out rigorous mathematical analysis for an obstacle of regular shape and apply it to explain
diffraction pattern of a circular aperture.
• Solve numerical problems.

5.1 INTRODUCTION

In the preceding two chapters, we learnt that when light from two coherent sources is made to
superpose, redistribution of energy manifests in the formation of beautiful fringes of different
shapes. We observe brilliant colours in the region of overlap. The interference phenomenon was
explained on the basis of wave theory of light. Another phenomenon associated with light was
reported by Francesco Maria Grimaldi, an Italian mathematician, when he observed the presence
of dark and bright bands near the edges of geometric shadows of objects. He termed the bending
of light across edges diffraction. A systematic explanation of the phenomenon was given by
Augustin Jean Fresnel, when he extended Huygens’ hypothesis of wave propagation to include
mutual interference between secondary wavelets from different parts of the same wavefront by
taking phase difference into account.
As we now know, diffraction is a universal wave phenomenon and a part of our common
experience. It is observed for sound, light or matter waves and occurs whenever a wavefront
or part of it is obstructed by an obstacle. It makes light passing through a small opening spread
into the region of geometrical shadow. Fresnel showed that the ease with which a wave
bends around corners depends on the size of the obstacle/aperture relative to its wavelength.
For example, music and speech wavelengths lie in the range of 1.7 cm to 17 m, whereas a door
155
156 Wave Optics

has an aperture of about 1 m. As a result, long wavelength waves bend more readily and make
sound audible behind edges of doors. (Does it give explanation for the Hindi proverb: Deewar
ke bhi kaan hote hain? Translated literally, it means even walls have ears and confidential
matters should be kept close to oneself.) On the other hand, the wavelength of light is about
600 nm and the obstacles used in ordinary experiments are 104–105 times bigger; being a few
centimetres in size. For this reason, diffraction of light is not visible so readily. Yet some
situations present themselves so naturally. In Section 5.2, we have discussed a few of these
which facilitate us to observe diffraction outside a physics laboratory.
For mathematical convenience and ease in understanding, diffraction is classified in two
categories: Fraunhofer diffraction and Fresnel diffraction. In Fraunhofer diffraction, the source of
light and the observation screen (or retina in human eye) are effectively at infinite distance from
the object (the diffracting aperture). For this reason, Fraunhofer diffraction is also known as far-
field diffraction. This condition is achieved by putting suitable (convex) lenses between the
source, the obstacle and the observation screen. The wave incident on the diffracting aperture
is a plane wave and the diffraction pattern is observed in the focal plane of a convex lens. (A
lot of good physics is involved in Fraunhofer diffraction and it is of particular interest to
understand the theory of optical instruments, whose resolution is diffraction-limited. That is,
diffraction places a fundamental restriction on optical instruments, including human eye, in
respect of resolution of objects.)
In Fresnel diffraction, also known as near-field diffraction, the source or the screen or both
are at a finite distance from the diffracting object. In Section 5.3, we begin by discussing Fresnel
diffraction and study its spatial evolution with reference to transition to Fraunhofer region.
You will note that Fraunhofer diffraction is a special case of Fresnel diffraction. In Section 5.4,
you will learn about Fresnel construction, which consists of dividing the wavefront into annular
spaces enclosed by concentric circles. We introduce the concept of Fresnel half period zones to
develop a qualitative understanding of diffraction pattern. You will note that Fresnel diffraction
at a point is obtained by summing contributions of wavefronts from different Fresnel zones,
unobstructed by the obstacle. A special optical device designed to block light from alternate
half-period elements, known as zone plate, is also discussed. It provided experimental evidence
in favour of Fresnel theory.
In Section 5.5, you will learn to use Fresnel construction to analyse diffraction patterns
produced by a circular aperture and a straight edge. We can use a graphical method based on
vector addition of amplitudes with some advantage to obtain their resultant at an external point
due to all secondary wavelets originating on a wavefront (spherical or cylindrical). This method
leads to the so-called Cornu’s spiral and helps to gain clearer physical insight into the origin
of the diffraction pattern. In Section 5.6, we have discussed it in some detail.
You will realise that Fresnel construction as well as Cornu’s spiral method are approximate.
In Section 5.7, you will learn a rigorous mathematical analytical analysis for a regular-shaped
obstacle and use it to explain the diffraction pattern of a circular aperture.

5.2 OBSERVING DIFFRACTION


We know that wavelength of visible light lies in the range of 400 nm to 700 nm and the
prerequisites to observe diffraction of light are:
Fresnel Diffraction 157
(i) a narrow and preferably monochromatic source of light,
(ii) a sharp-edged obstacle, and
(iii) an observation screen, which could be retina in human eye.
These pre-requisites are fulfilled in the following situations:
1. Look at a distant mountain just before the sun rises behind it. You will observe a
luminous border around the upper part of solar profile. This is due to the diffraction of
light by the mountain.
2. Look at a distant street light at night with nearly shut eye lashes. The light seems to
streak out from the bulb. This is because of bending of light around the corners of eyelids.
3. Take a piece of fine cloth, say fine handkerchief or muslin cloth. Stretch it flat and keep
it close to your eyes. Now focus your eyes on a distant street lamp through it. You will
observe a regular pattern of coloured spots arranged along a rectangle. It arises due to
diffraction of light by the mesh (criss-cross) of fine threads in the handkerchief. If you
rotate the handkerchief in its own plane, the pattern will be seen to rotate. If you take
a cotton blanket and look at the sun through it on a bright sunny day, you will observe
a similar pattern.
4. Take a small ball bearing on a glass plate with beeswax. Make sure that no wax spreads
beyond the rim of the ball. Place it in the path of monochromatic light diverging from
a pinhole. You should observe a bright spot, called Poisson spot at the centre of the
shadow. This observation proved unchallengeable evidence for diffraction of light.
5. Take a pair of razor blades and one clear glass electric lamp. Hold the blades so that
the edges are parallel and form a narrow slit in between. Bring this arrangement close
to your eyes and parallel to the bulb filament. By adjusting the width of the slit, you will
observe bright and dark bands due to diffraction of light by the slit.

The Poisson Spot

Poisson, the famous mathematician, was a member of the committee


appointed to judge Fresnel’s dissertation. He was a great supporter
of corpuscular theory of light. To disprove Fresnel, and hence wave
theory, he argued as follows:
Consider the shadow of a circular object formed by a point source,
O. According to wave theory, the spherical waves from the source
reach the periphery in phase since these cover equal distance from
the source. So, the waves starting from the rim PP¢ reach the screen
at C in phase at the centre of the shadow (Figure 5.1). This should
lead to formation of a bright spot at the centre of the shadow. Figure 5.1 The Poisson spot
Poisson considered this logic absurd and against common sense. and diffraction.
(Probably Poisson was not aware that the central bright spot had
been observed by Maraldi almost a century ago.) Immediately after Poisson’s objection, Arago performed
the experiment using 2 mm diameter disc and he also established the existence of the central bright
spot. This settled the issue in support of Fresnel and wave theory.
158 Wave Optics

5.2.1 Producing Diffraction Pattern

We now know that in Fresnel diffraction, the source or the screen or both are at a finite distance
from the diffracting object. Several researchers have observed and analysed Fresnel diffraction
patterns generated by obstacles of different shapes and sizes. In India, a systematic study of
Fresnel diffraction patterns from small spheres, discs and apertures of different shapes—circular,
elliptical, square and triangular—was undertaken by Y.V. Kathavate under the guidance of Prof.
C.V. Raman using a very simple arrangement shown in Figure 5.2. (This was an excellent
demonstration of the fact that nature likes simplicity!)
Kathavate used a nearly 5 m long light tight box with a fine pinhole at one end. A convex
lens was used to focus light from a 100 W incandescent lamp and a red filter was inserted
between the lens and the pinhole to obtain monochromatic light of wavelength 632 nm. Inside
the box, the obstacle was placed at about 2 m from the pinhole and the photographic plate was
mounted on a movable stand so that its distance from the obstacle could be varied. The spherical
obstacles used in these experiments were in the form of steel ball bearings of radii 1.58 mm,
1.98 mm, 2.37 mm and 3.17 mm and kept on a glass plate. They also worked with discs of these
sizes. The diffraction patterns were recorded on a photographic plate, which was kept at 5 cm,
10 cm, 20 cm, 40 cm and 180 cm from the obstacle.

Figure 5.2 Schematics of the experimental set up used by Kathavate to observe Fresnel diffraction.

Refer to Figure 5.3. It shows enlarged view of the


diffraction pattern obtained by Kathavate from a disc
of radius 1.58 mm when separation of the obstacle
from the photographic plate was 180 cm.
Note that the diffraction pattern shows markedly
clear circular fringes around the central spot. Similar
results were observed for spherical ball bearings.
The simultaneous existence of the Poisson spot and
diffraction rings around it suggests that light bends
around opaque obstacles and constitutes the most direct
evidence of non-rectilinear propagation of light.
It may be mentioned here that Fresnel diffraction Figure 5.3 An enlarged view of the
Fresnel diffraction pattern
patterns change with distance of the source and the screen
obtained by Kathavate for a
from the obstacle. Let us now understand how spatial disc of radius 1.58 mm.
evolution of Fresnel diffraction pattern takes place.
Fresnel Diffraction 159

5.3 SPATIAL EVOLUTION OF FRESNEL DIFFRACTION PATTERN

To study transition of Fresnel diffraction with distance, refer to Figure 5.4(a). A point source
of light is located in the focal plane of a converging lens L. It changes the spherical waves
originating from the source into plane waves. So, the wavefront is rendered parallel to the
diffracting screen with a narrow opening in the form of a long narrow slit, as shown in
Figure 5.4(b). On passing through the slit, we note that the diffracting waves are also plane and
have an angular spread. Let us now discover the shape, size and intensity distribution in the
diffraction pattern on the observation screen. We note that
• When the incident wavefront is parallel to the diffracting screen and the observation
screen is placed next to the aperture, a small portion of the screen is uniformly illuminated.
This manifests as a vertical patch of light whose size (in width and height) is equal to
that of diffracting slit. The remaining portion of the screen is dark. This is shown in
Figure 5.5(a). From P to A, the intensity is zero. At A, it abruptly rises to a maximum
value (I0) and remains constant from A¢ to B¢, i.e. over the width b of the slit before
dropping to zero again. So we can say that AA¢B¢B represents the edges of the geometrical
shadow, which implies that the law of rectilinear propagation of light holds when the
distance between the aperture and the screen is very small.

Figure 5.4 (a) Schematic depiction of the arrangement used to observe spatial evolution of Fresnel
diffraction, and (b) cross-sectional view of the geometry shown in (a) above.

• As we move the observation screen away from the diffracting aperture, the patch of light
AA¢B¢B begins to lose sharpness. If the distance d between the aperture and the screen
is large compared to the width of the slit (d > b), a careful observation shows a few
fringes at the edges of the patch of light. The intensity distribution shows diffraction
rippling effect somewhat like that shown in Figure 5.5(b).
160 Wave Optics

Figure 5.5 Spatial evolution of Fresnel diffraction pattern.

• When d » 1 m and b » 0.1 mm, i.e. d ? b, the fringes appearing close to the edge of the light
patch spread out and the geometrical image becomes increasingly obscured (Figure 5.5(c)).
If d is increased gradually, diffraction effects become progressively more pronounced.
• When d is very large, diffraction ripples stabilize. It means that we are now in the
Fraunhofer region. To observe the shape of the resultant pattern, you should put a convex
lens in-between the diffracting aperture and the observation screen and arrange it so that
the screen is in its second focal plane. You will observe the diffraction pattern as shown
in Figure 5.5(d–f). We will discuss about it in detail in the next chapter.
From the above discussion, we may conclude, and quite logically, that Fresnel diffraction
changes significantly as the separation between the obstacle and the observation screen changes.
We first explain these observations qualitatively on the basis of Fresnel’s (geometrical)
construction. This will be followed by a more rigorous mathematical analysis.

5.4 FRESNEL CONSTRUCTION

Consider a plane wavefront represented by WW ¢ propagating towards the positive x-axis,


as shown in Figure 5.6(a). We first calculate the field due to all disturbances reaching an
arbitrary point P0 on the screen placed at a distance d from the wavefront. Then, we will
examine the effect of an obstacle on the intensity at the point of observation using
Huygens–Fresnel principle. One way would be to write the equation of motion of each disturbance
(wavelet) reaching the point of observation from different portions of the wavefront and then
add them together. But this method is mathematically cumbersome because
(i) we have to consider an infinite number of points and each point acts as a source of
secondary wavelets, and
(ii) the phase of secondary wavelets reaching the point of observation is different as these
travel different distances before reaching there.
To overcome these difficulties, Fresnel devised a simple geometrical method, which provided
very useful insight and beautiful explanation of diffraction phenomenon from small obstacles.
Now we will learn about it.
Fresnel Diffraction 161

Figure 5.6 Fresnel construction (a) propagation of a plane wavefront and (b) division of wavefront
into annular spaces.

Fresnel argued that it is possible


1. to locate a series of points situated at the same distance from the point of observation
so that all the secondary wavelets originating from them travel the same distance, and
O
2. to locate the locus of the points from where the wavelets travel a distance d +ÿ ,
2
3O
d + l, d + , ... .
2
Fresnel construction consists of dividing the wavefront into annular spaces enclosed by
concentric circles, as shown in Figure 5.6(b). The resultant at the point of observation is
obtained by adding contributions of wavelets from these annular spaces, called half-period
zones or half-period elements. When an obstacle is introduced between the wavefront and the
screen, some of the half-period elements will be obstructed depending on its shape and size.
As a result, the secondary wavelets from the unobstructed elements only will reach the point
of observation on the screen and their resultant can be easily calculated by summing their
contributions. We now discuss Fresnel construction and half-period zones.

5.4.1 Half-period Zones


To discuss the concept of Fresnel’s half-period zones, we assume that monochromatic light is
coming from infinity so that we have to consider plane wavefronts (Figure 5.7). WW¢F ¢F is a
plane wavefront propagating along the positive x-axis. To determine the resultant amplitude of
the field at an arbitrary point P0 due to superposition of all secondary Huygens’ wavelets
originating from the wavefront, we divide the wavefront into half-period zones using the following
construction: From the point P0, we drop a perpendicular P0O on the wavefront. The point O
is called the pole of the wavefront with respect to the point P0. Let us denote the distance
between the foot of the perpendicular O and point P0 as d. Next, with P0 as centre, we draw
O 3O
spheres of radii d + , d + l, d + , ... and so on. These spheres will intersect the wavefront
2 2
WW¢F ¢F in a series of concentric circles with centre O and radii OQ1, OQ2, OQ3, ..., as shown
in Figure 5.7. This geometrical construction divides the wavefront into circular strips called
162 Wave Optics

Figure 5.7 Construction of Fresnel half-period zones on a plane wavefront.

zones. The first zone is the space enclosed by the circle of radius OQ1, the second zone is the
annular space between circles of radii OQ2 and OQ1. Similarly, the third zone is the annular
space between circles of radii OQ2 and OQ3 and so on. These concentric circles or annular rings
are called Fresnel zones or half-period elements. The genesis of this nomenclature is in the fact
that the path difference between the wavelets reaching P0 from the corresponding points in
successive zones is one-half wavelength.
To calculate the resultant amplitude at P0 due to all the secondary wavelets originating from
the entire wavefront, we first consider an infinitesimal area dA of the wavefront. We assume that
the amplitude at the point of observation due to the area under consideration is
1. directly proportional to the area, dA;
2. inversely proportional to the distance of dA from P0; and
3. directly proportional to the obliquity factor (1 + cosq) for Huygens secondary wavelets,
where q is the angle between the normal drawn from the wavefront at dA and the line
joining dA to P0.
As we move away from O, the value of q increases and takes the value p/2 for a point at
infinite distance on the wavefront. Physically it ensures that wavefront moves forward and there
is no backward wave. Moreover, for q = p/2, the amplitude falls to one-half and the intensity
drops to one-quarter of its maximum value, which occurs for q = 0. The obliquity factor takes
values 2, and 0 for forward direction and backward direction, respectively.
If u1, u2, u3, ..., un denote the resultant amplitudes at P0 due to all secondary wavelets from
the 1st, 2nd, 3rd, and nth zone, respectively, we can write
An
un constant – (1  cos T ) (5.1)
dn
Fresnel Diffraction 163
where An and dn respectively denote the area and the average distance of the nth zone from the point
of observation. q denotes the angle at which light leaves the zone. For simplicity, we write dn = d.
From Eq. (5.1) we note that to know the amplitude of secondary wavelets arriving at the
point of observation from any zone, we must know An. This, in turn, requires the knowledge
of the radius of the corresponding circle defining the boundaries of the Fresnel zone concerned.
Proceeding further, we denote the radii of different half-period zones as OQ1 = r1,
OQ2 = r2, OQ3 = r3, ..., OQn = rn. By applying Pythagoras’ theorem, we can write the expression
for the radius of the nth circle (zone) as
1
2
ËÈ nO Ø 2
Û2
rn = ÌÉ d  Ù d Ü
ÍÊ 2 Ú Ý
1
2 2Ø2 1
È n O nO Ø 2
= É nd O  Ù = nd O
È
É1  Ù
Ê 4 Ú Ê 4d Ú
Since wavelength of visible light is much less than the distances involved in practical systems and
ordinary measurements/experiments (l = d), we can neglect the term nl/4d in comparison to
unity, provided n is not very large. Hence, the expression for radius of the nth zone reduces to
rn nd O (5.2)
This result shows that
• for a fixed value of rn, d decreases as n increases;
• radii of half period zones are proportional to the square root of natural numbers, i.e.
1, 2, 3, ...
Thus, we can write
r1 = dO
r2 = 2d O 1.41 r1
r3 = 3d O 1.73 r1 (5.3)
To give you an idea about the number and size of Fresnel zones, we take a typical example. In a
particular experiment, d = 30 cm and wavelength of light used is 632.8 nm, which corresponds to
He-Ne laser. Then radii of the first few Fresnel zones are 0.436 mm, 0.614 mm, 0.754 mm, ...
Let us now calculate the area of the half-period zones. The area of the first half-period zone
is given by
A1 S r12 S d O
The area of the second zone, i.e. the annular region between the first and the second circles,
is given by
A2 S r22  S r12 S d O
Similarly, the annular region between the nth circle and the (n – 1)th circle constitutes the nth
half-period zone. Hence, the area of the nth half-period zone is given by

An S (rn2  rn21 ) S d O (5.4)


164 Wave Optics

Let us pause for a while and reflect as to what we have achieved so far. We find that within
the validity of the approximation d ? nl, all Fresnel zones are of equal area. Physically, it
implies that the amplitude of secondary wavelets starting from any two zones will be nearly
equal. Note that this result is approximate.
A more rigorous calculation shows that area of a zone increases with n gradually:

Ë È 1Ø O Û (5.5)
An SO Ì d  Én  Ù
Í Ê 2 Ú 2 ÜÝ

Ë 1Ø O Û
È
In this equation, the term in the square brackets Ì d  Én 
Ù denotes the average distance
Í 2 Ú 2 ÜÝ
Ê
of the nth zone from the point of observation P0, as shown in Figure 5.8. It may be mentioned
here that impact of increase in area of a zone with n is almost completely offset by an increase
in the average distance of that zone from the point of observation. It means that the ratio
An/dn = pl remains constant and independent of n. It means that the amplitude of secondary
wavelets reaching the point of observation will be determined by the obliquity factor alone; it
is actually responsible for the monotonic decrease in the amplitude of secondary wavelet reaching
P0 from the higher order zones (u1 > u2 > u3 ... > un).

Figure 5.8 Geometry of nth half-period zone and its average distance from the point of observation.

Now we proceed to know the phase of the resultant disturbance produced by the nth zone
with respect to the phase of the disturbance produced by the preceding or successive zone and
their impact on the resultant amplitude. To understand this, refer to Figure 5.8 and consider the
contributions of (n – 1)th and nth zones. Since both zones have nearly equal area, we can safely
assume that the amplitudes of secondary wavelets starting from these zones will also be equal.
But the points like R (Figure 5.9) on the inner periphery of (n – 1)th zone are situated at d +
(n – 2)l/2 from the point of observation P0, whereas points such as S on the inner periphery
of nth zone [or outer periphery of (n – 1)th zone] are situated at d + (n – 1)l/2 from P0. It means
that the path difference between the secondary wavelets reaching the point of observation from
R and S is one-half wavelength. With reference to Figure 5.8, we can write
O
Qn P0  Qn 1P0
2
Fresnel Diffraction 165

Figure 5.9 Geometry of (n – 1)th and nth half-period zones.

In terms of phase, we can say that the secondary wavelets reaching the point of observation from
points R and S are out of phase by p and annihilate each other. If we extend this logic to other
points between R and S in the (n – 1)th zone, we can identify one-to-one correspondence in points
between S and T in the nth zone, which have a phase difference of p and cancel the effect of each
other. Since the areas of these zones are nearly equal, we can conclude that secondary waves from
any two corresponding points in successive zones [nth and (n – 1)th or (n + 1)th], reach the point
of observation out of phase by p or half of a period. This is the genesis of the nomenclature half-
period zone. Note that their contribution to resultant amplitude is almost negligible.
The resultant amplitude at P0 due to the entire wavefront can be written as
u(P0) = u1 + u2eip + u3ei2p + L + unei(n–1)p
= u1 – u2 + u3 – u4 + L + (–1)n+1un (5.6)
where un denotes the net amplitude produced by secondary wavelets originating from the nth
zone. Note that the resultant amplitude at P0 due to the entire wavefront is a sum of an infinite
series whose terms are alternately positive and negative. We expect that the magnitudes of
successive terms to decrease monotonically because of increased obliquity and inverse dependence
on the average distance of the point of observation from the wavefront.
The series given in Eq. (5.6) can be summed up mathematically as well as geometrically.
We first discuss Schuster’s method and rewrite Eq. (5.6) as

u ( P0 )
u1
2
È u1

Ê 2
 u2 
u3 Ø È u3
Ù É
2Ú Ê 2
 u4 
u5 Ø u5
Ù 
2Ú 2
 " (5.7)

When n is odd, the last term would be un/2 and when n is even, the last term would be
È u n 1 Ø
É  un Ù . If the obliquity term is such that each term is less than the arithmetic mean of its
Ê 2 Ú
u  un1
preceding and succeeding terms, i.e. un < n 1 , the quantities in the square brackets in
2
Eq. (5.7) will be positive. It means that when n is odd, the minimum value of amplitude is
given by
(u  un )
u ( P0 ) ! 1 (5.8)
2
166 Wave Optics

To obtain the upper limit, we rewrite Eq. (5.6) as


È u2 Ø È u2 u4 Ø È u4 u6 Ø un 1
u ( P0 ) É u1 Ù É  u3  Ù É  u5  Ù  ...   un
Ê 2Ú Ê 2 2Ú Ê 2 2Ú 2
Following the argument used in obtaining the lower limit on the amplitude, we find that the
upper limit is
u u
u ( P0 )  u1  2  n 1  un (5.9)
2 2
Since the amplitudes for any two adjacent zones are nearly equal, we can take un–1  un. In this
approximation
u  un
u ( P0 )  1 (5.10)
2
The results contained in Eqs. (5.8) and (5.10) suggest that when n is odd, the resultant amplitude
at the observation point P0 is given by
u1  un
u ( P0 ) (5.11)
2
Following the same method, we can show that if n were even, the resultant amplitude at the
observation point P0 is given by
u1  un
u ( P0 ) (5.12)
2
Equations (5.11) and (5.12) suggest that the resultant amplitude produced by all the secondary
wavelets emanating from the entire wavefront is either half the sum or half the difference of the
amplitudes contributed by the first and the last zones. But if we allow n to be large so that
q » p for the last zone, the obliquity factor causes un to become negligible. Then we can ignore
un in comparison to u1 and the resultant amplitude produced by all the secondary wavelets
emanating from the entire wavefront is equal to one-half of the amplitude produced by secondary
wavelets emanating from the first zone:
u
u ( P0 )  1 (5.13)
2
Before proceeding further, you should answer a Practice Exercise.

Practice Exercise 5.1 Starting from Eq. (5.6), prove Eq. (5.12).

We can also obtain the result contained in Eq. (5.13) using a simple graphical construction.
Refer to Figure 5.10(a), which shows the magnitudes and positions of vectors AB, CD, EF,
GH, ... These respectively denote the amplitudes of resultant vectors u 1, u 2, u 3, u4, due to the
first, second, third, fourth, zones. Since these vectors are alternately positive and negative,
their addition is normally performed by drawing them along the same line segments so that
their mid points coincide with the mid point of u1 along the same straight line, as shown in
Figure 5.10(b). Note that the resultant amplitude due to first two zones is vector AD. But the
resultant of the first three zones is vector AF (>AD) and so on. The resultant of a given number
Fresnel Diffraction 167
of zones will be equal to the height of the final arrowhead. For infinitely large number of zones,
the resultant will be as given by Eq. (5.13).

Figure 5.10 Phasor diagram for Fresnel half-period zones. (a) Individual amplitudes, (b) collinear
arrangement of individual amplitudes, and (c) resultant amplitudes due to n = 2, 3, zones.

We hope that now you know how to construct Fresnel half-period zones and obtain the contribution
of a wavefront. Let us now learn to apply these to understand propagation of light waves.

5.4.2 Fresnel Construction and Rectilinear Propagation of Light

Refer to Figure 5.11. S is a point source and emits monochromatic light, which propagates
towards the right. Suppose that the distance between the source and the aperture is such that
the spherical waves reaching it can be treated as plane. For l = 500 nm and d = 0.50 m, the
radius of the first half-period zone is r1 (0.50 m) – (500 – 109 m) 5 – 104 m. It means that
diameter of the first zone will be about 1 mm. However, the radius of the hundredth half-period
zone will be r100 100 – (0.50 m) – (500 – 109 m) 5 – 103 m and the diameter will be 1 cm.
So if the diameter of the aperture is about 1 cm, it will accommodate nearly 100 Fresnel zones.
Since contribution of the 100th zone will be extremely small, the intensity at the point of
observation P0 will be one-half of that due to the first half-period zone. Did you not expect this
intensity at P0 when the aperture was completely removed? It means that even through a small
aperture, we get the original intensity at the point of observation P0. In other words, light travels
in a straight line for all practical purposes.

Figure 5.11 Fresnel construction and rectilinear propagation of light.


168 Wave Optics

Let us now understand how shadows are formed by an obstacle. Refer to Figure 5.12.
WW ¢ is a plane wavefront and points P2, P0 and P1 are three typical points on the observation
screen. Let us first consider point P2 whose pole is O2. In the above paragraph, we have seen
that if the diameter of the aperture is 1 cm, it will accommodate about 100 Fresnel half-period
zones when the screen is at a distance of 0.5 m and the source emits light of wavelength
500 nm. It means that if in the instant case, the distance between O2 and the edge A of the
obstacle is nearly 1 cm, 100 half-period zones will be accommodated and the intensity at P2 will
be nearly half of that due to the first half-period zone. That is, the obstacle AB has no effect
at the point P2. Similarly, at P1, which is 1 cm inside the geometrical edge of the shadow, over
100 half-period zones around O1 will be obstructed and the amplitude at P1 will be less than
u100/2, which is almost negligible. This implies almost complete darkness at P1. In other words,
we can say that the obstacle has completely stopped the light from the source and the region
around P1 is in the shadow. However, around point P0, which signifies the geometrical edge of
the shadow, we will observe fluctuations in intensity depending on the number of half-period
zones being obstructed. So we can conclude that the rectilinear propagation of light is observed
since Fresnel half-period zones are obstructed or allowed to pass through by the obstacles of
the size of a few millimetre for the typical distances considered here.

Figure 5.12 Fresnel construction and formation of shadows.

A beautiful application of the concept of Fresnel half-period zones is in the construction of


a zone plate. It is a special device designed to obstruct light from alternate half-period zones.
As a result, we can remove either all the positive or negative terms in Eq. (5.6). In either case,
the amplitude at P0 increases significantly. It provided experimental evidence in favour of
Fresnel’s theory. We will discuss it now.

5.4.3 Zone Plate

A zone plate is made by drawing concentric circles whose radii are proportional to the square
roots of natural numbers and alternate annular regions are shaded/blackened. The resultant
drawings are photographed on a reduced scale. The photographic transparency (negative) acts
as a Fresnel zone plate. Figure 5.13 shows two zone plates, where all even numbered and odd
numbered zones have been blacked out. These are respectively referred to as positive and
Fresnel Diffraction 169

negative zone plates. The radii of the circles can be expressed as 1K , 2K , 3K , ... , where
K is a constant and has dimension of length. When held in light from a distant source, a zone
plate produces a large intensity at a point on its axis at a distance determined by the size of the
zone and the wavelength of light used. That is, a zone plate acts as a lens.

Figure 5.13 The (a) positive and (b) negative zone plates.

The first zone plate was constructed by Lord Rayleigh in 1871. Nowadays, zone plates are
used to form images using even X-rays and microwaves which render ordinary lenses useless.
To understand the working of a zone plate, go through the following examples carefully.

EXAMPLE 5.1 Show that a zone plate acts like a multi-foci converging lens.
Solution Refer to Figure 5.14, which depicts the section of the zone plate normal to the plane
of the paper. Suppose that S is a point source which emits monochromatic light waves (spherical)
and is at a distance u from the zone plate.
We wish to study the effect of zone plate at the point P0. The distance between P0 and the
zone plate is v. Suppose that the centre of the zone plate is at O and the radii of various zones
are OQ1, OQ2, OQ3, ..., OQn such that each wave has to travel an additional distance of l/2
from successive zones. Then we can write

Figure 5.14 Zone plate as a multi-foci converging lens.


170 Wave Optics

O
SQ1 + Q1P0 = u  v 
2
SQ2 + Q2P0 = u + v + l
#
nO
SQn + QnP0 = u  v  (i)
2
By applying Pythagoras theorem to DSOQn, we can write

SQn = SO 2  OQn2
= u 2  rn2
If radius of the nth zone is much less than the distance of the source from the zone plate, we can
use binomial expansion and neglect terms of order higher than rn2 /2u. This leads to the result

SQn u
rn2
2u
 " (ii)

Similarly, by applying Pythagoras theorem to DOQnP0, we can write

P0 Qn v 2  rn2

As before, if the distance of the point of observation from the zone plate is much greater than
the radius of the nth zone, we get

P0 Qn v
rn2
2v
 " (iii)

By adding (ii) and (iii), we get

SQn  Qn P0 u
rn2
2u
r2
v n 
2v
"
On combining this result with (i), we get

rn2 r2 nO
u v n u v
2u 2v 2
On simplification, we get
È1 1Ø
rn2 É Ù nO
Êu vÚ
If we identify rn2 /nO as focal length of the zone plate and denote it by fn, i.e. if we put

rn2
fn (iv)
nO
Fresnel Diffraction 171
we get the lens equation:
1 1 1
 (v)
u v fn
This result clearly establishes that a zone plate acts like a multi-foci converging lens and forms
a real image of S at P0. If the observation screen is at a distance of one focal length from the
diffracting aperture, f1 = r12 /O. This corresponds to the most intense image (bright spot).
To get an idea of the numerical figures, go through the following example.

EXAMPLE 5.2 Consider a zone plate with radii rn 0.11 n cm illuminated by a


monochromatic light of wavelength 589 nm. Calculate the positions of various foci.
Solution Recall that focal length of a zone plate is given by

rn2
fn
nO
The first order focal point, which is most intense, is situated at

r12 (1.1 – 10 3 m) 2
f1 2.06 m
O 589 – 10 9 m

Other focal points will be located along the axis at f3 = 2.06/3 = 0.69 m, f5 = 2.06/5 = 0.41 m,
f7 = 2.06/7 = 0.29 m, and so on. Between any two consecutive foci, there will be dark points on
the axis corresponding to which the first circle will contain an even number of half-period zones.

In these examples, you have seen that a zone plate acts as a converging lens, but there are
differences in details. These are highlighted below.

Differences between a convex lens and a zone plate


1. A convex lens has one second principal focus but a zone plate has multiple foci between
r2
the plate and the brightest focus at a distance f1 = 1 .
O
2. A luminous point at a distance u from a zone plate (u is the object distance in the
geometrical optics parlance) will give rise to a series of images at distances v1, v2, v3,...
determined by the condition that for each ‘v’, a zone comprises an odd number of half-
period elements. v1, v2, v3,... correspond respectively to the focal lengths

r12 r12 f1 r12 f1


f1 , f2 , f3 , ...
O 3O 3 5O 5
3. For a convex lens, the rays after refraction through the lens reach the image point in the
same phase as that of the incident rays, whereas light from the consecutive clear zone
plate arrives at the image point after one complete period of the wave.
172 Wave Optics

4. For a convex lens, the focal length for red light (fR) is greater than that for violet light
(fV). But when white light is incident on a zone plate, we have
rn2 rn2
( f R ) zp and ( fV ) zp
nO R nOV
As lR > lV, (fR)zp < (fV)zp.

5.5 FRESNEL DIFFRACTION PATTERNS OF SIMPLE OBSTACLES

In section 5.3, we discussed Kathavate’s experimental arrangement to observe Fresnel diffraction


pattern of simple obstacles. In optical instruments such as a telescope, only a part of the
wavefront is incident on the objective, which comprises an achromatic convex lens with a
circular aperture fixed in front of it. Therefore, it will be instructive to use the preceding
analysis and understand the diffraction of a plane wave by a circular aperture or a straight edge.
We begin by considering Fresnel diffraction pattern of a circular aperture.

5.5.1 Diffraction by a Circular Aperture


(a) Axial point
Refer to Figure 5.15(a). It shows a plane wave incident normally on a thin metallic sheet having
a circular aperture of radius OA = r. The sectional view of the experimental arrangement is
shown in Figure 5.15(b). The plane of the wavefront is parallel to the plane of the sheet and
both are perpendicular to the plane of the paper. We wish to calculate intensity at a point P0
located at a distance d on the line passing through the centre of the circular aperture and
perpendicular to the wavefront.

Figure 5.15 (a) Diffraction of a plane wave incident normally on a circular aperture of radius r;
(b) Cross-sectional view of the experimental arrangement.

We can have aperture of any desired radius (r). Note that as r increases, the intensity at the
observation point will continue to increase till the first-half period zone fills the circular aperture.
This will happen for r = d. When these conditions are satisfied, the resultant amplitude at point
P0 will be equal to u1, which is twice the value of the amplitude for unobstructed wavefront
[see Eq. (5.13)]. It means that when a circular aperture obstructs the path of the wavefront, the
intensity, which is proportional to the square of amplitude at the observation point, will be four
times the intensity due to an unobstructed wavefront. This is a surprising result and is not
apparent in everyday experience dominated by rectilinear propagation of light.
Fresnel Diffraction 173
If the aperture has radius such that half-period zones beyond the first also begin to contribute,
the amplitude will tend to decrease. When the radius of the circular aperture equals to 2O d ,
the resultant amplitude is expected to be very small, since it will now contain first two half-
period zones. And the intensity at point P0 will become almost zero if the diameter of the
aperture equals 8O d . We can generalize this result by noting that if
r (2n  1)O d n = 0, 1, 2, ... (5.14)
the aperture will transmit contributions from odd number of half-period zones and the intensity
will be maximum. On the other hand, if
r 2n O d n = 1, 2, 3, ... (5.15)
the aperture will transmit contributions from even number of half-period zones and the intensity
will be minimum. Thus we may conclude that if the size of the aperture changes continuously,
intensity passes through maxima and minima along the axis of the aperture. To get an idea
about the numbers and to develop an appreciation for the size of aperture, we note that for
l = 550 nm and d = 0.30 m, the diameters of the first three zones will be 0.812 mm, 1.148 mm
and 1.406 mm, respectively.
We will observe a similar spatial variation in intensity if point P0 is moved towards
the diffracting screen for an aperture of fixed size. To understand this, we rearrange terms in
Eq. (5.14) and write
r2
d n = 0, 1, 2, ... (5.16)
(2n  1) O
This will correspond to a maximum.
By inverting Eq. (5.15), we get the condition for the minimum:

r2
d n = 1, 2, 3, ... (5.17)
2 nO
You will observe waxing and waning of intensity for an aperture of fixed size, if the observation
screen is moved towards the zone plate.
When P0 is close to the screen so that an infinite number of half-period zones contribute
to the amplitude, the resultant amplitude will be u1/2.
To appreciate the concept discuss here, go through the following example carefully.

EXAMPLE 5.3 Consider a circular aperture of radius 0.55 mm. A beam of parallel light of
wavelength 550 nm is incident on it normally. The shadow is observed on a movable screen.
At what distances will the aperture transmit 1, 2, 3 Fresnel zones?
Solution Suppose that the distances at which 1, 2, 3 Fresnel zones are transmitted by the
circular aperture are d1, d2 and d3. From Eq. (5.2), we can write

rn2
dn
nO
174 Wave Optics

Since radius of the aperture is fixed (rn = r), the distance at which circular aperture will transmit
only the first Fresnel zone is given by

r2 (5.5 – 104 m)2


d1 0.55 m
O 550 – 109 m
Similarly, we see that the distances at which circular aperture will transmit second and third
Fresnel zones are d2 = r2/2l = 0.275 m and d3 = 0.55/3 = 0.183 m, respectively.
The qualitative nature of the amplitudes along these axial positions is shown in Figure 5.16.

Figure 5.16 Qualitative plot of amplitudes of diffracted light at different axial positions.

You may now like to solve a Practice Exercise.

Practice Exercise 5.2 On the basis of Fresnel’s construction, discuss the diffraction pattern
of an opaque disc which obstructs the first p half-period zones.

Suppose that we just cover the first Fresnel zone of circular aperture by a circular disc or
such other round object. Do you expect the centre of the shadow to be dark? No we will obtain
a bright spot—the Poisson spot—on the axis behind the circular disc. This is because light will
reach the point of observation P0 from all zones except the first and the second zone becomes
the first contributing zone. Moreover, the intensity of light spot at the centre of the shadow of
the obstacle will be almost as bright as when the first zone was unobstructed.
We now advise you to answer a Practice Exercise.

Practice Exercise 5.3 A source of light emitting waves of wavelength 500 nm falls on a coin
of radius 1 cm. How many Fresnel zones will it cut-off, if a screen is
(a) placed 1 m away and
(b) moved to a distance of 4 m?
(c) In which case will the central spot be brighter?
[Ans. (a) 200; (b) 50; (c) In the latter, case]

Note that we have taken the observation point on the line passing through the centre of the
circular aperture and perpendicular to the wavefront. For off-axis points, we can, in principle,
use Fresnel half-period zones to calculate intensity distribution but the process is fairly
Fresnel Diffraction 175
cumbersome and result approximate. However, the symmetry of the problem suggests that the
diffraction pattern will be in the form of concentric circular rings with their centre at the point
P0. This is confirmed by a detailed and rigorous mathematical analysis, which we shall discuss
in a later section. For completeness, here we present a simple treatment.

(b) Off-axis point


Refer to Figure 5.15a again. AB is a circular aperture whose centre is at O. The perpendicular
drawn from on the observation screen meet it at P0. The point of observation on the screen is
P. Let PP0 = x, OP0 = d and radius of aperture OA = r. We draw perpendiculars AA¢ and BB¢
from A and B on the screen. The path difference between the secondary wavelets from A and
B reaching the point of observation P is given by
D = BP – AP (5.18)
On applying Pythagoras theorem to DBPB¢, we can write

BP [ d 2  ( x  r ) 2 ]1/2
If d ? (x + r), we can use binomial expansion and retain terms only up to first order to obtain
Ë ( x  r )2 Û
BP d Ì1  Ü (5.19)
2d 2 Ý Í
Similarly by considering DAA¢P and proceeding as before, we can easily show that
Ë ( x  r )2 Û
AP d Ì1  Ü (5.20)
2d 2 ÝÍ
On substituting for BP and AP from Eqs. (5.19) and (5.20) in Eq. (5.18), we get
1 2 xr
' [( x  r ) 2  ( x  r )2 ] (5.21)
2d d
O
The point of observation P will be dark if the path difference D = 2n . Hence, we can say
2
that the radius of the nth dark ring is given by
O 2rxn
2n
2 d
or
nd O n = 1, 2, 3, ... (5.22)
xn
2r
Similarly, we can show that the radius of the nth bright ring is given by
(2n  1)d O
xn n = 0, 1, 2, ... (5.23)
4r
Let us now apply these results to a practical situation and consider a telescope, whose objective
is in the form of an achromatic convex lens and a circular aperture is fixed in front of the lens.
176 Wave Optics

The light coming from a source at infinity implies that a plane wavefront is incident on the
objective so that d = f, the focal length of the objective. So, the diffraction pattern will consist
of a bright central spot surrounded by dark and bright rings of gradually diminishing intensity.
If the diameter of the objective is D = 2r, we can rewrite Eq. (5.23) as
nf O
xn n = 1, 2, 3, ...
D
Hence the radius of the first dark ring, i.e. the distance of the first secondary minimum from
the central bright maximum is given by
fO
x1 (5.24)
D
In Chapter 6, we will discuss Airy’s theory and show that the radius of the first dark ring is
given by
fO
x1 1.22 (5.25)
D
Note that the size of central spot depends on the wavelength of light used, the focal length of
the lens and the diameter of the aperture.
You must have noted that in our discussion so far, the diffraction patterns had axial symmetry:
the object (or aperture) was circular and the plane wavefront originated from a point source. If
the configuration of the diffracting screen involves straight edges, as those of a slit or a wire,
it is more convenient to use a slit source for symmetry considerations. The slit is set parallel
to these edges so that diffraction fringes obtained on the observation screen (and produced by
each element along its length) are straight. This helps to increase intensity of the pattern
considerably. We now consider a slit source. Note that to produce a cylindrical wavefront,
different points on the slit must emit coherently. But this may not be true in practice. Nevertheless,
it does not affect the resultant diffraction pattern. Therefore, we assume that the slit source is
illuminated by a parallel monochromatic beam so that it emits a truly cylindrical wave. We first
take the object to be a straight edge.

5.5.2 Diffraction by a Straight Edge

Refer to Figure 5.17(a). It shows a sectional view of the experimental arrangement used to
observe diffraction by a straight edge MN placed perpendicular to the paper and parallel to a
long narrow slit source S. The line joining S and O, the point on the cylindrical wavefront WW¢
touching the edge of the straight edge, when produced further meets the screen at P0, which
defines the geometrical boundary of the shadow.
Consider an arbitrary point P on the screen. A line joining P and S cuts the wavefront WW¢
at R. We wish to calculate the intensity variation on the screen LL¢ based on Fresnel construction.
The appropriate method of constructing half-period elements on a cylindrical wavefront consists
of dividing it into strips, whose edges are successively out of phase by p or one-half wavelength
farther from the point of observation. Therefore, to construct half-period elements in this case,
Fresnel Diffraction 177

O
we draw a set of circles with P0 as centre and radii d, d + , d + l, ... . These circles intersect
2
the circular section of the cylindrical wavefronts at points O, A and A¢, B and B¢, C and C ¢, ...
as shown in Figure 5.17(b). By referring to Figure 5.17(a) we can write

O
SA + AP – SRP =
2
2O
SB + BP – SRP =
2
3O
SC + CP – SRP = (5.26)
2

Figure 5.17 (a) Cross-sectional view of the experimental arrangement used to observe diffraction due
to a straight edge, (b) Fresnel construction divides the cylindrical wavefront in half-period
zones, and (c) Fresnel half-period strips.

If lines are drawn through A, A¢, B, B¢, etc., normal to the plane of the paper, the upper
as well as the lower half of the wavefront will be divided into a set of half-period strips.
These strips stretch along the wavefront parallel to the slit. That is, these are perpendicular to
the plane of the paper and have widths OA, AB, BC, ... in the upper-half wavefront and OA¢,
A¢B¢, B¢C¢, ... in the lower-half wavefront, as shown in Figure 5.17(c). Note that unlike
178 Wave Optics

the Fresnel half-period zones, the areas of half-period strips will not be equal; the areas of
half-period strips are proportional to their respective widths and these decrease rapidly as we
go outwards along the wavefront from O. As such, this effect is more pronounced than any
variation in the obliquity factor. (For this reason, we ignore the latter in this case.) Nevertheless,
this complicates the analysis considerably. However, we can draw the following general
conclusions:
1. Corresponding to the edge of the geometrical shadow, which is signified by P0 in
Figure 5.17(a), half of the wavefront is obstructed by the edge, and the amplitude is
given by
u0
u ( P0 ) (5.27)
2
where u0 is the amplitude corresponding to the unobstructed wavefront. It means that
the intensity will be one-fourth of that due to unobstructed wavefront.
2. Suppose that the point P satisfies the relation

O
SO  OP  SRP
2
Then only the first half-period strip of the lower part of the wavefront will contribute
and the resultant amplitude would very nearly be equal to

u1 u1 3 È u1 Ø 3
 É Ù u ( P0 ) (5.28)
2 4 2Ê 2 Ú 2
where u1/2 signifies the amplitude produced by the first half-period strip in the lower
portion and u1/4 is the resultant amplitude generated by the upper half of the wavefront.
9
The intensity will be I.
4 0
3. For a point P1 such that
SO + OP1 – SP1 = l
we obtain a minimum and the resultant amplitude is given by

È u1 u2 Ø u1
É
Ê 2
 Ù  (5.29)
2Ú 4
4. In general, any arbitrary point P will correspond to the maximum intensity (bright
band) if
O
SO  OP  SP (2n  1) n = 0, 1, 2, ... (5.30)
2
and a minimum (dark band) if
SO + OP – SP = nl n = 1, 2, ... (5.31)
Fresnel Diffraction 179
If we use Pythagoras theorem for DOPP0, we can write
OP = (d2 + x2)1/2
For x = d, we use binomial expansion and retain only the first order term in x2.
This gives
È x2Ø x2
OP d É1  2Ù
d 
Ê 2d Ú 2d
Similarly, from DSPP0, we can write

x2
SP [(a  d ) 2  x 2 ]1/ 2 (a  d ) 
2(a  d )
On substituting these results in Eq. (5.30), we get

x2 Ë x2 Û O
SO  OP  SP ad   Ì( a  d)  Ü (2n  1) n = 0, 1, 2, ...
2d Í 2( a  d ) Ý 2
On simplification, we obtain
a O
xn2 (2n  1) n = 0, 1, 2, ...
2(a  d ) d 2
Hence, we will observe nth maximum (bright band) on the screen when the distance
between P0 and P is
1/2
Ë d (a  d ) Û
xn ÌÍ (2n  1) a
OÜ ,
Ý
n 0, 1, 2, ... (5.32)

Similarly, the position of nth minimum (dark band) is given by


1/ 2
Ë d (a  d ) Û
xn ÌÍ 2n OÜ n = 1, 2, ... (5.33)
a Ý
Note that the distances of the dark band from the edge of the geometrical shadow are
proportional to the square root of natural numbers. It means that the bands get closer
as we go out from the shadow. This fact distinguishes the diffraction bands from the
interference bands, which are equidistant.
To illustrate this point, we consider a typical experiment in which a = d = 25 cm and
l = 550 nm. So, the first maxima and minima respectively occur at distances of
5.24 ´ 10–4 m and 7.42 ´ 10–4 m from the edge of the shadow. The second and third
maxima will occur at distances of 9.08 ´ 10–4 m and 11.73 ´ 10–4 m, respectively. You can
easily check that the second and third minima will occur at distances of 10.49 ´ 10–4 m and
12.85 ´ 10–4 m.
If in an experiment we know the locations of minima and maxima produced by a straight
edge, we can calculate the wavelength of light used to produce diffraction pattern. Though
precise calculation of spatial distribution of intensity distribution based on the above analysis
is difficult, a qualitative plot is shown in Figure 5.18.
180 Wave Optics

A very common and striking observation of the straight-edge pattern occurs when we view
a distant street lamp through rain-spattered spectacles. The edge of each drop on the glass
behaves like a prism and refracts light into the pupil of our eye. (This light would have
otherwise not entered our eye.) And the outline of the drop is seen as an irregular but bright
patch surrounded by very clear diffraction fringes.

Figure 5.18 Spatial variation of intensity in the diffraction pattern due to a straight edge.

Before proceeding further, we advise you to answer a Practice Exercise.

Practice Exercise 5.4 Use Fresnel construction to discuss diffraction pattern produced by
a circular opaque disc.

Let us pause for a while and ask: What have we achieved so far? We have used Fresnel
construction to analyse diffraction patterns produced by a circular aperture and a straight edge. We
can use graphical method based on vector addition of amplitudes with some advantage to obtain
their resultant at an external point due to all secondary wavelets originating on a wavefront (spherical
or cylindrical). This method leads to the so-called Cornu’s spiral and helps to gain clearer physical
insight into the origin of the diffraction pattern. We will now discuss it in some detail.

5.6 GRAPHICAL METHOD: CORNU’S SPIRAL

Refer to Figure 5.19(a). It shows the amplitude diagram obtained by dividing the first half-
period strip of incident wavefront WW ¢ into nine sub-strips so that the corresponding nine
amplitude vectors extend from O to M1. What can be said about the shape? It arises due to
gradual change of phase, which, in turn, is caused by continuous increase in the obliquity factor
from O to M1. The resultant amplitude at M1 due to the first half-period strip is given by
A1 = OM1. Similarly, if this process is continued for the second half-period strip, we obtain nine
sub-strips and the corresponding nine amplitude vectors will extend from M1 to M2, giving the
resultant A2 = M1M2. Note that A2 << A1. This arises due to rapid decrease in amplitudes and
Fresnel Diffraction 181
the difference in their phases will be much different from p. A repetition of this process of
subdivision for the succeeding strips on the upper half of the wavefront leads to Figure 5.19(b),
which is more complete. As may be noted, the vectors spiral in towards Z implying that the
resultant of all half-period strips above O is given by OZ.

Figure 5.19 Amplitude diagram for (a) two half-period strips, and (b) the upper half of the wavefront.

When half-period strips are divided into elementary strips of infinitesimal width, a smooth
curve is obtained. The complete curve for the entire wavefront will be carried through many
more turns. The Cornu’s spiral is characterised by the fact that for any point on curve, the phase
lag d from any element of the wavefront is directly proportional to the square of its distance
along the curve from O, which defines the origin in this case.
To understand the basis of this statement, refer to Figure 5.20. WW¢ is incident wavefront
and O is the pole of the wavefront with reference to P0. Let us denote the distances of the source
and the point of observation from the pole by a and b respectively. With P0 as centre, we draw
a circle of radius b, which touches the incident wavefront at O. The path difference between
waves emanating from S and travelling to P0 along SAP0 and SOP0 is given by
D = SA + AP0 – SOP0
= SA + AP0 – (SO + OP0)

Figure 5.20 Calculation of path difference D from the pole of a spherical wavefront.

We drop perpendiculars from A and B on the line joining S and P0 so that SA  SM and
BP0 = OP0. Then, we can write
182 Wave Optics

D = SA + AB + BP0 – (SO + OP0)


= a + AB + b – (a + b) = AB
Usually a and b are large compared to AM and BN; in fact these can be taken to be nearly equal
and path difference D can be expressed as
D = AB = MO + ON (5.34)
Using the properties of a circle, we can write

AM 2 h2
MO =
2 SO 2a
and
BN 2 h2
ON =
2OP 2b
Using these results in Eq. (5.34), we get

h 2 h 2 h 2 (a  b )
' 
2a 2b 2 ab
If it so happens that AM corresponds to the radius of the nth half-period zone, then D will be
equal to nl/2 and we can write
h 2 (a  b) nO
(5.35)
2 ab 2
We can now readily obtain the value of phase lag d, which is defined as

2S 2S h2 (a  b) S Ë 2h2 (a  b) Û S 2
G ' Ì Ü [ (5.36)
O 2abO 2 Í abO Ý 2
where we have introduced a new dimensionless variable x defined as

2( a  b)
[ h (5.37)
abO
Proceeding further, we express Cartesian coordinates of Cornu’s spiral in terms of the Fresnel
integrals. To obtain the desired result, we note that if the distance of the point of observation
P0 along the curve from the origin is x, then tangent to the curve at this point makes an angle
d with the x-axis. If this point undergoes a small displacement dx along the curve, the changes
in the x and y coordinates are given by
È S[ 2 Ø
dx cos G d [ cos É Ù d[
Ê 2 Ú
and
È S[ 2 Ø
dy sin G d [ sin É Ù d[
Ê 2 Ú

By integrating these expressions in the range 0 to t, we can express the (x, y) coordinates of
any point on the Cornu’s spiral in terms of Fresnel integrals:
Fresnel Diffraction 183
W
È S[ 2 Ø
x Ô cos É
Ê 2 Ú
Ù d[ C (W ) (5.38)
0
and
W
È S[ 2 Ø
y Ô sin É
Ê 2 Ú
Ù d[ S (W ) (5.39)
0

The Fresnel integrals cannot be integrated in closed form but yield infinite series. So we will
not go into these details here. However, it will suffice to include a table of their numerical
values for typical values of t (Table 5.1).

Table 5.1 Table of Fresnel integrals


t C(t ) S(t ) t C(t ) S(t )
0.00000 0.00000 0.00000 2.4 0.55496 0.61969
0.1 0.100000 0.00050 2.6 0.38894 0.54999
0.2 0.19992 0.00419 2.8 0.46749 0.39153
0.3 0.29944 0.01410 3.0 0.60572 0.49631
0.4 0.39748 0.03336 3.2 0.46632 0.59335
0.6 0.58110 0.11054 3.4 0.43849 0.42965
0.8 0.72284 0.24934 3.6 0.58795 0.49231
1.0 0.77989 0.43826 3.8 0.44809 0.56562
1.2 0.71544 0.62340 4.0 0.49843 0.42052
1.4 0.54310 0.71353 4.2 0.54172 0.56320
1.6 0.36546 0.63889 4.4 0.43380 0.46227
1.8 0.33363 0.45094 4.6 0.56724 0.51619
2.0 0.48225 0.34342 5.0 0.56363 0.49919
2.2 0.63629 0.45570 ¥ 0.5 0.5

Properties of Fresnel Integrals

Refer to Eqs. (5.38) and (5.39). Note that the integrands are even functions of t and the Fresnel
integrals C(t) and S(t) are odd functions of t. That is,
C(–t) = –C(t) and S(–t) = –S(t ) (5.40)
Further, if the entire wavefront is exposed to the point of observation, the value of t ® ¥. Then using
the result
‡
2 S
Ô exp( ax ) a

we can write
‡
È iS[ 2 Ø S È iS Ø
Ô exp ÉÊ 2 ÙÚ d [ iS
2 exp É Ù (1  i)
Ê 4Ú
(5.41)

2
(Contd...)
184 Wave Optics

(Contd...)

Properties of Fresnel Integrals

Hence,
‡
È iS[ 2 Ø Ë‡ È S[ 2 Ø
‡
È S[ 2 Ø Û
exp É Ù d [ = 2 Ì
Ô cos É Ù d [  i Ô sin É Ù d[ Ü
Ô Ê 2 Ú Ì Ê 2 Ú Ê 2 Ú Ü
‡ Í0 0 Ý
= 2[C(¥) + iS(¥)] (5.42)
The factor of 2 arises because we have changed limits of integration from –¥ to ¥ to 0 to ¥ since
the integrand is an even function of x.
On comparing the right-hand sides of Eqs. (5.41) and (5.42), we get

1
C (‡ ) S ( ‡)
2
Thus, the important properties of Fresnel integrals are:

1
C(¥) = S (‡ ) ; C (0) S (0) 0
2
C(–t) = –C(t) and S(–t) = –S(t) (5.43)

Now refer to Figure 5.21, which is a parametric representation of Fresnel integrals and is
known as the Cornu’s spiral. We now examine some of its qualitative features. The horizontal
and the vertical axes represent C(t) and S(t), respectively. The numbers written on the spiral
give the values of particular upper limit, t. Note that the spiral passes through the origin since
C(0) = S(0) = 0. Moreover, the curve is symmetrical with respect to the origin. Further,
C(–t) = –C(t) and S(–t) = –S(t). This implies that the curve is asymmetrical with respect to
both axes. At any intermediate point on the spiral, the tangent to the curve makes an angle f
with the C-axis, i.e.
dy dS È S[ 2 Ø
tan I tan É Ù
dx dC Ê 2 Ú

where we have substituted the values of dy and dx. Hence, we can write

S[ 2
I (5.44)
2
This result shows that f increases monotonically with x. The tangent to the curve is parallel to
the C-axis when [ 2m and parallel to the S-axis when [ 2m  1, where m is an integer.
On differentiating Eq. (5.44), we can write

df = pxdx
Fresnel Diffraction 185

Figure 5.21 Plot of Fresnel integrals: Cornu’s spiral.

Hence, the radius of curvature of the spiral at the point under consideration is given by

d[ 1
dI S[
This result shows that as the value of x increases, the radius of curvature of the curve gradually
decreases and takes the shape of a spiral. The points W 1, 2, 2 on the curve respectively
correspond to one-half, one and two Fresnel half-period strips. You can check this and convince
yourself by calculating the corresponding values of d using Eq. (5.36).
Note that the length of the chord of a segment of the curve gives amplitude due to any given
portion of the wavefront and square of this length gives the intensity at the point of observation.
For instance, the amplitude due to the first strip on one side of the pole is given by A1 and that
of two zones by A2. Also C(¥) = S(¥) = 1/2 so that the two branches of the curve approach
the end points Z and Z¢ with coordinates (1/2, 1/2) and (–1/2, –1/2) respectively. The quantity
C(t) + iS(t) can be interpreted as the complex phasor joining the origin to the point t on the
spiral.
186 Wave Optics

Let us now discuss how Cornu’s spiral construction helps us in interpreting and understanding
the results in Fresnel diffraction for the simple case of a straight edge.

5.6.1 Straight Edge

The study of diffraction pattern of a straight edge [Figure 5.22(a)] is the simplest application
of Cornu’s spiral. The amplitude from an unobstructed wavefront is given by the vector ZZ¢ in
Figure 5.21. From Figure 5.22(a), we note that the geometrical edge of the straight edge is on
the line joining the pole to the point of observation. The half-period strips corresponding to
point P0 situated at the edge of the geometrical shadow are marked on the wavefront WW¢ as
AB. The amplitude at P0 is given by the vector OZ, which is half of the vector ZZ¢. (The length
of the vector OZ is 1/ 2 .) It means that the intensity will be one-fourth of the intensity due
to the unobstructed wave.
Next we consider the intensity at the point P at a distance x above P0. That is, the point
P lies in the direction SA, where A is the upper edge of the first Fresnel half-period strip.
For this point, the centre O of the half-period strips lies on the straight line joining S with P
and the figure has to be reconstructed as in Figure 5.22(b). The straight edge now lies at point
A¢. In this case, one half-period strip below O, in addition to all half-period strips above O,
will be exposed. The resultant amplitude is represented on the spiral of Figure 5.23 by a straight
line joining M¢ and Z. Note that this amplitude is more than twice that at P and obviously the
intensity will be more than four times as large.

Figure 5.22 Two different positions of the half-period strips relative to a straight edge.

If we move the point of observation steadily above the edge of the geometrical shadow, the
tail of the amplitude vector moves to the left along the spiral but its head remains fixed at Z.
The amplitude is maximum at b¢, minimum at c¢, d¢, etc., before attaining the value Z¢Z for the
unobstructed wave. If we move into the region of geometrical shadow, the tail of the vector
moves to the right from O and the amplitude will decrease monotonically before approaching zero.
A qualitative measure of the intensities in the diffraction pattern for a straight-edge can be
obtained from Cornu’s spiral by measuring the length of the chord as a function of x. The square
Fresnel Diffraction 187

Figure 5.23 Cornu’s spiral depicting resultant amplitudes for diffraction due to a straight-edge.

of this length gives intensity. Plots of amplitude and the intensity versus x are shown in
Figure 5.24. Note that at the point O, which corresponds to the edge of the geometrical shadow,
the intensity falls to one-fourth of that for large negative values of x, where it approaches the
value for the unobstructed wave. The other letters correspond to points representing the exposure
of one, two, three, half-period strips below O. These are labelled as M¢, C ¢, D¢, ... on the
spiral. Note that the maxima and minima of these diffraction fringes occur a little before these
points are reached.

Figure 5.24 Amplitude and intensity contours for Fresnel diffraction at a straight edge.

Let us pause for a minute and ask: What have we achieved so far? We see that Fresnel
construction as well as Cornu’s spiral method are approximate. It is, therefore, desirable to
discuss more rigorous mathematical analysis.
188 Wave Optics

5.7 FRESNEL DIFFRACTION: A RIGOROUS ANALYSIS

Refer to Figure 5.25, which shows a plane wave incident normally on an aperture of arbitrary shape
and size. The observation screen SS ¢ is placed at a distance d from the aperture. We consider
an infinitesimal area on the aperture plane. The field at the point of observation P0 due to the
A exp(ikr )
waves generated by the area dxdz will be proportional to dxdz, where r = MP0.
r

Figure 5.25 Schematics of the arrangement used to observe diffraction of a plane wave by an aperture
of arbitrary shape.

To calculate the total field at the point of observation, we integrate over the area of the
aperture by summing up all contributions from infinitesimal areas to obtain
A exp(ikr )
u ( P0 ) K ÔÔ A
r
d [ d] (5.45)

where K is the constant of proportionality. Note that the integration is to be carried out over the
entire aperture. The constant K is given by
ik i
K  
2S O
Equation. (5.45) now takes the form
i A exp(ikr )
u ( P0 ) 
O ÔÔ A
r
d [ d] (5.46)

Let us now consider some simple cases. We start with a circular aperture for comparative
mathematical ease.
Fresnel Diffraction 189

5.7.1 Circular Aperture

Refer to Figure 5.26. It shows a plane wave incident normally on a circular aperture of radius a.
We choose z-axis normal to the plane of the aperture and the screen SS ¢ is assumed normal to
the z-axis. The geometry of the arrangement suggests that it will be more convenient to use the
circular system of coordinates. Therefore, the coordinates of the point M on the aperture will
be (r, f), where r denotes the distance of the point from the centre O and f is the angle that
OM makes with the x-axis. From your knowledge of coordinate systems, you will recall that a
small element of area dS surrounding the point M can be represented as rdrdf.

Figure 5.26 Diffraction of a plane wave by a circular aperture.

For a circular aperture, Eq. (5.46) takes the form


2S a
iA exp(ikr )
u ( P0 ) 
O Ô dI Ô r
UdU (5.47)
0 0

By referring to Figure 5.26, we can write


r 2 + d2 = r2 (5.48)
where d is the distance between the observation plane and the aperture plane.
From Eq. (5.48), we can write
rdr = rdr

The limits of integration change from 0 to a to d to a 2  d 2 so that Eq. (5.47) takes the form

2S a2 d 2
iA
u ( P0 ) 
O Ô dI Ô
0 d
exp(ikr ) dr

.
190 Wave Optics

The integration over f gives 2p and integration of exponential function is simply


exp(ikr)/ik and in the limits of integration, we get

Ë Û
iA Íexp ik a 2  d2  exp(ikd ) Ý
u ( P0 )  2S
O ik
Since wave number k = 2p/l, this expression simplifies to

u(P0) = A exp(ikd) [1 – exp(ipp)] (5.49)

where
k a2  d 2  d pS (5.50)
This relation implies that
pO
QP  OP (5.51)
2
where Q is an arbitrary point on the periphery of the circular aperture.
Since intensity of a wave is proportional to the square of amplitude, it readily follows from
Eq. (5.49) that
pS
I ( P0 ) 4I 0 sin 2 (5.52)
2
where I0 = A2 is the intensity associated with the incident wave.
From Eq. (5.52) we note that the intensity of the diffracted wave at the point of observation
is zero when p is an even integer and maximum when p is an odd integer, i.e. the intensity
fluctuates between minimum and maximum values when QP – OP is an even or odd multiple
of l/2. This can be understood physically in terms of Fresnel half-period zones discussed earlier
(Section 5.5.1): If an aperture contains an even number of half-period zones, the intensity at the
point of observation will be negligibly small. Conversely, if the aperture contains an odd number
of half-period zones, the intensity at the point of observation will be the maximum.

u(P0) = A exp(ikd ) [1  exp(ipS )]


I = |u(P0)|2 = A2 [1  exp(ipS )] [1  exp( ipS )]
pS
= I 0 – 2(1  cos pS ) 4 I 0 sin 2
2
Since a = d, we can use binomial expansion to rewrite Eq. (5.50) as

k Ë È a2 Ø Û
p= Ì d É1  2Ù
 dÜ
S Í Ê 2d Ú Ý
or
a2k a2
p= (5.53)
2S d Od
Fresnel Diffraction 191
This is known as the Fresnel number of the aperture.
You may now like to answer a Practice Exercise.

Practice Exercise 5.5: Consider a plane wave incident normally on a straight edge.
By defining r [[ 2  (]  y ) 2  d 2 ]1/2 in Eq. (5.46), discuss the diffraction pattern produced
on a screen.

We now give chapter summary.

5.8 SUMMARY

• Fresnel diffraction changes significantly as the separation between the obstacle and the
observation screen changes. Spatial evolution of Fresnel diffraction leads to Fraunhofer
diffraction.
• Fresnel zones or half-period elements refer to concentric circles on a rectangular wavefront
such that the path difference between the wavelets reaching a point of observation from
the corresponding points in successive zones is one-half wavelength.
• The radius and area of the nth zone of a half-period zone are respectively given by
rn = nd O ; n = 1, 2, 3, ... and An S (rn2  rn21 ) S d O .
• The resultant amplitude produced by all the secondary wavelets emanating from the entire
wavefront is either half the sum or half the difference of the amplitudes contributed by the
first and the last zones. But if n is large, the resultant amplitude produced by all the
secondary wavelets emanating from the entire wavefront is equal to one-half of the amplitude
produced by secondary wavelets emanating from the first zone:
u1
u ( P0 ) 
2
• Fresnel zone plate acts as a multi-foci converging lens.
• For a circular aperture, the radii of the nth dark and bright rings are respectively given by
nd O (2n  1) d O
xn , n = 1, 2, 3, ... and xn , n = 0, 1, 2, ...
2r 4r
• The (x, y) coordinates of any point on the Cornu’s spiral in terms of Fresnel integrals are
given as
W W
È S[ 2 Ø È S[ 2 Ø
x Ô
0
cos É
Ê 2 Ú
Ù d[ C (W ) and y Ô sin É
Ê 2 Ú
Ù d[ S (W )
0
• For a plane wave incident normally on a circular aperture of radius a, the intensity at a
point P0 on the axis is given by

pS
I ( P0 ) 4I 0 sin 2
2
192 Wave Optics

where I0 = A2 is the intensity associated with the incident wave. The intensity of the
diffracted wave at the point of observation will be zero when p is an even integer and
maximum when p is an odd integer.

REVIEW EXERCISES

1. Calculate the radius of the first, fourth, ninth, sixteenth and thirty-sixth half period zones of
a plane wavefront for a point 0.2 m ahead of it. The wavelength of light used is 500 nm.
[Ans. r1 = 0.0316 cm; r4 = 0.0632 cm;
r9 = 0.0948 cm; r16 = 0.1264 cm; r36 = 0.1896 cm]
2. Monochromatic light of wavelength l is normally incident on a screen with a circular hole
of radius R. The intensity of light vanishes at an axial point directly behind the hole.
Calculate the distance between the screen and the point of zero intensity. Assume l = R.
[Ans. R2/2l]
3. Light of wavelength l, wave number k and intensity I0 is incident normally on (a) an
opaque circular disc of radius a, and (b) an opaque screen with a circular hole of radius
a. For both cases, derive an expression for the intensity of light at a distance l (? a)
behind the obstacle on an axis passing through the centre of the circle.
[Ans. (a) I0; (b) 4I0 sin2(ka2/4l )]
4. A parallel beam of monochromatic light of wavelength l is incident normally on an
opaque screen with a circular hole of radius a. It is detected on the axis of the hole at a
distance l . The intensity is seen to oscillate as a increases from 0 to ¥. Calculate
(a) the radius of the hole for the first maximum and the first minimum;
(b) the ratio of the intensity corresponding to the first maximum and that for the first
minimum;
(c) the intensity if the screen is replaced by an opaque disc of radius equivalent to that
for the first maximum. [Ans. (a) OA , 2O A ; (b) 4; (c) I¥]
5. An opaque sheet has a hole of radius 0.6 mm in it. Plane waves of wavelength 600 nm
are allowed to fall on it. Calculate the maximum distance from the sheet at which a screen
should be placed so that the light is focused to a bright spot. Also calculate the relative
intensity of this spot when the opaque sheet has been removed. [Ans. 0.6 m, 0.25]
6. A Fresnel zone plate is made by dividing a photographic image into five separate zones.
The first zone consists of an opaque circular disc of radius r1. Thereafter, we have alternate
transparent and opaque rings included between radii r1 and r2; r2 and r3; r3 and r4 and the
final zone is from r4 to ¥ and is black. The ratios between the radii of various zones can
be expressed as r1 : r2 : r3 : r4 :: 1: 2 : 3 : 4 . The zone plate is illuminated by plane
monochromatic light of wavelength 500 nm. The most intense spot of light behind the
plate is seen on its axis 1 m away.
(a) Calculate r1.
Fresnel Diffraction 193
(b) If I is the intensity of the spot and I0 is the original intensity, show that I = 16 I0.
and
(c) Locate the positions of the intensity maxima on the axis.
[Ans. (a) r1 = 0.707 mm; (c) 0.333 m, 0.2 m; 0.143 m]
7. Light from a monochromatic point source of wavelength l is focused on a point image
by a Fresnel half-period zone plate having 100 open odd half-period zones (1, 3, 5, ...,
199). All even zones are opaque. Compare the intensity of the point image with that at the
same point by replacing the zone plate by a lens of the same focal length and diameter
corresponding to 200 half-period zones of the zone plate. Assume the diameter of the
opening is small as compared to the distance between the source and the image.
[Ans 4 times]
8. In a straight edge diffraction pattern, a plane wave of wavelength 600 nm is used. If the
distance between the screen and the straight edge is 1 m, calculate the distance between
the most intense maximum and the next maximum. [Ans. 1.8 cm]
9. In a straight edge diffraction pattern, the most intense maximum occurs at a distance of
1 mm from the edge of the geometrical shadow on a screen 3 m away. Calculate the
wavelength of light used. [Ans. 448 nm]
10. For problem no. 8 above, if the distance between the eye piece and the straight edge is
increased to 4 m and the distance between the straight edge and the slit source is 6 m,
calculate the positions of (a) first three maxima and (b) the separation between them.
[Ans. (a) 0.2 cm, 0.35 cm, 0.45 cm; (b) 0.15 cm, 0.1 cm]
CH A P T E R 6
FRAUNHOFER DIFFRACTION

EXPECTED LEARNING OUTCOMES


After reading this chapter, the student will acquire the capability to:
• Describe experimental arrangement for observing Fraunhofer diffraction from a narrow vertical
slit and a circular aperture.
• Explain the observed single slit diffraction pattern on the basis of simple theoretical analysis.
• Compare single slit and double slit diffraction patterns.
• Discuss double slit and N-slit diffraction patterns.
• Derive the expression for the intensity distribution for the double slit diffraction pattern and
extend it for N equally spaced slits.
• Describe the use of diffraction gratings in spectral analysis.
• Explain how diffraction limits image forming ability of optical devices.
• Apply Rayleigh criterion to discuss resolving power of some optical devices.
• Solve numerical problems.

6.1 INTRODUCTION

In the preceding chapter, we studied how Fresnel diffraction evolves as a function of distance
between the aperture and the screen as well as the source. A particularly interesting observation
was its transition to Fraunhofer diffraction. When a plane wavefront is incident at a diffracting
aperture, this transition is determined by the ratio of the size of the aperture to its distance from
the source and/or the point of observation. A qualitative analysis based on Fresnel construction
as well as rigorous analysis for diffraction produced by simple obstacles was particularly exciting.
In this chapter, we will discuss Fraunhofer diffraction, also known as far-field diffraction.
In Section 6.2, we begin by describing an experimental arrangement that can be used to
observe Fraunhofer diffraction and discuss the salient features for a single slit illuminated by
a point source. You will learn that when a plane wave is incident on a small opening, the
Fraunhofer diffraction pattern consists of a series of spots situated symmetrically about a central
spot along a horizontal line. The central spot is the brightest and intensity decreases rapidly as

194
Fraunhofer Diffraction 195
we move away from it. A simple theoretical analysis of the observed results is also presented
in this section. You will note that theoretical analysis of Fraunhofer diffraction is particularly
simple when a plane wavefront from a point or a slit source is incident on the diffracting
aperture.
We are aware that most optical instruments use lenses which are circular in shape.
To understand the physics of how diffraction limits the ability of optical instruments, including
human eye, we discuss Fraunhofer diffraction by a circular aperture in Section 6.3. You will
learn that diffraction pattern consists of a central bright disc surrounded by concentric dark and
bright rings and diffraction places a fundamental restriction in respect of sharpness and
distinctiveness, i.e. resolution of objects. At one point in time this restriction hampered the
spectroscopic work for substances whose spectrum consisted of doublets. This problem was
overcome by increasing the number of diffraction slits, which led to the development of a
diffraction grating.
There is a general perception that wave phenomena such as interference and diffraction can
be observed in physics laboratory only. To dispel this misconception, in the preceding chapters
we listed a number of examples of interference and Fresnel diffraction which could be observed
in daily life. To demonstrate that it is in fact possible to observe diffraction at macroscopic level,
in Section 6.3, we have highlighted the case of diffraction halos, which arise due to diffraction
of light by suspended water droplets (n = 1.33) in air (n = 1). You will also learn that when
observed through less dense cloud cover around the sun or the moon, the diffraction halos are
referred to as coronas. (You should not confuse diffraction halos with ice crystal halos, which
arise due to refraction and dispersion of light and have red colour on the inner side of the ring.)
In Section 6.4, we begin by listing qualitative features of the observed double slit diffraction
pattern and compare these with those of a single slit. A distinctive feature of double slit Fraunhofer
diffraction pattern is that it consists of equally spaced bright and dark fringes as seen in
interference experiments. This is followed by calculation of resultant intensity distribution by
extending the analysis used for the single slit. You will discover that the intensity of the central
maximum is four times the intensity due to either slit at that point. These results also suggest
that sharpness of the principal maximum should increase if the number of slits is increased.
The results for double diffracting slit are generalized for N equally spaced slits in Section 6.5.
You will discover that with an increase in the number of slits, interference maxima
get sharper and as N ® ¥, these maxima become narrow lines. This makes a diffraction grating
an excellent tool for spectral analysis. In nature, we observe several diffraction grating effects.
The layered structure in the retina of a cat acts as a reflection grating and is responsible for the
metallic green reflection at night. Similarly, the green on the neck of a male mallard duck and
blue appearance of wings of the peacock feathers are manifestations of diffraction grating effect.
One of the major concerns while working with a plane transmission grating, which uses
lenses to focus light, arises due to chromatism. This problem was overcome by Rowland using
a concave grating through its two-fold action—reflection and diffraction. We have also discussed
it in this section. For completeness, we will also discuss echelon gratings, which not only help
to overcome shortcomings arising due to unequal spacing between rulings on a plane transmission
grating but also improve resolving power.
A common feature of diffraction limited image of an object is that it is fringed, even if we
use an aberration-free converging lens. That is, the image of a point object is spread over a small
196 Wave Optics

area on the observation screen. It means that diffraction limits the ability of optical devices to
transmit perfect information about any object. Broadly speaking, diffraction limited systems can
be classified into two groups:
(i) Systems which enable us to see two objects as distinct. Human eye, microscope and
telescope belong to this group.
(ii) Systems which form a spectrum and enable us to see two or more colours (wavelengths)
as distinct. A diffraction grating and prism belong to this group.
In principle, in both types of instruments two close fringed (diffraction) patterns are obtained
on an observation screen. The ability of an optical instrument to distinguish two close but
distinct diffraction images of two objects or wavelengths is measured in terms of what is known
as resolving power. In Section 6.6, we will discuss resolving power of optical instruments using
Rayleigh criterion, which states that two diffraction images are said to be just resolved when
the first minimum of diffraction pattern of one object falls on the same position where the
central maximum of the diffraction pattern of the other object lies. This has vast implications
for space as well as sub-atomic research. In this section, you will also learn about ways to
improve resolution so as to see fainter stellar bodies deeper into space or microscopic details
of materials for technological applications.

6.2 SINGLE SLIT DIFFRACTION PATTERN

We know that to observe Fraunhofer diffraction pattern, the prerequisites are:


• a narrow and preferably monochromatic light source far away (effectively at infinity)
from the diffracting aperture, and
• an observation screen placed at an effectively infinite distance from the aperture.
As such, it is not practical to put the source of light and/or the observation screen at an
infinite distance from the diffracting aperture because
– the amount of diffracted light that reaches the observation screen will be almost negligible
due to inverse square law, and
– the physical dimensions of the equipment and the laboratory required to observe Fraunhofer
diffraction will have to be infinitely large.
You may now ask: Is there any way out of these challenges? We can readily overcome
these difficulties, if we use converging lenses—the source and the observation screen are placed
respectively at the first and second principal foci of the lenses. Let us discuss it now.

6.2.1 Point Source

Refer to Figure 6.1, which shows the experimental arrangement used to observe Fraunhofer
diffraction of a single vertical slit. This ensures that the source and the observation screen are
at infinity with respect to the diffracting aperture. To understand how this is achieved, let us
assume that S is a monochromatic point source of light, which emits spherical waves. It is
Fraunhofer Diffraction 197
placed in the first focal plane of converging lens L1, which allows only plane waves to be
incident on a long narrow vertical slit in the diffracting screen. The slit is essentially a rectangular
aperture whose length is significantly large compared to its breadth (b). The observation screen,
which could also be a photographic plate, is kept in the second focal plane of another converging
lens L2 placed on the other side of the slit. So, we can say that light focused at any point on
the observation screen is due to parallel diffracted wavelets originating from different parts of
the vertical slit. The observation screen and the diffraction screen are kept parallel to one
another and normal to the common axis of converging lenses L1 and L2. It is important to
mention here that the common axis is made to pass normal to the length and through the middle
of the slit, both in height and width. Note that the beam parallel to the axis of the lens is brought
to the focal point in the back focal plane, whereas the beam inclined to the axis of the lens will
be focused away from the focal point.

Figure 6.1 Schematic representation of experimental arrangement used to observe Fraunhofer diffraction
of a single vertical slit illuminated by a point source.

In physics laboratory, we can use a spectrometer to observe diffraction pattern. And to


obtain a point source, we should replace collimator slit by a fine pinhole. (This can be readily
done by covering the collimator slit with a paper having a fine hole.) The pinhole is placed at
the focal plane of the collimator lens. The observation screen is placed at the back focal plane
of the telescope. The diffracting screen is placed between the two lenses on the turn table.
Let us pause for a while and speculate on the nature of diffraction pattern that is produced
by a vertical slit. You may imagine that the pattern will consist of one or a series of vertical
lines. This is not surprising; we observe a series of horizontal streaks of light in the form of
bright elongated spots connected by faint streaks. It means that after passing through the vertical
slit, light spreads out along a horizontal line perpendicular to the length of the diffracting slit.
We can interpret this as a horizontally spread out image of the point source. How does width
of the slit influence the horizontal spread and what will happen if the slit is completely removed?
Let us now discover answers to these questions.

A. Observed diffraction pattern


Refer to Figure 6.2, which shows the salient features of observed Fraunhofer diffraction pattern
of a single vertical slit illuminated by a point source emitting monochromatic light. We note that
• The diffraction pattern consists of a horizontal streak of light in the form of bright
elongated spots along a line perpendicular to the length of the slit.
198 Wave Optics

Figure 6.2 Observed Fraunhofer diffraction pattern of a single vertical slit.

• The spot at the central point P0, which lies at the intersection of the axis of L1 and L2
with the observation screen, is the brightest. A few bright spots on either side of the
central spot are symmetrically situated with respect to P0.
• The intensity of the central spot peaks at P0. The peak intensities of other spots at Pq
decrease rapidly as we move away from P0. The central maximum is called principal
maximum whereas other maxima are called secondary maxima.
• The width of the central maximum (spot) is greater than the width of secondary maxima
(spots).
• The principal maximum is symmetrical but secondary maxima are asymmetrical.
We now discuss the theoretical basis of these observations and obtain an expression for the
intensity distribution of the Fraunhofer diffraction pattern of a single vertical slit.

B. Intensity distribution of single slit diffraction pattern


We choose the plane of the paper, which is defined by the axis of the lens L2, as horizontal.
To calculate intensity distribution, recall that the diffraction pattern lies on a horizontal line,
which is normal to the common axis of L1 and L2. So we can say that the diffracted wavefronts
will be vertical planes normal to the plane of the paper. That is, after passing through the
vertical slit, the incident plane waves are replaced by a system of vertical plane wavefronts,
which proceed in different directions. For our analysis here, we assume that the aperture (slit)
consists of N equally spaced coherent point sources such as at A, A1, A2, A3, ..., B
[Figure 6.3(a)] and are all identical, even to their polarization. When a plane wavefront is
incident on the diffracting slit, each point becomes a source of Huygens’ secondary wavelets,
as shown in Figure 6.3(b). These waves interfere amongst themselves and give rise to diffracted
plane waves.
Note that
• diffracted waves have no existence in the domain of geometrical optics, and
• the diffracted waves arise due to interaction between light and edges of the slit.
The trace of optical paths between diffracting slit and observation screen is shown in
Figure 6.3(c).
Fraunhofer Diffraction 199

Figure 6.3 (a) Cross-sectional view of the geometry for single slit diffraction, (b) Huygens wavelets
emitted across the aperture, and (c) trace of optical paths between diffracting slit and
observation screen illuminated by monochromatic plane waves.
A very clear idea about the positions of principal maximum as well as minima in the
Fraunhofer single slit diffraction pattern can be obtained by very simple qualitative arguments.
By referring to Figure 6.3(a), we can say that in the limit of a distant screen, the central point
is equidistant from each point from the slit and all diffracted waves reach the point P0 in phase
and interfere constructively. This maximum is also referred to as the principal maximum.
For any off-axis point, the path difference between waves diffracted by extreme points of
the slit is BD = b sinq. Suppose that the path difference is an integral multiple of l. Then for
m = 1, the angle q satisfies the relation
b sinq = l
200 Wave Optics

We now divide the slit into two equal halves AM and MB, as shown in Figure 6.4. Consider
the waves starting from two point sources A and M separated by a distance b/2. Obviously, the
path difference between the disturbances originating from A and M and reaching the point Pq
b O
will be AM sin q = sin q = . Recall that phase difference corresponding to path difference
2 2
of l/2 is p. Therefore, on reaching the point under consideration, these waves will interfere
destructively and lead to zero resultant intensity. Similarly, for a point A¢ just below A, we can
identify a point M¢ just below M such that path difference between wavelets emanating from
there is also l/2. On superposition, this pair will also lead to zero intensity in the direction q.
It means that when path difference between waves diffracted by extreme points of a slit is equal
to l, we can pair off all the points in the upper half of the slit with the corresponding points
in the lower half of the slit such that the resultant intensity at the point of observation is zero.
This explains why we get minimum intensity at a point on the screen in the direction q defined
by b sinq = l, i.e. when the path difference between the rays from extremes equals one
wavelength.

Figure 6.4 Division of slit into two halves to derive the condition for first minimum.

Let us now consider the case m = 2 so that the path difference b sin q between the waves
emanating from the points A and B is equal to 2l. In this case, we imagine that the slit is divided
into four equal parts. Then by similar pairing, we can show that wavelets originating from points
in the first and second quarters have a path difference of l/2 and annihilate each other. By the
same argument, the wavelets emanating from points in the third and fourth quarters cancel out
each other so that the resultant intensity at the point of observation in the focal plane of the
converging lens is zero.
Similarly, when m = 3, the path difference between the waves emanating from extreme
points will be b sin q = 3l. In this case, we divide the slit into six equal parts and by similar
reasoning show that waves starting from corresponding pairs annihilate the effect of each other
leading to zero intensity at the point of observation in the focal plane of the lens. Thus we may
conclude that when the path difference between the parallel diffracted waves emanating from
extreme points of a slit in a particular direction is an integral multiple of l, the resultant
diffracted intensity in that direction is zero.
Let us now calculate the resultant field at an arbitrary point Pq on the observation screen
(i.e. focal plane of the lens). By referring to Figure 6.3(a) again, we note that we have to sum
Fraunhofer Diffraction 201
up contributions of different wavelets reaching the point of observation from the diffracting
aperture. For this, we must know the amplitudes and phases of the wavelets. Since the distance
of observation screen from the diffracting screen is appreciably greater than the width of the
aperture, we can take their amplitudes to be nearly equal. It means that we have to consider only
the phases of the disturbances reaching the point of observation.
We note that the points A, A1, A2, A3, ..., B
• form a series of coherent sources, since these have originated from the same point
source; and
• are in the same phase since these lie on the same wavefront.
Let us now see how the phase difference between different diffracted waves arises.
From the basic knowledge of wave phenomena, recall that phase difference between different
waves arises due to different path lengths traversed by them in reaching the point under
consideration. To know the path difference, we draw a plane from A normal on to the parallel
diffracted rays. The trace of this plane in the plane of the paper is AD. It means that diffracted
waves
• traverse different path lengths to reach the trace AD and hence the off-axis observation
points in different phases, in spite of starting in the same phase from points at A, A1, A2,
A3, ..., B, and
• cover equal optical path lengths from the plane AD to the point of observation Pq.
This is because in an aberration free converging lens, the optical paths of all rays between
any plane intersecting the parallel beam of light perpendicularly and the point where rays
converge after passing through the lens are equal [Figure 6.3(c)]. It means that the wavelets that
reach point Pq have the same relative phases that existed at the trace AD.
To proceed further, let us divide the aperture AB into n – 1 equal parts so that
b
AA1 = A1A2 = A2A3 = AnB = = D. It means that we have a series of point sources from
n 1
A to B. But strictly speaking, the aperture has a continuous distribution of points from A to B
and only in the limiting cases n ® ¥ and/or D ® 0, we can write (n – 1)D ® b. We will now
calculate the resultant field produced by these n sources at an arbitrary point Pq on the focal
plane of the lens at an angle q with the normal to the slit. If we consider two rays starting from
two neighbouring points A and A1, the path difference between them will be AA1 sin q, where
q is the angle between the diffracted rays and the normal to the diffracting slit. The corresponding
phase difference f is given by

2S 2S È b Ø 2S
I ( AA1 sin T ) sin T Ù ' sin T (6.1)
O O ÉÊ n  1 Ú O
Suppose the field at the point Pq due to the disturbance emanating from the point A is a0 cosw t.
Then Eq. (6.1) implies that the field due to the disturbance emanating from A1 will be
a0 cos(w t – f). Note that we have taken the amplitudes of disturbances from different points
to be equal. If the phase difference originating from successive points differ by f, the fields
due to disturbances from A2, A3, ..., B can respectively be written as a0 cos(wt – 2f),
202 Wave Optics

a0 cos(wt – 3f) ..., a0 cos[wt – (n – 1)f]. Hence, the sum of the interfering wavelets yield
resultant electric field at Pq , given by
E ( PT ) a0 cos Z t  a0 cos(Z t  I )  a0 cos(Z t  2I )  "  a cos[Zt  (n  1)I ]
0

To evaluate this sum, we use complex number representation and take the real part of
E(Pq) = a0eiwt + a0ei(wt–f) + a0ei(wt–2f) + L + a0ei[wt–(n–1)f]
= a0ei[wt–(n–1)f] [1 + eif + e2if + L + ei(n–1)f]
The terms within the square brackets form a geometric series with common ratio eif and a = 1.
1  einI
Hence, its sum has the value so that we can write
1  eiI
1  einI
E ( PT ) a0 ei[Z t ( n 1)I ] A0 eiZt
1  eiI
where we have put
i ( n 1)I 1  einI
A0 = a0 e
1  eiI
einI / 2 (e inI / 2  einI /2 ) i ( n 1)I
= a0 e
eiI / 2 (e iI / 2  eiI /2 )
È nI Ø È nI Ø
sin É Ù sin É Ù
i ( n 1)I / 2 Ê 2 Ú Ê 2 Ú
= a0 e e i ( n 1)I a0 e i ( n 1)I /2
ÈIØ ÈIØ
sin É Ù sin É Ù
Ê 2Ú Ê 2Ú

Sum of a finite geometric series


The sum of a finite geometric series having n-terms with common ratio r is given by
1  rn
S a
1 r
where a is the first term.
Hence, the resultant field at Pq can be expressed as
È nI Ø
sin É Ù
Ê 2 Ú
E ( PT ) a0 ei[Z t ( n 1)I / 2]
ÈI Ø
sin É Ù
Ê 2Ú

If we retain only the real part of the resultant field, we can write
È nI Ø
sin É Ù
Ê 2 Ú Ë IÛ
E(Pq) = a0 cos ÌZ t  ( n  1)
ÈIØ Í 2 ÜÝ
sin É Ù
Ê 2Ú
Fraunhofer Diffraction 203

Ë IÛ
= ET cos ÌZ t  ( n  1) (6.2)
Í 2 ÜÝ
where Eq is the amplitude of the resultant field at Pq :

È nI Ø
sin É Ù
Ê 2 Ú
ET a0 (6.3)
ÈIØ
sin É Ù
Ê 2Ú

In the limit n ® ¥ and D ® 0 in such a way that nD ® b, from Eq. (6.1), we have
nI n 2S S S
– ' sin T (n') sin T  b sin T (6.4)
2 2 O O O
so that
2S b sin T
I
O n
will be very small for n ® ¥. We may therefore write
ÈIØ I S b sin T
sin É Ù  (6.5)
Ê 2Ú 2 nO
On substituting the values of (nf/2) and (f /2) from Eqs. (6.4) and (6.5) respectively in Eq. (6.3),
we get on simplification
È S b sin T Ø
sin É Ù
ET na0
Ê O Ú na sin E A È sin E Ø
0 É E Ù
È S b sin T Ø E Ê Ú
É Ù
Ê O Ú
where
A = na0
and
b sin T
E S (6.6)
O
Note that in the limit n ® ¥ and a0 ® 0, the product na0 tends to a finite limit. Moreover, for
a given wavelength, b signifies half of the phase difference between disturbances originating
from the extreme points A and B.
The expression for resultant field at Pq takes the form
sin E
E ( PT ) A cos(Z t  E ) ET cos(Z t  E ) (6.7)
E
The corresponding intensity distribution at Pq is given by
2 2
È sin E Ø È sin E Ø
IT A2 É Ù I0 É Ù (6.8)
Ê E Ú Ê E Ú
204 Wave Optics

Let us pause for a while and reflect on what we have achieved so far. Note that Eq. (6.8) gives
intensity of diffracted light in different directions. It shows that the intensity will be maximum,
equal to I0, when (sin b/b) ® 1. This can happen only for b ® 0. It readily follows from
Eq. (6.6) that b = 0 for q = 0. So we can write
Iq =0 = I0 = A2 (6.9)
Did you not expect this result expected on geometrical considerations? We may now ask: Will
there be more than one maximum? Let us now discover answer to this question.

Positions of maxima and minima


Refer to Figure 6.5, which depicts a plot of Eq. (6.8) for intensity distribution as a function
of b. Note that

Figure 6.5 Intensity distribution corresponding to the single slit Fraunhofer diffraction pattern.

• for b = 0 = q, the intensity distribution displays a maximum as predicted by Eq. (6.9);


• on either side of the principal maximum, the intensity falls gradually and becomes zero
for b = ±p, since sin(±p) = 0. This defines the first minimum.
One may now logically ask: What is the angular width of the principal maximum? Since
first minimum on either side of the principal maximum occurs at b = ±p, the angular width of
the principal maximum is from b = –p to b = p.
The second minimum on either side of the principal maximum occurs at b = ±2p. So we
can write the condition for occurrence of minima as
b = ±p, ±2p, ±3p, ... = mp; m = ±1, ±2, ±3, ... (6.10)
Fraunhofer Diffraction 205
Note that we have not included the value of m = 0 in Eq. (6.10) because it corresponds to the
principal maximum.
We can express the condition of minima in terms of the physical characteristics of the
diffracting arrangement (slit width and wavelength of light) using Eq. (6.6) as
b sin q = ±ml; m = ±1, ±2, ±3, ... (6.11a)
or
1 È mO Ø
T “ sin É Ù (6.11b)
Ê b Ú

This result shows that for a slit of given width, the angle of diffraction, i.e. the spread in
diffraction pattern varies directly and linearly with the wavelength. It means that red light will
be diffracted more than any other wavelength. Similarly, we can say that as slit becomes
narrower, the spread of principal maximum obtained with a given wavelength increases.
Conversely, wider the slit, sharper will be the diffraction pattern. What will happen if slit
aperture is varied relative to the wavelength of light used? We expect the diffraction pattern to
spread out.
So far we have considered that a monochromatic source of light illuminates the slit.
Now a logical question to ask is: What will happen if white light is made to illuminate the slit?
If you think that each wavelength will be diffracted independently, you are on the right track.
We expect to observe coloured fringes; in fact, the central spot will be white in this case also
but it will be surrounded by coloured fringes. We have seen that red colour is diffracted more
than any other wavelength. It means that the outer part of this pattern will tend to be reddish.
You can easily observe this pattern by looking through the tines of a dinner fork at a candle in
a dimly illuminated room. On rotating the fork about its handle slowly, you will observe the
diffraction pattern as soon as the cross-sectional area becomes small enough.
Note that Eq. (6.11) gives the conditions for minima on either side of the principal maximum.
If more than one minimum is observed in the spectrum, it is obvious that there will be maxima
in between. These are referred to as secondary maxima. To determine the directions and positions
of secondary maxima, we differentiate Eq. (6.8) with respect to b and equate the resulting
expression to zero. This gives

dIT È sin E Ø È E cos E  sin E Ø ÈE cos E  sin E Ø


2I 0 É Ù 2 I 0 sin E É 0
dE Ê E Ú ÉÊ E 2 Ù
Ú Ê E3 Ù
Ú
or
sin b (b cosb – sinb) = 0 (6.12)
From this result, we get two conditions: sin b = 0 and b – tanb = 0. The first of these conditions
implies that b = ±mp, where m is any integer. As discussed before, this condition corresponds
to minima and gives no further information. It is therefore a trivial condition.
The conditions for maxima are roots of the transcendental equation
(b cosb – sinb) = 0
or
tanb = b (6.13)
206 Wave Optics

The root b = 0 defines the central or principal maximum. The other roots of this equation are
obtained by a graphical method. All that we have to do is to recall that an angle equals its
tangent at intersections of the curves defined by y = b and y = tanb. While y = b is a straight
line passing through the origin, y = tan b is represented by a family of curves having for
asymptotes b = ±p/2, ±3p/2, ±5p/2, ... . The plots of these curves are shown in Figure 6.6.
The points of intersection, excluding b = 0, occur at b = 1.43p, 2.46p, 3.47p and so on.
These respectively give the positions of the first, second, and third subsidiary or secondary
maxima on either side of the principal maximum. Note that secondary maxima do not fall
midway between two minima. For instance, the first secondary maximum occurs at 1.43p rather
than 1.50 p. Similarly, the second secondary maximum occurs at 2.46p rather than 2.50p and
so on. We may therefore conclude that intensity curves are asymmetrical and the positions of
secondary maxima are slightly shifted towards the centre of the pattern. We observe this
asymmetry while performing diffraction experiments in physics laboratory.

Figure 6.6 Graphical method for determining the roots of the equation tan b = b.

You may now like to know the intensities of the secondary maxima. The intensity of the
2
sin1.43S Ø
first secondary maximum is given by I 0 ÈÉ Ù = 0.0496 I0. It means that the intensity of
Ê 1.43S Ú
the first maximum, which is closest to the central maximum, is about 4.96% of the central
maximum. Similarly, you can check that the intensities of the second and third secondary
maxima are about 1.68% and 0.83% of the central maximum. This shows that most of the
diffracted light is concentrated in the central maximum.
We now advise you to answer a Practice Exercise and calculate the intensities of the
secondary maxima by another method.

Practice Exercise 6.1 Calculate the values of b at half-way positions of first and second
secondary maxima and show that the intensities at these positions are 0.0453 and 0.0162
respectively.
Fraunhofer Diffraction 207
Another important characteristic of the principal maximum is that its width is twice the
width of the secondary maximum. To show this, we recall that the angular spread of the central
maximum is from q = sin–1(l/b) to q = –sin–1(l/b). For small q, sinq » q. It means that principal
maximum is spread from q = l/b to q = –l/b.
But the spread of the first secondary maximum on the positive side, in small angle
approximation, extends from q = l/b to q = 2l/b and on the negative side from q = –l/b to
q = –2l/b. Thus we see that the principal maximum is twice as wide as the secondary maximum.
To give you a feel of the numerical values and help fix the ideas discussed in this section,
we now give a few solved examples. You should go through these carefully.

EXAMPLE 6.1 In an experiment, a parallel beam of light from a He-Ne laser source
(l = 630 nm) is made to fall on a narrow slit of width 0.3 ´ 10–3 m. The Fraunhofer diffraction
pattern is observed on a screen placed in the focal plane of a convex lens of focal length
0.3 m. Calculate the distance between the first two minima and the first two maxima on the
screen. Assume that the lens is placed very close to the slit.

O 630 – 109 m
Solution 2.1 – 103
b 3 – 104 m
The conditions for minima are given by
b sinq = ml m = ±1, ±2, ±3, ...
For small values of q, we take sin q » q. So we can rewrite the conditions for minima as
O
T m
b
Hence, the diffraction angles for the first two minima are
q1 = 2.1 ´ 10–3 rad
and
q2 = 2 ´ 2.1 ´ 10–3 rad
These minima are separated by a distance
d1 = f (q2 – q1) = (0.3 m) ´ 2.1 ´ 10–3 = 0.63 ´ 10–3 m
That is, the first two minima are separated by 0.63 ´ 10–3 m on the focal plane of the lens.
The first two secondary maxima occur at b = 1.43p, and 2.46p, respectively. So, we can write

b sin T1 1.43O À (T1 ) max 1.43 – Ob 1.43 – 2.1 – 10 3


rad
and

b sin T 2 2.46 O À (T )
2 max 2.46 – Ob 2.46 – 2.1 – 10 3
rad

Hence, the first two secondary maxima will be separated by a distance


d2 = (2.46 – 1.43) ´ 2.1 ´ 10–3 ´ 0.3 m = 0.90 ´ 10–3 m
208 Wave Optics

Note that the distance between the first two secondary maxima is greater than the corresponding
distance between the first two minima.
If we calculate the separation between the second and third minima and maxima, we will
observe that
• the distance between the second and third minima is equal to the distance between the
first two minima; and
• the distance between the second and third secondary maxima is less than the distance
between the first two secondary maxima.
EXAMPLE 6.2 In Example 6.1, the slit width is changed to 0.1 ´ 10–3 m. How are the
positions of the first two minima influenced?
Solution In this case
O 630 – 109 m
6.3 – 103
b 1 – 104 m
Hence, the diffraction angles for the first two minima are
q1 = 6.3 ´ 10–3 rad
and
q2 = 2 ´ 6.3 ´ 10–3 rad
These minima are separated by a distance
d = f (q2 – q1) = (0.3 m) ´ 6.3 ´ 10–3 = 1.89 ´ 10–3 m
Recall that the corresponding separation for a slit of width 0.3 ´ 10–3 m was 0.63 ´ 10–3 m.
It means that for a given wavelength, the spread of secondary maximum increases as slit width
decreases. Moreover, the first minimum on either side of the principal maximum occurs at a
distance given by fq1. For slits of width 0.3 ´ 10–3 m and 0.1 ´ 10–3 m, it is 0.63 ´ 10–3 m and
1.89 ´ 10–3 m. It means that the spread of principal maximum increases from 1.26 ´ 10–3 m to
3.78 ´ 10–3 m as the slit becomes narrower.

You may now like to answer a Practice Exercise.

Practice Exercise 6.2 For a slit of width 8l, 4l and l, calculate the angles of diffraction
for which intensity of the principal maximum is the highest. [Ans. 7.2°; 14.5°; 90º]

On solving this exercise, we may conclude that as the aperture of the slit decreases,
the diffraction pattern spreads out. For b = l, the diffraction pattern will show no ripples.
So far we have discussed Fraunhofer diffraction produced by a slit aperture (with width
much less than its length) illuminated by a point source. We would now like you to apply this
knowledge to answer Practice Exercise 6.3 and predict the nature of diffraction pattern.

Practice Exercise 6.3 For a rectangular slit, predict the nature of diffraction pattern and
depict its qualitative nature.
Fraunhofer Diffraction 209
For a slit characterized by b » h, we expect that the emerging wave will spread along the
length as well as width of the slit. We shall discuss it more rigorously in a later section.
Proceeding further, we consider diffraction pattern produced by a diffracting aperture
illuminated by a slit source. You can consider a slit source as an extension of point source.

6.2.2 Line Source

Refer to Figure 6.7, which depicts the experimental arrangement in which a line source illuminates
the diffracting slit. (It is a modified version of the arrangement shown in Figure 6.1.) For a
qualitative discussion, it is fairly reasonable to assume that
(i) A line source is made up of a series of independent point sources such as O1, O2,
O3, ..., one above the other.
(ii) Each point source gives its own diffraction pattern in the form of a horizontal streak
of light, as in the case of a vertical slit.

Figure 6.7 Experimental arrangement for observing diffraction pattern of a vertical narrow single slit
illuminated by a line source.

In physics laboratory, and in fact in most practical diffraction experiments, a slit as in a


spectrometer acts as a line source (Figure 6.8). (As a first order approximation, an ordinary
clear, straight filament display bulb is a reasonable source.) The slit of the collimator arm is
illuminated by a monochromatic source of light. In view of the discussion in the preceding
paragraph, we can say that for a given diffracting slit and the converging lenses L1 and L2,

Figure 6.8 Experimental arrangement for observing diffraction pattern of a slit illuminated by a slit
source in physics laboratory.
210 Wave Optics

• the intensity along any horizontal line will be the same;


• the central diffraction pattern due to all point sources will lie one over the other resulting
in a central bright vertical fringe; and
• minima and secondary maxima will result in the formation of a series of vertical fringes
situated at equal intervals on either side of the central fringe.
We may therefore be tempted to conclude that superposition of a series of horizontal
diffraction streaks stacked over each other in a vertical direction leads to a diffraction pattern
comprising vertical fringes. This is essentially because each point on the slit source acts as an
independent and effectively incoherent source. We will show that this is confirmed by rigorous
mathematical analysis. However, it is important to mention here that we obtain clear fringes
only when the width of the source slit is small. As the width of the source slit increases, the
width of the fringes obtained on the observation screen also increases gradually. And when the
width of slit source becomes comparable to inter-fringe separation, the diffraction pattern gets
blurred and the fringes become less clear.
Let us pause for a while and reflect on the importance of what we have discussed so far.
We now know that when a plane wave is incident on a slit-like aperture
• the emergent wave spreads out along the width of the slit with angular divergence
» l/b;
• the diffraction pattern consists of a central maximum surrounded on both sides by minima
and secondary maxima of gradually diminishing intensity; and
• the shape and/or size of the diffracting slit as well as the shape of the source influences
the spread of diffraction pattern.
You may have observed that in winter, the window glasses of a car get fogged, particularly
at night. If a motorcycle follows such a car and light from its headlight falls on the fogged
window, brilliant halos can be observed. You can produce such halos by breathing on the side
of a clear glass and then looking through the fogged area at a small source such as an ignited
match stick, penlight or a distant bulb. As a student of physics, you are familiar with the
construction of human eye; it has a lens of spherical shape, just like a lens in an optical device
such as microscope and a telescope. It suggests that Fraunhofer diffraction by a circular aperture
is an extremely important effect as it can influence our visual contact with the outside world
as well as limit the ability of optical instruments to observe sharp and distinct images. We now
discuss these and other related phenomena in the following section.

6.3 DIFFRACTION BY A CIRCULAR APERTURE

Refer to Figure 6.9, which schematically depicts the experimental arrangement for observing
Fraunhofer diffraction by a circular aperture. A stream of plane waves impinges normally
on a screen containing a circular aperture through lens L1. The emergent waves fall on lens L2
whose diameter is large compared to that of the diffracting aperture and is placed near it.
The diffraction pattern is observed in the back focal plane of lens L2. Note that the system
has rotational symmetry. Therefore, we expect the diffraction pattern to consist of concentric
dark and bright rings. (This pattern is known as Airy pattern.) You will discover a little later
Fraunhofer Diffraction 211
in this section that this guess, which is based on qualitative arguments, is well supported by
experimental observations.

Figure 6.9 Experimental arrangement for observing Fraunhofer diffraction by a circular aperture.

To obtain an expression for intensity distribution of diffraction pattern for a circular aperture,
refer to Figure 6.10. It shows a circular aperture of radius q illuminated by a point source.
Suppose that the distance between an arbitrary point P on the aperture and the point of observation
Pq (x, y, z) is R. Similarly, let us denote the distance between the centre of the aperture and the
point of observation Pq by r¢. In view of the geometry, it is more convenient to use the plane
polar coordinate system. Let us denote the coordinates of point P by (r, f), where r is its
distance from the centre O and f is the angle that OP makes with the horizontal. From the
knowledge of coordinate systems, recall that an element of area dA in cylindrical coordinates
can be expressed as dA = rdrdf.

Figure 6.10 Diffraction of a plane wave incident on a circular aperture.

The electric field at point Pq due to waves emanating from the element of area dA will be
a eikR
proportional to 0 dA, where a0 is the amplitude of the plane wave in the plane of the
R
212 Wave Optics

diffracting aperture. The resultant field at the point of observation Pq will be obtained by
summing up contributions from all such infinitesimal areas, we get
a0 eikR
u ( PT ) K Ô R
dA (6.14)

where K is the constant of proportionality and it is equal to –i/l. Note that the integration has
to be carried out over the entire area of the aperture.
Since the diffraction pattern will be in the form of concentric rings with their centres at the
point O ¢, we will calculate the intensity distribution along the x-axis.
Since r varies from 0 to q and f varies from 0 to 2p, we can rewrite Eq. (6.14) as
q 2S
u ( PT ) Ô
C Ud U
0
Ô exp(ik U sin T cos I )dI
0
ia eikR
where C =  0 . To evaluate the integral over r, we introduce a change of variable by
OR
defining s = kr sinq so that ds = k sinq dr and the limits of integration change to 0 and
kq sinq. Hence, the resultant field at the point of observation Pq is given by
kq sin T 2S
C
u ( PT )
( k sin T ) 2 Ô sds Ô exp(is cos I )d I (6.15)
0 0

Note that the integral associated with the variable f occurring in Eq. (6.15) arises quite frequently
in physics. It denotes zeroth order Bessel function of the first kind:
2S
1
Ôe
“is cos I
J 0 ( s) dI
2S
0

Bessel function of the first kind


The Bessel function of the first kind and of order m is defined as
2S
i m
J m ( s) exp[ (mI  s cos I )]d I
2S Ô
0
Numerical values of J0(s) and J1(s) are tabulated for a large range of values of s and are available in
most mathematical handbooks. Just like sine and cosine functions, Bessel functions have series
expansions. In Mathematical Physics course, you must have got opportunity to learn about the special
functions. In particular, you may recall Bessel function of the first order which varies with its argument
like a damped sine curve.
A general property of Bessel functions, referred to as a recurrence relation, is

d m
[ s J m ( s )] s m J m 1 (s )
ds
Fraunhofer Diffraction 213
Using this result in Eq. (6.15), we can express the resultant field at the point of observation as
kq sin T
2S C
u ( PT )
( k sin T ) 2 Ô s J 0 ( s )ds (6.16)
0

If we now use the relation connecting Bessel functions of zeroth order and first order

d
[ s J1 ( s )] s J 0 (s )
ds
Equation (6.16) simplifies to
2S C kq sin T
u(Pq ) = s J1 ( s ) 0
( k sin T ) 2
Ë 2 J1 ( D)Û
= u0 Ì
Í D Ü
Ý

2S SD
where D kq sin T q sin T sin T and u0 = 2pq2C = 2AC is the amplitude; l is the
O O
wavelength of light used, q is the angle of diffraction and D and A are respectively the diameter
and area of the circular aperture.
The intensity distribution of the diffraction pattern is given by

Ë 2 J1 ( D ) Û2 Ë 2 J1 (kq sin T ) Û
2
(6.17)
I ( PT ) I0 Ì Ü I0 Ì
Í D Ý Í kq sin T ÝÜ
Note that I0 is the intensity at q = 0 and corresponds to the central maximum.
The intensity distribution for a circular aperture corresponding to Eq. (6.17) is depicted in
Figure 6.11(a) as a function of a. Note that
• The intensity is maximum at the centre of the pattern since
2 J1 (D )
lim 1
D 0 D
The central maximum corresponds to very bright circular spot known as the Airy disc
and may be regarded as the diffracted image of the point source by a circular aperture
[Figure 6.11(b].
• The Airy disc is surrounded by a dark ring that corresponds to the first zero of the
function J1(a) = 0 for a = 3.832. It means that the first dark fringe appears for
3.832O 1.22O
sin T # (6.18)
SD D
where D is the diameter of the aperture.
• As diameter of the circular aperture approaches the wavelength of light used, the Airy
disc can be very large and the circular aperture begins to resemble a point source
emitting spherical waves.
214 Wave Optics

Figure 6.11 (a) Plot of intensity distribution as a function of a for a circular aperture, and (b) Diffracted
image of a point source by a circular aperture.

• Higher order zeros of the function J1(a) occur at a = 7.02, 10.17, ... and signify successive
dark fringes in the diffraction pattern, as shown in Figure 6.11(b). A detailed mathematical
analysis shows that about 84% of the total energy is contained within the first dark ring,
i.e. in the Airy disc, and 91% within the second dark ring.
• Bright fringes of gradually diminishing intensity are located where a satisfies the condition

d Ë 2 J1 (D ) Û
0
d D ÌÍ D ÜÝ
This is equivalent to J2(a) = 0 and occurs for a = 5.14, 8.42, 11.6 and so on.
Let us now pause for a while and compare Eq. (6.18) with the corresponding result for the
narrow slit. We note that the angular half-width of the central disc, i.e. the angle between the
central maximum and the first minimum for a circular aperture differs from that for the slit
pattern through a factor of 1.22. This result finds an interesting application in determining the
dimensionality of laser beams. For instance, for a He-Ne laser beam of wavelength 630 nm,
we can easily calculate the divergence semi-angle as 2.1 ´ 10–5 rad, if the diameter of the
aperture is 3 ´ 10–3 m. It means that on traversing a distance of about 103 m, the beam width
will be only about 2 cm. Such a high directionality of a laser beam facilitates us to focus it on
an extremely small region/target.
Another simple application of Eq. (6.18) is in the determination of the aperture of a loud
speaker. On the basis of common perception, one would expect that a loudspeaker of smaller
aperture will generate more directed beam of sound waves. However, this is not true; in fact,
it causes significant diffraction divergence leading to a greater loss of energy; only a small
fraction of energy will reach the listener/audience. That is why large aperture loudspeakers are
used in large public meetings and the use of small aperture loudspeakers is confined to smaller
gatherings.
Let us now discuss an activity by which you can observe diffraction pattern produced by
a circular aperture.
Fraunhofer Diffraction 215

Activity 6.1 Take a sheet of aluminium foil and make a small pinhole in it. Look through
it at a distant light bulb or a lighted candle in a poorly illuminated room. What do you
observe? We expect you to observe an Airy disc at the centre of the diffraction pattern.
Next take a random array of small circular apertures and illuminate it by a white point source.
The plane waves incident on each of these apertures will generate an Airy type diffraction
pattern. What will you observe if the apertures are small and close together? We expect
the answer that diffraction patterns will be large and overlap with the central white disc
surrounded by circular coloured rings. It is known as halo. Which colour do you expect to
observe on the outermost rim of the halo? According to us, red colour will be on the outermost
rim of the halo.

You may now ask: Can we observe similar halos when the diffraction takes place due to
a random array of circular obstacles? You can check by doing a small activity and try to observe
similar halos yourself.
The example given below will give you an idea of the size of Airy disc.

EXAMPLE 6.3 Plane waves from a sodium source emitting yellow light of wavelength
589 nm are incident on a circular aperture of radius 2 ´ 10–3 m. The diffraction pattern is
observed in the back focal plane of a convex lens of focal length 0.3 m. Calculate the diameter
of the Airy disc.
Solution To calculate the diameter of the Airy disc, we have to determine the angular location
of the first dark ring. From Eq. (6.18), we know that
1.22O 1.22O
sin T
D 2r
On substituting the given values, we get

1.22 – (589 – 10 9 m)
sinq =
2 – (2 – 10 3 m)
= 1.796 ´ 10–3
In the small angle approximation, sin q » q, so that
q = 1.796 ´ 10–3 rad = 0.103º
Since the diffraction pattern is observed in the focal plane of the convex lens of focal length
0.3 m, the radius of the first dark ring
x = f q = (0.3 m) ´ 1.796 ´ 10–3 = 0.54 ´ 10–3 m
This value of x essentially signifies the radius of the Airy disc. Hence the diameter of the Airy
disc is nearly 1.08 ´ 10–3 m.
You are now advised to answer a Practice Exercise.
216 Wave Optics

Practice Exercise 6.4 A He-Ne laser emits a beam of diameter 2 ´ 10–3 m and wavelength
630 nm. It is directed towards an aeroplane flying at a height of 39 ´ 103 m. Calculate the
diameter of the light patch produced on the surface of the aeroplane. Assume that light is not
scattered in earth’s atmosphere. [Ans. 30 m]

From the answer of this question, we note that almost half of the aeroplane will be covered
by the patch created by the laser light. This forms the guiding principle of laser-guided weapons
programme.
So far, we have discussed the Fraunhofer diffraction produced by a slit or a circular aperture.
When a narrow vertical slit is illuminated by a point source, the Fraunhofer diffraction consists
of a series of spots along a horizontal line and situated symmetrically about a central spot. The
intensity of the central spot is maximum. For a circular aperture, the diffraction pattern consists
of concentric rings around a bright central disc. We now discuss Fraunhofer diffraction pattern
when we increase the number of slits.

6.4 FRAUNHOFER DIFFRACTION FROM TWO VERTICAL SLITS

Refer to Figure 6.12. It shows the experimental arrangement for observing diffraction from two
vertical parallel slit apertures in an opaque screen. Lens L focuses light from a monochromatic
source S on the slit source, which is placed in the focal plane of the converging lens L1. As a
result, a parallel beam of light emerging from it falls on two long vertical diffracting slits. (Sometimes
we refer to it as diffraction by a double slit.) The converging lens L2 focuses the diffracted light
on the observation screen placed in its back focal plane. Recall that this arrangement corresponds
more closely to the conditions under which an experiment is performed in laboratory. In other
words, the diffraction pattern from a slit source has greater practical significance.

Figure 6.12 Experimental arrangement for observing diffraction from two identical vertical slits.

We assume that both slits are identical having the same width, say b and height h. If the
width of the intervening opaque space between the slits is a, the distance between two similar
points in these slits will be a + b. Let us denote it by d. Before deriving an expression for the
intensity distribution of diffraction pattern of a double diffracting slit, we discuss it on physical
considerations.
Suppose that one of the diffracting slits, say slit 1 in Figure 6.12 is blocked. Obviously,
we expect to observe the single slit diffraction pattern, which in the instant case will be due to
slit 2. Next we uncover slit 1 and block slit 2. Do you expect any shift in the lateral position
Fraunhofer Diffraction 217
of the pattern? Common sense suggests that as long as light is diffracted by only one slit,
we should get exactly the same intensity distribution, irrespective of whichever slit is blocked.
However, the patterns in the two cases may be laterally shifted from one another. Surprisingly,
the diffraction patterns are identical and located at the same position. You may now logically
ask: Why the effect of the location of the slit does not show up? The patterns are not laterally
displaced with respect to one another because of the presence of converging lens L2. The
diffracted wavefronts originating from either slit travel along the axis of lens L2 and are brought
to focus at P0 to form the peak of the central spot. However, the wavelets originating from either
slit at an angle q are focused at Pq .
Now uncover both the slits so that each slit gives its own diffraction pattern. The salient
features of the resultant diffraction pattern are shown in Figure 6.13.

Figure 6.13 Qualitative features of observed Fraunhofer diffraction due to a double slit.

Note that
• unlike the single slit diffraction pattern, the double slit diffraction pattern consists of a
number of equally spaced fringes, as observed in interference pattern;
• the intensity of the fringes gradually decreases as we move away from the centre;
• the intensity at the maximum of double slit pattern is greater than the intensity of
principal maximum in single slit diffraction pattern; and
• fringes reappear three or four times with gradually diminishing intensity before they
become too faint to be observed.
You may now like to know: How bright are double slit fringes compared to those in single
slit pattern? What explains the salient features of double slit diffraction pattern? And so on. Let
us seek answers to these and similar other questions.

6.4.1 Intensity Distribution in Two Slit Diffraction Pattern

To calculate the intensity distribution in double slit diffraction pattern produced using the
arrangement shown in Figure 6.12, we extend the procedure used for a single slit illuminated
by a point source. A logical question now arises: Is it sufficient for us to consider a point source
though we are using a slit source? The answer to this question is in the affirmative. This is
because a slit source can be considered to be made up of a series of point sources placed one
above the other and a point source gives the intensity distribution along a section perpendicular
to the vertical fringes formed by a slit source.
Refer to Figure 6.14. Slit 1 can be assumed to act as a source of Huygens’ secondary
wavelets which appear as diffracted plane wavefronts originating from a large number of
points such as A1, A2, A3, ... . As before, we can represent these by a0 coswt, a0 cos(wt – f),
218 Wave Optics

a0 cos(wt – 2f), ..., a0 cos[wt – (n – 1)f], where f is the constant phase difference.
The magnitude of field E1 produced by slit 1 at the point Pq is given by Eq. (6.7):
sin E
E1 A cos(Z t  E )
E
b sin T
where E S .
O

Figure 6.14 Fraunhofer diffraction of a plane wave incident normally on a double slit.

For every point like A1, A2, A3, in slit 1, we can identify a corresponding point like
B1, B2, B3, , in slit 2 at a distance d. The phase difference between diffracted wavefronts
reaching Pq from any two corresponding pairs of points like (A1, B1), (A2, B2), (A3, B3), ...,
on these slits is given by
2S 2S
G ( a  b)sin T d sin T (6.19)
O O
So we can represent the diffracted plane wavefronts originating from points B1, B2, B3, , as
a0 cos(w t – d), a0 cos(w t – d – f), a0 cos(w t – d – 2f), ..., and the magnitude of field E2
produced by slit 2 at point Pq will be given by
sin E
E2 A cos[(Z t  E )  G ]
E
Since the points A1, A2, A3, and B1, B2, B3, are coherent, the magnitude of the resultant
field at Pq is given by

Eq = E1 + E2
sin E
= A [cos(Z t  E )  cos(Z t  E  G )]
E
sin E È GØ ÈG Ø
= 2A cos É Z t  E Ù cos É
E Ê 2Ú Ê 2Ù
Ú
Fraunhofer Diffraction 219

sin E
= 2A cos(Z t  E  ] ) cos ]
E
= E0 cos(wt – b – z) (6.20)
G S sin E
where ] d sin T and E0 2 A cos ] .
2 O E
Since intensity is directly proportional to the square of the amplitude, we get
2 2
2 È sin E Ø 2 È sin E Ø
IT 4A É Ù cos ] 4I0 É Ù cos2 ] (6.21)
Ê E Ú Ê E Ú

From this result, we note that


• For q = 0, b as well as z vanish suggesting that intensity of the central bright fringe is
Iq = 0 = 4A2 = 4I0. That is, the double slit provides four times more intense central
maximum compared to the single slit.
• I0(sin2b/b 2) corresponds to the intensity distribution produced by a single slit.
• The intensity distribution is a product of two terms: (sin2b/b 2) represents the diffraction
pattern produced by one of the slits of width b and cos2z represents the interference
pattern produced by two diffracted wavelets (of equal intensity) having phase
difference d.
Now refer to Figure 6.15. It depicts separate plots of cos2 z as a function of z
[Figure 6.15(a)], sin2b/b 2 versus b [Figure 6.15(b)], and their product [Figure 6.15(c)] for the
particular case of z = 3b . For cos2z, we obtain a set of equidistant maxima of equal intensity
located at b = 0, ±p, ±2p, ±3p, . The curve for sin2b/b 2 shows a maxima at b = 0 and minima
at b = ±p, ±2p, ±3p, ... . The product of these functions shown in Figure 6.15(c) shows that
modulation by the diffraction factor sin2b/b 2 results in the brightest central fringe and the
intensity of successive two fringes decreases until we reach the point b = p, where the intensity
is zero.
For very narrow diffracting slits (a > b), the (sin2b/b 2) ® 1. It means that this term will
show essentially no variation with angle of diffraction, q, and the resultant pattern will be
similar to Young’s interference pattern. On the other hand, for a = 0, the two slits coalesce into
one so that z = 0 and Eq. (6.21) reduces to Iq = 4I0(sinb/b )2. This is equivalent to Eq. (6.8)
for single-slit diffraction with the source strength doubled.
Let us now reflect on the physics involved in formation of the double slit diffraction. Recall
that the diffracted light waves emerging from the two slits were assumed to be given out by
coherent pairs of sources in them. When these waves reach the point of observation, there is
a finite phase difference between them and their superposition leads to formation of interference
fringes on the screen. Recall that the intensity of a fringe is determined by the intensities of
interfering waves and their phase difference at the point of observation. We know that the
intensities of diffracted beams are determined by the conditions under which diffraction occurs
as also the direction of observation (determined by the angle q). It means that the intensities of
interference fringes will not be the same at all points on the screen. In particular, the directions in
which the intensities of diffracted beams are large, the constructive interference will produce
220 Wave Optics

Figure 6.15 Intensity curves for double slit diffraction for z = 3b.

brighter fringes. On the other hand, the directions in which the intensities of diffracted beams
are low, even constructive interference will produce less bright fringes.
In the above discussion, we have explained the formation of fringes as a result of interference
between two diffracted beams. It means that in diffraction, secondary wavelets originating from
different parts of the same wavefront are made to superpose. (In interference, the wavelets arise
from various sources situated in the aperture between the two jaws of a slit.)
Now consider that the monochromatic source of light is replaced by a source such as
sodium lamp, which emits a doublet (l1 = 589.0 nm and l2 = 589.6 nm). Will the double slit
diffraction pattern get modified? If so, how? We can check that both wavelengths will give rise
to independent diffraction patterns with interference fringes in between. However, these diffraction
patterns will be superimposed on one another and coincide at q = 0.
Let us now investigate the positions of maxima and minima.

A. Positions of minima and maxima


To determine the positions of minima and maxima in double slit diffraction pattern, we refer to
Eq. (6.21) again. Note that the intensity of diffraction pattern will vanish when either
Fraunhofer Diffraction 221
(sin2b/b 2) (the diffraction term) or the cos2z (the interference term) is zero. The diffraction term
will be zero for
b sin T
E S mS m¹0
O
or
b sinq = ml m¹0 (6.22)
This relation specifies the directions along which the available intensity of either beam is zero
due to diffraction at each slit.
The interference term will be zero when

d sin T È 1Ø
z= S Én  ÙS
O Ê 2Ú
or
È 1Ø
d sinq = ÉÊ n  ÙO n = 0, 1, 2, ... (6.23)

This result specifies the directions along which intensity will be zero due to destructive interference
between the two beams emanating from the slits.
Unlike in case of minima, we cannot specify the exact positions of maxima by a simple
relation. Do you know the reason for this? It is because intensity is a product of two terms. But
if we assume that variation in sin b/b is almost negligible over a given region, it is possible to
locate the approximate positions of maxima. (Note that this approximation is reasonable for
narrow slits only.) The positions of maxima are then solely determined by the cos2z factor and
occur for
z = np
or
d sin q = nl n = 0, 1, 2, 3, ... (6.24)
Recall that d sinq denotes the path difference between any two corresponding points in the
diffracting slits. When this path difference is an integral multiple of l, two beams interfere
constructively leading to the formation of a series of bright fringes. The central fringe corresponds
to d sinq = 0. The nth fringe on either side of the central fringe occurs at d sinq = nl. For this
reason, n is referred to as the order of interference.

B. Missing orders
S b sin T S d sin T
In Eq. (6.21), b = and z = . It means that b and z are mutually dependent.
O O
To discover the relation connecting these, we calculate their ratio:

] S d O sin T d ab
(6.25)
E S bO sin T b b
Let us now consider the case when d is an integral multiple of b and we can write d = pb, where
p is an integer (>1). This will happen when the opaque portion is such that we can write a = b,
222 Wave Optics

a = 2b, a = 3b, so that d/b = p = 2, 3, 4, ... . You can easily check that under these conditions,
the directions of diffraction minima and interference maxima will necessarily coincide.
Note that if the directions of diffraction minima are given by
b sinq m = ml (6.26)
we will automatically have interference maxima in these direction when d = pb since
d sin q = (pb) sin q = p(b sin q) = pml = nl (6.27)
where we have put n = pm. Since the possible values of p are 2, 3, 4, ..., and those of m are
1, 2, 3, ..., the nth order interference fringe which satisfies the condition n = pm will have zero
intensity. That is, the intensity of interference as well as diffracted beams will be zero because
of diffraction condition. As a result, their constructive interference also leads to net zero intensity.
These are usually referred to as missing orders. For example, for p = 2, the missing orders will
correspond to 2, 4, 6, ... for m = 1, 2, 3, ... . What will be the missing orders for p = 4?
In the special case d/b = 1, the two slits will merge into one another since a = 0. Then all
interference effects will disappear. Physically it means that we have a single slit of twice the
width.
To fix these ideas, carefully go through the following solved example.

EXAMPLE 6.4 A double slit arrangement with b = 8.0 ´ 10–5 m and d = 4.0 ´ 10–4 m is
illuminated by light of wavelength 640 nm. Calculate the number of interference minima that
will occur between the two diffraction minima on either side of the central maximum. In the
experimental arrangement, the screen is placed at a distance of 3 m from the diffracting aperture.
Also calculate the fringe width.
Solution From Eq. (6.22) we recall that the first diffraction minima on either side of the
central maximum will occur at b sin q 1 = ±l so that

O 640 – 109 m
sin T1 “ 8 – 103
b 8 – 105 m

The interference minima will occur when Eq. (6.23) is satisfied:

È 1Ø
d sin T n Én  ÙO
Ê 2Ú
On substituting the given values, we get

È 1Ø O È 1 Ø 640 – 109 m È 1Ø 3
sin T n ÉÊ n  ÙÚ ÉÊ n  ÙÚ ÉÊ n  ÙÚ – 1.6 – 1 0 n = 0, 1, 2, ...
2 d 2 4.0 – 104 m 2

Hence, the values of sinqn will be 0.8 ´ 10–3, 2.4 ´ 10–3, 4.0 ´ 10–3, 5.6 ´ 10–3, and 7.2 ´ 10–3.
This shows that there will be ten interference minima between the two first order diffraction
minima.
Fraunhofer Diffraction 223
In small angle approximation, sinqn » qn and we can write q1 = 0.8 ´ 10 –3 rad,
q2 = 2.4 ´ 10 –3 rad, q3 = 4.0 ´ 10–3 rad, q4 = 5.6 ´ 10–3 rad and q5 = 7.2 ´ 10–3 rad. From these
values, we note that the angle between successive interference minima Dq = 1.6 ´ 10–3 rad.
Thus, fringe width is given by
Z = DDq
= 3 ´ 1.6 ´ 10–3 = 4.8 ´ 10–3 m

You should now answer a Practice Exercise.

Practice Exercise 6.5 If we use a white light source in the arrangement as shown in
Figure 6.12, what changes will occur in the double slit diffraction pattern?

Recall that the diffraction factor sin2b/b 2 results in the brightest central fringe and the
intensity of the two successive fringes decreases until we reach the point b = p, where the
intensity is zero. Thus the third fringe corresponding to cos2z = ±3p falls at b = p or –p and
their product is zero. It means that the third fringe on either side of the central maximum has
zero intensity and its location at angle q simultaneously satisfies the conditions
b = ±p
and
z = ±3p
or
b sinq = ±l
and
d sinq = ±3l
Therefore, the third fringe will be missing. We will observe 4th and 5th fringes.
We can similarly argue that the 6th fringe for which
b = ±2p
and
z = ±6p
will have zero intensity and thus be missing.
We may therefore conclude that in double slit diffraction, diffraction term governs the
intensities of interference fringes, whereas the interference term controls their width.
We now advise you to answer a Practice Exercise.

Practice Exercise 6.6 A double diffraction slit, having slits of width 5 ´ 10–4 m and separated
by a distance d = 0.1 ´ 10–2 m, is illuminated by light of wavelength 632.8 nm. If a convex
lens of focal length 0.1 m is placed beyond the double slit, calculate the positions of the
maxima inside the first diffraction minimum.
[Ans. 3.16 ´ 10–5 m; 9.4 ´ 10–5 m]
Before discussing the diffraction pattern produced by N identical parallel slits, we discuss
an activity by which you can observe diffraction pattern produced by a double slit.
224 Wave Optics

Activity 6.2 Take a microscope slide and coat it with India ink or a colloidal suspension of
graphite in alcohol. Allow it to dry and then scratch a pair of slits on the slide with a razor
blade. Now keep it at about 3 m from a line source such as a bulb filament and hold the slits
close to your eyes, keeping these parallel to the filament. Interpose red or blue cellophane and
observe the change in the width of the fringes. Observe what happens when one of the slits
is covered. Next, move the slits horizontally, then hold them fixed and move eye horizontally.
How does the position of the centre of the pattern behave?

Now we know that interference of waves diffracted by individual slits determines the
intensity distribution in a double slit pattern. Let us now consider the diffraction pattern produced
by N identical parallel slits.

6.5 FRAUNHOFER DIFFRACTION FROM N IDENTICAL SLITS

Refer to Figure 6.16. It shows cross-sectional view of N identical (long, narrow and parallel)
slits, which are illuminated by a parallel beam of monochromatic light emanating from a slit
source, which is parallel to the length of diffracting slits. For simplicity, we assume that
1. the width of each slit b is much less than its length h (i.e. b = h).
2. all slits are parallel to each other.
3. the thickness of opaque space between any two successive slits is the same, equal to a,
say.

Figure 6.16 Fraunhofer diffraction of a plane wave incident normally on a multiple slit aperture.

As before, we assume that the distance between two consecutive slits is d = a + b. This is
also the distance between any two equivalent/corresponding points in any two consecutive slits.
In deriving an expression for the intensity distribution of diffraction pattern in this case, we will
use the same procedure as used for the double slit.
Fraunhofer Diffraction 225

6.5.1 Intensity Distribution in N-Slit Diffraction Pattern

Consider a point source of light, which emits monochromatic waves. As in case of the double
slit, a converging lens is used to focus this light on the source slit placed in its focal plane.
It means that a plane wave will be incident normally on the diffracting slits. The light emanating
from the N-slits after diffraction at each slit results in N diffracted beams, which are focussed
on the observation screen by another converging lens. Since these diffracted beams are coherent,
they interfere leading to formation of fringes. From the discussion of double slit diffraction, we
recall that diffraction, not interference, controls the intensity of the pattern in a given direction.
We can extend this argument to the case of N-slits as such.
As before, we assume that each slit consists of n equally spaced point sources with spacing
D. Then electric field at an arbitrary point Pq , where q is the angle between the diffracted rays
and the normal to the diffracting slit, will be a sum of N terms:
E = E1 + E2 + E3 + L + EN
= A
sin E
cos(Z t  E )  A
E
sin E
cos(Z t  E  G )  A
E
sin E
cos(Z t  E  2G ) 
E
"
sin E
A cos[Z t  E  ( N  1)G ] (6.28)
E
We rewrite it as

E A
sin E
E
{cos(Z t  E )  cos(Z t  E  G )  cos(Z t  E  2G )  "  cos[Zt  E  ( N  1)G ]}
To sum up this series, we express the terms within the curly brackets in complex notation and
obtain

E= A
sin E
E
Re[ei (Z t  E )  ei (Z t  E G )  ei (Zt  E  2G )  " e Z i{ t  E  ( N 1)G }
]

= A
sin E
E
Re[ei (Z t  E ) {1  e iG  e 2iG  " e i ( N 1)G
}]

The terms within the curly brackets form a finite geometric series with a = 1 and common ratio
e–id. Hence, it can be easily summed to obtain

sin E Ë È 1  e iNG Ø Û
E= A Re Ìei (Z t  E ) É iG Ù Ü
E ÌÍ Ê 1 e ÚÜÝ

sin E Ë e  iN G /2 È eiNG /2 e iN G / 2 Ø Û


= A Re Ì ei (Z t E ) É iG / 2 Ù Ü
E ÍÌ e iG / 2 Ê eiG /2 e ÚÝÜ

Ë È NG Ø Û
Ì sin É Ù
sin E Ê 2 ÚÜ
= A Re Ìei{Z t  E ( N 1)G /2} Ü
E Ì ÈG Ø
sin É Ù Ü
Ì
Í Ê 2Ú Ü Ý
226 Wave Optics

sin E sin N ]
= A cos[Z t  E  ( N  1)] ] (6.29)
E sin ]
G S d sin T
where ] .
2 O
The intensity of the N slit diffraction pattern is given by
È 2 sin 2 E Ø sin 2 N]
IT ÉA Ù (6.30)
Ê E2 Ú sin 2 ]
Recall that A2(sin2b/b 2) represents the intensity distribution produced by a single slit. Further, as
in the case of double slit diffraction, the intensity for N-slit diffraction pattern is a product of
diffraction term (sin2b/b 2)due to a single slit and the interference term (sin2Nz/sin2z ) due to N-slits.
For N = 1, Eq. (6.30) reduces to Eq. (6.8). For N = 2, we can write

È 2 sin 2 E Ø sin 2 2]
Iq = ÉA Ù
Ê E2 Ú sin 2 ]
sin 2 E (2 sin ] cos ] ) 2
= A2
E2 sin 2 ]
sin 2 E
= 4 A2 cos 2 ]
E2
This is identical to Eq. (6.21) for intensity of double slit diffraction pattern.
Let us now investigate the positions of maxima and minima. We first consider principal
maxima.

A. Positions of principal maxima


To determine the positions of maxima, we recall that in Eq. (6.30) the diffraction
term (sin2b/b 2) controls the intensities of interference fringes, whereas the interference term
(sin2Nz/sin2z) determines their width. The maximum value of interference term is N2. To
understand this, we note that the interference term becomes indeterminate for z = 0, p, 2p, ...,
np. From elementary calculus we recall that to maximize a function which becomes indeterminate
for certain values of the argument, we calculate the first order derivatives of the numerator as
well as the denominator separately and then substitute the value of the argument. (This is known
as L’Hospital’s rule.) Following this procedure, we obtain
sin N] N cos N]
lim lim “N
] nS sin ] ]  nS cos ]
Hence, for z = 0, p, 2p, ..., np, we get
2
È sin N] Ø
lim N2
]  nS É
Ê sin ] Ú
Ù

That is, the maximum value of interference term is N 2 and it occurs for z = 0, p, 2p, ..., np.
Fraunhofer Diffraction 227
On using this result in Eq. (6.30), the expression for intensity for z = 0, p, 2p, ..., np takes
the form
sin 2 E
IT N 2 A2 (6.31)
E2
S b sin T S b mOmbS
where E – .
O O d d
We may therefore conclude that for N identical, narrow but long slits, the positions of
maxima are obtained for
z = 0, p, 2p, ..., np
or
Nz = 0, Np, 2Np, ..., nNp (6.32)
Physically speaking, the fields produced at these maxima by each slit are in phase and the
resultant field is N-times the field due to any individual slit. As the number of slits increases,
the intensity of the maxima increases, provided the diffraction term is not very small. Such
maxima are referred to as principal maxima.
The condition for principal maxima can be rewritten as
d sin qmax = nl (6.33)
This result is identical to Eq. (6.24). It implies that
• The principal maxima in N-slit diffraction pattern corresponds in position to those of the
double slit. Moreover, these are more intense in the ratio of the square of the number of
slits.
• The relative intensities of different orders are modulated by the single slit diffraction
envelope.
d
• The order of the maximum cannot be greater than since sinq £ 1. This condition
O
implies that only a few principal maxima can exist. These are designated as the first,
second, third order of diffraction.
• There will be as many first order principal maxima as the number of wavelengths in the
incident wave.
• The relation between b and z obtained for the double slit in terms of slit width and slit
separation holds for N-slits as well.

B. Positions of minima and secondary maxima


While identifying the positions of principal maxima in the diffraction pattern,
we located the maxima of the interference term (sin2Nz/sin2z). In the same way, to locate the
positions of minima in the diffraction pattern, we have to locate the minima of this term.
In doing so, we note that the numerator of the interference term will vanish more often as
compared to the denominator. We can easily verify that the numerator of the interference term
vanishes for
Nz = 0, p, 2p, ..., pp
228 Wave Optics

i.e.
pS
] p = 0, 1, 2, 3, ... (6.34)
N
It means that the sine function in the denominator of the interference term will not become zero
for all integral values of p. In fact, sin z = sin pp/N will become zero for the special cases when
p = 0, N, 2N, ... so that the permitted values of z are 0, p, 2p, ... . Note that for these special
values of z, both sin Nz and sinz vanish and the interference term defines the positions of
principal maxima, as discussed in the previous sub-section. However, for all values of p other
than 0, N, 2N, ..., only the numerator vanishes; denominator is finite. That is, the intensity
vanishes when p, though an integer, is not an integral multiple of N. Hence, we can express the
condition of minimum as z = pp/N, except when p = nN; n being the order of maximum. These
values correspond to

Nz = [S , 2S , ..., ( N  1)S ], [( N  1)S ,( N  2)S , ..., (2 N  1)S ], [(2 N  1)S , ...]


or
Ë S 2S
z= Ì ,
( N  1)S Û Ë ( N  1)S ( N  2)S (2 N  1)S Û Ë (2 N  1)S Û
(6.35)
, ..., ÜÝ , ÌÍ N , , ..., ÜÝ , ÌÍ , ...Ü
ÍN N N N N N Ý
In terms of path difference, these values of z respectively correspond to

Ë O 2O ( N  1)O Û Ë ( N  1) O ( N  2)O (2 N  1)O Û Ë (2 N  1)O Û


d sin T min ÌÍ N , N ,..., ÜÝ , ÌÍ N , ,..., ÜÝ , ÌÍ ,...Ü (6.36)
N N N N Ý
We can rewrite the condition of minima in the general form as

AO
d sin T min (6.36a)
N

Note that l takes all integral values except 0, N, 2N, 3N, .... In other words, l = [1, 2, ...,
(N – 1)], [N + 1, N + 2, ..., (2N – 1)], [2N + 1, ...]. For l = 0, N, 2N, ..., nN, d sin q takes values
0, l, 2l, ..., nl, which represent principal maxima. These values are omitted from the values
in the minima.
From the above discussion, it is clear that between any two successive principal maxima,
there will be (N – 1) positions of minima. The first square bracket in Eq. (6.36) lists the
positions of (N – 1) minima between the central maximum and the first principal maximum.
Similarly, the second square bracket in Eq. (6.36) lists the positions of (N – 1) minima between
the first and the second principal maxima. In other words, the first minimum on either side of
the nth principal maximum defined by d sin q max = nl will be at

O (6.36b)
d sin T min nO “
N
Further, recall that between any two consecutive minima, there has to be a maximum.
Such maxima are said to be secondary maxima or subsidiary maxima. You may now ask:
Fraunhofer Diffraction 229
How many secondary maxima will there be between any two consecutive principal
maxima? The number of secondary maxima between any two consecutive principal maxima
will be (N – 2). As in single slit diffraction pattern, the secondary maxima are not symmetrical
and their intensity is too small to be important. Figure 6.17 shows the intensity pattern for

Figure 6.17 Typical intensity pattern for a six slit diffracting aperture with a = 4b.
230 Wave Optics

six slit aperture with a = 4b. The principal maxima corresponding to n = 0, 1, 2, 3 and
the corresponding secondary maxima between any two adjacent principal maxima are
easily visible.

C. Width of the principal maxima


From the above discussion, we now know that
• For the nth order principal maxima to occur in the diffraction pattern produced by N-slits,
the condition in terms of path difference and angle of diffraction is d sin qmax = nl;
n = 0, 1, 2, ... .
• On either side of the principal maxima, we have a minima and the condition in terms of
path difference and angle of diffraction is d sinqmin = nl ± l/N.
Since the principal maximum extends from minimum on one side to minimum on the other,
the separation between qmax and qmin defines the angular half-width of the principal maximum.
(You should not confuse it with half-width of a spectral line, which is defined as full-width at
half maximum.) Let us denote it by dq, i.e. dq = |qmax – qmin|. To calculate the value of dq, we
choose qmin – qmax and insert the value qmin = qmax + dq in the condition for minima to obtain
O
d sin(T max  GT ) nO 
N
On using the relation sin(a + b) = sina cosb + cosa sin b, we get
O
d (sin T max cos GT  cos T max sin GT ) nO 
N
In the limit dq ® 0, cosdq ® 1 and sindq ® dq. Hence, the above relation simplifies to
O
d sin T max  d cos T maxGT nO 
N
Using the condition for occurrence of principal maximum (d sinqmax = nl), this equation
simplifies to
O
d cos T max GT
N
Hence, the angular half-width of the principal maximum is given by
O
GT (6.37)
Nd cos T max
From Eq. (6.37), we note that angular half-width of the principal maximum is directly proportional
to the wavelength of light and inversely proportional to the number of slits in the aperture and
the grating element. It means that as the number of slits increases, interference maxima
get sharper as well as more intense and in the limit N ® ¥ these maxima become narrow lines.
This led to the realisation of a diffraction grating which was subsequently used for precise
spectral analysis. We now discuss it in some detail.
Fraunhofer Diffraction 231

6.5.2 Plane Diffraction Grating

In the preceding section, we discussed the diffraction pattern produced by N parallel, equidistant
vertical slits. A repetitive array of a large number of equidistant, narrow, vertical parallel diffracting
elements (slits/apertures/obstacles) which can induce periodic alterations in the phase, amplitude
or both of an emergent wave is said to be a diffraction grating. You might have seen or even
worked with a plane transmission diffraction grating. If you have not worked with a plane
transmission grating so far, you can easily create one on your own. Do you know who made
the first grating and how? The first gratings were made by Fraunhofer, who used fine silver
wires and stretched these on a frame. His grating had about 200 wires per centimetre.
Subsequently, a diamond pen was used for ruling grooves on an optically transparent sheet. (The
transparent part between the grooves behaved as a slit and the grooves as the opaque part.) An
important requirement for a good quality grating is that the lines should be equally spaced, the
pitch of the screw used to control movement of the sheet must be constant. Rowland achieved
a nearly perfect screw in 1882 and ruled gratings on a metallic surface. He produced plane as
well as concave gratings with 14438 lines to an inch. These gratings produced very sharp
spectral lines but were difficult to produce and were expensive as well.
Commercial gratings are produced by taking the cast of an actual grating on a transparent
film such as that of cellulose acetate. Then a solution of cellulose acetate of a particular
concentration is poured on the ruled surface and made to dry to form a strong thin film.
The film carries the impression of the grating. It is then detached from the parent grating and
preserved by mounting between two glass sheets. In physics laboratory, we get celluloid replicas,
which are not so expensive and have 15,000 lines per inch. (We can make a simple coarse
grating by drawing equidistant and parallel scratches on a photographic emulsion. Look through
it at a distant electric lamp and record what you observe. You may verify your conclusions while
working in physics laboratory.) Nowadays, holographic gratings, which record the interference
pattern between two plane/spherical waves, are produced. Such gratings have a large number
of lines per inch and are definitely superior to ruled gratings.

(a) Grating spectrum


From Section 6.4, we recall that the exact positions of the principal maxima in the diffraction
pattern depend on the wavelength of light used and the angle of diffraction. It means that if we
use a monochromatic light source, which emits wavelength l, the principal maxima are specified
by the so-called grating equation:
d sinq n = nl n = 0, 1, 2, 3, ... (6.38)
The values of n specify the order of various principal maxima. The grating equation depends
on l. So if a source of light emits two (as in sodium lamp) or more spectral lines (as in mercury
lamp) corresponding to wavelengths l1, l2, ..., li, we observe as many lines in each order.
This is because, except the central maximum which corresponds to n = 0, the angles of diffraction
for these colours will be different in different orders. (In the central maximum, q0 = 0 for all
wavelengths and different colours are not separated from each other. For a white light source,
the central maximum will be white.) Such a diffraction pattern produced by a diffraction grating
is known as grating spectrum.
232 Wave Optics

From Eq. (6.38), we note that if we know d, qn and n, we can determine the wavelength
of light. Since d is known for a grating and qn can be measured (in physics laboratory) using
a spectrometer, a grating provides an excellent tool to measure l for any particular colour.
If we differentiate Eq. (6.38) and rearrange terms, we can write
'T n n
(6.39)
'O d cos T n
From this result, we can conclude that
• For small values of qn , i.e. when cosqn » 1, Dqn /Dl is constant for a given order of
spectrum. Such a spectrum is said to be normal. However, for large values of qn , angular
dispersion will be greater at the red-end of the spectrum.
• Since Dqn is inversely proportional to the grating element, angular dispersion will be
large, if grating element is small.
To appreciate these aspects, you must observe grating spectrum. In physics laboratory,
you will use a simple spectrometer. Let us now learn about it.

(b) Observing grating spectrum


Refer to Figure 6.18, which depicts the use of a simple spectrometer to observe grating spectrum.
Light from a (monochromatic or polychromatic) slit source is allowed to fall on the collimator
slit, which has been focused at infinity for parallel beam of light by Schuster’s method.
We know that for observing Fraunhofer diffraction, the diffracting aperture/obstacle should be
at an infinite distance from the source as well as the observation screen. (In this case, the cross-
wires of the telescrope act as observation screen.) For this reason, the collimator is adjusted
so that parallel rays are incident on the diffracting aperture. Similarly, the telescope is adjusted
for parallel rays. A pre-requisite for Schuster’s method is optical levelling of the spectrometer.
It essentially means that the collimator axis and the telescope axis lie on the same plane,
which, in turn, is normal to the axis of the turn table. By so doing, we make sure that the parallel
beam of light reaching the objective of the telescope is always focused at the cross-wires,

Figure 6.18 Schematics of a typical laboratory arrangement to observe diffraction of a plane wave
incident normally on a grating.
Fraunhofer Diffraction 233
which are in the focal plane of the eye-piece. Next, we mount the grating on the turntable such
that the light is incident on it normally and the diffracting slits are parallel to the collimator slit.
(The telescope carries a circular scale on which a vernier is made to slide. This arrangement is
used to note down the angular dispersion of the spectral lines.) The diffracted light from the
grating enters the telescopic arm and you can observe sharp spectral lines.
To mount the grating for normal incidence, we record the position of the turntable when the
telescopic arm is in line with the collimator. Let us denote it by q. Next we rotate the telescope
through an angle of 90º. Then we mount the grating on the turntable and rotate it till we see the
image of the collimator slit through the grating. We can say that in this position, the surface of
the grating is inclined at 45º to the parallel beam of light emerging from the collimator. If we
now turn the grating further through 45º, the light will be incident normally on the grating surface.
Once grating has been mounted for normal incidence, the telescope is rotated to the left or
the right. We expect you to observe the first order spectrum in the field of view of the telescope.
For sodium, mercury or argon source, the spectrum will consist of a series of sharp spectral
lines. (If the spectral lines are not sharp, rotate the grating in its own plane.) Note that each
spectral line is diffracted image of the slit corresponding to different wavelengths emitted by the
source. It may be mentioned here that spectrum of an atom consists of sharp lines, whereas the
spectrum of a molecule is band-like. An incandescent lamp will give out a continuous spectrum,
where various colours merge into one another.
To measure the wavelength of a spectral line, the vertical cross-wires are set at its centre
and we note the corresponding position of the telescope. This procedure is repeated for each
spectral line. The difference between this position of the telescope and that corresponding to the
direct position gives the angle of diffraction for the spectral line under consideration. To minimize
experimental error, it is desirable to record the positions of the telescope on both sides of the
direct image and half of this angle gives the angle of diffraction. This procedure is repeated for
each spectral line.
Depending on the grating element, you may observe the first, the second and possibly the
third order spectrum. However, the intensity and sharpness of different spectral lines will be
seen to diffuse/fade gradually.
For numerical appreciation of the above results, we now give a few solved examples.
You should go through these carefully.

EXAMPLE 6.5 Consider a diffraction grating having 15,000 lines to an inch. If we use a
white light source, calculate the dispersion suffered due to diffraction by red (lR = 700 nm) and
violet (lV = 400 nm) colours in the first order spectrum. How this spread is influenced in the
second and third order spectra?
Solution The grating element
2.54 – 10 2 m
d 1.69 – 106 m
15000
Let us denote the angle of diffraction for red colour by qR. Then using grating equation for a
maximum, we can write
d sinqR = nl
234 Wave Optics

For the first order, we get

O 700 – 10 9 m
TR sin 1 sin 1 sin 1 (0.414)
d 1.69 – 10 6 m
Hence
qR = 23.71º
Similarly, for the violet colour (lV = 400 nm), we get

400 – 109 m
TV sin 1 sin 1 (0.237)
1.69 – 106 m
Hence
qV = 13.55º
It means that the visible spectrum in the first order extends from q = 13.55º to q = 23.71º, i.e.
the spread is about 10º.
In the second and third order spectra, we can easily verify that

1 2 – 400 – 109 m
(qV )2 = sin  sin 1 (0.473) 28.2º
1.69 – 106 m
1 2 – 700 – 10 9 m
(qR)2 = sin  sin 1 (0.828) 55.90º
1.69 – 106 m
1 3 – 400 – 109 m
(qV )3 = sin 6
 sin 1 (0.710) 45.23º
1.69 – 10 m
Note that the first and the second order spectra will not overlap when a grating is used to
analyse light in the wavelength region 400 nm and 700 nm. However, since (qR)2 > (qV)3, the
second and third order spectra will overlap. In actual experimental situations, such a situation
is avoided by using suitable colour filters.
If we calculate the value of sinqR for the third order, it will come out to be greater than one.
This obviously is trivial as the maximum possible value of sinq is unity. Physically, this
result implies that in the third order spectrum, red colour will not be observed. We can even
say that in the third order, only a part of the visible spectra will be exhibited by the grating in the
instant case.
You can easily note that the angular spread increases with the order of the spectra.

In your school science, you have learnt about formation of spectrum by a prism. You may
recall that prism spectrum is of a single order and the entire energy (intensity) belongs to it.
However, in the case of grating, we observe spectra up to a few orders. Thus, the prism
spectrum has superior intensity, whereas the grating spectrum is superior in respect of spectral
purity.
Fraunhofer Diffraction 235

EXAMPLE 6.6 Consider a diffraction grating having 10,000 lines per cm. If we use a lens
of focal length 1.0 m, calculate the linear separation of wavelengths 500 nm and 520 nm in the
first order spectrum.
Solution Here grating element a + b = 10–4 cm, l1 = 5 ´ 10–5 cm, l2 = 5.2 ´ 10–5 cm and
n = 1. Using the grating equation, we can write
(a + b)sinq1 = nl 1
so that the angle of diffraction corresponding to wavelength l1 = 5 ´ 10–5 cm is given by

nO1 1 – 5.0 – 10 5 cm
sin T1 0.5 À T1 30º
( a  b) 10 4 cm
Similarly, we can calculate the angle of diffraction corresponding to wavelength l2 =
5.2 ´ 10–5 cm as
nO 2 1 – 5.2 – 105 cm
sin T 2 0.52 À T 2 31.3º
(a  b) 104 cm
Hence the linear separation between these colours is given by
Dx = x2 – x1 = f (tanq 2 – tanq1) = 100(0.6087 – 0.5774) = 3.13 cm

We would now like you to answer a Practice Exercise.

Practice Exercise 6.7 Monochromatic light of wavelength 500 nm is diffracted by a plane


transmission grating having 6000 lines per cm. Determine the highest order spectrum that can
be seen. [Ans. n = 3]

6.5.3 Concave Reflection Grating

One of the major concerns, while working with a plane transmission grating to determine the
wavelength of light, arises due to chromatic aberration present in collimating and objective
lenses used to focus light. To overcome this problem, Rowland, who dispensed with both the
lenses, used a concave reflection grating. It performed two-fold action—reflection and
diffraction—simultaneously. We now discuss it in some detail.
In a concave reflection grating, rulings are made on a concave spherical highly polished
metallic mirror instead of a plane surface. The concave grating completely eliminates the effect
of chromatic aberration. Moreover, a concave grating can be used to study diffraction of ultraviolet
light, which is not transmitted by glass lenses.
Refer to Figure 6.19(a). It shows the horizontal section SOS ¢ of a concave grating whose
rulings are kept vertical. The dotted circle represents the Rowland circle of diameter OC = R.
We choose the coordinate system with origin at the vertex of the concave surface, x-axis along
the horizontal through the centre of curvature C of the grating, z-axis vertical, parallel to the
rulings of the grating, while y-axis lies in the horizontal plane and is perpendicular to the rulings
236 Wave Optics

of the grating. A1(a1, b1, 0) is a narrow slit perpendicular to the plane of the paper and
illuminated by light. We assume that it acts as a point source of monochromatic light. Suppose
that the image is formed in the same plane at A2(a2, b2, 0) by reflection from an element of
grating at S1(x, y, z). By referring to Figure 6.19(a), we note that the equation of the concave
surface of the grating is given by
(S1C)2 = R2 = (S1L)2 + (LC)2 = (S1L)2 + (LM)2 + (MC)2
= z2 + y2 + (R – x)2
We can simplify this equation as
x2  y 2  z 2
x (6.40)
2R
From Figure 6.19(a), we can also write
(A1S1)2 = (x – a1)2 + (y – b1)2 + z2
= (x2 + y2 + z2) – 2a1x – 2b1y + a12  b12
È a Ø
= r12  ( x 2  y 2  z 2 ) É1  1 Ù  2b1 y (6.41a)
Ê RÚ
where we have put r12 2 2
a1  b1 .

Figure 6.19 (a) Focussing action of a concave grating, and (b) depiction of angles of incidence
and reflection.

For large R, x will be small and we can ignore the term containing its second power.
Similarly, if the vertical aperture of the concave grating is not large, the z2 term can also be
ignored. Then Eq. (6.41a) takes a very simple form:

2 È a1 Ø
(A1S1)2 = r1  y 2 É1  Ù  2b1 y
Ê RÚ
2
Ë y 2 È a12  b12 a1 2b1 Ø Û
= r1 Ì1  É   Ü (6.41b)
ÌÍ r12 Ê r12 R y ÙÚ ÜÝ
Fraunhofer Diffraction 237

a12 + b12
where we have used the relation 1. Hence, the expression for A1S1 takes the form
r12
1/2
Ë y2 È a2  b2 a 2b1 Ø Û
A1S1 r1 Ì1  2 É 1 2 1  1  Ü
ÍÌ r1 Ê r1 R y ÙÚ ÝÜ

To simplify this expression, we expand the term within the square brackets using binomial
expansion and retain terms in the first power of y 2 /r12 . This gives

A1S1
Ë
r1 Ì1 
y2
È a12  b12
2 É
2r1 Ê r1 2
a
 1 
R
2b1 Ø
y ÙÚ
 " Û
Ü (6.41c)
Ì
Í Ü
Ý

For normal or nearly normal incidence, b12 can be ignored in comparison to a12 so that the
expression for A1S1 reduces to
y 2 a1 È a1 1Ø b1 y
A1S1 r1    (6.42)
2r1 ÉÊ r12 R ÙÚ r1
Similarly, we can obtain the expression for A2S1 as

y 2 a2 È a2 1Ø b2 y
A2 S1 r2    (6.43)
2r2 ÉÊ r22 R ÙÚ r2
where we have put r22 a22  b22 .
Hence the total optical path travelled by light generated by the source at A1 to travel to the
image position at A2 via S1 is obtained by combining Eqs. (6.42) and (6.43):

y 2 Ë a1 È a1 1Ø a2 È a2 1 ØÛ È b1 b Ø
A1S1  A2 S1 p1 r1  r2     Ü  yÉ  2 (6.44)
2 ÌÍ r1 ÉÊ r12 R ÙÚ r2 ÉÊ r22 Ù
RÚÝ Ê r1 r2 ÙÚ

If the source and the image are located on a circle of diameter R whose centre lies on x-axis
at a distance R/2 from the origin O, we will have r12 a1R and r22 a2 R . Then the term
containing second power of y will disappear. Such a circle is called Rowland circle
[Figure 6.19(b)]. Physically, it means that the source of light and its spectrum (image) formed
by a concave grating lie on the Rowland circle. Under these conditions, the third term on the
right hand side of Eq. (6.44) drops out and we obtain a compact expression for total optical path.

È b1 b2 Ø
p1 r1  r2  yÉ  (6.45)
Ê r1 r2 ÙÚ
Let us choose another point source at S2 on the grating surface separated from S1 through a
distance equal to grating element d(width of the ruling plus width of the polished strip) measured
along the y-axis. Following the procedure followed in arriving at Eq. (6.45), we can readily
238 Wave Optics

derive the expression for total optical path travelled by light generated by the source at A1 to
travel to the image position at A2 via S2 as
È b1 b2 Ø
p2 r1  r2  ( y  d) É  (6.46)
Ê r1 r2 ÙÚ
From Eqs. (6.45) and (6.46), we note that the consecutive values of the y-coordinates of the
lines are differing by grating constant, d. It implies that ruling in a concave grating should be
equi-spaced along the chord rather than the arc of the surface.
By combining Eqs. (6.45) and (6.46), we can calculate the path difference between waves
proceeding from the source at A1 to the image position at A2 after reflections from S1 to S2:
È b1 b2 Ø
p p1  p2 dÉ  (6.47)
Ê r1 r2 ÙÚ
For maxima, p should be an even multiple of half-wavelength:
È b1 b2 Ø
dÉ  mO (6.48a)
Ê r1 r2 ÙÚ
To express the condition of maxima in terms of the angles of incidence and diffraction, refer
to Figure 6.19(b). It shows the angle of incidence as i and angle of diffraction as q. From a basic
knowledge of trigonometry, we can write sin i = b1/r1 and sinq = b2/r2 so that the condition for
maxima can be expressed as
d(sini + sinq) = ml (6.48b)
Let us pause for a while and ponder as to what we have achieved so far. In arriving at
Eq. [6.48(b)], we have derived the condition for reinforcement of diffracted beams emitted by
a monochromatic source in terms of the angles of incidence and diffraction. You may now ask:
How will this condition modify if source at A1 emits light of different wavelengths? Though
angle of incidence will be the same for all colours, we expect different values of angle of
diffraction. Therefore, from Eq. (6.48b) we can write
d cosq dq = mdl (6.49)
Note that m signifies the order of spectrum. If the wavelength range dl covers an arc ds of
Rowland circle, these will be related through the relation dq = ds/R so that Eq. (6.49) can be
rewritten as
ds
d cos T md O
R
so that
ds mR
(6.50)
d O d cos T
If the diffracted wave is close to the normal (OC), we can take cosq » 1 with fair degree of
confidence. Then Eq. (6.50) simplifies to
ds mR
(6.51)
dO d
Fraunhofer Diffraction 239
For a given order of spectrum, m is constant and the change in length of the arc of Rowland
circle will be proportional to the wavelength of light incident on the grating. It means that a
concave grating produces normal spectrum.
Before proceeding further, you should answer a Practice Exercise

Practice Exercise 6.8 Suppose that the source and the image are on the opposite sides of
the centre of curvature C. How will Eq. (6.48b) modify? [Ans. d(sini – sinq) = ml]

6.5.4 Echelon Gratings

We now know that with a plane transmission grating, we can observe only a few orders of
spectrum and the intensity diffuses beyond three or possibly four orders, though the total
number of rulings in the grating is very large. But due to mechanical difficulty of the ruling
machine, N cannot be increased beyond a certain value. In 1898, Michelson conceived the idea
of making N small, say between 20 and 40, but n very high, say 20,000. By so doing, he could
increase the resolving power of the grating. To increase n, the path difference between the
diffracted waves from two corresponding adjacent points is made very large. Like plane gratings,
echelon gratings are also of transmission type and reflection type. But the latter is more useful.
Refer to Figure 6.20(a). It shows an echelon grating, which consists of a number of plane-
parallel plates of glass or quartz of thickness, t. In a transmission grating, the plates are
homogeneous and no air film is left between any pair of plates. Each plate is of the same
breadth, b, but their lengths decrease by a constant factor, s, so that the width of the step is s.
When the plates are wrung together, these take the form of flight of steps, each of width s,
which is of the order of 1 mm. This gives it the nomenclature echelon. Each step acts as an
aperture of width s. In reflection echelon grating, the front steps are aluminized to achieve good
reflection. The experimental arrangement used to observe diffraction pattern formed by an
echelon grating is shown in Figure 6.20(b). L1 and L2 are two achromatic doublets, which help
in minimising chromatic abberation.

Figure 6.20 (a) Schematic representation of action of an echelon grating; (b) experimental arrangement
to observe diffraction pattern formed by an echelon grating.
240 Wave Optics

When the breadth b is vertical, thickness t crossed and width s is horizontal, the aperture-
steps of the grating would be vertical and the grating dispersion will be horizontal. But if the
breadth b is crossed, thickness t is horizontal while the step width s is vertical, aperture-steps
would be horizontal and the grating dispersion will be vertical.
Suppose that a parallel beam of monochromatic light of wavelength l is incident normally
on the largest plate of the echelon. Let m be the refractive index of the material of the plate.
The light will be diffracted in various directions by each step, which acts as an aperture of
width s. Due to large width of the step, the diffracted light is confined within a small angle. The
diffracted light from all the steps is collected in the focal plane of the objective of the camera,
where we obtain diffraction pattern due to a single aperture of width s. This diffraction pattern
produces an envelop below which is formed a pattern due to the interference of secondary
waves originating from different steps.
The formation of maxima and minima is determined by the path difference between secondary
waves diffracted at an angle q, say from the corresponding points such as P and R of any two
adjacent steps. From Figure 6.20(a), we note that light waves cover equal distance beyond R and
L. Thus the path difference between diffracted waves is given by
p = S1R – PL = mt – PL (6.52)
Note that PL = PM cosq = (PN – MN)cosq = (t – s tanq) cosq = t cosq – s sinq. For small
q, we can use the approximation sin q » q fairly confidently. Then the value of PL can be
expressed in terms of the characteristics of the echelon:
PL = t – sq
Using this result in Eq. (6.52), we note that the path difference between waves diffracted by the
points such as P and R of any two adjacent steps is given by
p = mt – (t – sq) = (m – 1)t + sq (6.53)
If this path difference is an integral multiple of l, all secondary waves travelling in this direction
reinforce each other. Therefore, for mth order maximum, we can write
(m – 1)t + sq = ml (6.54)
For small values of q, i.e. if observations are made close to or along the normal, Eq. (6.54)
reduces to
m = (m – 1)t/l (6.55)
For glass, m = 1.5. If t = 10–2 m and light of wavelength 500 nm is incident on echelon grating,
we get
0.5 – 102 m
m 104
500 – 109 m
In the next section, we will discover that resolving power of a grating is equal to the product
of the order of spectra that can be observed and the number of rulings per cm. It means that
for an echelon grating, resolving power will be of the order of 105, if the number of plates used
is 10 or more. Obviously, this is a very large value suggesting that the resolving power of an
Fraunhofer Diffraction 241
echelon grating is very high. It means that if the incident light is not truly monochromatic, any
two nearby spectral lines will appear well separated. For this reason, an echelon grating is used
in the study of hyperfine structure as in the splitting of spectral lines in Zeeman effect.
Now you may like to know the angular separation between two successive orders of
spectra. To do so, we differentiate Eq. (6.54) by considering m and q as variables, taking m, l
and t as constant. This gives
dT
s O (6.56a)
dm
We rewrite it as
O
dT dm (6.56b)
s
For dm = 1, it simplifies to
O
dq = (6.56c)
s
This result shows that the linear separation between two successive orders is given by Dx = fl/s,
where f is the focal length of the objective lens. If the width of the step is large in comparison
to the wavelength of light, the angular or linear separation between two successive orders will
be small.
Before proceeding further, we advise you to answer a Practice Exercise.

Practice Exercise 6.9 Starting from Eq. (6.54), show that the dispersion in an echelon
grating is proportional to the thickness of the plates and is inversely proportional to the width
of the steps.
Ë dT bt dP Û
ÌÍ Ans. d O ; b ( P  1)  O
Os d O ÜÝ

Another parameter of interest is spectral range, which specifies the wavelength difference
between two spectral lines whose angular separation is equal to the angular separation of two
consecutive orders. We therefore equate the values of dq given by Eq. (6.56b) and that obtained
in Practice Exercise 6.9. This gives
bt O
'O
Os s
Hence, spectral range is given by
O2
SR 'O (6.57)
bt
We know that due to diffraction, image of a point object is fringed and spreads over a small
area on the observation screen. For instance, if we consider two stars, which can be regarded
as point sources, the light reaching the objective of a telescope results in the formation of a
bright Airy disc surrounded by a number of alternate dark and bright rings in the back focal
plane of the telescope objective, as shown in Figure 6.21.
242 Wave Optics

Figure 6.21 Formation of Airy patterns of two distant stars in


the focal plane of the objective of a telescope.

We know that the diameters of Airy rings are determined by the diameter of the objective,
its focal length and the wavelength of light. As long as the Airy patterns are far apart from each
other, the two objects will be seen as distinct and are said to be resolved. In case of a telescope
of smaller diameter, it is quite possible that these patterns may overlap. Then the objects cannot
be seen as distinct, i.e. the objects appear as one. You may now like to know: How close can
the two point objects be to be just resolved? We will seek answers to this and such other related
questions in the following section. However, such diffraction-induced constraints in an optical
device arise due to finite sizes of its components.
To appreciate this, you can do a simple activity.

Activity 6.3 Take a metre scale and concentrate on the millimetre lines. Stand at a distance
of about 2.5 m and look at these markings. Do you see them as distinct? Move back and forth
till you see them blurring and merge into one another. Measure the distance between your eye
and the metre scale markings. We expect that you will be at a distance of about 3 m. So we
can say that 1 mm is barely resolved from about 3 m. The angle subtended at the eye by the
markings 1 mm apart and kept 3 m away is (1/3000) rad or about one minute of arc. So we
can say that unaided human eye can just resolve two bright points with an angular separation
of about a minute of arc. That is, the resolution limit of human eye is about 1 minute of arc.
Sometimes this is also referred to as the resolving power of eye. You can indeed verify this
result by considering the diameter of the pupil to be 2 ´ 10–3 m.

6.6 RESOLVING POWER OF OPTICAL INSTRUMENTS

We now know that image of a point object formed by an optical device is fringed and spreads
over a small area on the observation screen. Refer to Figure 6.22, where broken curves represent
the intensity distributions of two distant point sources for three angular separations (6l/pD,
2l/pD, 1.22l/D) between the central maximum and the first minimum. Note that both sources
have been assumed to independently give rise to the same intensity at their respective central
spots and the intensities have been plotted on the line joining the centres of the Airy patterns
Fraunhofer Diffraction 243
(curves 1 and 2). Since the sources are independent, the resultant intensity will be obtained by
adding individual intensities (curve 3). Note that when angular separation Dq = 6l/pD, the two
intensity curves are well separated and the objects are said to be well resolved. On the other
hand, for Dq = 2l/pD, the resultant intensity distribution has only one peak and we say that the
1.22 O
objects are unresolved. However, for 'T , the maximum of the intensity curve of first
D
pattern coincides with the first minimum of the intensity curve of second pattern and the objects
are said to be just resolved. This criterion for the limit of resolution is called the Rayleigh
criterion of resolution. Though it may appear somewhat arbitrary, yet it has the virtue of
simplicity. In Rayleigh’s words, ‘The rule is convenient on account of its simplicity and is
sufficiently accurate in view of the uncertainty as to what exactly is meant by resolution.’

Figure 6.22 Intensity distribution produced by two independent point sources when angular separation
between them is (a) 6l/pD, (b) 2l/pD, and (c) 1.22l/D. The broken curves (1 and 2) correspond
to intensities of individual sources and the solid curves represent the resultant intensity.

For human eye, if we take the diameter of pupil to be 2 ´ 10–3 m, the angular resolution
for green colour (l = 600 nm) is

1.22 – 600 – 10 9 m
'T 3
3.66 – 10 4 rad
2 – 10 m
244 Wave Optics

Hence, at a distance of 3 m, the eye should be able to resolve two points separated by
3.66 ´ 10–4 ´ (3 m) = 1.1 ´ 10–3 m. We had indeed arrived at this result in Activity 6.3.
An astronaut is orbiting at a height of 450 km. She claims that she could see individual
houses of her city as she passed above them. To know whether or not this claim is justified,
we note that the lateral width for resolution rDq = (450 ´ 103 m) ´ (3.66 ´ 10–4) = 164.7 m.
Since it is much greater than the width of individual houses, we have to believe the astronaut.
Before applying Rayleigh criterion to particular optical instruments, it may be pointed out
here that when the diffraction patterns of two equally intense stars are just resolved, i.e. when
the maximum of one intensity curve coincides with the
first minimum of the second intensity curve, the two
curves cross each other at 0.4 of the maximum intensity.
So, at this point, the resultant intensity will be equal to
0.81. That is, we should expect a dip of about 20% in
the resultant intensity curve. Such a dip can be easily
detected and the resolving powers of different instruments
can be fixed.
For two equally luminous stars (sources), we can
easily verify this number if we calculate the intensities
of (sinb/b)2 curves at the point of intersection. Since
these curves are symmetrical, have finite lateral width Figure 6.23 Resolution of two central
and same intensity, Rayleigh criterion suggests that these maxima according to
will cross at b = p/2, as shown in Figure 6.23. Then, we Rayleigh criterion.
can write
2
È SØ
sin
É 2Ù 4
I É S Ù 0.4053
É Ù
S2
Ê 2 Ú

At b = p/2, the resultant intensity will be obtained by adding the individual contributions and
therefore equals 0.8106. It means that the resultant intensity will show a dip of about 20%.
However, if the intensities of central maxima were not equal, the dip will have a different value.
In fact, in such a situation, we may need some other criterion so that the two maxima stand out.
We would now like you to answer a Practice Exercise.

Practice Exercise 6.10 Estimate the dip in the resultant intensity of two (sinNz/sinz)2
curves on the basis of Rayleigh criterion. [Ans. 20%]

6.6.1 Resolving Power of an Astronomical Telescope


Refer to Figure 6.21 again. A telescope points towards two close luminous stars, which subtend
an angle f on the objective. We assume that these stars are equally bright. Since the stars are
effectively at an infinite distance from us, the plane waves from these stars reach the objective
of the telescope and result in the formation of Airy diffraction patterns in the back focal plane
of the objective. These are observed through the eyepiece. For these stars to be just resolved,
Fraunhofer Diffraction 245
the angle f must be equal to the angular half-width of the Airy disc. A telescope can resolve
two distant stars if the angular separation is given by
1.22O
T min (6.58)
D
From this equation we can say that, if f > qmin, the stars will be resolved and vice versa.
If the focal length of the objective is f, the centre-to-centre linear separation between two
just resolved stars is given by
1.22O
s f T min f (6.59)
D
The reciprocal of resolving limit defines resolving power. It is generally abbreviated as R.P. From
Eq. (6.59), we note that the ability of a telescope to resolve two objects is inversely proportional
to the size of the aperture and directly proportional to the wavelength emitted by the source. It
means that for a fixed l, the resolving power of a telescope can be improved by increasing the
diameter of the objective; the larger the aperture of the objective of telescope, the higher will be
its resolving power. Moreover, the larger aperture allows a telescope to collect more light so that
it can be used to ‘see’ deeper in space. In fact, large telescopes have been constructed by the US
and other advanced countries to observe remote and faint celestial bodies clearly.
It is important to realize that when observations are taken on astronomical bodies using
terrestrial telescopes, light from a distant star first reaches the top of the atmosphere as a plane
wavefront and then travels through the atmosphere to reach the objective of ground based
telescope. The atmosphere may distort the plane wavefront, affecting the quality of image.
This can occur due to temperature fluctuations, which, in turn, change the atmospheric refractive
index. To minimize such ‘atmospheric turbulence’, terrestrial telescopes are mounted on high
mountain peaks situated in a flat area.
Another important aspect that we need to take note of is the fact that a large part of
electromagnetic spectrum is absorbed or scattered by earth’s atmosphere and is lost as far as a
ground-based telescope is concerned. Therefore, in order to utilize the entire information coming
from a star, it would be desirable to put up a telescope above earth’s atmosphere. You may now
ask the question: How can we do so? One alternative is to mount the telescope in a satellite.
In fact, the Hubble Space Telescope, launched in 1990, has been used with considerable success.
Since it is above earth’s atmosphere, it receives complete incoming radiation, which can be
beamed to the earth station and is free from atmospheric distortions in the image.
As a student in kindergarten, while gazing the night sky you must have uttered—
Twinkle, twinkle little star, How I wonder what you are? But as a student of physics, while
gazing the day sky, you may now like to rephrase—Twinkle, twinkle little star, How I wonder
where you are? Alternatively, you may ask: Why do stars fade away and become invisible as
the sun rises? Going a step further, we can raise the question: What measures will help researchers
to perform astronomical experiments in the day time? (You should not be scared by the literal
meaning of the Hindi Phrase: Din mein taare dikhna.) If you ponder for a while and think
logically, you should hit the point: As aperture of the telescope is increased, more light is
concentrated in the diffraction pattern. So, we can achieve a stage when image of the star will
become brighter than the background and be visible. (This essentially becomes possible because
246 Wave Optics

the intensity of a star is proportional to the fourth power while the background sky light
increases as the area of the aperture.) It means that astronomical observations are possible
during the daytime, if researchers use a large objective telescope.
To illustrate the concepts developed in this section and help develop appreciation for numerical
values, we now give some solved examples.

EXAMPLE 6.7 An astrophysics research laboratory has option to mount a telescope with an
objective of aperture either 1 m or 2 m. Both objectives, however, have the same focal length.
Which telescope will you recommend and why?
Solution As such, you should be able to outright state your preference for the larger aperture
telescope. But to support your choice with sound argument, you should calculate the resolving
powers and brightness of the images formed.
For a telescope, the minimum angle of resolution is given by Eq. (6.58). So for the first
telescope, qmin = 1.22l/D = 1.22l. Similarly, for the second telescope, the minimum angle of
resolution will be half of that for the first telescope. It means that resolving power of second
telescope will be twice the resolving power of the first telescope.
To compare the performances of these telescopes, we should also compare the brightness
of the diffraction patterns formed.
Recall that radius of the Airy disc is given by Eq. (6.59). So for the first telescope,
r1 = fqmin = 1.22lf, where l and f are expressed in metre and the area of the Airy disc

A1 S r12 S (1.22 f O )2
S
The area of the telescope objective which collects light falling on it is pr2 = m 2 . As seen
4
before, this light is essentially contained in the central maximum. If we assume that light is
distributed uniformly over the disc, its brightness, i.e. light collected by the objective per unit
area of the disc
S
I1 µ š S (1.22 f O ) 2
4
1
= m4
4 f (1.22) 2 O 2
2

For the second telescope, area of the Airy disc


2
1.22O Ø È
A2 S ÉÊ f
Ù
2 Ú
It means that the area of Airy disc of the first telescope is four times more than the area of the
second telescope.
The brightness of Airy disc of the second telescope
2
È 1.22O Ø 4
I2 S š S ÉÊ f Ù = m4
2 Ú f O (1.22) 2
2 2

= 16I1
Fraunhofer Diffraction 247
It means that the second (larger) telescope will concentrate more light over a smaller area and
lead to formation of 16 times brighter image. That is, doubling the aperture increases intensity
24 times. So we may conclude that
• the ability of a telescope to resolve two nearby objects depends on the diameter of its
objective.
• as size of aperture is increased, it collects more light and concentrates it over an area
which is much less. So the brightness of the pattern increases. In the instant case, it
increases by a factor of 16. It means that a distant star, which is too faint to be observed
by a smaller objective telescope, becomes visible by a larger telescope. That is why
larger aperture telescopes are being used to penetrate deeper into space.
So you should recommend the bigger telescope.

Another important question relating to the quality of telescope is the magnifying power of
its eyepiece. The magnification of the telescope should not be greater than the ratio of the
minimum angle of resolution of unaided eye to the minimum angle of resolution of the telescope.
For a telescope of aperture 2 m,
1.22O 1.22 – 600 – 109 m
(T min ) telescope 3.66 – 107 rad
D 2m

Hence, magnification of the telescope


3.66 – 10 4
m 1,000
3.66 – 107
Any further magnification will increase size of the primary image.
An astronomical telescope is used to view very far off objects and their exact distances
are not known. It is for this reason that to resolve them, we consider their angular separation
at the telescope objective. Another physical situation of great interest is to see very small objects
such as microbes, which may be very close. The optical device used for imaging nearby
small objects is a microscope. Let us now derive an expression for the resolving power of a
microscope.

6.6.2 Resolving Power of a Microscope

Refer to Figure 6.24, which shows two closely spaced point objects S and S¢. We assume that
these objects are self-luminous so that the light emitted by these has no constant phase relationship
and their intensities can be added up. The fringed images (diffraction patterns) of S and S¢ are
formed respectively at I and I¢. (Note that in practice, the objects are illuminated by an external
source and the waves emanating from such objects will be out of phase. In such cases,
the intensities will not be additive.) We expect that each image will consist of a central Airy disc
surrounded by a system of very faint fringes.
248 Wave Optics

Figure 6.24 Resolving power of an optical microscope. (a) Airy patterns of objects S and S¢ formed
by a microscope and separated through a distance s; (b) Enlarged ray diagram for calculating
path difference S¢B – S¢A.

If we compare Figures 6.21 and 6.24, we note that


• The angle subtended by the stars on the telescope objective is small. This is because of
the huge astronomical distances.
• The angle subtended by the objects on microscope objective, represented by the vertical
line AB, is comparatively large. This is because a microscope is used to observe closely
spaced objects. For this reason, while applying Rayleigh criterion, we express resolving
power of a microscope in terms of the smallest distance between the two objects when
their images are just resolved.
According to Rayleigh criterion, the central maximum of I should lie at the same position
as the first minimum (dark ring) of I¢. The angular separation between the Airy discs will be
1.22O
sin T min  T min
D
Refer to Figure 6.24(a). Since the first dark fringe of I¢ lies at I, we can say that when the two
objects are just resolved, the wave(s) emanating from S¢ and diffracted to I results in zero
intensity and the path difference S ¢BI – S¢AI = 1.22l. Next, we note that S ¢BI = S¢B + BI and
S¢AI = S ¢A + AI. Then using the result BI = AI, we find that the condition on path difference
takes the form S¢B – S¢A = 1.22l.
Now refer to Figure 6.24(b), which shows enlarged ray diagram of part (a). We can easily
verify that the path difference S ¢B – S B = s sin i and SA – S ¢A = s sin i, where s is the minimum
resolvable distance between S and S¢. On adding these expressions, we get S ¢B – S ¢A = 2s sin i
since SA = SB. On equating this result to 1.22l, we get the required expression for the minimum
separation between two points in an object that can be resolved by a microscope:
2s sin i = 1.22l
or
0.61O
s= (6.60)
sin i
Suppose that the space between the object and the objective is filled with a material such as oil
say, of refractive index µ. Then Eq. (6.60) modifies as
Fraunhofer Diffraction 249

0.61O
s (6.61)
P sin i
It may be pointed out here that Eq. (6.61) will hold reasonably well even if the objects S and
S¢ were illuminated by an external source.
The quantity m sini is referred to as the numerical aperture (N.A.) of the microscope.
The maximum possible value of i is p/2 and N.A. = m. Therefore, the minimum distance
between two objects that can be resolved by a microscope is 0.61l/m.
Since R.P. = s–1, we can say that resolving power (R.P.) increases with an increase in
numerical aperture and/or decrease in wavelength. It means that using ultraviolet rather than
visible light for illumination will allow better perception of finer details.
In your Modern Physics course, you have learnt about matter waves and wave-particle
duality proposed by de Broglie. Using the mass-energy equivalence as the basis, he argued that
matter should also exhibit interference and diffraction effects. The evidence in favour of this
hypothesis was provided by the experiments of Davison and Germer on electron diffraction.
The deBroglie wavelength of an electron is given by

12.3 – 1010
O m (6.62)
V
For electrons accelerated to 1 MV, the de Broglie wavelength is given by

12.3 – 1010
O 12.3 – 1013 m 123 – 10 5 nm
6
10
Note that de Broglie wavelength of an electron accelerated to 1 MV is several orders of
magnitude less than the wavelength of visible light. For this reason, R.P. of an electron microscope
is very high. In fact, as high resolution application device, an electron microscope helps us to
examine viruses, microbes and finer details of crystal structures, which would otherwise have
remained completely obscured due to diffraction effects in the visible spectrum.

The de Broglie wavelength of an electron is defined as


h h
O (i)
p mev
where h is Planck’s constant, me is mass of electron and v is speed of electron. If an electron is
accelerated through a potential difference V, the work done (eV) on it changes into kinetic energy.
So we can express electron speed in terms of V as

2eV
v (ii)
m
On combining (i) and (ii), we can write
h 1
O – (iii)
2mee V

(Contd...)
250 Wave Optics

(Contd...)

We now substitute the standard values of different constants (h = 6.6 ´ 10–34 J s, me = 9.11 ´ 10–31 kg
and e = 1.6 ´ 10–19 C) and simplify to obtain
12.3 – 1010
O m
V

6.6.3 Resolving Power of a Diffraction Grating

We now know that with the help of a telescope, we can resolve two close stars and view a faint
star deep in space. Similarly, with an optical microscope we can resolve two small and closely
spaced objects. However, these devices are not useful if we wish to resolve two nearby
wavelengths which may occur as a doublet, say. For example, sodium lamp gives out D1 and
D2 lines at l1 = 589 nm and l2 = 589.6 nm. To resolve such a doublet, we use a diffraction grating.
We know that a grating will form principal maxima corresponding to wavelengths l1 and
l2 at diffracting angle q1 and q2(>q1) respectively. Since these angles are different, let us think
that these wavelengths should be separated and appear to be resolved always. But this situation
is not obtained experimentally. You may like to know: What makes these colours to overlap?
It is the finite widths of the principal maxima. (If the spectral lines were ideal geometric lines,
they would not have overlapped.) Therefore, a logical question to raise is: How close can the
spectral lines be brought and yet be seen separate? The answer to this question is determined
by the ability of a grating and is referred to as its resolving power. It is defined as

O
R.P. (6.63)
( 'O ) min

where (Dl)min is the minimum separation of two wavelengths that the grating can just resolve
and l is the mean wavelength. It is sometimes also called chromatic resolving power.
As before, we use the Rayleigh criterion to determine the limit of resolution of two spectral
lines by a diffraction grating. According to this criterion, if the principal maximum corresponding
to wavelength l2 falls on the first minimum (on either side of the principal maximum)
corresponding to wavelength l1, as shown in Figure 6.25(b), the two wavelengths are said to
be just resolved. Let us assume that this position corresponds to common diffraction angle q.
From Eqs. (6.33) and (6.36b), we recall that the conditions of principal maxima of l2(=l + Dl)
and the first minimum of l1(=l) adjacent to the principal maximum for wavelength l2 are
d sinq = n(l + Dl)
and
O
d sinq = nO 
N
On equating these, we get
O
R.P. nN (6.64)
'O
Fraunhofer Diffraction 251

Figure 6.25 Resolution of two spectral lines according to Rayleigh’s criterion.

From Eq. (6.64), we note that the resolving power of a diffraction grating depends on the
total number of lines in the grating and the order of the spectrum. It means that to resolve the
D1 and D2 lines of sodium in the first order, the minimum number of lines in the grating
should be

5893 Å
N 982

Similarly, in second order, the resolving power of a one inch grating having 15,000 lines is

15,000
R.P. 2– 11811
2.54
From Eq. (6.64) we may conclude that R.P. of the grating will increase indefinitely, if the
number of lines is increased. This line of thought fails because the width of the grating is finite
and as N increases, the grating element will decrease.
To help you to fix these ideas, we now give some solved examples. You should solve the
in-between steps yourself.

EXAMPLE 6.8 An experimentalist uses a grating with 15,000 lines per inch and a lens of
focal length 1.5 m to observe D1(l = 589 nm) and D2(l = 589.6 nm) lines emitted by a sodium
source. Calculate the linear separation between the two lines in the first order.
Solution The grating constant d is given by

2.54
d 1.69 – 104 cm
15,000

We know that the condition of principal maximum is


d sinq = nl
252 Wave Optics

Since n = 1, for D1-line, we can write


589 – 109 m
sin T 348.5 – 103
1.69 – 10 6 m
or
q1 = sin–1(0.3485) = 0.3560 rad
Similarly, for the D2-line, we get

589.6 – 109 m
T2 sin 1 sin 1 (0.3489) 0.3564
1.69 – 106 m
Hence
Dq = q2 – q1 = 0.0004 rad
The linear separation between these lines is given by

x = fDq = (1.5 m) ´ 4 ´ 10–4 = 6 ´ 10–4 m

It means that 60 nm is separated by 0.6 mm in the first order. Alternatively, the linear separation
is 100 nm per millimetre in the first order.
EXAMPLE 6.9 In the minimum deviation position of a diffraction grating, the first order
spectrum corresponds to an angular deviation of 30º. For l = 600 nm, calculate the grating
element.
Solution We know that the condition of principal maximum is
d sinq = nl
Here, q = 30º, n = 1 and l = 600 nm. Hence

600 nm
d 1200 nm
sin 30º

We now sum up what you have learnt in this chapter.

6.7 SUMMARY

• To observe Fraunhofer diffraction, we must have a monochromatic light source and an


observation screen placed at an effectively infinite distance from the aperture.
• Fraunhofer diffraction pattern of a single vertical slit illuminated by a point source emitting
monochromatic light consists of a horizontal streak of light in the form of bright elongated
spots along a line perpendicular to the length of the slit. The spot at the centre is the
brightest. A few bright spots on either side of the central spot are symmetrically situated
about it. The central maximum is called principal maximum, whereas the other maxima
are called secondary maxima. The width of the central maximum (spot) is greater than the
width of secondary maxima (spots).
Fraunhofer Diffraction 253
• When a plane wavefront is incident on a diffracting slit, each point becomes a source of
Huygens’ secondary wavelets. These waves interfere amongst themselves and give rise to
diffracted plane waves.
• The intensity of diffraction pattern due to a vertical slit illuminated by a point source at
an arbitrary point making an angle q with the normal to the slit is given by
2 2
È sin E Ø È sin E Ø
IT A2 É Ù I0 É Ù
Ê E Ú Ê E Ú

b sin T
where A = na0 and E S .
O
• The secondary maxima correspond to points of intersections of the curves defined by
y = b and y = tan b. The points of intersection, excluding b = 0, occur at b = 1.43p,
2.46p, 3.47p and so on. On the other hand, the condition for occurrence of minima is
b = ±p, ±2p, ±3p, ... = mp; m = ±1, ±2, ±3, ...
• The intensity distribution of the diffraction pattern due to a circular aperture of diameter
D and illuminated by point source is given by

Ë 2 J1 ( D ) Û2 Ë 2 J1 ( kq sin T ) Û
2
I ( PT ) I0 Ì Ü I0 Ì Ü
Í D Ý Í kq sin T Ý

2S SD
where a = kq sinq = q sin q = sinq, l is the wavelength of light used, q is the
O O
angle of diffraction and q is radius of the circular aperture. The diffraction pattern consists
of concentric rings around a bright central disc.
• The intensity distribution of a double slit illuminated by a point source is given by
2 2
È sin E Ø È sin E Ø
IT 4 A2 É Ù cos 2 ] 4I 0 É Ù cos 2 ]
Ê E Ú Ê E Ú

G S b sin T
where ] d sin T and E S .
2 O O
• The intensity distribution of a double slit is a product of two terms: (sin2b/b2) represents
the diffraction pattern produced by one of the slits of width b and cos2z represents the
interference pattern produced by two diffracted wavelets (of equal intensity) having phase
difference d. The diffraction term governs the intensities of interference fringes, whereas
the interference term controls their width.
• When path difference between any two corresponding points in the diffracting slits is an
integral multiple of l, two beams interfere constructively leading to formation of a series
of bright fringes. The central fringe corresponds to d sinq = 0. The nth fringe on
either side of the central fringe occurs at d sinq = nl. n is referred to as the order of
interference.
254 Wave Optics

• The intensity of the N slit Fraunhofer diffraction pattern is given by


È 2 sin 2 E Ø sin 2 N]
IT ÉA Ù
Ê E2 Ú sin 2 ]
The condition for principal maxima can be rewritten as
z = np n = 0, 1, 2, ...
or
d sinqmax = nl
This is usually referred to as the grating condition.
On either side of the principal maxima, we have a minima and the condition in terms of
path difference and angle of diffraction is
O
d sin T min nO “
N
The angular half-width of the principal maximum is given by
O
GT
Nd cos T max
• For a concave transmission grating, the condition for maxima is expressed as
d(sin i + sin q) = ml
where i is the angle of incidence and q is the angle of diffraction.
• In echelon grating, the resolving power is increased by reducing the number of lines per
inch but increasing the order of spectrum substantially.
• When the maximum of the intensity curve of one object coincides with the first minimum
of the intensity curve of the second object, they are said to be just resolved. This criterion
for the limit of resolution is called the Rayleigh criterion of resolution.
• An astronomical telescope can resolve two stars if their angular separation is given by
1.22O
T min
D
If the focal length of the objective is f, the centre-to-centre linear separation between two
just resolved stars is given by
1.22 O
s f T min f
D
The ability of a telescope to resolve two nearby objects depends on the diameter of its
objective.
• The minimum separation between two points in an object that can be resolved by a
microscope is given by
0.61O
s
P sin i
Fraunhofer Diffraction 255
• The resolving power of a diffraction grating is given by
O
R.P. nN
( 'O ) min
where N is the total number of lines in the grating.

REVIEW EXERCISES

1. A He-Ne diffraction-limited laser beam (l = 630 nm) of diameter 4 ´ 10–3 m is directed


towards the earth from a space station orbiting at a height of 500 km. Calculate the area
that will be illuminated by the beam. [Ans. 0.029 km2]
2. A plane wave of wavelength 600 nm falls on a long narrow slit of width 0.6 mm.
(a) Calculate the angles of diffraction for the first two minima.
(b) How are these angles influenced if the size of slit is changed to 0.2 mm?
(c) If a convex lens of focal length 0.2 m is now placed after the slit, calculate the
separation between the second minima on either side.
[Ans. (a) 0.0573º, 0.1146º; (b) 0.1719º, 0.3438º; (c) 2.4 ´ 10–3 m]
3. For the arrangement described in 2(a) above, repeat the calculations for the first two
maxima. [Ans. 0.0819º, 0.1638º]
–4
4. A circular aperture of radius 5 ´ 10 m is placed in front of a converging lens of focal
length 0.2 m. The aperture is illuminated by a parallel beam of light of wavelength
600 nm. Calculate the radii of first two dark rings.
[Ans. 1.46 ´ 10–4 m, 1.96 ´ 10–4 m]
5. A laser source emits light of wavelength 500 nm and circular cross-section of diameter
3 ´ 10–2 m. Calculate the beam diameter after it has traversed a distance of 3 km.
[Ans. 0.12 m]
6. For two slit diffraction arrangement, the first primary maximum occurs at 30º. Calculate
the slit separation. [Ans. 2l]
7. Light of wavelength 680 nm falls normally on a diffraction grating that has 600 lines
per mm. Calculate the angle separating the central maximum from first principal maximum.
What is the largest order that can be observed with this grating and this wavelength?
[Ans. 24.1º, Central maximum and two orders on both sides]
8. Calculate the diameter of a telescope objective if a resolution of 0.1 s of arc is obtained
for l = 600 nm. [Ans. 1.5 m]
9. An astronomical telescope has diameter 40¢¢. Calculate the angular half-width of Airy disc
for l = 600 nm. [Ans. 0.15 s of arc]
10. In a double slit arrangement, each slit has width b = 5 ´ 10–4 m. The distance between
them is twice the width of either slit. If these are illuminated by He-Ne laser light
(l = 630 nm) and a converging lens of focal length 0.1 m is placed next to the double
slit arrangement, determine the positions of maxima inside the first diffraction minimum.
[Ans. 3.16 ´ 10–5 m, 9.4 ´ 10–5 m]
256 Wave Optics

11. In an experimental arrangement with a plane diffraction grating, having 4,000 lines per
cm, the collimator slit of a spectrometer has an air film enclosed between two glass plates.
On making a parallel beam of white light pass through the air film, it is found that the
angles of diffraction for a particular dark band and the tenth dark band are 17º20¢ and
20º40¢, respectively corresponding to the first order. Calculate the thickness of the air film.
[Ans. 2.39 ´ 10–3 cm]
12. White light with wavelength range from 400 nm to 800 nm is made to fall on a grating
having 4,000 lines per cm. Will the first and second order spectra overlap? Also calculate
the angular width of the first order spectrum. [Ans. No, 9º23¢35¢¢]
13. An equilateral glass prism characterized by m1 = 1.6545 and m2 = 1.6635 for l1 = 656.3 nm
and l2 = 527 nm, respectively can just resolve two sodium lines of wavelengths 588.9 nm
and 589.5 nm. Calculate the length of the base of the prism. [Ans. 14.11 cm]
14. A transmission etalon grating with 30 steps, each step of height 1 cm, has a resolving
power of 3.75 ´ 105. Calculate the refractive index of its material for a wavelength of
400 nm. [Ans. 1.5]
15. Examine whether two sodium lines of wavelengths 589 nm and 589.6 nm can be resolved in
(i) the first order, and
(ii) the second order by one inch grating having 300 lines per cm.
[Ans. No, yes]
CH A P T E R 7
DISPERSION AND
SCATTERING OF LIGHT

EXPECTED LEARNING OUTCOMES


After reading this chapter, the student will acquire the capability to:
• Explain the normal and anomalous dispersion of light.
• Discuss the electron theory of dispersion and obtain the expression for refractive index as a
function of wavelength.
• Derive the expression for intensity of scattered light.
• Explain the natural phenomena like blue of the sky and red of the rising as well as setting sun.
• Explain the Raman Effect and discuss its applications.

7.1 INTRODUCTION

In Chapter 1, we discussed how Newton formulated his ideas on corpuscular nature of light
based on his experiments on dispersion: the celebrated phenomenon of colours. He concluded
that when white light is made to pass through a triangular glass prism, it splits into its constituent
colours. That is, white light was seen to be composed of a mixture of independent colours.
In your elementary physics classes, you have learnt that dispersion is said to be normal if
• Order of the principal colours in the visible region follows the acronym VIBGYOR.
(It stands for Violet, Indigo, Blue, Green, Yellow, Orange and Red.)
• Violet light undergoes maximum deviation and red light suffers the least deviation.
We know that different substances are characterizd by different values of refractive index
and all transparent substances show normal dispersion in the visible region. Recall that the
refractive index of a material n = c/v, where c is the velocity of light in free space and v is the
velocity of light in a material medium. So we can say that dispersion is intimately connected
to the speed of light in material medium and its variation with wavelength (frequency).
This explains why refractive index of violet light is higher than that of the red colour (wavelength)
257
258 Wave Optics

for any substance exhibiting normal dispersion. It is observed that in general, the refractive
index decreases with increasing wavelength and the rate of change increases for shorter
wavelengths. Moreover, the variation of refractive index with wavelength is unique for every
material. The first qualitative explanation of the variation of refractive index with wavelength
for normal dispersion was attempted by Cauchy. However, the origin of refractive index as well
as physical basis of its frequency dependence could be understood in terms of the oscillations
of atomic dipoles. We have discussed these aspects in Section 7.2. In Chapter 9, you will get
an opportunity to learn the importance of dispersion for communication of information, its
quality and content using optical fibres.
You may now ask: Does the nature of dispersion curve (plot of refractive index with
wavelength) remain the same at all frequencies for every material? The answer to this question
reveals that refractive index changes sharply and increases with wavelength in the visible region
and beyond. This is termed anomalous dispersion. In Section 7.3, you will learn that dispersion
curves display resonance-like shape. Qualitative explanations by Sellmeier and Helmholtz provided
fairly reasonable explanation of the observed behaviour. However, in actual practice, the behaviour
of refractive index with wavelength is determined by the interplay of absorption (damping) and
restoring force (electrostatic forces between electrons and nucleus). A quantitative discussion of
dispersion based on electromagnetic theory is presented in Section 7.4.
When we analyse interaction of light with matter, light is either absorbed or scattered.
When the size of the particles is less than the wavelength of incident light, spectacular effects are
observed due to scattering of light. The laws governing scattering of light were studied in detail
by Lord Rayleigh. He observed that when monochromatic light is scattered by a transparent
substance, the scattered light has the same wavelength as the incident light. This is known as
Rayleigh scattering. The mathematical investigation of Rayleigh scattering led to a general law
for intensity of scattered light: the intensity varies inversely as fourth power of wavelength.
Rayleigh scattering is responsible for blue of the sky, the red of the setting sun as well as the
surrounding sky and the light emerging from the side of an optical fibre. You will learn about
it in Section 7.5.
When monochromatic light is incident on a system of molecules, as in a fluid medium,
several lines appear in the scattered light on either side, in addition to the incident frequency.
If the frequency of the incident light is n, the scattered light has frequencies n ± n1, n ± n2,
n ± n3, ... . The frequencies n1, n2, ..., are characteristic of the scattering molecule and quite
independent of the incident frequency. The appearance of modified wavelengths is known as
Raman Effect, after Indian physicist Sir C.V. Raman, who reported it with his co-worker
K.S. Krishnan on 28th February, 1928 in organic vapours. (You may be aware that, February
28 is observed as National Science Day in India. Raman got the Nobel Prize in 1930 for this
work.) In Section 7.6, you will learn that change of wavelength in light scattering arises due
to excitation of vibrational and/or rotational modes of the molecules. And a careful analysis of
Raman spectra yields very useful information about the structure of molecules.

7.2 NORMAL DISPERSION


We now know that when white light is incident on a triangular glass prism, it splits into seven
constituent colours, which, in turn, undergo different amounts of deviation. Physically, it means
Dispersion and Scattering of Light 259
that different colours propagate with different speeds. Now refer to Figure 7.1, which depicts
variation of refractive index with wavelength of light for some typical materials used for prisms
and lenses. For each material, these dispersion curves characterize normal dispersion and show
that in the visible region
• the refractive index decreases as wavelength increases;
• the rate of change of refractive index with wavelength is more rapid at shorter wavelengths;
• for a given wavelength, the dispersion curve is steeper for a material having higher
refractive index; and
• the shape of dispersion curves is different for different substances, i.e. dispersion curve
for every substance is unique.

Figure 7.1 Dispersion curves for a few typical transparent materials.

The first of these points explains the experimental observation that the violet colour deviates
more than the red colour when white light is made to undergo refraction through a prism.
An important consequence of the second point is that in the prism spectrum, the violet end of
the spectrum spreads out more than the red-end. We know that a substance having greater
density has higher refractive index and shows greater dispersion. For example, the refractive
indices of air, crown glass, and diamond for the violet colour are respectively 1.0003, 1.5380
and 2.4580. This is borne out in several cases, with some exceptions.
Proceeding further, we also note that a simple consequence of the last point is that we
cannot reproduce the dispersion curve for any material by adjusting the coordinates corresponding
to any other material, i.e. spectra from prisms of different materials will not have identical
relative spacing of their spectral lines. For this reason, dispersion of different substances is said
to be irrational.
260 Wave Optics

7.2.1 Cauchy Equation

In Chapter 1, we learnt that according to Maxwell’s theory, the velocity of light in free space
is given by
1
c
H 0 P0
For a homogeneous, isotropic dielectric, e0 and m0 should be replaced by e and m, respectively.
Then, the expression for velocity of light modifies to
1
v
HP
The ratio of the speed of an electromagnetic wave in vacuum to that in matter defines refractive
index n and is given by
c HP
n Ke K m (7.1)
v H 0 P0
where Ke and Km, respectively denote relative permittivity and relative permeability of the medium.
For most gases and great majority of other substances, with the exception of ferromagnetic
materials, Km is of the order of unity. Then the expression for refractive index simplifies to
n Ke (7.2)
This result shows that refractive index is an absolute dimensionless constant, independent
of frequency. But this does not conform to experimental observations, which predict that refractive
index depends on frequency. This is depicted in the form of normal dispersion curves in Figure 7.1.
To deal with this problem, Cauchy proposed an empirical equation of the form
B C
n A  (7.3)
2
O O4
Here n is refractive index and l is wavelength of light. The constants A, B, C are characteristic
of a substance.
Before proceeding further, we would like to know: Are you familiar with any other equation
in physics that is qualitatively similar to Eq. (7.3) and has been used to explain the behaviour
of a physical system? The Kammerling Onnes equation used to study the behaviour of real
gases in kinetic theory of gases has the similar form.

Kammerling Onnes Equation


To explain the behaviour of real gases at high pressures, Onnes used an empirical equation of the form
pV = A + Bp + Cp2 + Dp3 + L
Here p and V are pressure and volume respectively of the gas, and A, B, C, D, ... are constants. Onnes
used up to 25 constants for different gases. At a given temperature, A = RT. If we include only the
first term, we obtain the perfect gas equation. For a real gas, the constant B = [¶(pV)/¶p]T.
Dispersion and Scattering of Light 261
Note that
(i) the values of Cauchy constants will be different for different materials, and
(ii) the refractive index of a medium decreases as wavelength increases.
That is, the refractive index is a function of frequency. Such media are called dispersive
media. If we observe propagation of a highly peaked pulse in a dispersive medium, then we will
see that it emerges as a flattened pulse. This is because different monochromatic components
propagate with different speeds. Do you know of any non-dispersive medium? The vacuum is
non-dispersive medium for electromagnetic waves and air is non-dispersive to sound waves.
It may be mentioned here that Eq. (7.3) describes dispersion curves fairly well for many
transparent materials in the visible region. However, for most purposes, it is sufficient and
customary to retain only the first two terms in Cauchy’s equation:
B
n A (7.4)
O2
Equation (7.4) is known as Cauchy equation for normal dispersion. The coefficients A and B
are called Cauchy constants. These respectively give measure of refraction and dispersion.
You will get an opportunity to determine these in your physics laboratory. If you measure the
wavelengths of different colours in the light from a mercury source using a prism-spectrometer
arrangement, calculate the corresponding values of refractive indices and plot n versus 1/l2,
you will obtain a straight line, as shown in Figure 7.2. The intercept OP of the straight line on
the y-axis gives A and its slope gives B.

Figure 7.2 Plot of n versus 1/l 2.

The expression for dispersion is obtained by differentiating Eq. (7.4) with respect to l:
dn 2B
 (7.5)
dO O3
This result shows that (normal) dispersion is inversely proportional to the third power of
wavelength. It means that dispersion will be nearly five times at the lower end of the
visible spectrum (violet colour, l = 400 nm) as compared to that at the upper end (red colour,
l » 700 nm). The minus sign in Eq. (7.5) signifies that the slope of the dispersion curve is
negative. It may be mentioned here that Cauchy equation is of significant practical interest.
262 Wave Optics

7.3 ANOMALOUS DISPERSION

We now know that the Cauchy equation satisfactorily explains normal dispersion curves for
many transparent media in the visible region. But when refractive index of a transparent medium
was measured in the infrared region, the dispersion curve began to show marked deviation from
the normal behaviour. This is illustrated in Figure 7.3. Note that beyond a particular wavelength,
say corresponding to point E in the dispersion curve, the value of refractive index drops more
rapidly than is predicted by Cauchy equation. For the visible region, Eq. (7.4) predicts that as
l ® ¥, n ® A and that too gradually. On the contrary, the measured value of refractive index
decreases more and more rapidly as we approach the infrared wavelengths. In fact, in the
infrared region all light is blocked completely. This is due to selective absorption, which is
characteristic of a material. The increase in refractive index with wavelength is known as
anomalous dispersion. Note that this nomenclature has genesis in the existence of a large
discontinuity—absorption band—in the dispersion curve. However, these results could not be
explained by Cauchy equation.

Figure 7.3 Anomalous dispersion curve for a transparent material.

It may be mentioned here that for the first time anomalous dispersion was observed in the
visible region for materials such as iodine vapour, whose absorption bands lay in the visible
region. The ‘iodine prism’ disperses red light more than the violet colour, which is very different
from normal spectrum. Sodium vapour exhibits anomalous dispersion in the neighbourhood of
the D1 and D2 lines at wavelengths 589 nm and 589.6 nm. Later studies showed that transparent
materials such as glass and quartz exhibit selective absorption in the infrared and ultraviolet
regions, respectively. (It is now known that every substance exhibits selective absorption at
some wavelength and hence use of the term ‘anomalous’, though not very appropriate, is being
continued for historical reasons!) We now discuss Wood’s innovative experiment to observe
anomalous dispersion exhibited by the sodium vapour.

Wood’s experiment
Refer to Figure 7.4, which shows the schematics of the experimental arrangement used by Wood
to produce an equivalent of a prism of sodium vapour by vaporizing the metal in a partially
evacuated cylindrical steel tube. The tube is provided with water cooled glass windows at the
ends and an outlet for pumping. A number of pieces of metallic sodium are placed along the
Dispersion and Scattering of Light 263
bottom of the tube. As the tube was heated at the bottom, sodium began to vaporize gradually
and diffuse upwards through the residual gas. At a pressure of about 2 cm of Hg, the density
gradient gave rise to an equivalent of a sodium prism with base at the bottom of the tube and
refracting edge perpendicular to the plane of the paper.

Figure 7.4 Wood’s experimental arrangement for observing anomalous dispersion.

White light from a narrow horizontal slit S1 was made parallel by a lens L1 and after passing
through the tube formed an anomalous spectrum on the vertical slit S3 of a prism spectroscope.
Since the refractive index of light shorter than the yellow drops below one, such wavelengths
deviated upwards, whereas those longer than yellow deviated sharply downwards. In the
spectroscope, the spectrum was seen to be deviated upward on the green side of the D lines and
downwards on the red side. Two typical spectra corresponding to different densities of the
vapour are shown in Figure 7.5. Note that these spectra give a qualitative plot of n versus l and
the spectrum is continuous, except at the positions corresponding to D1 and D2 lines.

Figure 7.5 Anomalous dispersion of sodium at two different concentrations.

7.3.1 Sellmeier Equation

We now know that Cauchy equation fails to explain anomalous dispersion. To understand the
physical mechanism and derive a more general formula, Sellmeier proposed that elastic forces,
which hold the atoms/molecules together, alter the velocity of light in a medium.
We know that a body held by an elastic force has its characteristic vibrational frequency, which
is termed natural frequency, n0. When light propagates through a medium, it exerts a periodic
force and forces them to vibrate with its frequency, n. For n ¹ n0, the atom/molecule executes
small amplitude forced vibrations of frequency n. But as n ® n0, the amplitudes of vibrations
264 Wave Optics

begin to increase gradually and at n = n0, it increases to a large value due to resonance. In this
process, the velocity of light undergoes a change. Based on this mechanism, Sellmeier proposed
the following formula:
AO 2
n2 1 (7.6)
O 2  O02
where A and l0 are constants. While A is determined by the number of oscillators vibrating
with frequency n0, l0 = c/n0 is wavelength of light in vacuum corresponding to frequency n0.
Equation (7.6) is known as Sellmeier equation. Note that
1. As l ® l0 from the shorter wavelength side, n ® –¥ and if l ® l0 from the upper
wavelength side, n ® ¥.
2. As l ® 0, n ® 1 suggesting that velocity of light in the medium becomes the same as
in the free space.
It is quite possible that a molecule may have several different natural frequencies. To allow
for the existence of such a possibility, we rewrite Eq. (7.6) as

n 2
1
A0 O 2
O 2  O02

A1O 2
O 2  O12

A2 O 2
O 2  O22
 " 1 ÇO
Ap O 2
2
 O 2p
(7.7)
p

Note that l0, l1, l2, ... denote wavelengths corresponding to natural frequencies n0, n1, n2, ...,
respectively. For two natural frequencies, plot of n versus l according to Eq. (7.7) is shown in
Figure 7.6.

Figure 7.6 Plot of Sellmeier equation for a medium having two natural frequencies.

It is important to mention here that Sellmeier equation is a significant improvement over


Cauchy equation. In particular, it takes account not only of anomalous dispersion but also
explains observed behaviour of refractive index more accurately far away from the absorption
bands with the same number of constants as Cauchy equation. However, note that Cauchy
equation is an approximation of Sellmeier equation. To show this, we simplify the second term
in Eq. (7.6) by dividing the numerator as well as the denominator by l2 and obtain
1
È O02 Ø
n2 1  A É1  Ù
Ê O2 Ú
Dispersion and Scattering of Light 265
For O02 /O 2  1, we can use binomial theorem to expand the second term on the right-hand side
of this expression and retain only the first order term in (O02 /O 2 ) ignoring all higher terms. Then,
we obtain

n2
È
1  A É1 
Ê
O02
O2
 "Ø
Ù
Ú
(1  A)  A
O02
O2
 " ]
F
O2
where we have put z = 1 + A, and c = AO02 . Hence, the expression for refractive index can
be expressed as
n (]  FO 2 )1/2 ] 1/2 (1  F] 1O 2 )1/ 2
Using binomial expansion, we get
1/ 2
n= ] 
F
2] 1/ 2 O 2

F2
8] 3/ 2 O 4
 " (7.8)

This equation is analogous to Cauchy equation.


Proceeding further, we note that though Sellmeier equation successfully explains dispersion
of light in regions far removed from the absorption bands yet it fails completely at those
wavelengths where medium exhibits appreciable absorption. This follows from the fact that the
curve of Figure 7.6 shows abrupt discontinuity and goes to infinity on either side of each li.
However, such behaviour is neither borne out by experiments nor is physically possible. In fact,
very careful measurements of refractive indices of materials with absorption bands in the visible
region have been made using thin films and Michelson interferometer. The form of the observed
curves around any li is seen to be very different from that predicted by Sellmeier’s equation.
This discrepancy was resolved by German physicist H.L.F. von Helmholtz (1821–1894)
when he pointed out that in the mechanical model used by Sellmeier, the oscillator was assumed
to be free from all energy dissipative mechanisms. That is, Sellmeier made no allowance for
absorption of energy of the wave by the oscillator or any frictional resistance to its motion.
By assuming the frictional force to be proportional to velocity of the oscillator, Helmholtz
derived an expression for refractive index which accounted for absorption. The expression for
refractive index based on such a purely mechanical model can be written as
Ap O 2
n2  N 02 = 1  Çp J pO
2
(7.9a)
( O 2  O p2 )  2
O  O 2p
and
Ap O 3 J p
2nk 0 = Ç (O
p
2
 O 2p ) 2  J p O 2
(7.9b)

Here constant g p is a measure of the strength of the frictional force and k0 is related to the
absorption coefficient a and wavelength of light in vacuum l through the relation
DO
N0
4S
266 Wave Optics

Physically, we say that the intensity of light falls exponentially with distance and to a value
exp(–4pk0) of its initial value in traversing one wavelength in the medium. It may be mentioned
here that
1. In regions far from the absorption bands, the values of k0 and gp become essentially zero
and Eq. (7.9b) reduces to Sellmeier equation [Eq. (7.7)].
2. Due to friction, there will be moderate changes in the value of refractive index at
l = lp. That is, in actual practice, the dispersion curve will not exhibit any sharp
discontinuity.
So we expect that Helmholtz equations should hold for all wavelengths, including those in
the absorption band. The qualitative features of dispersion and absorption curves are depicted
in Figure 7.7 for different values of friction and absorption. Curves in Figure 7.7(a, b) respectively
signify high absorption-high damping and high absorption-weak damping. The amplitude of the
curve for a given absorption coefficient is determined by the number of oscillators causing
absorption.

Figure 7.7 Dispersion curve for an oscillator subjected to (a) variable damping, and (b) absorption.

Before proceeding further, you should answer a Practice Exercise.

Practice Exercise 7.1 Depict the dispersion curves for the entire electromagnetic spectrum.
Discuss your answer in the light of Sellmeier and Helmholtz equations.

7.4 ELECTROMAGNETIC THEORY OF DISPERSION

In order to explain the physics of the phenomenon of dispersion and frequency dependence of
refractive index on the basis of electromagnetic theory, we have to take note of the molecular
structure of matter and examine how an electromagnetic wave interacts with an atom or a
Dispersion and Scattering of Light 267
molecule. We know that an atom has a positively charged nucleus surrounded by a cloud of
negatively charged electrons. When an incident harmonic electromagnetic wave is incident on
an atom or a molecule in a dielectric medium, the internal charge distribution experiences a
periodic force, which causes the bound charges to execute oscillatory motion with the frequency
of the wave. An electron is subject to the Lorentz force F in an electromagnetic field:
F = e[E + v ´ B]
where e(=1.6 ´ 10–19 C) is charge of an electron and v is its velocity. Since velocity of an
electron is small in comparison with the speed of light, the forces arising from the magnetic
component of the field are usually negligible compared to the electric component. Further, for
polar dielectrics, the molecules undergo rapid rotations and align themselves with the E-field.
But these molecules are relatively large and so is their moment of inertia. As a result, polar
molecules fail to follow the changes in the electric field at high driving frequencies. On the
other hand, the moment of inertia of electrons is small and these continue to follow the field,
contributing to Ke even at optical frequencies, which are about 5 ´ 1015 Hz. So we can conclude
that the dependence of refractive index on frequency is governed by the interplay of the electric
polarization mechanisms occurring at a particular frequency. With this understanding, we now
derive an analytical expression for n(w).
First of all, we note that the electron cloud of an atom is bound to the positively charged
nucleus by an attractive force, which maintains an equilibrium configuration. For simplicity, we
can picture them as a mechanical system in which a mass (electron cloud here) is bound elastically
at one end of a spring of spring constant k0 and a net restoring force F comes into play whenever
there is any fluctuation from the equilibrium configuration. For small displacements, let us assume
that Hooke’s law is applicable and the restoring force can be represented as
F = –k0x (7.10a)
A mechanical representation of an atomic forced oscillator, which can oscillate equally in all
directions in an isotropic medium, is shown in Figure 7.8. Note that the negatively charged
electron cloud is fastened to a stationary positive nucleus by identical springs. Moreover, the

Figure 7.8 Mechanical oscillator model for an isotropic medium: All springs are identical and the
oscillator can vibrate equally in all directions.
268 Wave Optics

amplitude of the oscillations will seldom exceed 10–17m. From your basic knowledge of
oscillations, recall that such a system will execute simple harmonic motion with frequency
Z 0 k0 /P e , where me is the reduced mass of the electron and is nearly equal to the mass of
the electron me.
A material medium can be thought of as made up of a large number of polarisable atoms
whose size as well as inter-atomic distances is small in comparison to the wavelength of light.
When a light wave impinges on such a medium, each atom may be construed as a classical
forced oscillator driven by a sinusoidal electric field of the wave. The force exerted on an
electron of charge e by the field E(t) due to a harmonic wave of frequency w travelling along
+z-direction is of the form
F = eE(t) = ẑ eE0 cos(kz – wt) (7.10b)
where ẑ is unit vector along the z-axis.

7.4.1 Undamped System

For simplicity, let us first consider that there is no damping in the system. We know that at any
time t, an electron in an isotropic medium will be simultaneously subject to a
(i) driving force and
(ii) restoring force –k0x(t).
These are given by Eqs. (7.10b) and (7.10a), respectively. Hence, the equation of motion of an
undamped oscillator, whose displacement is in the direction of the electric field, can be written
using Newton’s second law of motion as

d 2 x (t )
me =  k0 x (t )  eE0 cos(kz  Z t )
dt 2
or
d 2 x(t ) E0
 Z 02 x(t ) = e cos( kz  Z t ) (7.11)
dt 2 me
where x(t) denotes the instantaneous position of the electron, me is its mass, k0 is force constant,
Z 0 k0 /me is natural frequency of the oscillator and k = 2p/l is wave vector of the incident wave.
To solve Eq. (7.11), we can use operator method, which we have illustrated in the box for
the more general case of a damped system, or the use of complex numbers. (You may also be
familiar with it from the course on Waves and Oscillations.) For brevity, here we shall illustrate
the use of complex numbers.
Note that the solution of homogeneous part of Eq. (7.11) will correspond to free oscillations
of frequency w0 and the motion will be simple harmonic. However, the solution of inhomogeneous
equation will correspond to frequency w and in phase with the incident wave. So we can write
x(t) = A exp[i(kz – wt)] (7.12)
where A is amplitude of the wave.
Dispersion and Scattering of Light 269
On differentiating this twice with respect to t, we get
dx
= iZ A exp[i( kz  Z t )]
dt
and
d 2x
 Z 2 A exp[i (kz  Z t)]
dt 2
On substituting these expressions in Eq. (7.11), we get

eE0
(Z 02  Z 2 ) A exp[i ( kz  Z t )] exp[i ( kz  Z t )]
me
On rearranging terms, we get the expression for A:

eE0
A
me (Z 02  Z 2 )
so that the instantaneous displacement is given by

eE0 e
x(t ) exp[i (kz  Z t ] E (t ) (7.13)
me (Z  Z )
2
0
2
me (Z  Z 2 )
2
0

As mentioned earlier, in the simplest model of the atom, a cloud of negatively charged electrons
surrounds a positively charged nucleus but the centre of the negative charges is assumed to be
at the centre of the nucleus. In a neutral molecule, the number of electrons equals the number
of protons and the negative charges compensate the charge of nucleus. But when an electric
field is applied, the internal charge distribution in the medium is modified under its influence
in such a way that each atom acquires a finite value of dipole moment. (Each pair of the positive
and negative charge centres separated through a finite distance acts as a dipole.) The dipole
moment will be equal to the product of charge and its displacement. And if N electrons per unit
volume contribute to dipole moment, the electric polarization, i.e. dipole moment per unit
volume is given by

e2 N
P (t ) eNx (t ) E (t ) (7.14)
me (Z 02  Z 2 )
Ideally, the dipole moment is defined as the product of the electronic charge and the separation
between the centres of +ve and –ve charges. We can visualise the centre of –ve charges as the
negative-end of the dipole and its distance from the centre of the +ve charge, i.e. nucleus, can
be conceived as its displacement, x(t).
By definition, P(t) = (e – e0) E(t). Therefore, on rearranging terms, we can write

P (t ) e2 N
H H0  H0 
E (t ) me (Z 02  Z 2 )
270 Wave Optics

Using the relation n2 = Ke = e/e0, we obtain the required expression for frequency-dependence
of refractive index:
1
n (Z ) 1 
2 Ne 2 È 1 Ø
1
Ne2 È
1
Z


(7.15)
É 2 2Ù É Ù
H 0 me Ê Z 0  Z Ú meH 0Z 02 Ê Z 02 Ú

This result shows that as w ® w0, n2(w) ® ¥. You may now ask: Is this conclusion realistic?
Have we neglected some quantity of physical interest in our derivation? What physical parameter
have we not accounted for and how it influences the behaviour of refractive index? We will
discover answers to these and such other questions a little later in this section. But for now, we
continue with the discussion of Eq. (7.15) and learn if there is any system of pragmatic interest,
for which predictions of this expression for refractive index are valid.
As an example, let us consider the situation in stratosphere, where electrons exist in the free
state and there is no elastic restoring force. Obviously, we have to put w0 = 0 in Eq. (7.15).
Then, the expression for the refractive index simplifies to

Ne 2
n2 (Z ) 1  (7.16)
H0 meZ 2
where N is density of free electrons. This expression is known to fairly well describe the index
of refraction for radio waves in stratosphere.
If we define plasma frequency as

Ne2
Zp (7.17)
me H 0
Eq. (7.16) can be rewritten as
ÈZ p Ø2
n2 (Z ) 1  É Ù (7.18)
Ê Z Ú

For w < wp, Eq. (7.18) predicts that n2(w) < 0 and the refractive index will be purely imaginary.
This implies attenuation of a radio wave. On the other hand, for w > wp, the refractive index
will be real. We illustrate these concepts with an example. Go through it carefully.

EXAMPLE 7.1 Suppose that refractive index of sodium arises primarily due to free electrons.
Calculate its plasma frequency using the following data: Density of sodium r = 971.2 kg m –3,
atomic weight of Na = 22.99 kg, m e = 9.1 ´ 10 –31 kg, e = 1.6 ´ 10 –19 C and
e0 = 8.854 ´ 10–12 C N–1 m–2.
Solution For sodium, we take one free electron per atom. Therefore, its number density N is
given by
6 – 1026 – (971.2 kg m 3 )
N 2.53 – 1028 m 3
22.99 kg
Dispersion and Scattering of Light 271
Hence plasma frequency of sodium

Ne2
wp =
me H 0
1/ 2
È (2.53 – 1028 m 3 ) – (2.56 – 1038 C2 ) Ø
= É 31 Ù 8.967 – 1015 Hz
Ê (9.1 – 10 kg) – (8.854 – 1012 C N 1m 2 ) Ú
Hence, wavelength corresponding to plasma frequency

2S c 2 – 22 – (3 – 108 m s 1 )
Op 210.3 nm
Zp 7 – (8.967 – 1015 s 1 )

This result shows that for w > 8.967 ´ 1015 Hz or l < 210.3 nm, the refractive index of sodium
will be real and the metal will become transparent.

Proceeding further with the discussion of Eq. (7.15), we note that if w/w0 < 1, i.e.
the natural frequency lies in the ultraviolet region, the quantity in the square bracket
1
È Z2 Ø
É1  will be positive definite when incident light has frequency anywhere in the visible
Ê Z 02 ÙÚ
region. And as w increases (or l decreases), n2 will increase. This corresponds to normal
dispersion and explains the dispersion curves shown in Figure 7.1.
Next, if we assume that w/w0 = 1, we can write
1
È Z2 Ø Z2 "
É1  1 
Ê Z 02 ÙÚ Z 02
Note that we have retained terms only up to second order in w. In this approximation, Eq. (7.15)
can be rewritten as

n2(w) = 1 
Ne 2 È
1 

Z
É Ù
meH 0Z 02 Ê Z 02 Ú
Ne 2 4S 2 c 2 Ne2 1
= 1  (7.19)
me H 0Z 02 me H 0Z 04 O02
where l0 = 2pc/w is wavelength of incident light wave in free space.
Equation (7.19) can be recast in the form
B
n 2 (Z ) A (7.20)
O02
Ne 2 4S 2 c 2 Ne 2
where A = 1  and B = .
me H 0Z 02 me H 0 H 04
272 Wave Optics

For hydrogen, the observed behaviour of n2 with l0 at 273 K and 0.76 m of Hg pressure
is given by the equation
2.11 – 10 18
n2 1  2.721 – 104  (7.21)
O02
where wavelength is expressed in metre. On comparing Eqs. (7.20) and (7.21), we get
Ne 2
2.721 – 104
me H 0Z 2
0
and
4S 2 c 2 Ne 2
2.11 – 1018 m 2
me H 0Z 04
On dividing the first of these relations by the second, we get

Z 02 2.721 – 10 4
1.29 – 1014
4S 2 c 2 2.11 – 10 18
Hence, the natural frequency of dipole is given by

Q 1.29 – 1014 c 3.4 – 1015 Hz


From Solved Example 7.1, we note that this frequency lies in the ultraviolet region.

7.4.2 Damped System

So far we have considered an undamped system. But as you know, every physical system
experiences damping and loses energy wastefully, though we make every effort to minimise it.
(In some engineering systems, we knowingly introduce damping. A familiar example is that of
brakes in automobiles. When we apply brakes, we increase friction between the tyres and the
road. This helps to reduce the speed of a vehicle in a short time.) You may have learnt about
Milliken’s oil drop experiment. A charged oil drop experiences viscous drag when it falls freely
through an electric field. The damping force is proportional to velocity of the body (Stokes’ law)
and is referred to as the viscous damping force. The direction of the resistive force is opposite
to that of velocity. Similarly, when an aircraft begins to descend or a spacecraft enters the
atmosphere of the earth, the magnitude of upward thrust, which acts as resistive force, can be
very large. Usually, the spacecraft experiences large stress, which raises its outside body
temperature. (Recall the disaster that hit spacecraft Columbia; all astronauts on board, including
American-Indian Kalpana Chawla, lost their lives. It was caused by the impact of viscous drag.)
In general, it is difficult to quantify damping exactly. However, for oscillations of sufficiently
small amplitude, it is fairly reasonable to model the damping force on Stoke’s law. That is, we
take Fd to be proportional to the (magnitude of) velocity and write
Fd = –g v (7.22)
Dispersion and Scattering of Light 273
In the instant case, we assume that electrons experience a damping force, which is proportional
to the velocity. So at any time t, an electron in an isotropic dielectric will be subject to a
(i) driving force given by Eq. (7.10b),
(ii) restoring force –k0x, and
dx
(iii) damping force –g .
dt
Hence, the equation of motion of a forced damped oscillator, whose displacement is in the
direction of the electric field, can be written using Newton’s second law of motion as

d 2 x (t ) dx(t )
me 2 =  k0 x(t )  J  eE0 cos( kz  Z t )
dt dt
or
d 2 x(t ) dx(t ) E
2
 2b  Z 02 x (t ) = e 0 cos( kz  Z t ) (7.23)
dt dt me

where x(t) denotes the displacement of the electron, me is its mass, k0 is the force constant,
2b = g /me characterizes damping, Z 0 k0 /me is the natural frequency of the oscillator and
k = 2p/l is the wave vector of the incident wave.
You came across this equation in the Oscillations and Waves course also. Recall that its
solution has two parts: (i) a transient part of frequency wd, which decays very rapidly in time
and is of a little interest to us here and (ii) A steady state part. Recall that a forced system gets
energy from the external source continuously and ultimately oscillates with the frequency of the
driving force. The steady-state forced solution of Eq. (7.23) will specify the instantaneous
displacement between the negative electron cloud and the positively charged nucleus. We can
solve it using operator method (see box on the next page for details) or complex numbers. For
continuity in our discussion, we will adhere to the latter approach here and write

d 2 x(t ) dx (t ) E0
2
 2b  Z 02 x (t ) e exp[i (kz  Z t )] (7.24)
dt dt me

As before, we assume a solution corresponding to frequency w and write


x(t) = K exp[i(kz – wt)]
where K is the amplitude of the wave.
On differentiating this twice with respect to t and substituting the expressions so obtained
in Eq. (7.24), we get
eE0
(Z 02  Z 2  2ibZ ) K
me
so that
eE0
K
me (Z 0  Z 2  2ibZ )
2
274 Wave Optics

Thus, for a damped system, the expression for polarisation modifies to

Ne 2
P (t ) E0 exp[i( kz  Z t )] (7.25)
me (Z 02  Z 2  2ibZ )

Before proceeding further with our discussion, we advise you to solve a Practice Exercise.

Practice Exercise 7.2 Rationalise the expression for polarisation and take the real part of the
resultant expression to show that

eE0
x(t ) cos(kz  Z t  T )
me [(Z 02  Z )  4b2Z 2 ]1/ 2
2 2

Note that the displacement lags behind the field by an angle q, which is governed by damping
in the system.

As before, using this relation, we can write the expression for n2 as

Ne 2
n2 1 (7.26)
me H 0 (Z 02  Z 2  2ibZ )
Recall that an oscillating charge radiates energy, the electromagnetic wave loses energy as it
propagates in the medium. From Eq. (7.26) we note that the refractive index is complex, and we
can express it as
n = s + iz (7.27)
where s and z are real numbers.

Solving an Inhomogeneous Second Order Differential Equation


We know that
1. An undamped free system (which is not subjected to any external force) oscillates with its natural
frequency w0.
2. A weakly damped (b < w0) free system oscillates with an angular frequency, wd (= Z 02  b2 ),
which is less than the natural frequency of the oscillator. Moreover, the amplitude of oscillations
decreases continuously due to loss of energy in overcoming damping.
3. When a driving force is applied, it tends to compensate for the energy loss in damping. In the
process, the oscillator begins to acquire the frequency of the applied force and the initial motion
of a weakly damped forced oscillator arises from superposition of damped oscillations (of
frequency wd ) and those of the driving force (of frequency w).
For w ¹ w0, the general solution of Eq. (7.23) is written as
x(t) = x1(t) + x2(t) (i)
where x1(t) is the solution of the homogeneous part and is given by
x1(t) = a0 cos(wd t + f) (ii)

(Contd...)
Dispersion and Scattering of Light 275
(Contd...)
Solving an Inhomogeneous Second Order Differential Equation
Since x1(t) decays exponentially, after sometime, it ceases to exist. For this reason, it is also referred
to as the transient solution. (The transient state persists from the moment the driving force is applied
till the time oscillator acquires completely the frequency of the driving force.) Note that the transient
motion corresponds to frequency wd, which is less than the natural frequency of the oscillator as well
as the frequency of the driving force. However, the system continues to oscillate since the driver feeds
energy continuously. Mathematically, we say that the general solution of Eq. (7.23) will not decay
with time. Physically, it means that a forced system gets energy from external source continuously and
will ultimately oscillate with the frequency of the driving force. The system is then said to be in
steady-state. (The transient part has no role once steady-state has been reached.)
We can solve Eq. (7.23) for the steady-state solution by assuming a harmonic solution for
instantaneous displacement, differentiate it twice with respect to time and substitute the results in the
given equation. Then trigonometric manipulations lead to the values of amplitude and phase of the
displacement. Alternately, we can use more elegant operator method or complex analysis. Here we
will demonstrate the operator method. Since x1(t) given by Eq. (ii) specifies the solution of homogenous
part of Eq. (7.23), only x2(t) is relevant in steady-state and we can write
d 2 x2 dx
 2b 2  Z 02 x2 (t ) f0 cos( kz  Z t ) (iii)
dt 2 dt
Note that we have put f0 = eE0/me.
In the operator notation (D = d/dx) and we rewrite it as

(D2 + 2bD + Z02) x2(t) = f0 cos(kz – w t)


On rearranging terms, we can write

1 1
x2(t) = f0 cos( kz  Z t ) = f 0 cos( kz  Z t )
( D2  2bD  Z 02 ) ( D2  2bD  Z 02 )
D 2  Z 02  2bD
= f0 cos(kz  Zt )
( D 2  Z 02 )2  4b 2 D 2
f0
= ( D 2  Z 02  2bD) cos(kz  Z t )
(Z  Z )  4b2Z 2
2
0
2 2

f0
= [(Z 2  Z 02 ) cos(kz  Z t )  2bZ sin(kz  Z t )]
(Z  Z )  4b
2
0
2 2 2
Z 2

If we now put Z 02  Z 2 a cos T and 2bw = a sinq, we can get

f0
x2(t) = a cos( kz  Z t  T ) (iv)
[(Z 02  Z )  4b2Z 2 ]
2 2

where
a= (Z 02  Z 2 )2  4b2Z 2

(Contd...)
276 Wave Optics

(Contd...)
Solving an Inhomogeneous Second Order Differential Equation
and
È 2bZ Ø
q = tan 1 É
Ê Z 02  Z 2 ÙÚ
Hence, the steady-state displacement is given by
x2(t) = A(w) cos(kz – wt – q) (v)
where
f0
A(w) =
[(Z  Z )  4b2Z 2 ]1/2
2
0
2 2

eE0
= (vi)
me [(Z 02  Z 2 )2  4b2Z 2 ]1/ 2
Note that:
• The frequency of oscillations of a forced weakly damped oscillator in steady-state corresponds to
the frequency (w ) of the driving force rather than its natural frequency (w0). This is because the
oscillator gradually loses its initial energy.
• The steady-state displacement lags the driving force in phase by q. However, it does not depend
on the initial conditions. In other words, the motion of a forced weakly damped system in steady-
state does not depend in any way how it began to oscillate.

By definition, the wave number k = w/v = nw/c, where c is the velocity of light. Using
Eq. (7.27), we can rewrite it in terms of refractive index as

k (V  i] )
Z
c
To obtain expressions for s and z, we combine Eqs. (7.26) and (7.27) and rationalise the
denominator of the second term. This gives

Ne 2 (Z 02  Z 2  2ibZ )
(V  i] )2 V 2  ] 2  2iV] 1
me H 0 [(Z 02  Z 2 ) 2  4b 2Z 2 ]
On comparing the real and imaginary parts, we get

Ne 2 (Z 02  Z 2 )
s2 – z2 = 1  (7.28a)
me H 0 [(Z 02  Z 2 ) 2  4b 2Z 2 ]
and
Ne2 2bZ
2sz = (7.28b)
me H 0 [(Z 0  Z 2 ) 2  4b 2Z 2 ]
2

Note that 2sz signifies absorption coefficient of the medium. Since an oscillating charge radiates
energy, the incident energy is depleted as it propagates in the medium. The qualitative variation
of s 2 – z 2 and 2sz with w are shown in Figure 7.9. Note that these curves show sharp variation
in the immediate neighbourhood of the natural (resonance) frequency of the oscillator.
Dispersion and Scattering of Light 277

Figure 7.9 Qualitative frequency variation of s 2 – z 2 and 2s z.

As mentioned earlier, an atom can in general execute oscillations corresponding to several


different resonant frequencies w0, w1, w2, ..., wp and we should make allowance for these all.
If f p denotes the fractional number of electrons per unit volume, whose resonant frequency is
wp, Eq. (7.26) for n2 modifies as
Ne2 1
n2 1
me H 0 Çf
p
p
(Z 2p 2
 Z  2ibpZ )
(7.29)

where bp signifies damping constant corresponding to frequency wp.


Note that Eq. (7.29) is similar to Helmholtz equation [Eq. (7.9)] and is expected to hold in
the entire range of electromagnetic spectrum, including X-rays.
In our discussion so far, we have used the mechanical analogy for electromagnetic treatment.
In Practice Exercise 7.1, we advised you to draw dispersion curves for the entire range of
wavelengths (frequencies). You may obviously like to seek answers to question such as: How
does damping arise when an electromagnetic wave propagates through a transparent material?
What types of charged particles are involved and how do they contribute to observe discontinuities
in the dispersion curves?
The damping in the far ultraviolet region arises due to absorption by the outer electrons in
atoms and molecules. As such, these are not shielded and in solids and liquids, an extensive
region of continuous absorption is produced. For molecular gases, the bands may consist of
quite sharp individual rotational lines. But these are so many in number that these bunch
together and damping due to collisions begins to become important. The collisions predominate
particularly at higher wavelengths. The near infrared absorption bands represent the different
natural frequencies of the atoms as a whole or even of the molecule. Since atoms/molecules are
significantly heavier than electrons, their vibrational frequencies are lower. In the far infrared
region, other molecular vibrations of lower frequency may be involved. In gases, rotational
modes also tend to become important.
Can you now guess the damping mechanism in the X-ray region? The X-ray absorptions are
attributed to the electrons in the various shells. These electrons are shielded from the effects of
collisions and electric fields due to neighbouring atoms, since these are deep inside the atom.
However, radiation damping contributes to broadening of spectrum lines. But this is quite
insignificant and absorption discontinuities are sharp.
278 Wave Optics

It is now possible to explain the phenomenon of dispersion and frequency dependence of


refractive index on the basis of electromagnetic theory. Another class of spectacular effects that
fascinated human mind whenever he gazed the open sky was the red of rising (as well as setting)
sun and the blue of the sky on a clear day. Have you noticed that the colour of the smoke rising
from the lighted end of a cigarette is blue? Do you know that following eruptions in the volcanic
island Krakatoa in the Sunda Strait west of Java, Indonesia, huge quantities of fine dust spewed
out high into the atmosphere and drifted over vast regions of the earth in 1883?
And the sun as well as the moon repeatedly appeared green or blue and the sunrise and sunset
were abnormally coloured for a few years. The scientists strived continuously to discover the
underlying principles of physics.
We know that when visible light encounters obstacles of size much greater than its wavelength,
we observe phenomena such as reflection and refraction, which are studied in the realm of
geometrical optics. Similarly, when size of the obstacle is of the order of wavelength of light,
it exhibits phenomena such as interference and diffraction, which fall in the domain of physical
optics. Since blue of sky and red of setting sun do not belong to either of the above-mentioned
two classes of phenomena, it was considered prudent to look for other alternatives. And one
viable possibility that presented itself was to consider isolated obstacles of size smaller than the
wavelength of light. A natural explanation of these optical spectacles, as also the after-effects
of pollutants released in our environment, was given by Lord Rayleigh in terms of scattering
of light by atoms and molecules. This is known as Rayleigh scattering. We first discuss the
mechanism of scattering of light from physical considerations.

7.5 RAYLEIGH SCATTERING

From the preceding section, we recall that when light propagates through a material medium,
it interacts with the bound electron cloud and imparts energy to the atom. As a result, the atom
may begin to oscillate and the frequency of oscillation, n, will be equal to the frequency of the
light wave. The amplitude of oscillations increases appreciably when n is nearly equal to the
resonance frequency of the atom.
Quantum-mechanically speaking, an atom in the ground state makes a transition to an
excited state on absorbing a photon and the absorption probability is the largest when the
frequency of the incident photon is equal to one of the excitation energies of an atom. In dense
gases, liquids and solids, absorption takes place over a range of frequencies and the usual
mechanism by which an atom de-excites to return to the ground state is thermal dissipation of
excess energy in intermolecular collisions. On the other hand, an excited atom in low pressure
rarefied gases reradiates excess energy predominantly in the form of a photon of the same
frequency in a random direction. This process was first observed by R.W. Wood in 1904 and
is known as resonance radiation. (We can observe this effect by putting a bit of pure metallic
sodium in an evacuated glass bulb. On heating the bulb gradually, the sodium vapour pressure
increases. If a portion of the vapour is illuminated with a strong beam of light from a sodium
arc, it glows with the characteristic yellow resonance radiation of sodium.)
At frequencies above and below the resonance frequency, the electron cloud oscillating with
respect to the nucleus behaves as an oscillating electric dipole and the emitted photon may have
Dispersion and Scattering of Light 279
the same frequency as that of the incident light. The absorption of energy from an incident wave
and re-emission of some part of it subsequently is known as scattering. The amplitude of an
oscillation, and thus the amount of energy removed from the incident wave, increases as the
frequency of the incident wave approaches a natural frequency of the atom. For low density
gases, we can ignore atomic interactions and hence absorption. So the scattered wave will carry
increasingly more energy as the driving frequency approaches a resonance. When natural
frequencies of the atom are in the ultraviolet region and incident wave is in the visible region,
some interesting effects are observed. In such a situation, an increasing fraction of incident
wave will be scattered elastically as the frequency of the incoming light increases.
From quantum mechanical point of view, you can visualize the mechanism as follows:
At frequencies other than those corresponding to the stationary energy levels of an atom, a
photon may be re-radiated without any appreciable time delay and most often with the same
energy as that of the absorbed quantum. The process is known as elastic or coherent scattering
because there is a definite phase relationship between the incident and the scattered fields.
The scattering of light by small particles whose linear dimensions are considerably smaller than
the wavelength of incident light is known as Rayleigh scattering. This is depicted pictorially in
Figure 7.10.

Figure 7.10 Rayleigh scattering: Quantum mechanical depiction.

The quantitative investigations made by Rayleigh led to a general result: The intensity of
light scattered by small particles is directly proportional to the fourth power of the driving
frequency:
1
I *Q 4 K 4 (7.30)
O
It may be mentioned here that Eq. (7.30)
• applies to any particle of refractive index different from that of the surrounding medium;
and
• for a given size of the particles, longer wavelengths are less scattered than shorter
wavelengths.
Since wavelength of red light (l = 700 nm) is nearly 1.75 times the wavelength of violet
light (l = 400 nm), according to Eq. (7.30), the scattering of the violet light is nearly 9.4 times
as great as for red light for the same incident intensity. Alternatively, we can say that red light
is scattered nearly one-tenth of the violet light.
280 Wave Optics

Proceeding further, we give a brief account of Rayleigh scattering. We first assume


that each scattering centre behaves independently as for a gas where average inter-atomic
spacing is greater than the wavelength. As discussed in Section 7.4, if we ignore damping,
the electric field E(t) of an electromagnetic wave of frequency n incident on an atom generates
an electric dipole. Using Eq. (7.14), we can write the expression for the dipole moment as

e2
P (t ) E (t ) (7.31)
me (Z 02  Z 2 )

where w0 is natural frequency of the atom.


From your electricity classes, you may recall that if the electric field is sinusoidal, an
oscillating electric dipole it produces also exhibits sinusoidal behaviour:
P(t) = p0 exp(–iwt)
And the rate at which energy is radiated by the dipole is given by

7
Z4 p02
Z4 e4
E02 (7.32)
12SH 0 c3 12SH 0 c 3 me2 (Z 02  Z 2 ) 2
If the number of dipoles produced in the medium per unit volume is N, the total energy radiated
per unit volume per unit time is given by NT.
Now let us suppose that the light wave is propagating along the x-axis. The intensity of the
wave is given by [see Eq. (1.34)]:
1
I H 0 cE02 (7.33)
2
The change in the intensity of an electromagnetic wave when it covers a distance dx can be
expressed as
dI = –NT dx
On combining this result with Eq. (7.32), we can write

dI NZ 4 e4
 dx  W dx (7.34)
I 6SH 02 c 4 me2 (Z 02  Z 2 ) 2
where
NZ 4 e4
W (7.35)
6SH 02 c 4 me2 (Z 02  Z 2 ) 2
is the attenuation coefficient. For most atoms, the values of natural frequency of vibration lie in the
ultraviolet region. Thus, if w = w0, the expression for attenuation coefficient takes a compact form:

NZ 4e4 8 N S 3e 4 1
W (7.36)
6SH 02 c 4 me2Z 04 3H 02 me2Z 04 O 4
Dispersion and Scattering of Light 281
where l = c/n. It shows that attenuation coefficient is directly proportional to fourth power of
frequency or inversely proportional to the fourth power of wavelength:
t µ w4
or
1
tµ (7.37)
O4
This is the famous Rayleigh scattering law.
On integrating Eq. (7.34), we obtain
I = I0 exp(–tx) (7.38)
Note that I0 signifies intensity at x = 0.
It may be mentioned here that Rayleigh studied scattering of light by the atmosphere and
successfully explained the observed visual effects such as blue colour of sky and red colour of
rising and setting sun. The deep red colour at sunrise and sunset indicates high concentration
of particulate matter in the environment.
For a gas, we can rewrite Eq. (7.15) as
Ne 2
n 2 (Z )  1 2
H 0 me (Z 0  Z2)
For air, the refractive index is very nearly equal to one. Therefore, we can write
n2 – 1 = (n – 1) (n + 1) ; 2(n – 1) so that the expression for refractive index can be expressed as

Ne2
n 1 (7.39)
2H 0 me (Z 02  Z 2 )
On combining Eqs.(7.35) and (7.39), we obtain

2 ÈZ Ø
4
2 2k 4
W É Ù ( n  1) (n  1) 2 (7.40)
3S N Ê c Ú 3S N
where k = w/c is wave number.
The attenuation length, L, is defined as the distance in which intensity of a wave drops to
e–1 times its initial value. From Eq. (7.38), note that it is reciprocal of attenuation coefficient
(L = t –1). Using Eq. (7.40), we can easily calculate the values of attenuation coefficient for
different colours. For air at STP, n – 1 = 2.78 ´ 10–4 in the visible region. With N = 2.69 ´ 1025
molecules per m3, the attenuation length comes out to be 30 km for l = 410 nm (violet colour)
and 188 km for l = 650 nm (red colour).
We know that the density of air molecules decreases exponentially with altitude. The intensities
of different colours at earth’s surface relative to that incident on the top of the atmosphere for
wavelengths corresponding to violet and red colours when the sun is at zenith and sunrise–
sunset are:
Colour Zenith Sunrise–Sunset
Red (650 nm) 0.96 0.21
Violet (410 nm) 0.76 0.000065
282 Wave Optics

From this we note that at sunrise and sunset, the light is predominantly red and the colour
deepens towards the zenith as well as with the altitude. In practice, the attenuation is greater
because of the presence of water vapour and ozone gas, which strongly absorb ultraviolet light,
apart from dust and pollutants in the atmosphere near earth’s surface.
Now refer to Figure 7.11. It shows the solar energy spectrum incident on earth’s atmosphere
(curve A) and at sea level when the sun is at the zenith (curve B) as a function of photon energy.
The expected sea-level spectrum for Rayleigh scattering by a dry and clean atmosphere with sun
at the zenith is shown by the dashed curve.

Figure 7.11 Solar energy spectrum incident on earth’s atmosphere (curve A) and at sea level when the
sun is at the zenith (curve B) as a function of photon energy.

We can now explain why the sky is blue and what makes the sunset red. We know that the
atmosphere is the mixture of gas molecules and other materials surrounding the earth. It is made
mostly of nitrogen (78%), and oxygen (21%) gases. Argon gas and water (in the form of vapour,
droplets and ice crystals) are the other constituents. There are traces of other gases plus extremely
small particulate matter and pollutants like dust, soot, ashes, and pollen. The exact composition
of the atmosphere varies, depending on the location, the weather, and such other conditions.
For example, there may be more water in the air after a rainstorm or near the ocean.
Volcanic eruptions such as experienced in April 2010 in Iceland and October 2010 in Indonesia
put large amounts of dust particles high (up to 30,000 m) into the atmosphere. Pollution adds
different gases, dust and soot at different altitudes in the atmosphere.
Light travels through space unhindered. But as light enters earth’s atmosphere, it bumps into
a dust or a gas molecule. Then what happens to the light depends on its wavelength and the
Dispersion and Scattering of Light 283
size of the obstacle it hits. The size of dust particles and water droplets is an order of magnitude
larger than the wavelength of visible light. When light is incident on such large particles,
it is reflected in different directions. The different colours are however reflected by such a
particle in the same way. Consequently, light appears white because it still contains all the seven
colours. On the other hand, the size of gas molecules and smoke particles is a fraction of the
wavelength of visible light. When light interacts with them, some of it may be absorbed by the
molecule and re-radiated in a different direction after a short while. It may be mentioned
here that although all the colours can be absorbed, higher frequencies (blues) are absorbed
(and re-radiated) more than the lower frequencies (reds).

Why is the sky blue?


The blue colour of the sky on a clear day can be explained as being due to Rayleigh scattering.
As light moves through the atmosphere, most of the longer wavelengths (red, orange and
yellow) remain unaffected but much of the shorter wavelength light is absorbed by the gas
molecules. The absorbed blue light is then radiated in different directions. It gets scattered
all around the sky. (We can similarly reason out the blue colour of the smoke of a lighted
cigarette.) Whichever direction we may look, some of this scattered blue light reaches us.
Since we see the blue light from everywhere overhead, the sky looks blue. This is depicted
schematically in Figure 7.12. And the blue of the sky is more saturated when you look further
from the sun.

Figure 7.12 Predominant scattering of longer wavelengths in


earth’s atmosphere gives the sky blue colour.

It is common experience that the sky appears pale in colour closer to the horizon. This is
because before reaching us, the scattered blue light has to pass through more air and some of
it gets scattered away again in other directions. As a result, less blue light reaches our eyes.
We now know why the sun appears yellow on the earth. But if we were out in space or
on the moon, the sun would look white. You should not be surprised because there is no
atmosphere to scatter the sunlight. Also, out in space, the sky looks dark and black because
there is no atmosphere and no scattered light reaches us.
284 Wave Optics

Why is the sunset red?


As the sun begins to set, the light has to travel farther through the atmosphere before it gets
to us. As a result, more light is reflected and scattered. That is, less light reaches us directly and
the sun appears less bright. The colour of the sun itself appears to change, first to orange and
then to red. This is because even more of the short wavelength blues and greens are now
scattered. Only the longer wavelengths are left in the direct beam that reaches our eyes.
This is depicted in Figure 7.13.

Figure 7.13 Sunrise and sunset are red due to reflection of light and scattering of shorter
wavelengths by small particles and air molecules in earth’s atmosphere.

The sky around the setting sun may take on many colours. The most spectacular shows occur
when the air contains many small particles of dust or water. These particles reflect light in all
directions. Then, as some of the light heads towards us, different amounts of the shorter wavelength
colours are scattered out. We receive predominantly longer wavelengths, and the sky appears
red, pink or orange. This can also give us an idea about pollution in the environment/atmosphere.

7.6 RAMAN EFFECT


In the preceding section, we discussed how blue colour of the sky or red colour of the rising
and setting sun were explained by Rayleigh in terms of scattering of light by air molecules.
While on his way to Bombay from London in 1921, Sir C.V. Raman was fascinated by the deep
blue colour of the Mediterranean sea. He did not accept Lord Rayleigh’s explanation that the
colour of the sea was just a reflection of the colour of the sky. Instead, he conjectured that the
colour of the sea had genesis in the scattering of sunlight by water molecules. To establish this,
a series of experiments on scattering of light by liquids and solids were undertaken at Indian
Association for Cultivation of Science (IACS), Kolkata under his guidance.
The fundamentals of Raman’s crucial but simple experiment comprised isolating the violet
light of the solar spectrum using a violet filter. The violet light was made to pass through a
liquid sample. Most of the light emerging from the liquid sample was predominantly of the same
wavelength as the incident violet beam—the so-called Rayleigh scattered light. However, C.V.
Raman and K.S. Krishnan showed that some of the scattered light was of a different wavelength,
which they could isolate using a green filter placed next to the sample. In his paper to Nature,
titled “A New Type of Secondary Radiation,” Raman indicated that approximately 60 different
liquids had been studied, and all showed the same result—some scattered light had a different
Dispersion and Scattering of Light 285
colour than the incident light. “It is thus,” Raman wrote, “a phenomenon whose universal nature
has to be recognised.”
Raman Effect is a very weak effect; only one in a million photons actually exhibits change in
wavelength. In fact, the Raman signal intensity is orders of magnitude weaker than the elastic
scattering intensity (Rayleigh scattering). For diatomic molecules, Raman scattering signal is more
than 1000 times weaker than the Rayleigh signal. For solids, this difference can be more than 106.
In all of his early studies, Raman used sunlight as the excitation source and it lacked the
desired intensity. In early 1928, he switched to more intense mercury arc lamps and measured
the exact wavelengths of the incident as well as Raman scattered lights. They used a quartz
spectrograph to photograph the spectrum of the scattered light. These quantitative results were
first published in the Indian Journal of Physics on March 31, 1928.
The significance of Raman Effect was recognised quickly by other scientists the world over.
R.W. Wood cabled Nature to report that he had verified Raman’s “brilliant and surprising
discovery ... . It appears to me that this very beautiful discovery is one of the most convincing
proofs of quantum theory.” In his 1930 Nobel Prize address, Raman remarked that “... the
character of the scattered radiations enables us to obtain an insight into the ultimate structure
of the scattering substance.”
Before proceeding further, it is desirable to appreciate how the phenomenon operates.
A qualitative picture of the Raman Effect can be obtained using electromagnetic theory. However,
it cannot predict the actual magnitudes of the changes in frequency. When an electromagnetic
wave interacts with matter, the electron orbits within the constituent molecules are perturbed
periodically with the same frequency (n0) as the electric field of the incident wave. The oscillation
of the electron cloud results in a periodic separation of charges within a molecule, producing
an induced dipole moment. The oscillating induced dipole moment acts as a source of
electromagnetic radiation, thereby resulting in scattered light. The majority of light scattered is
emitted at the same frequency (n0) as that of the incident light. This process corresponds to
elastic scattering. However, additional light is also scattered at different frequencies due to so
called inelastic scattering. Raman scattering is an example of inelastic scattering.
As discussed in Section 7.4, an incident electromagnetic wave induces a dipole moment
during the light-matter interaction. The strength of the induced dipole moment is given by
a 0E0 cosw0t, where a0 is the polarisability and E0 is the strength of electric field of the incident
electromagnetic wave. (The polarisability is a material property that depends on the molecular
structure and nature of the bonds.) For the incident electromagnetic wave, the electric field is
sinusoidal so that the time-dependent induced dipole moment is given by
M(t) = a E0 cosw0t
Note that the ability to perturb the local electron cloud of a molecular structure depends on the
relative location of the individual atoms. Therefore, polarisability is a function of the instantaneous
position of constituent atoms. For any molecular bond, the individual atoms are confined to
specific vibrational modes which correspond to quantized vibrational energy levels. (This is
similar to electronic energies.) The energy of a particular vibrational mode is given by

È 1Ø
Evib É j Ù hQ vib (7.41)
Ê 2Ú
286 Wave Optics

where j denotes the vibrational quantum number and takes values 0, 1, 2, , nvib is frequency
of the vibrational mode, and h is Planck’s constant.
The displacement dy of the atoms about their equilibrium position due to the particular
vibrational mode can be expressed as
dy = y0 cosw vibt (7.42)
where y0 is amplitude of vibration about the equilibrium position. For a typical diatomic
molecule such as N2, O2 and H2, the maximum displacement is about 10% of the bond length.
For such small displacements, the polarisability may be approximated by a Taylor series
expansion. Then we can write
˜D
D D0  d\ (7.43)
˜\
where a0 is the polarisability of the molecule at equilibrium position.
On substituting for vibrational displacement dy from Eq. (7.42), we can rewrite Eq. (7.43)
for polarisability as
˜D
D D0  \ 0 cos Z vib t (7.44)
˜\
On inserting this result in the expression for dipole moment, we get
˜D
M (t ) D 0 E0 cos Z 0 t  E0\ 0 cos Z 0t cos Z vib t
˜\
Using the trigonometric relation 2 cosA cosB = cos(A + B) + cos(A – B), we can rewrite this
result as

M (t ) D 0 E0 cos Z 0t 
˜D E0\ 0 [cos(Z  Z )t  cos(Z  Z )t ] (7.45)
˜\ 2 0 vib 0 vib

From this equation we can say that induced dipole moments are created at three distinct
frequencies—n0, (n0 – nvib) and (n0 + nvib)—and the corresponding spectral lines are observed
in the spectrum. Of these, the first scattered frequency corresponds to the incident frequency—
the elastic scattering or Rayleigh scattering. The latter two frequencies are shifted to lower or
higher values and therefore represent inelastic scattering or Raman scattering; the lower frequency
(longer wavelength) is referred to as Stokes scattering, and the higher frequency (shorter
wavelength) is referred to as anti-Stokes scattering.
Note that the necessary condition for Raman scattering to occur is that the term ¶a/¶y must
be non-zero. This condition may be physically interpreted to mean that the vibrational displacement
of atoms corresponding to a particular vibrational mode results in a change in the polarisability.
To illustrate, let us consider a diatomic molecule A-B, with maximum vibrational displacement
y0. Recall that the ability of an incident electric field to perturb the electron cloud depends on
the relative positions of the atoms. For example, when diatomic molecule A-B is at maximum
compression, the electrons from a given atom feel the effects of the nucleus of the other atom
and are therefore not perturbed as much. It means that the polarisability is less for smaller bond
Dispersion and Scattering of Light 287
length. On the other hand, when atoms in the diatomic molecule have greater separation,
the electrons are more readily displaced by the electric field of the incident electromagnetic
wave. That is, polarisability increases at the maximum bond length. Therefore, ¶a/¶y about the
equilibrium position (¶y = 0) will be finite. This implies that the fundamental vibrational mode
of the diatomic molecule under consideration would be Raman active and give rise to inelastically
scattered light of frequencies (n0 – nvib) and (n0 + nvib).
The condition that the polarisability must change with vibrational displacement can be
thought of as the Raman selection rule. It may be mentioned here that this discussion provides
a classical framework for understanding Raman scattering. However, it is also useful to describe
Raman scattering in terms of the discrete vibrational energy states of each molecular vibrational
mode. This is commonly done by considering a vibrational energy well, as shown in
Figure 7.14. (Each discrete vibrational state corresponds to the vibrational quantum numbers
defined by Eq. (7.41). The adjacent energy levels differ by one quantum number, hence DEvib
is equal to hnvib.)

Figure 7.14 Vibrational energy states of a molecule.

Raman scattering may also be interpreted as a shift in vibrational energy state due to the
interaction with an incident photon. The incident electromagnetic wave induces an oscillating
dipole moment, as discussed above, thereby putting the molecular system into a virtual energy
state. The energy level of the virtual state is generally much greater than the vibrational quanta,
but is not necessarily (and generally not) equal to any particular electronic quantum energy.
Therefore, the molecule stays in its ground electronic state. However, during the interaction with
the incident photon (i.e. energy of the electromagnetic wave), an amount of energy equal to the
vibrational mode may be imparted to the molecule, as shown in Figure 7.15. As a result,
the residual photon energy (which is now less than the energy of the incident photon) leaves
the molecule as Raman scattered (i.e. inelastically scattered) radiation.

Figure 7.15 Conservation of energy for Raman scattering (Stokes shift).


288 Wave Optics

As such, a molecular system is more likely to exist in the ground vibrational state.
Therefore, the Stokes scattering as depicted above, is the dominant Raman scattering process.
For anti-Stokes scattering, the molecule has to be in an excited vibrational state (e.g. j = 1)
before interacting with the incident electromagnetic wave. During the interaction of the incident
photon with the molecule, the vibrational energy state goes back to a lower level (Dj = 1), and
the vibrational quantum of energy associated with the change in vibrational state is then taken
up by the incident photon. This results in a scattered photon of energy h(n0 + nvib). For a large
system of molecules, both Stokes and anti-Stokes Raman scattering occurs simultaneously.
However, the intensity of Stokes scattering is, in general, greater than the intensity of anti-
Stokes scattering.
We now know that a molecule can absorb energy from the sun. If the energy is absorbed
• in the visible and ultraviolet regions, the mechanism of electron transitions, much like
those of an atom, comes into play;
• in the far-infrared and microwave regions, it is converted to rotational kinetic energy; and
• in the infrared region, it transforms into vibrational motion of the molecule.
Now refer to Figure 7.16. It shows vibrational modes of a molecule, which we denote
as j = 0, 1, 2, ... . The molecule may or may not be an excited state. Suppose that an
incident photon of energy hni is absorbed and the system goes to some intermediate state.
Subsequently, it makes a Stokes transition, emitting a (scattered) photon of energy hns < hni.
The principle of energy conservation demands that the difference in energies hni – hns = hn21
is utilised in exciting the molecule to a higher vibrational state j = 2. On the other hand, a
molecule initially in an excited state may, on absorbing and emitting a photon, fall back to a
lower state. This gives rise to anti-Stokes transition. In this case, hns > hni, which means that
some vibrational energy of the molecule has been converted into radiant energy. In both the
cases, the difference between ns and ni corresponds to the energy difference corresponding to
the specific energy-levels for the material under study and affords a unique tool to ‘look’ into
its molecular structure.

Figure 7.16 Vibrational energy states and Raman scattering.


Dispersion and Scattering of Light 289
Note that Raman scattering is somewhat similar to fluorescence*. However, there are two
important differences:
(i) the light incident on the scatterer should not correspond to any of its absorption band/
line, and
(ii) the intensity of Raman scattered light is much less intense. (That is why Raman Effect
is rather difficult to detect and we need high intensity source such as a laser to
observe it.)

7.6.1 Raman Spectroscopy


Soon after its discovery, the Raman Effect was used to study vibrational and rotational modes
of molecules and relate these to molecular structure. The Raman Effect was also adopted by
chemists as an analytical and research tool. In fact, the Raman Effect became the principal
method of non-destructive chemical analysis of both organic and inorganic compounds. As we
now know, the unique spectrum of Raman lines for any particular substance can be used for
qualitative analysis, even in a mixture of materials. Further, the intensity of the spectral lines
is determined by the amount of substance. It may be mentioned here that Raman spectroscopy
can be applied not only to liquids but also to gases and solids. And unlike many other analytical
methods, it lends itself naturally to the analysis of aqueous solutions. It has been a ubiquitous
technique to get information on what and how much substance is present in a sample. However,
non-availability of strong sources and the requirement of large samples seemed to hamper its
applicability initially. But the situation changed dramatically after the discovery of the laser.
(Laser is an intense, collimated, quasi-monochromatic light source available in a wide range of
frequencies.) Yet researchers continued to actively work on Raman Effect from fundamental
research to applied solutions.
Raman spectroscopy with a laser as the excitation source enables measurement of relatively
small Raman shifts with improved spatial resolution and signal-to-noise ratio.
Raman spectroscopy has also been used to monitor manufacturing processes in the
petrochemical and pharmaceutical industries. Illegal drugs captured at a crime scene can also
be analysed instantaneously without breaking the evidence seal. Using a fibre-optic probe, it is
now used to analyse nuclear waste material from a safe distance. Photo-chemists and
photo-biologists use laser Raman techniques to record the spectra of transient chemical species
with lifetimes as small as 10–11 s. Raman spectroscopy can be used to obtain specific biochemical
information that may foreshadow the onset of life-threatening illnesses, including malignancy
in any part of body. In fact, the applications of Raman Effect are so many and so varied that
volumes have been written on them. However, due to limitations of space and scope, here we
shall discuss applications only for determination of the structure of molecules.
Raman spectra are determined by the factors which influence the nature of vibrations the
most. These include the number of atoms in a molecule, the mass of the atoms and the strength of

*When a photon is absorbed by an atom, it is excited to a higher energy state. If the excited atom emits a

photon almost instantaneously (within roughly 10–7 s) but fails to regain its initial state, the process is called
fluorescence. However, if there is appreciable time gap between absorption and emission of a photon (in some
cases up to hours), the process is known as phosphorescence.
290 Wave Optics

the chemical bonds between them. We know that oscillations of a diatomic molecule can be
analysed as a single body oscillation and the natural frequency of oscillations is given by
1 F
n0 = , where F is restoring force and m is reduced mass of the molecule. It means that
2S P

(i) there is only one frequency of oscillation for a diatomic molecule;


(ii) a lighter molecule will have higher frequency of oscillation; and
(iii) a molecule having stronger binding force between the atoms will have higher
characteristic frequency.
The binding force depends on the nature and strength of inter-atomic bonds. It means that
a diatomic molecule with double bond should have a higher frequency than that containing only
a single bond.
The Raman lines are expected to appear with relatively greater intensity for molecules
having covalent bonding as in homo-polar or electrically non-polar molecules. On the other
hand, for polar or hetero-polar molecules, which have electrovalent bonding, the Raman lines
will have low intensity. This is essentially because Raman lines depend on the symmetry of
molecules and the extent to which oscillations influence polarisability. In covalent molecules,
the binding electrons are shared by the atoms so that polarisability of a molecule is significantly
modified by nuclear oscillations and this variation gives rise to Raman lines. On the contrary,
in electrovalent molecules, the binding electrons transfer from one atom to another in the
formation of the molecule so that nuclear oscillations do not affect molecular polarisability of
the molecule implying non-appearance of Raman lines.
Raman lines are more complex for poly-atomic molecules. For example, a tri-atomic molecule
will, in general, have three frequencies. Moreover, the arrangement of atoms—linear or non-
linear, symmetrical or unsymmetrical—also determines the intensity of Raman lines. Hence, in
conjunction with infrared data, it is possible to draw useful inferences about molecular structure
based on the number and the intensity of the observed lines in Raman spectra. According to the
rule of mutual exclusion, transitions that are allowed in the infrared region are forbidden in the
Raman spectrum for molecules with a centre of symmetry and vice versa. To illustrate these
remarks let us consider diatomic and tri-atomic molecules.
We first consider diatomic molecules such as H2, D2, N2, O2, HCl, HBr and HI. Of these,
the first four molecules are homo-nuclear, i.e. made up of identical atoms, whereas the other three
molecules are hetero-nuclear, i.e. made up of different atoms. In all these cases, there is only
one oscillation frequency and its value can be obtained from Raman spectra. It is observed that
• the heavier the molecule, the lower is the vibration frequency; and
• the value of restoring force per unit displacement, which is a measure of the binding
strength, may be classified as 3 : 2 : 1 indicating the existence of three different types
of bonds—triple, double, or single—between the atoms in a molecule.
For tri-atomic molecules, let us consider CO2, N2O, H2O as typical examples. For CO2, very
strong bands are observed in its IR absorption spectrum at 668 cm–1 and 2349 cm–1, and only
one strong band exists at 1389 cm–1 in its Raman spectrum. Since none of these occur in Raman
as well as IR spectrum, the rule of mutual exclusion suggests that the carbon dioxide molecule
must have a centre of symmetry. It implies that the molecular structure is linear and symmetric.
Dispersion and Scattering of Light 291
So molecular structure of CO2 is O – C – O. The analysis of the IR and Raman spectra of
nitrous oxide (N2O) reveals that this molecule is linear and asymmetrical. The three fundamental
frequencies of N2O at 2224 cm–1, 1285 cm–1 and 589 cm–1 have been recorded in the IR
absorption spectrum. Of these, two (2224 cm–1 and 1285 cm–1) appear in Raman spectrum also.
But the line at 589 cm–1 has not been recorded due to weak intensity.

Life of Sir C.V. Raman


Raman was born in a highly educated and scholarly family in Trichy, Tamil Nadu (India) in 1888.
His father was a professor of physics and mathematics and mother came from a family of sanskrit
scholars. Young Raman exhibited glimpses of brilliance at an early age. He received a B.A. degree
from Presidency College, Madras (now Chennai) with gold medal in physics. While studying for his
M.A. degree, he published his first research paper in Philosophical Magazine at the young age of 18.
(It was the first research paper ever published from Presidency College.) He could not continue his
further education for several reasons, including poor health. In 1907, he joined the Financial Civil
Service and was posted at Kolkata (earlier Calcutta) as Assistant Accountant General. However, his
fascination for research continued to charm him.
Soon after arriving in Kolkata, Raman began pursuing research after-hours at the Indian Association
for Cultivation of Science (IACS). In the first ten years, working almost alone, he published 27
research papers. This helped him to put the IACS on the world map as a vibrant research centre. Much
of this early work was on the theory of vibrations in musical instruments. He received research prizes
in 1912 and 1913 while he was still a full-time civil servant. He lectured extensively on popular
science and held the audience spellbound with his lively demonstrations, superb diction and rich
humour.
At the age of 29, he resigned from his civil service job and joined Calcutta University as Palit Chair
Professor on the invitation of Sir Ashutosh Mukherjee. This was a defining milestone in his career.
Raman continued to use the IACS as the research arm of the University. In 1924, he was elected
Fellow of the Royal Society.
After the discovery of the Raman Effect in 1928, he was knighted by the Government of British
India and received the Nobel Prize in physics in 1930. Three years later, Raman left Kolkata for
Bangaluru (earlier Bangalore), where he served as Head of the Indian Institute of Science. While
continuing his work on the Raman Effect, he became interested in the structure of crystals, especially
diamond. In 1934, he founded the Indian Academy of Science and began the publication of its
Proceedings.
Soon after India became free, Pt. Jawaharlal Nehru began restructuring Indian science and research.
To accomplish his avowed objectives, he created a number of research institutions and offered
Sir C.V. Raman appointment as Director of the newly established Raman Research Institute in 1948.
Raman remained continually active, delivering his last lecture just two weeks before he set out for
heavenly abode in 1970. His research interests changed in later years when he primarily investigated
the perception of colour. He was so dedicated to his research students that he declined an invitation
from the President of India to participate in a function held at Rastrapati Bhawan, New Delhi, for one
of his students was to submit PhD Thesis on that day and he thought that he should be available to
the student in Bangaluru.
Raman did outstanding work at a time when only a few Indians made science a career. In fostering
Indian science, Raman emerged as one of the heroes of the Indian political and cultural renaissance.
Through his deceptively simple experiment, Raman demonstrated that Mother Nature likes simplicity
and its mysteries can be best unfolded with logical thinking.
(Contd...)
292 Wave Optics

(Contd...)

Life of Sir C.V. Raman


The American Chemical Society and the Indian Association for the Cultivation of Science designated
the discovery of the Raman Effect as an International Historic Chemical Landmark in a ceremony on
December 15, 1998 at the IACS. The plaque commemorating the event reads:
“At this institute, Sir C. V. Raman discovered in 1928 that when a beam of coloured light entered a
liquid, a fraction of the light scattered by that liquid was of a different colour. Raman showed that the
nature of this scattered light was dependent on the type of sample present. Other scientists quickly
understood the significance of this phenomenon as an analytical and research tool and called it the
Raman Effect. This method became even more valuable with the advent of modern computers and
lasers. Its current uses range from the non-destructive identification of minerals to the early detection
of life-threatening diseases. For his discovery Raman was awarded the Nobel Prize in physics in 1930.”

Let us now summarise what you have learnt in this chapter.

7.7 SUMMARY

• In normal dispersion, order of the principal colours in the visible region follows the
acronym VIBGYOR; the red colour deviates the least and the violet colour the most.
• For normal dispersion in the visible region, the refractive index decreases as wavelength
increases; the rate of change of refractive index with wavelength is more rapid at shorter
wavelengths, and at a given wavelength, the dispersion curve is steeper for a substance
having higher refractive index.
• Cauchy’s empirical equation to explain frequency dependence of refractive index for
normal dispersion is of the form
B
n A 2  4 
O
C
O
"
where l is the wavelength of light and A, B, C, ... are characteristic constants of a
substance.
• The increase in refractive index with wavelength is known as anomalous dispersion.
The Sellmeier’s equation for anomalous dispersion is
Ap O 2
n2 1 ÇO 2
 O p2
p

where lp corresponds to pth natural frequency of vibration of the substance.


• According to electromagnetic theory of dispersion, interaction of light with the constituents
of a material medium generates electric dipoles, which can be modelled as a mechanical
system analogous to a spring-mass configuration. If there are no damping forces, the
refractive index can be expressed as

Ne 2
n2 (Z )  1
H0 me (Z02  Z 2 )
Dispersion and Scattering of Light 293
• If fp denotes the fractional number of electrons per unit volume, whose resonant frequency
is wp, and damping present in the system can be modelled using Stokes’ law, the expression
for refractive index can be expressed as

Ne2 1
n2 1
me H 0 Çf
p
p
(Z 2p  Z 2  2ibpZ )

• The scattering of light by small particles whose linear dimensions are considerably smaller
than the wavelength of incident light is known as Rayleigh scattering. The intensity of
light scattered by small particles is directly proportional to the fourth power of the driving
frequency or inversely proportional to the fourth power of wavelength.
• When a beam of coloured light is incident on the surface of a liquid, it is scattered
inelastically by its molecules and distinct lines appear around the incident frequency in the
energy spectrum. This is known as Raman Effect. The nature of scattered light depends
on the type of sample present. It has been put to diverse uses in research and development.

REVIEW EXERCISES

1. Shallow water is said to be non-dispersive where the waves travel at a speed which is
proportional to the square root of the depth. In deep water, the propagation of wave cannot
be felt all the way down to the bottom. The behaviour in deep water is as though the depth
is proportional to the wavelength. As a matter of fact, the distinction between ‘shallow’
and ‘deep’ depends on the wavelength. If the depth is less than the wavelength, the water
is shallow and if it is much greater than wavelength, the water is deep. Show that the wave
velocity of water waves in deep water is twice the group velocity.
2. The table below gives a set of corresponding values of n and l for pure silica:

l ´ 106 (m) n
0.70 1.45561
0.75 1.45456
0.80 1.45364
0.85 1.45282
0.90 1.45208
0.95 1.45139
1.00 1.45075
1.05 1.45013
1.10 1.44954
1.15 1.44896

Plot n versus l and estimate the values of A and B using Eq. (7.4).
[Ans. A = 1.44508, B = 4.67 ´ 10–15 m2]
294 Wave Optics

3. Using Eq. (7.26), calculate the width of the anomalous dispersion region for the case of
a single resonance at w = w0 by assuming that b = w0.
With reference to the values of absorption coefficient, where does the refractive index
assume maximum and minimum values? [Ans. 2b; (w0 – b), (w0 + b)]
4. Light of frequency n is incident on a sample of molecules of moment of inertia, I. Assume
that only rotational transitions are tenable and determine the frequencies that would be
present in the scattered light in terms of J and I, where rotational energy is given by
=2
E J ( J  1)
2I
It is given that the selection rule, DJ = ±2, applies for rotational transitions in the Raman
spectrum.
Ë h h Û
Ì Ans. Q  (2 J  3) 2S I , Q  (2 J  1) 2S I Ü
Í Ý
5. For air at STP, n = 1.000278 in the entire visible spectrum. Calculate the attenuation
coefficient in km for l = 410 nm, 589.3 nm, 648 nm and comment on the values so
obtained. Take N = 2.69 ´ 1025 molecules m–3.
[Ans. 29.75 km, 127.14 km, 187.93 km]
P ART III
MODERN OPTICS
CH A P T E R 8
LASERS AND THEIR APPLICATIONS

EXPECTED LEARNING OUTCOMES


After reading this chapter, the student will acquire the capability to:
• Relate the temporal coherence to width of a spectral line and the spatial coherence to visibility
of fringes.
• Derive the expressions for Einstein’s A and B coefficients.
• Explain the principle of lasing action and discuss different ways of pumping.
• Describe the different types of lasers and discuss their important applications.
• Solve numerical problems.

8.1 INTRODUCTION

We now know that light is part of electromagnetic spectrum in the visible region and exhibits
phenomena such as interference, diffraction and polarisation. These lend themselves more
naturally to wave nature of light. While discussing Young’s double slit experiment in Chapter 3,
we highlighted that to obtain interference pattern, the sources of light must be coherent.
That is, two sources must have the same frequency and the phases of waves emanating from
these must bear a constant relationship in time. In general, the phase relationship can be in
space or time. This gives rise to spatial or temporal coherence. The temporal coherence refers
to the phase relationship at different times at a point in space, whereas spatial coherence relates
to phase relationship at different points in space at a particular time. In Section 8.2, you will
learn about temporal coherence and its effect on the width of spectral lines (spectral purity).
You may recall that for temporal coherence, Young used an ingenious method to divide a
single wavefront into two and these behaved as if they originated from two sources having a
fixed phase relationship. When such waves were allowed to superpose, a stationary interference
pattern was obtained. Section 8.3 is devoted to the discussion of spatial coherence and its
correlation to the visibility of interference fringes. Michelson’s ingenious method wherein he
used the concept of spatial coherence to measure the angular diameter of distant stars is also
discussed here.
297
298 Wave Optics

You may now ask: Is there a way to obtain a source with high degree of coherence? In
fact, this quest led to the invention of a LASER, which is an acronym for Light Amplification
by Stimulated Emission of Radiation. Because of its high degree of coherence, lasers find
diverse applications such as in basic research, communications, defense, holography, industry,
space and surgery, among others. (The now supposedly abandoned Star War programme was
also based on the use of highly directed laser beams.) The basic principle of laser involves
stimulated emission of radiation. It was predicted by Einstein in 1917 when he gave the theory
of stimulated emission and realised in 1954 by C.H. Townes and his co-workers for microwaves.
Despite the pioneering work of Townes, the first successful operation of a laser was demonstrated
by Maiman in 1960 using a ruby crystal.
Einstein had argued that an atom in an excited energy state can make transition to a lower
energy state either by spontaneous emission or by stimulated emission of radiation. You will learn
about these in Section 8.4 and derive expressions for Einstein’s A and B coefficients. (As such,
Laser is outgrowth of MASER, which stands for Microwave Amplification by Stimulated Emission
of Radiation and uses microwaves instead of light waves in the visible region. The first maser was
developed by Gordon J.P. and Townes C.H. and his associates at Columbia University in the period
1951–54. It is also reported in literature that A. Prokhorov and N.G. Basov independently invented
MASER at Lebedev Laboratories, Moscow. Prokhorov shared 1964 Nobel Prize with Townes.)
The main components of a laser are: The active medium, the pumping source and an
optical resonator. The active medium consists of atoms, molecules or ions, which can amplify
light waves. In general, the number of atoms in lower energy states is more than in excited
states. As a result, an electromagnetic wave passing through such an assembly is attenuated.
Therefore, to achieve amplification, the active medium has to be kept in a state of population
inversion, i.e. in a non-equilibrium state wherein the number of atoms in a higher energy level
is more than the number of atoms in the lower energy level. For a pair of energy levels, such
a state of population inversion is achieved by pumping mechanism. In Section 8.5, we will
learn the physical principles involved in the operation of a laser.
Since the time ruby laser was demonstrated by Maiman in 1960, lasers have covered
unchartered territories and their spectacular applications have propelled use of different types
of active media—solid, liquid, gas or ions. Even semiconductor based lasers have been designed
now. Some of the characteristics which are responsible for so many and so varied applications
of lasers are mono-chromaticity (infinitely small spectral line width), coherence, high
directionality (low spread of beam as it propagates) and intensity (brightness). Though detailed
discussion of various applications of lasers is beyond the scope of this book, typical applications
in basic research, communications, healthcare and industry are discussed in Section 8.6.
You will realise that the developments associated with lasers put the field of optics at the
forefront of basic research in the second-half of the 20th century.

8.2 TEMPORAL COHERENCE

In the study of interference and diffraction phenomena, we assumed that light waves are perfectly
sinusoidal at all times (–¥ < t < ¥) and the associated electric field could be represented as
E = E0 cos(kx – wt + f). (8.1)
Lasers and Their Applications 299
But this is an idealised situation. You may now ask: What then is the actual situation?
To visualize the actual situation, we reconsider the basic mechanism of emission of light by
an ordinary source. We know that it involves atoms, each of which radiates electromagnetic
waves in the form of photons for a very short time (» 10–9 s or less). Though each wave is
sinusoidal in itself, the collection of wave trains is not sinusoidal. The average time for which
the electric field exhibits sinusoidal behaviour is known as coherence time. We denote it by
tc. Physically, it means that at a given point, the electric fields at times t and t + Dt will, in
general, exhibit a definite phase relationship, if Dt = tc. It may be mentioned here that for red
cadmium line (l = 643.8 nm), the coherence time is about 10–9 s and for neon (l = 632.8 nm),
the coherence time is about 10–10 s.
Note that the finite value of tc can arise due to
• finite lifetime of an excited atom in the higher energy level from which it radiates and
goes to a lower energy state;
• random motion of atoms; and
• collision of the radiating atom with another atom.
The path length corresponding to tc defines coherence length Lc and is given by
Lc = ctc (8.2)
where c is the speed of light.
For the motion of electrons in a solid, Lc and tc are analogous to electron mean free path
and relaxation time, respectively. You can now easily estimate the coherence lengths
corresponding to red cadmium line and the neon line; these will be 0.3 m and 0.03 m,
respectively. For commercially available lasers, the coherence length is significantly greater
than that for ordinary light sources. For example, the coherence length for commercial He-Ne
laser is about 15 m, which corresponds to coherence time of about 50 ns.
To understand the concept of temporal coherence (or coherence time), we reconsider
Young’s double slit experiment as well as Michelson’s interferometer experiment. Refer to
Figure 8.1. It shows Young’s experimental arrangement, which we have reproduced for
completeness. You will recall that interference pattern is obtained when waves from two
coherent sources (S1 and S2) arrive at a point, say P, in different phases and travel different
distances. Alternatively, we can say that interference fringes observed around a point at time
r r
t arise due to superposition of waves originating at S1 and S2 at times t – 1 and t – 2 ,
c c
r2  r1
respectively. Here r1 and r2 denote the distances S1P and S2P. If = tc, the waves
c
arriving at the point of observation from the coherent sources exhibit definite phase relationship

Figure 8.1 Young’s double slit arrangement with a point source to obtain interference pattern.
300 Wave Optics

and are said to be temporally-coherent. These temporally-coherent waves give rise to good
quality interference fringes. On the other hand, if waves arriving at the point of observation
are not temporally-coherent, we will not obtain good quality interference pattern. We can
generalize this statement and say that the central fringe will have excellent contrast but as we
move away from it towards higher order fringes, the contrast will gradually degenerate.
Now refer to Figure 8.2, which shows experimental arrangement used in Michelson
interferometer. S is a neon source which emits light of wavelength 632.8 nm. To make sure
that all other wavelengths are filtered out, an optical filter F is placed in front of the neon lamp.
The light falling on the plate P1 is split into two parts of equal intensity through partial
reflection and partial transmission. These beams travel along two mutually perpendicular paths.
The reflected wave undergoes second reflection at M2. This wave is partially transmitted
through P1. To equalise path difference between waves reflected by M1 and M2, a compensating
plate P2 is introduced between P1 and M1. The transmitted wave undergoes reflection at M1 and
is partially reflected by P1 towards the eye (telescope).

Figure 8.2 A schematic diagram of Michelson interferometer. S is an extended source, mirror M1 is


fixed and mirror M2 is moveable, P1 is a beam splitter and P2 is compensating plate.

From Chapter 4, you may recall that nature of fringes formed in a Michelson Interferometer
depends on the inclination of mirrors M1 and M2. When mirrors M1 and M2 are perfectly
perpendicular to each other and the path difference P1M2 – P1M1 is small, circular fringes are
formed due to superposition of waves reflected by these mirrors. If mirror M2 is moved away
from plate P1, the contrast becomes poorer and eventually, the fringe pattern disappears. We
can understand disappearance of fringes in terms of temporal coherence of light waves emitted
by the neon source. When mirror M2 is moved through a distance d, the beam reflected by it
travels an additional distance equal to 2d. It means that the wave reflected by mirror M2
interferes with the wave reflected by mirror M1 that had originated from the source (2d/c)
seconds earlier. So if this time difference in the emission of waves is greater than the coherence
time, the waves reaching the eye do not exhibit any definite phase correlation. That is, the
waves reflected by the mirrors M1 and M2 are not temporally-coherent i.e., they are incoherent.
On the other hand, if time difference (2d/c) is much less than the coherence time tc, the waves
Lasers and Their Applications 301
reaching the eye after reflection from mirrors M1 and M2 can be regarded as temporally
coherent and exhibit a definite phase relationship leading to the formation of a well defined
interference pattern.
If the source used in the Michelson interferometer experiment is replaced by a commercial
laser, the interference fringes can be seen even for a path difference of a few metres.

8.2.1 Spectral Linewidth


We now know that emission of spectral lines owe their origin to transition of an electron from
an excited (higher energy) state to a lower energy (ground) state in an atom. The frequency
(and hence wavelength) of the spectral line is determined by the difference in energies of the
energy states within which transition takes place. You may recall that each element/material
has its characteristic energy states and hence spread of spectral lines. In your Thermal Physics
course, you have learnt that spread of energies of atoms in a substance leads to finite width
of spectral lines. (In fact, this constituted indirect evidence in favour of Maxwell’s law for
distribution of velocities.) That is, the emission (as well as absorption) of spectral lines are not
sharp geometrical lines but exhibit a spread of frequencies in a very narrow interval. For
example, the width of red cadmium line is about 7 ´ 10–4 nm.
A logical question to ask now is: What determines the width of a spectral line? Can we
correlate it to temporal coherence? To know answers to these questions, let us revisit Michelson
interferometer experiment and examine if we can interpret loss of contrast of interference
fringes as being due to emission of frequencies over a narrow band. For zero or very small
optical path difference between the interfering waves, the different wavelengths emitted by a
source form different fringe systems but these are superimposed over each other leading to
good contrast. On the other hand, when the optical path difference is large, slightly laterally
displaced fringe systems are formed by different wavelengths leading to poor fringe contrast.
In other words, we can say that poor fringe visibility for large optical path differences arise
due to lack of spectral purity or non-monochromaticity of the source of light.
To appreciate the equivalence of these approaches, let us consider that a sodium lamp, which
emits two closely spaced spectral lines (l1 = 589.6 nm and l2 = 589 nm), illuminates the beam
splitting plate P1 in Michelson interferometer experiment. By appropriately positioning mirror
M2 with respect to plate P1, we can make the bright fringe corresponding to l1 coincide with
the dark fringe corresponding to l2 so that the fringe pattern will disappear. Then we can write
2d = ml1 (bright fringe) (8.3a)
È 1Ø
and 2d = É m  Ù O2 (dark fringe) (8.3b)
Ê 2Ú
where m is an integer.
We rearrange Eqs. (8.3a) and (8.3b) as
2d
m=
O1
2d 1
and m= 
O2 2
302 Wave Optics

Note that 2d denotes the path difference between the waves reflected by mirrors M1 and M2.
On equating these values of m, the condition for disappearance of the interference pattern
can be written as
2d 1 2d
 =
O2 2 O1
or
2d 2d 1
 =
O2 O1 2
In terms of the path difference, the condition for disappearance of the interference pattern can
be written as
O1O2 O2 O2
2d  (8.4)
2( O1  O2 ) 2( O1  O2 ) 2'O
where l is the average wavelength.
In practice, the light beam may consist of a large number of wavelengths between l and
l + Dl. If so, we divide this interval in two equal parts of width Dl/2 each. It means that for
disappearance of interference pattern produced by waves of wavelength l1 = l + (Dl/2) and
l2 = l, we must have
O2 O2
2d (8.5)
È1 Ø 'O
2 É 'O Ù
Ê2 Ú
'O
Note that for each wavelength lying between l and l + , there will be a corresponding
2
'O
wavelength between l + and l + Dl such that the minimum of one falls on the maximum
2
of the other. This makes the fringe pattern to disappear. It means that the contrast of interference
fringes will be dismally poor for
O2
2d • (8.6a)
'O
From this equation we note that as the spread of wavelengths becomes smaller and smaller,
i.e. the source tends to be monochromatic, the path difference for disappearance of fringes
becomes larger and larger. And larger the path difference for which fringes do not disappear,
more temporally coherent the light waves will be. That is, monochromaticity is strongly related
to its temporal coherence.
We rewrite this condition for the spread of wavelengths as

O2
'O • (8.6b)
2d
This result shows that the contrast of interference fringes becomes poor when the spectral
width of the source is greater than or equal to l2/2d.
Lasers and Their Applications 303
You now know that when path difference 2d is more than the coherence length Lc, no fringe is
observed. So we can say that the spectral width of the source is related to coherence length
as
O2
'O (8.7a)
Lc
Using Eq. (8.2), we can rewrite as
O2
'O (8.7b)
cW c
This result shows that wavelength spread is inversely proportional to coherence time.
Since n = c/l, the corresponding spread in frequencies emitted by a source is given by
c c 1
'Q 'O (8.8)
O 2
Lc Wc
That is, the frequency spread is reciprocal of coherence time.
To get a feel for the numbers, go through the following example carefully.

EXAMPLE 8.1
(a) The wavelength of red cadmium line is 643.8 nm. Calculate the spectral spread of the
source. Take coherence time as 10–9 s and c = 3 ´ 108 m s–1.
(b) If a continuous beam of wavelength given in (a) above is chopped into 0.05 ns pulses
using a shutter, calculate the coherence length, bandwidth and resultant linewidth.
Solution
(a) From Eq. (8.7b), we recall that spectral spread of a source is given by

O2
'O
cW c
On substituting the given values, we get

(643.8 – 109 m) 2
'O 1.38 – 10 3 nm
(3 – 108 m s 1 ) – (109 s)

The frequency of red cadmium lines is given by

c 3 – 108 m s 1
Q 4.66 – 1014 Hz
O (643.8 – 10 9 m)
From Eq. (8.8), we know that frequency spread is reciprocal of coherence time. Hence,
the monochromaticity or spectral purity is

'Q 109 s 1
2.14 – 106
Q (4.66 – 1014 s 1 )
304 Wave Optics

It shows that even for an ordinary source of light, spectral purity is very small.
For a commercial laser
'Q
4 – 108
Q
(b) Since Dt = 0.05 ´ 10–9 s, coherence length

Lc c 't (3 – 108 m s 1 ) – (0.05 – 109 s) 1.5 – 102 m


1 1
Bandwidth = 2 – 1010 Hz
't 0.05 – 10 9 s

O2 (643.8 – 10 9 m) 2
Linewidth = 1 9
0.026 – 10 9 m
c 't (3 – 10 m s ) – (0.05 – 10 s)
8

Now you should answer a Practice Exercise.

Practice Exercise 8.1 Calculate spectral spread and monochromaticity for


(a) sodium light characterized by l = 589 nm and tc = 10–10 s, and
(b) orange krypton line defined by l = 605.8 nm and coherence length of 0.20 m.

'Q
[Ans. (a) 'O 1.16 – 102 nm; 2 – 105 ;
Q
'Q
(b) 'O 1.83 – 10 3 nm; 3.03 – 10 6 ]
Q

Let us now pause for a while and reflect as to what we have achieved so far. We now know
that temporal coherence relates constancy of phases of waves (fields) emitted by a point source
and traversing different optical paths in space. We also discovered that these waves produced
good quality interference pattern only if they were of the same frequency and the path difference
was less than coherence length. Now you may ask: What happens when waves originate from
an extended source? From your previous knowledge, you may conclude that good contrast
interference pattern is not obtained with an extended source. But you may not be fully justified.
We now discuss the effect of finite size of the source on the interference pattern.

8.3 SPATIAL COHERENCE

Refer to Figure 8.3(a), which depicts Young’s double slit experiment with an extended source
SS ¢. Let us assume that the width of the source is h and it is located at a distance r from the
plane containing the slits. Suppose that light from a point s in the source illuminates the slits.
If this was the only luminous spot in the source, we would obtain good contrast interference
Lasers and Their Applications 305

Figure 8.3 (a) Young’s double slit experiment with an extended source SS¢;
(b) Angle subtended by the extended source at the slits.

fringes on the observation screen. Do you know the reason? This is because the arrangement
will essentially act as an idealized Young’s experiment. However, with an extended source,
which can be assumed to be made up of infinite point sources, we obtain mutually displaced
interference patterns due to each point. This spread in fringes leads to reduction in
fringe visibility.
To quantify our argument, let us assume that the extreme points S and S¢ of the extended
source act as independent sources. These point sources will produce their own interference
patterns. We assume that slits S1 and S2 are located symmetrically with respect to S as well
as the point O on the observation screen so that SS1 = SS2 and S1O = S2O. It means that point
O will be the point of maximum due to source S. But intensity at O due to S ¢ will depend on
the path difference S ¢S2 – S ¢S1. Now let us assume that
O
S ¢S2 – S ¢S1 = (2n  1) n 0, 1, 2, ... (8.9)
2
The minima of the interference pattern due to S will coincide with the maxima of the interference
pattern due to S ¢. This will lead to zero fringe visibility, i.e. interference pattern will not be
observed. To relate this result to the dimensions of the extended source, refer to Figure 8.3(a)
again and note that
S ¢S2 – S ¢S1 = S2P = S1S2 tana ; ad
From Ds QO¢S2 and S¢SQ, we note that
d/2 h
tana ; a =
r2 r1
2h
Þ r1 = r2
d
Èd Ø 2r2
and r1 + r2 = É  hÙ
Ê2 Ú d
Èd ØÈ 1Ø
or r= É  hÙ É
Ê2 ÚÊ D ÙÚ
È dØ d
Hence ad = ÉÊ h  Ù
2Ú r
306 Wave Optics

If we ignore terms of the order of d 2, we get


hd
S¢S2 – S¢S2 =
r
Thus fringes will disappear, if
O hd
S „S2  S „S1 (2n  1)
2 r
For the lowest order, we must have
hd O
=
r 2
or
Or
h= (8.10)
2d
This result shows that if the linear dimensions of an extended source are nearly equal to
lr/d, no interference pattern will be observed. Alternatively, we can say that for an extended
source whose linear dimension is »lr/d, there is a point at a distance lr/2d which produces
fringes which are shifted by half a fringe width. Therefore, the fringe pattern will be obscured.
It means that for an incoherent extended source, interference fringes of good contrast will be
obtained only if
Or
h  (8.11)
d
Now refer to Figure 8.3(b). If the source subtends an angle q at the mid-point of the slits S1
and S2, we can take q » h/r. Then, condition for getting good contrast fringes can also be
expressed as
O
d  (8.12)
T
On the other hand, fringes of very poor contrast will be obtained if
O
d (8.13)
T
The relation
O
Aw (8.14)
T
defines the maximum lateral distance between slits S1 and S2 at which light waves from the
extended source will exhibit some degree of spatial coherence. That is, if Eq. (8.14) is satisfied,
the light waves from an extended source, after passing through the slits S1 and S2, will combine
and produce interference pattern, which has distinct fringes with high visibility. The quantity
l/q is known as lateral coherence width.
Note that coherence width has linear dimension and is almost perpendicular to the direction
of wave propagation. On the other hand, the coherence length defined in relation to temporal
coherence is along the direction of wave propagation. For this reason, temporal coherence is
also referred to as longitudinal coherence and spatial coherence is said to be transverse
coherence.
Lasers and Their Applications 307
Note that Eqs. (8.13) and (8.14) hold only when the extended source is linear. If the source
is in the form of a uniform circular disc, the critical value of lateral coherence at which fringes
disappear changes to
O
Aw 1.22 (8.15)
T
The genesis of the factor 1.22 was discussed while deriving Eq. (6.18).
To help you appreciate the concepts developed here, we now give a simple numerical
example.

EXAMPLE 8.2 In Young’s double slit experiment, light from He-Ne laser (l = 632.8 nm)
is made to pass through a small circular hole of diameter 0.15 mm. If the distance between the
source and the slit planes is 1.5 m, calculate the separation between slits at which fringe pattern
will disappear.
Solution We know that diameter of the hole is 0.15 mm = 1.5 ´ 10–4 m and distance between
the source and slit planes is 1.5 m. Hence, the angle subtended by the source on the slits

h 1.5 – 104 m
T 104 rad
r 1.5 m
Since l = 632.8 nm, the lateral coherent width for a circular extended source is

1.22O 1.22 – (632.8 – 109 m)


Aw 7.72 – 103 m
T 104
Thus, interference fringe pattern will not be visible if the slits are separated by more than
7.7 mm. In other words, if the separation between the fringes is less than 7.7 mm, good contrast
interference fringes will be observed.
You may now like to answer a Practice Exercise.

Practice Exercise 8.2 In Young’s double slit experiment, the distance between the source
and the slit planes is 1.0 m. Calculate the angular diameter of the hole in source plane which
will produce a good contrast interference fringes on the observation screen. Take l = 589 nm
and lateral coherent width as 7.186 ´ 10–3 m. [Ans. 10–4 m]

8.3.1 Visibility of Fringes


In the preceding sections, we have discussed the importance of coherence for observing
interference pattern. In particular, as the slits are moved apart, the fringes move closer and may
ultimately merge together with complete loss of visibility. This signifies that degree of visibility
is a measure of spatial coherence.
Consider two sources which produce waves of the same frequency. When such waves are
in-phase and made to superpose, we obtain good contrast interference fringes with well defined
maxima and minima. As a result, the visibility is of the high quality. You may now ask: Do
we obtain interference fringes even when such waves overlap partially? The visibility will drop
308 Wave Optics

significantly. That is, the degree of visibility of fringes is determined by the extent of overlap.
You may now ask a logical question: How much visibility do we need to observe an interference
pattern? Let us now discover answer to this question.
To facilitate the discussion of visibility, we define the term arena as the amount of radiated
power incident on a surface per unit area. The ratio of the difference between maximum arena,
Imax, and minimum arena, Imin to the sum of maximum and minimum arenas is referred to as
visibility (V). Mathematically, we write
I max  I min
V (8.16)
I max  I min
Note that visibility essentially facilitates comparison. Let us now assume that Imax takes an
arbitrary value but Imin = 0. Then visibility will be maximum, equal to one and the fringes will
exhibit perfect contrast. But if Imax = Imin, the visibility will be zero implying that we cannot
observe interference pattern (fringes). In general, for ordinary light sources, V = 0.8 corresponds
to ‘almost coherent’ light, whereas V = 0.2 defines incoherent component and fringe pattern
from such a light is barely visible.
We now pause for a minute and ask: Is visibility related to coherence? If so, how? To
discover answer to this question, we recall that light waves emitted by an ordinary source consist
of coherent as well as incoherent components and interference pattern is obtained due to the
coherent component. The arena due to coherent and incoherent components can be expressed as
( I A1 )coh ] I0
and
( I A2 )incoh (1  ] ) I 0
where z denotes the degree of coherence.
From Chapter 3, you will recall that energy is redistributed in interference and intensity at
the maximum is four times the contribution from individual waves, whereas intensity at the
minimum is zero. In terms of arena, we can write
I Amax
1
= 4zI0
and
I Amin = 0
Let us now consider two points on a distant screen illuminated by two ordinary light sources,
which produce equal arenas I0. In the interference pattern, uniform distribution due to incoherent
part superimposes over coherent part. The arena due to incoherent component will be twice as
high as the contribution I A2 because it comes from two sources. Hence

I Amax
2
2 I A2 2(1  ] ) I0 I Amin
2

The arena in the maxima is given by

I max = I Amax
1
 I Amax
2

= 4] I 0  2(1  ] ) I 0 2(1  ] ) I 0
Lasers and Their Applications 309
Similarly, the arena in the minima is given by

I min = I Amin
1
 I Amin
2
= 0  2(1  ] )I 0
= 2(1 – z)I0
On substituting these values in Eq. (8.16) and simplifying the resultant expression, we get
V=z (8.17)
This result shows that the degree of visibility or contrast of fringes produced by two light
waves equals the degree of coherence between them. The highest visibility and hence the
highest degree of coherence occurs for zero minimum arena.
While discussing spatial coherence, we considered an extended source and discovered that
for a critical value of lateral coherence length, interference fringes disappear. This concept was
used by Michelson to design an ingenious method to measure angular diameter of stars. It is
based on the result that for a distant star, which can be treated as a circular source, the interference
fringes will disappear if the distance between the pinholes (slits) S1 and S2 is given by
O
d 1.22
T
Here q is the angle subtended by the circular source on the slits. If a star whose angular
diameter is about 10–7 rad emits light of wavelength l = 600 nm, the fringes will disappear for

1.22 – (600 – 109 m)


d 7.32 m
107
For such a large value of lateral coherence width, fringes will be extremely close. Moreover,
we will need a very large lens, which also presents manufacturing difficulties. In the design
of stellar interferometer, Michelson circumvented these difficulties by using movable mirrors.
Let us learn about it now.

8.3.2 Spatial Coherence and Angular Diameter of Stars:


Michelson Stellar Interferometer

Consider the arrangement shown in Figure 8.4. It shows Michelson’s stellar interferometer. M1,
M2, M3 and M4 is a symmetrical system of mirrors mounted on a rigid support in front of the
telescope. The mirrors M1 and M2 are movable and can be separated out symmetrically normal
to the lengths of the slits S1 and S2. On the other hand, mirrors M3 and M4 are fixed. Light from
a distant star reaches the slits after reflection from the system of these mirrors and converges
in the focal plane PP ¢ of the telescope. The paths O1M1M3S1P1 and O1M2M4S2P1 traversed by
light from one edge of a star are shown by solid lines. These will form interference fringes with
the angular separation equal to l/d with the central fringe at P1. Light from the other edge of
the star travels along O2M1M3S1P2 and O2M2M4S2P2. These are shown by dashed lines and
give rise to a similar fringe pattern with the central fringe at P2.
Since optical path lengths M1M3S1 and M2M4S2 are equal, the light from two ends of the
star reaches the slits S1 and S2 with the same path difference as at M1 and M2. So when the path
310 Wave Optics

Figure 8.4 Schematic representation of Michelson stellar interferometer.

difference at M1 and M2 is an integral multiple of wavelength l, the fringe pattern in the focal
plane will shift through a distance equal to one fringe width and the resultant intensity pattern
will show uniform intensity. That is, the fringes will disappear.
Note that light travelling along O1M1 and O1M2 arrives in phase at mirrors M1 and M2.
However, light travelling along O2M1 and O2M2 is inclined to the axis of the objective lens at
an angle q so that
O
T
D
where D denotes the separation of the mirrors M1 and M2.
This result shows that stellar interferometer magnifies the resolving power of the telescope
in the ratio D/d, where d is diameter of the objective of the telescope.
You may recall that if an object is in the form of a circular disc, the fringes disappear when
q = 1.22 l/D. This implies that for a circular star disc, the mirrors M1 and M2 will have to be
moved outward somewhat.
It is interesting to point out here that Michelson stellar interferometer was first used
in Wilson laboratory to measure the angular size of Betelgeuse in a-Orion constellation.
The separtion between the slits was about 114 cm and the distance between the mirrors was
increased up to 6.1 m. The value of angular size was measured as 0.047¢¢. Its distance from
the earth was measured using parallax method and then linear diameter was estimated to be
4.1 ´ 108 km, which is nearly 300 times the diameter of the sun.
The main drawback of the Michelson stellar interferometer lies in its instability to vibration
and thermal changes. It is extremely difficult to maintain the path difference for large mirror
separations to a fraction of wavelength. The mirrors can be separated by a few metres only.
Brown and Twiss modified the Michelson stellar interferometer and obtained much larger
separation between the slits using two photo-detectors to receive light from the distant star.
You should now answer a Practice Exercise.
Practice Exercise 8.3 The sun subtends an angle of 0.007 seconds of an arc. If the light
of wavelength 550 nm is used, calculate the separation between outer mirrors for disappearance
of fringes. [Ans. 20 m]
Lasers and Their Applications 311
So far we have considered temporal and spatial coherence and discussed their implications
for spectral line width and visibility, respectively. Now we study Einstein’s formulation of
spontaneous and stimulated emission of radiation, which proved vital for realisation of a laser.

8.4 SPONTANEOUS AND STIMULATED EMISSION OF


RADIATION: EINSTEIN’S FORMULATION
In atomic physics, you have learnt Bohr’s theory of hydrogen spectrum. It stipulates that
• in an atom, electrons revolve around the nucleus in stationary states having discrete
energy levels;
• the transition of an electron from one energy level to another level occurs in quantum
jumps; and
• frequency of radiation emitted by an atom when an electron makes a transition from an
excited state to a lower energy state is given by
hn = Ei – Ef (8.18)
where h is Planck’s constant, n is frequency of emitted radiation and Ei and Ef respectively
denote energies of higher (excited) and lower (ground) states. On the basis of these
assumptions, Bohr successfully explained the observed spectral series emitted by a
hydrogen atom.
To understand the physics of atomic processes involved in emission and absorption of
light, we consider a two energy level system, as shown in the Figure 8.5(a). Let the energy of
the lower and upper energy levels be E1 and E2 respectively. An atom in the lower energy level
can absorb a photon of energy hn (= E2 – E1) and get excited to the higher energy level E2.
This is shown as absorption or excitation in Figure 8.5(b).

Figure 8.5 Absorption and spontaneous emission of a photon in a two energy level system.

On the other hand, an atom in higher energy level will tend to emit energy and attain lower
energy level. Such a transition is accompanied by spontaneous emission of photon of energy hn and
is termed spontaneous emission [Figure 8.5(c)]. The frequency of the emitted photon is given by
E2  E1
Q (8.19)
h
Note that due to spontaneous transition, photons are emitted randomly, i.e. there is no definite
phase relationship amongst them. The emitted light is incoherent and has a broad spectrum.
This is the basic mechanism of emission of light by ordinary sources.
312 Wave Optics

Next we consider an atom in an excited state. It is irradiated by a photon of energy hn.


Instead of being absorbed, this photon may trigger the atom to de-excite and emit radiation.
As illustrated in Figure 8.6, two photons, each of energy hn, will be emitted simultaneously.
Such a transition is said to be stimulated or induced.

Figure 8.6 Stimulated emission of radiation.

Note that the inducing as well as induced photon has the same energy and both are in
phase, i.e. these are coherent as well as co-directional and therefore add to amplify the incident
beam. These unidirectional photons can de-excite two other atoms in their path producing four
and this cascading process of photon multiplication will continue. It means that when a large
number of atoms are involved, the stimulated emission generates an intense, highly coherent
and directional beam. In the optical range this process is facilitated by a laser. (The device
which facilitates stimulated emission in the microwave range is known as maser.) Note that
stimulated emission is the fundamental process responsible for the development of a laser.
Before proceeding further, we give a simple numerical example.

EXAMPLE 8.3 A sodium lamp emits a spectral line of wavelength 590 nm. Consider it as
a two level system so that the spectral line corresponds to a transition from the first excited
state (3p) to the ground state (3s). Calculate the corresponding energy.
Solution From Eq. (8.18), we recall that

hc È cØ
E2  E1 hQ É since Q Ù
O Ê OÚ
On substituting the values of h(= 6.6 ´ 10–34 J s), c(= 3 ´ 108 m s–1) and l = 590 ´ 10–9 m,
we get
(6.6 – 10 34 J s) – (3 – 108 m s 1 )
E2  E1 3.4 – 1019 J = 2.1 eV
590 – 109 m
This corresponds to the energy of the first excited state of sodium.

Suppose that the number of atoms in the upper and lower non-degenerate energy levels is
N2 and N1 respectively. The absorption of photons by atoms in the lower energy level and
emission of photons in the upper energy level is known to be proportional to the intensity of
incident light. In your Thermal Physics course, you have learnt that if such an assembly of
atoms is in thermal equilibrium at temperature T and are assumed to obey Maxwell–Boltzmann
distribution law, we can write
Lasers and Their Applications 313

È  E1 Ø
N1 — exp É Ù
Ê kBT Ú
and
È  E2 Ø
N2 — exp (8.20)
Ék TÙ
Ê B Ú

where kB is Boltzmann constant. Its value is 1.38 ´ 10–23 J K–1.


On combining these results, we can write

N2 È E2  E1 Ø È hQ Ø
exp É  exp É  (8.21)
N1 Ê k BT ÙÚ Ê k BT ÙÚ

For degenerate energy levels, this expression will modify as

N2 g2 È hQ Ø
exp É  Ù (8.21a)
N1 g1 Ê kBT Ú

You may now like to know: How will the ratio of the populations in energy levels change,
if the radiation of energy hn is introduced in the system? The answer to this important question
was provided by Einstein, who argued that for the given system of energy levels and the
radiation to be in thermal equilibrium, the rate of transitions due to spontaneous and stimulated
emissions (downward transitions) must be equal to the rate of transitions due to absorption
(upward transitions). Before we discuss the implications of this argument further, go through
the following example carefully.

EXAMPLE 8.4 For the sodium lamp considered in Example 8.3, calculate the fraction of
sodium atoms in the first excited state if the lamp is at 327°C. Assume it to be non-degenerate.
Solution In this case, we use Eq. (8.21):

N2 È hQ Ø È hc Ø
exp É  exp É 
N1 Ê k BT ÙÚ Ê k BT O ÙÚ

Here l = 590 nm, and T = 273 + 327 = 600 K. On substituting the numerical values
of physical constants, we get

N2 È (6.6 – 1034 J s) – (3 – 108 m s 1 ) Ø


= exp É  23 1 9 Ù
N1 Ê (1.38 – 10 J K ) – (590 – 10 m) – (600 K) Ú
= exp (–40.53)
= 2.5 ´ 10–18
This result shows that ordinarily, the number of atoms in a higher energy state is extremely small.
314 Wave Optics

We now advise you to answer a Practice Exercise.

Practice Exercise 8.4 A two-level ruby laser emits light of wavelength 693 nm. If
temperature of the system is 27º C, calculate the ratio of populations in the two levels.
[Ans. 8.67 ´ 10–31]

Following Einstein, let us suppose that the energy density of the incident radiation of
frequency n is u(n). Then the rate of spontaneous emission will be proportional to N2; the
number of atoms in the higher energy state. However, the rate of stimulated emission will be
determined by the energy density of the incident radiation as well as the number of atoms in
the higher energy state. Hence, we can express the rate of spontaneous and stimulated
emissions as
P21 = A21 N2 + B21 N2 u(n) (8.22)
Here A21 and B21 respectively denote constants of proportionality for spontaneous and stimulated
emission of radiation from higher state (2) to lower state (1). Note that u(n) does not appear
in the first term on the right-hand side of Eq. (8.22). This is because no photon is needed for
spontaneous emission.
Similarly, we can write the expression for the rate of absorption as
P12 = B12 N1u(n) (8.23)
where B12 is the constant of proportionality for absorption of radiation.
If the lower and higher energy levels correspond to respective single non-degenerate states,
we take constant A21 = A, and B12 = B21 = B. These are known as Einstein’s A and B
coefficients. However, for generality, we will continue to work with A21, B12 and B21.
For equilibrium, we equate Eqs. (8.22) and (8.23) to obtain
N2 [A21 + B21 u(n)] = N1 B12 u(n)
or
N2 B12 u (Q )
(8.24)
N1 A21  B21u (Q )
On combining this result with Eq. (8.21), we get
B12 u (Q ) È hQ Ø
exp É 
A21  B21 u (Q ) Ê k BT ÙÚ
We now cross-multiply and rearrange terms to solve for energy density. This leads to the result

È A21 Ø 1
u (Q ) ÉB Ù (8.25)
Ê 12 Ú exp ( hQ /k BT )  B21/B12

In your Thermal Physics or Statistical Mechanics course, you will learn about Planck’s law for
black body radiation. Mathematically, we express it as
8S hQ 3 dQ
u (Q ) dQ (8.26)
c 3
exp (hQ /k BT )  1
Lasers and Their Applications 315
As may be noted, Eqs. (8.25) and (8.26) are strikingly similar, though these have been obtained
from completely different arguments. (This indicates the soundness of both thought processes.)
On comparing these expressions, we obtain
B21 = B12 (8.27a)
and
A21 8S hQ 3
= (8.27b)
B12 c3
This result shows that the proportionality coefficient for stimulated emission, B21, is inversely
proportional to the third power of frequency. (This suggests that laser action will become more
difficult at higher frequencies.)
For degenerate energy levels, these expressions respectively modify as
B21 g1
= (8.27c)
B12 g2
and
A21 g 8S hQ 3
= 1 (8.27d)
B12 g 2 c3
Note that we would not have obtained the expression for energy density similar to Planck’s law
without considering stimulated emission. In fact, Einstein predicted stimulated emission in
1917 but it was realised in 1954 when Townes and his co-workers developed a microwave
amplifier using ammonia. In 1958, Schawlow and Townes extended the concept to visible
region and Maiman succeeded in constructing a Ruby laser in 1960.
Let us pause for a minute and reflect as to what we have achieved so far. From Eq. (8.27a),
we note that the probabilities of absorption and stimulated emission are the same. It means that
when an atomic system is in thermal equilibrium, absorption and emission occur simultaneously.
As shown in Example 8.4, N2 < N1 in general. It means that from physics point of view, under
normal conditions, absorption dominates stimulated emission. However, if we devise a
mechanism to ensure that N2 > N1, stimulated emission may begin to dominate absorption.
Such a condition in an atomic system is known as population inversion and the system is said
to lase.
Proceeding further, we combine Eqs. (8.25) and (8.27a) to obtain

A21 È hQ Ø
exp É Ù
1 (8.28)
B21u (Q ) Ê k BT Ú

Do you recognize this result? Physically, it gives the ratio of the number of spontaneous
emissions to stimulated emissions for a system in thermal equilibrium. For hn = kBT, it will
suffice to retain only the linear term in the expansion of exponential function in Eq. (8.28).
Then it simplifies to

A21 hQ
(8.29)
B21u (Q ) k BT
316 Wave Optics

This result shows that for hn/kBT = 1, stimulated emission will dominate spontaneous emission.
You may recall that stimulated emission is coherent because forced atomic oscillations bear a
constant phase relation to the incoming radiation. The process of stimulated emission is central
to the operation of lasers. This, however, is very uncommon in nature.
It is clear that for hn/kBT ? 1, spontaneous emission will dominate stimulated emission.
This works well for electronic transitions in atoms and molecules as also in the case of
radiative transitions in nuclei. To appreciate this, go through the following solved example.

EXAMPLE 8.5 Consider a tungsten filament lamp. The thermionic emission is assumed to
take place at T = 1.2 ´ 103 K. Calculate the ratio of spontaneous emission to stimulated
emission. Take l = 550 nm, kB = 1.38 ´ 10–23 J K–1 and h = 6.67 ´ 10–34 J s. Assume the
energy levels to be non-degenerate.
Solution From Eq. (8.28) we recall that the ratio of spontaneous emissions to stimulated
emissions is given by
A21 È hQ Ø (i)
exp É 1
B21u (Q ) Ù
Ê k BT Ú
Since l = 550 nm, the corresponding frequency emitted by the tungsten lamp is given by
c 3 – 108 m s 1
Q 
5.45 – 1014 Hz
O 550 – 10 m 9

On inserting this value of frequency along with the required data in (i), we get
A21 È (6.67 – 10 34 J s) – (5.45 – 1014 s 1 ) Ø
= exp É Ù 1
B21u (Q ) Ê (1.38 – 1023 J K 1 ) – (103 K) Ú
= [exp (26.14)] – 1
= 2.25 ´ 1011
From this result, we can conclude that at optical frequencies, emission is predominantly due
to spontaneous transitions and that is why light from such sources is incoherent.

In the foregoing paragraphs, we have discussed the concept of stimulated emission which
constitutes the working principle of a laser through population inversion. You may now ask:
What are the prerequisites for operation of a laser? Do we need a medium with peculiar
characteristics or can lasing action be achieved for any medium? How is population inversion
achieved? We will discuss these and other related questions now.

8.5 CONSTRUCTING A LASER: THE PREREQUISITES

There are three basic prerequisites for constructing a laser:


• Active laser medium
• Laser pumping mechanism
• Laser cavity
We now discuss the roles and functions of each of these components in some detail.
Lasers and Their Applications 317

8.5.1 Active Laser Medium

The active medium consists of a collection of atoms, molecules or ions, which can increase the
intensity of light passing through it. An active medium is the core of a laser and can be in any
one state of matter—solid, liquid, or gas. As of now, thousands of materials are known to
support lasing. But the most commonly used materials are solid crystals—ruby, Nd: YAG,
silicate or phosphate glasses and plastics, gases—CO2, He-Ne, N2, Ar or metallic vapour,
liquid dyes and semiconductors—gallium arsenide (GaAs), indium gallium arsenide (InGaAs),
or gallium nitride (GaN). In Nd: YAG laser, the active medium is a rod of yttrium aluminium
garnate (YAG) containing neodymium (Nd) ions. In a dye laser, it is a solution of a fluorescent
dye in a solvent and in a He-Ne laser, it is a mixture of helium and neon gases.
You may like to know the characteristics of an active medium. The most general requirement
for an active medium is that it should have atoms whose electrons can be excited to a meta-
stable energy level. In other words, an atom should have at least three energy states; an upper
energy state in which atoms can be ‘pumped’, a lasing state and a lower energy state to which
they can return (accompanied by spontaneous emission). Also, it must support population
inversion between any two states. In actual practice, the number of atoms of a particular
species that take part in lasing action in an active medium is only a fraction of the whole.
(Such atoms are known as Active Centres and the remaining material acts as Host.) For example,
in He–Ne laser, neon is the active centre and helium acts as the host. It may be mentioned
here that the typical number of active species is 1025–1026 m–3 in solids and liquids and
1021–1023 m–3 in gases.
You may now like to know: How does an active medium amplify light beam when it passes
through it? To discover answer to this question, we discuss the phenomenon of population
inversion.

Population inversion
We know that natural tendency of every system is to go to the state of minimum energy, i.e.
ground state. As a result of this characteristic, under normal circumstances, the occupancy of
higher energy levels is significantly less than that of ground state (see Example 8.4). But we
know that for the lasing action to be initiated, more atoms/molecules should be in an excited
state than in a lower energy state. That is, the pre-condition for amplification of light intensity
by stimulated emission is that in a system, the number of atoms in an upper energy level
exceeds the number of atoms in a lower energy level. This so-called population inversion is
an artificial condition but fundamental for the working of a laser. The mechanism used to attain
population inversion is known as pumping. (A mechanical analogue of population inversion is
pumping of water from the ground floor to the top of a high rise, i.e. from a state of lower
potential energy to a state of higher potential energy.) It may be mentioned here that population
inversion cannot be achieved with just two levels. This is because the probabilities of absorption
and spontaneous emission are exactly equal for this case.
To understand why population inversion is necessary for operation of a laser, we consider a
collection of atoms (active medium) and study the rate of change of intensity of a beam of light
as it propagates through it. We choose the direction of propagation of the beam of light to be
along x -axis. Let us consider an active medium between two planes P1 and P2 having area
318 Wave Optics

of cross-section A perpendicular to x-axis at x and x + Dx, as shown in Figure 8.7. The volume
of the medium between these planes is ADx. Let N1(n)dn denote the number density of atoms
which can absorb radiation of frequency lying between n and n + dn. Hence, the rate of
stimulated absorptions, i.e. the number of atoms that absorb energy and move to higher energy
level (from E1 ® E2) per unit time, in the medium between planes P1 and P2 is given by
[N1(n )dnB12 u(n)] ADx (8.30)

Figure 8.7 A collimated beam of light of intensity In moving through


an active medium along the x-axis.

Since a photon of energy hn is absorbed in each transition, the rate of loss of energy from
the incident beam is
hn [N1(n)dnB12 u(n)] ADx. (8.31)
Suppose that N2(n)dn denotes the number density of atoms which undergo stimulated emission
and fall to a lower energy level. If the frequency of these photons lies between n and n + dn,
the rate of stimulated emission in volume element ADx will be [N2(n)dnB21u(n)] ADx. You will
agree that in each transition, the emitted photon of energy hn will reinforce the propagating
beam. Thus the rate at which incident beam gains energy is given by
hn [N2(n)dnB21u(n)] ADx (8.32)
The net amount of energy absorbed per unit time in volume element ADx is obtained by
subtracting Eq. (8.32) from Eq. (8.31). This gives
DEab = hn [N1(n)dnB12 u(n)] ADx – hn [N2(n)dv B21u(n)] ADx
Note that we have not considered radiation arising due to spontaneous emission. This is
essentially because photons emitted in spontaneous emission emerge in random directions and
as such, do not contribute much to the intensity of the incident beam.
To discover what happens to the intensity of the light beam as it propagates, let us denote
the intensity of the beam at the plane P1 by In (x). Then the total energy entering the volume
element ADx per unit time is obtained by multiplying the intensity incident on the plane P1 with
the area of the plane under consideration. This gives In (x) A for total energy incident on the
volume element. Similarly, if In (x + Dx) is the intensity in the plane P2, the rate at which
energy leaves the volume element under consideration can be expressed as
Lasers and Their Applications 319

IQ ( x  'x) A IQ ( x) A 
˜ IQ
˜x
A'x  " (8.33)

Hence, the net rate at which energy leaves the volume element is given by the second term on
the right-hand side of Eq. (8.33). This obviously is equal to the negative of the net energy
gained by the medium enclosed between planes P1 and P2. Hence
˜ IQ
A'x  hQ {[ N1 (Q ) dQ B12 u (Q )] A'x  [ N 2 (Q ) dQ B21 u (Q )] A'x} (8.34)
˜x
On simplification, we can write
˜ IQ
 hQ > B12 N1 (Q )  B21 N 2 (Q )@ u (Q )dQ (8.35)
˜x
We know that energy density u(n) dn and the intensity In are connected though the relation
In = cmu(n) dn
where cm (= c/µm) is velocity of the light in the active medium and mm is refractive index of
the medium (Note that this relation is analogous to J = rv. Here r denotes the number density
of the particles moving with velocity v and J denotes the number of particles crossing a unit
area normal to the direction of propagation per unit time.)
On using this result in Eq. (8.35), we can express the spatial rate of change of intensity
as
˜ IQ IQ
 hQ > B12 N1 (Q )  B21 N 2 (Q )@ (8.36)
˜x cm
We write B12 = B21 = B and cm = c/µm. Then on rearrangement, Eq. (8.36) takes the form

1 ˜ IQ hQPm
 [ N1 (Q )  N 2 (Q )] B (8.37)
IQ ˜ x c
If the light beam is propagating in an absorbing medium, the decrease in intensity will be
proportional to instantaneous intensity as well as the range over which loss takes place.
Mathematically, we can write
dIn = –bn In dx
or
dIQ
 EQ IQ (8.38)
dx
where bn is absorption coefficient and is assumed to be a function of frequency.
Integration over x leads to the result
In = I0 exp(–bn x) (8.39)
Here I0 is the initial intensity and x signifies the distance travelled by light in the active
medium. Do you recognize this equation? We come across an analogous equation in the study
320 Wave Optics

of free paths of gaseous molecules in thermal physics and radioactive decay in atomic and
nuclear physics. Equation (8.39) suggests that if bn > 0, the intensity of the beam will drop
exponentially and reduce to 1/e of its initial value over a distance of 1/bn.
By combining Eqs. (8.37) and (8.38), we can relate absorption coefficient with N1(n) and
N2(n):
hQP m
EQ [ N1 (Q )  N 2 (Q )] B (8.40)
c
It means that bn will be positive when the number of atoms in the lower energy state is greater
than the number of atoms in the higher energy state.
On combining Eqs. (8.39) and (8.40), we can relate intensity of light beam with the number
of particles in different energy levels:

IQ ^
I 0 exp 
BhQP m
c `
[ N1 (Q )  N 2 (Q )]x (8.41)

When the system is in thermal equilibrium and N1 > N2, the intensity of the beam will decrease
exponentially as it propagates through the active medium. Do you know what happens to the
lost energy? It is utilised to excite atoms to higher energy states. It points to the possibility
where population in the upper energy level could exceed the population in the lower energy
level (N2 > N1). As such, it will correspond to a non-equilibrium state.
For N2 > N1, bn < 0 and the material is said to be in the state of population inversion.
However, Eq. (8.41) implies that the intensity of light beam will improve. That is, the intensity of
the incident beam will get amplified as it propagates through the active medium. This process
is termed light amplification. From this discussion we may conclude that for light amplification
by stimulated emission of radiation, population inversion is an essential prerequisite.
We now pause for a while and ponder as to how population inversion is achieved in actual
practice. The mechanism used for this purpose is known as pumping. The process of achieving
population inversion is also known as excitation. You will learn about it now.

8.5.2 Pumping

We now know the importance of population inversion in an active medium for lasing action
to be initiated. We can attain population inversion by energising the active medium by
– pumping atoms into the upper energy levels; and/or
– de-populating lower energy levels (other than the ground state) faster than the upper
energy level.
A few typical excitation mechanisms used to activate the lasing medium include:
• Optical pumping or excitation by photons using a light source.
• Electrical pumping using gas discharge.
• Inelastic atomic collisions (semiconductor laser).
• Chemical pumping using energy from chemical reactions.
Lasers and Their Applications 321
We now briefly discuss these one by one.
In optical pumping, a high energy light source such as a Xenon flash lamp or another laser
can be used to supply energy to the active medium (Figure 8.8). This energy comes in the form
of short flashes of light and atoms are raised to the appropriate excited state by selective
absorption of radiation in an atomic/molecular system. In this process, strong deviations from
thermal equilibrium populations of selected states may take place. This method is particularly
suited for solid state or liquid lasers whose absorption bands are wide enough.

Figure 8.8 Xenon flash lamp in ruby laser.

As such, optical pumping is a resonant phenomenon; the energy of the incident photon
must be equal to the difference of energies of excited and normal states. If X and X* respectively
denote a normal and excited atom, we can express optical pumping symbolically as
X + hn ® X*
If we have a laser which gives out light whose wavelength lies in the absorption bands of the
active medium (solid, liquid or a gas), we can use it for pumping. Since the bandwidth of a
laser is very narrow, its pumping efficiency will be very high. It may be mentioned here that
this method was first used by Maiman for Ruby Laser but even today, it is widely used in
solid-state lasers.
In electrical pumping, electrons are usually excited by high voltage electrical discharge.
(One can also use a continuous dc current, radio frequency current or pulsed current.)
The electric field is of the order of a few kVm–1 and it accelerates electrons emitted by the
cathode towards the anode. Some of these electrons may collide with the atoms of the active
medium and raise them to the excited state. This is a preferred method for gas lasers such as
argon laser for two reasons:
(i) a gas has to be contained in an enclosure or a tube, and
(ii) the absorption band of a gas is narrower as compared to a liquid or a solid.
Obviously, the use of a wide band lamp light will be inefficient as much of its energy will
remain unused and dissipate as heat.
Electrical pumping is non-resonant pumping as excitation is impacted by an electron.
We can symbolically express it as
X + ei ® X* + ef
322 Wave Optics

Here X, X*, ei and ef respectively denote normal atom, excited atom, incident electron and
final electron.
Electrical pumping is construed more convenient and efficient for semi-conducting lasing
media as well.
In inelastic atomic collision method, an electric discharge is passed through a gaseous
medium such as He-Ne. As a result, atoms of one type are excited to higher energy states and
these subsequently collide inelastically with the other types of atoms. The energy transferred
in inelastic collisions raises the latter type of atoms to their excited sates and these in turn bring
about population inversion. This process is particularly useful for He-Ne lasing medium.
In chemical pumping, energy generated in an exothermic chemical reaction is used to
produce an atom/molecule in an excited state. It is used for materials in gaseous phase and
generally requires highly reactive gas mixtures. A familiar example is CO2 laser, in which HF
molecule is produced in an excited state when H and F molecules combine. The number of
these molecules is significantly more than the number of molecules in the normal state. Such
lasers are used as directed energy lasers.
In semi-conductor lasers, light emitting diodes are used for pumping by direct conversion
of electrical energy into radiation.
So far we have considered only two-level atomic systems to achieve population inversion.
But an efficient inversion mechanism uses three level pumping scheme as in ruby laser, or four
levels pumping scheme as in Nd: YAG laser. Let us learn about these pumping schemes now.

Pumping schemes
To discuss pumping schemes, it is convenient to indicate the pumping transition by upward
arrow, the lasing transition by downward arrow and non-radiative decay by slanted arrow.
(In non-radiative transition, the excess energy does not appear in the form of electromagnetic
radiation. It is, in general, utilized in raising the vibrational energy of the host medium,
resulting in its heating.)
Refer to Figure 8.9(a). It shows a three-level pumping scheme. Note that the ground state
is represented as 1, the lasing state as 2 and the pumping state as 3. Consider an assembly of
N atoms, each of which can exist in any of the three energy states having energies E0, E1 and
E2 and populations N0, N1 and N2, respectively. We assume that to begin with, the system is
in thermal equilibrium so that most of the atoms will be in the ground state, i.e. N0 » N and

Figure 8.9 Pumping schemes: (a) three-level and (b) four level.
Lasers and Their Applications 323
N2 » N1 » 0. If a radiation of frequency np = E2 – E1 is now made to fall of the system,
the atoms will begin to get excited from the ground state to the pumping state due to optical
absorption. And in due course of time, an appreciable number of atoms may be excited to the
pumping level.
In a medium suited for laser operation, the pumped atoms will be expected to decay
quickly to the lasing level 2 through a non-radiative transition. (This decay is very fast;
the lifetime is of the order of 10–8 s and is shown by slanted arrow in Figure 8.9(a). You may
recall that lifetime signifies the average time that an atom spends in an excited state.) This state
is generally metastable. (A metastable state is an excited state of an atom/nucleus/molecule
that has a lifetime longer than the ordinary excited state but shorter than the lowest and
stable ground energy state. From quantum mechanical point of view, a metastate is not a truly
stationary state but we can say that it is almost stationary. For example, the lifetime of a lasing
state »10–3 s or even more, and is much greater than the lifetime of excited or pumping state 3.)
An atom in state 2 may decay to the ground state by spontaneously emitting a photon of
frequency nL = (E1 – E0)/h. This is shown by downward arrow in Figure 8.9(a). If the lifetime of
this transition is much greater than the lifetime of the non-radiative transition, the population of
the pumping level 3 will drop very rapidly and become essentially zero. As a result, a population
of excited atoms will crowd level 2 in a three level pumping scheme. If more than half the total
population accumulates in level 2, this will exceed the population of the ground state population.
We then say that population inversion has occurred between the metastable lasing state 2 and
the ground state 1 and this may lead to optical amplification at the frequency nL.
Note that in three level pumping scheme, more than half the atoms have to be excited from
the ground state to obtain population inversion. This feature proves too demanding because
ordinarily, occupancy of the ground state is higher and in the instant case, the laser medium
has to be pumped very strongly. This makes a three level pumping scheme rather inefficient
and unattractive.
Now refer to Figure 8.9(b). It shows a four-level pumping scheme. The energy levels
correspond to energies E0, E1, E2 and E3 and populations N0, N1, N2 and N3, respectively.
Note that in this pumping scheme, we have two upper levels, instead of one. As before, atoms
are excited from the ground state to the pumping state (level 4). From level 4, atoms undergo
fast non-radiative transition to the upper lasing (metastate or level 3). Since lifetime of meta-
state is longer compared to the lifetime of pumping state, atoms tend to accumulate in level
3, which may relax by spontaneous or stimulated emission into level 2 (lower lasing level).
Atoms in this level again decay non-radiatively and reach the ground state.
Due to non-radiative transition, population in the pumping state (level 4) depletes fast, as it
happens for level 3 in the case of three level pumping scheme. Similarly, atoms in the lower lasing
level (level 2) also de-excite rapidly and the number of atoms drops to a negligible number
(N1 » 0). As a result, any appreciable accumulation of atoms in the upper lasing level leads to
population inversion between levels 2 and 3. That is, if N2 > 0, N2 will be greater than N1. Thus
optical amplification and laser operation takes place at a frequency nL = (E2 – E1)/h.
You may now like to know: Which pumping scheme is more efficient? Since excitation
of only a few atoms into the upper level helps achieve population inversion, a four-level
scheme is more efficient than a three-level scheme. Primarily for this reason, four-level pumping
scheme is used in a wide variety of practical lasers. But a serious drawback in four-level
324 Wave Optics

scheme arises due to loss of a lot of energy in non-radiative transitions between the pumping
state (4) and the upper lasing state (3), as also the lower lasing state (2) and ground state (1).
We may, therefore, conclude that each pumping scheme has its own limitations but the ultimate
choice is dictated by the active media, type of usage, and such other factors. Before we discuss
these aspects, you should now go through the following solved example.

EXAMPLE 8.6 For lasing action, transition takes place from an excited state to a zero
energy state. If light of wavelength 600 nm is produced, calculate the energy of excited state.
Solution Since lower energy state corresponds to zero energy, the frequency of the emitted
radiation is given by hn = E2. Hence

c (3 – 108 m s 1 )
E2 h (6.67 – 10 34 J s) – 3.34 – 10 19 J = 2.09 eV
O (600 – 10 9 m)

We now know that a medium that supports population inversion can amplify light of a
particular frequency when stimulated emission dominates spontaneous emission. This is the
basic principle of an optical amplifier. But in a laser, we have amplification, coherence and
mono-chromaticity; a spontaneously emitted photon from an upper lasing state stimulates
emission of new photons. These, in turn, effect light amplification. However, it is important
to know how coherence of amplified light arises in a laser. This is done through positive
feedback wherein a part of the light emerging from the active medium is reflected back into
it. An amplifier with positive feedback is known as an oscillator. You may now like to know
as to how we achieve this condition in a practical situation. This is done by placing the active
medium between a pair of mirrors facing each other. The space between the two mirrors is
known as laser cavity. Note that the mirrors in the cavity help increase the length of the active
medium and amplify the beam strength through multiple reflections. Moreover, these determine
the boundary conditions for the electromagnetic fields inside the cavity and allow only particular
frequencies. For this reason, it is also termed optical resonant cavity. Let us learn about it now.

8.5.3 Optical Resonant Cavity

Refer to Figure 8.10. It shows a simple optical resonator, which consists of two plane mirrors
M1 and M2 placed parallel to each other. The active medium is placed between these mirrors.

Figure 8.10 An optical resonator.


Lasers and Their Applications 325
While M1 is a perfectly reflecting mirror, M2 is a partially reflecting mirror, i.e. a part, 2% to
80% of the beam is transmitted by mirror M2 depending on the type of the laser. (The transmitted
light constitutes the output laser beam.) The mirrors are set normal to the optic axis, which
defines the direction of the laser beam.
In the process of achieving population inversion, spontaneous and stimulated photons are
initially emitted in all possible directions. However, to generate coherent output, it is essential
that photons of only a specific direction are selected. The photons emitted spontaneously along
the optic axis (PQ) or infinitely close to it travel a relatively longer distance in the active
medium. (Other photons are lost as their motion is weird.) This is because most of these
photons travel back and forth in the medium as these are reflected by the mirrors M1 and M2
several times and spend longer time. (After reflection by mirror M2, the photons de-excite
atoms while moving in the medium and augment their number. In this way, the impact of
transmission losses is minimized.) Therefore, the chances of their interaction with atoms in the
upper lasing level are significantly high. As a result, identical photons will be added in the
same direction due to stimulated emission and an ever increasing population of coherent
photons will bounce back and forth between the mirrors. On the other hand, spontaneous
photons and the corresponding stimulated emissions in other directions will travel comparatively
shorter distances and hence spend lesser time in the active medium before being lost. So we
can say that an optical resonant cavity provides desired directional selectivity and ensures
spatial coherence of the laser beam.
Let us now consider monochromaticity aspect of laser light. On the basis of physical
arguments, we can say that monochromaticity lends itself to the origin of laser—the stimulated
emission. This is because spontaneously emitted photons, whose frequency does not match
with the frequency difference between lasing levels, will not stimulate emissions. Thus the
band of wavelengths emitted during spontaneous emission is narrowed down. The
monochromaticity of a laser is further enhanced by the optical resonant cavity. We can understand
it mathematically by considering standing wave formation in an optical resonator.
Let us consider a cavity of length L. Suppose that light of wavelength l starts from mirror
M1 and travels towards mirror M2. In propagating through the active medium, it will get amplified.
Since mirror M2 is partially transparent, a small fraction is transmitted but a major fraction is
reflected back towards M1. In its journey back to M1, the beam is again amplified and regains
its origin level. For simplicity, we assume that two plane waves travel in opposite directions
in the cavity. Their superposition leads to interference and only certain patterns—stationary
waves—and frequencies will be stable and sustained by the resonator; other formations and
frequencies are suppressed by destructive interference. The frequency of stable mode is given by
nOn
L n = 1, 2, 3, ... (8.42)
2
That is, only those wavelengths can exist inside the resonant cavity in the steady state which
satisfy Eq. (8.42); the waves of other wavelengths will interfere destructively as they travel in
the medium.
In terms of frequency, we can rewrite Eq. (8.42) as
c c
Qn n n = 1, 2, 3, ... (8.43)
O 2L
326 Wave Optics

where c is the velocity of light and nn is the frequency of the nth mode. It may be mentioned
here that for an actual resonator, n is very large and enhances monochromaticity of the beam.
You can easily verify that for l = 600 nm or n = 5 ´ 1014 Hz, n = 2 ´ 106 for L = 0.6 m.
Also note that we have assumed that waves inside the cavity move undamped/unmodified. But
this is not true in practice; due to finite transverse dimensions of the mirrors, some energy is
lost due to diffraction effects. This loss is referred to as diffraction loss. There could be losses
due to absorption in the medium as well.
From Eq. (8.43) we note that the frequency difference between two consecutive modes is
c
'Q (8.44)
2L
Out of this spread of frequencies, the resonator selects and amplifies a particular narrow band.
(Other frequencies are attenuated rapidly.) This accounts for the extreme monochromaticity
(spectral purity) of a laser beam. That is why resonant cavity is the most vital component of
a laser for giving out highly coherent light. The mechanical analogue of such behaviour is a
vibrating guitar string in that a particular string vibrates only at a certain frequency.
It may be mentioned here that another widely used laser resonant cavity has spherical
mirrors with the same radius r separated by a distance L = 2r so that their centres are coincident.
Such a resonator is referred to as concentric resonator. Confocal resonator and ring resonators
are other important class of laser resonators.
So far, we have discussed the basic principle of laser and its components. Since the
development of ruby-based laser in 1960, there have been significant developments due to
intense research in this field essentially driven by vast applications of lasers. Different types
of lasers are now available and we discuss these in the following paragraphs. It may, however,
be mentioned that due to paucity of space, it is not possible to discuss all of them in detail.

8.6 TYPES OF LASERS


Lasers can be classified in many ways depending on the state of the active medium—solid,
liquid, gas—pumping technique used—optical, electrical, chemical—the purpose—research,
industrial, commercial, spectral range of the laser wavelength—visible, infrared and so on. The
material of the active material determines the laser size, wavelength, pumping scheme, output
and efficiency. In this section, we have discussed some of these with particular reference to
the physical state of the active medium.

8.6.1 Solid State Lasers


The most common solid state lasers are the Ruby lasers and the Nd:YAG (neodymium yttrium
aluminium–garnet) lasers. In these lasers, the active material is essentially an insulator doped
with transition element ions in the host structure and population inversion is obtained using
optical pumping. It means that excitation of the active medium is achieved by absorption of
light. For optical pumping, discharge flashtubes or a continuously operating lamp is employed
as the source. It is important to highlight here that transition element ions act as active centres
in oxide-based host materials.
Lasers and Their Applications 327
Ruby laser
T.H. Maiman heralded the beginning of laser age when he constructed a ruby laser in 1960.
Ruby is Al2O3 crystal doped with 0.05% (by weight) triply ionized chromium atoms (Cr3+).
Al3+ ions are substituted by the Cr3+ ions, which constitute the active centres, whereas the
aluminium and oxygen atoms act as the host.
Now refer to Figure 8.11(a). It depicts the schematics of ruby laser. It is shaped into a rod
from a single cylindrical ruby crystal, whose ends are flat and polished; one end is completely
silvered and the other end is partially silvered. These ends behave as completely reflecting and
partially reflecting mirrors and the arrangement works as a resonant cavity. The ruby laser rod
is surrounded by a xenon flash lamp (an external source). The light from this flash lamp enters
the laser rod and excites the Cr3+ ions.
The energy levels of the chromium ion are shown in Figure 8.11(b). Note that the bands
marked as E1 and E2 have a lifetime of about 10–8 s, whereas the metastable state M has a

Figure 8.11 (a) Schematic representation of a ruby laser, and (b) energy levels
of the chromium ion and lasing action in a ruby laser.
328 Wave Optics

lifetime of about 3 ´ 10–3 s. Depending on the energy of the photon absorbed, a Cr3+ ion in
ground state can make transition to one of the states in band E1 or E2. In either case,
it immediately goes to the metastable M state via non-radiative transition. Since the lifetime
of the metastate is very high (several orders higher than the states in bands E1 and E2), the
number of atoms continue accumulating in it and in due course we obtain population inversion
between states M and G. Once population inversion takes place, light amplification is achieved
with the reflecting ends of the ruby rod forming a cavity.
In the original set up of Maiman, the xenon flashlamp was connected to a capacitor, which
was charged to a few kilo-volts. The energy stored in the capacitor discharged through the
xenon lamp in a few milliseconds. This generated a few mega-watt power. Some of this energy
was absorbed by chromium ions leading to their excitation and subsequent lasing action. It may
be mentioned here that the first ruby laser produced intense pulses of red light of wavelength
of 694.3 nm. The duration of pulses ranged from fraction of a millisecond to a few milliseconds.
To increase the pumping efficiency of a ruby laser, the source and the crystal are kept at
the foci of an elliptical reflector. The active material, in the form of rod, is placed at one focus
and pumping source (in the shape of right cylinder) at another focus of an elliptical reflector.
In this arrangement, light leaving one focus of the ellipse is made to pass through the other
focus after reflection from the silvered surface of the pump cavity. In this way, entire radiation
is maximally focussed on the active material.

Nd: YAG laser


The Nd: YAG laser employs four level pumping scheme. In Figure 8.12, the energy levels of
neodymium have been marked as E0, E1, E2, and E3. The yttrium-aluminium-garnet (YAG)
constitutes the active medium with trivalent Nd3+ (neodymium) ions as impurities. So, in this
laser, YAG is used as the host and ionised Nd as lasing agent.

Figure 8.12 Energy level diagram of Nd in Nd: YAG and lasing action.

The optical pumping (using high pressure gas-discharge lamps or diode laser) raises the
neodymium atoms from the ground state to a few excited states. The energy levels marked as
E2 and E1 are the lasing levels. As in a ruby laser, the atoms in the excited state undergo non-
radiative transition to the upper lasing level (E2). Out of the group of lower lasing levels,
the major portion of energy is emitted in the transition E2 ® E1. The Nd: YAG laser operates
in pulse mode with average power outputs up to 1 kW. This laser has two distinct advantages:
It has (i) a low excitation threshold, and (ii) high thermal conductivity. Because of these, this
laser can be used to generate light pulses at a high repetition rate.
Lasers and Their Applications 329
In solid state lasers, the pulsed beam is invariably made of spikes of high intensity emissions.
This difficulty is overcome by using gas lasers. Some gas lasers can emit continuous beam for
years, whereas a few others emit pulses lasting for a very short duration (» 10–9 s). Moreover,
gas lasers can be designed to produce output beams over a wide range of wavelengths which
may vary from deep ultraviolet through the visible and infrared to millimetre waves.
Except cesium vapour laser, gas lasers are pumped electrically, unlike solid state lasers.
You may now like to know the advantages offered by gas lasers. Some of these are:
• Active material is available in abundance and is relatively inexpensive.
• The heat generated in the cavity can be removed conveniently and quickly.
• The active material is not damaged easily.
• High degree of optical perfection due to uniform density of the gas.
It may be mentioned here that as of now, a wide variety of gas lasers with different active
media have been designed. These include He–Ne, Ar, Kr, Xe, N2, and CO2. You may now like
to ask: What is the mechanism of pumping in gas lasers and why? We use electrical pumping
in gas lasers as it offers several advantages. Some of these are as follows:
• The electrical discharge satisfies the condition for amplification by stimulated emission
for almost every gas at some wavelength.
• The optical pumping shows considerable spread in energy but a gaseous active medium
absorbs radiation in a very narrow range. It means that most of the pumped energy goes
waste and the mechanism will prove very inefficient.
It is pertinent to mention here that the most widely used gas laser is the Helium-Neon laser.
We now discuss it in some detail.

8.6.2 Helium–Neon Laser

The helium–neon (He–Ne) laser was the first successfully operated gas laser. It was developed
in 1960 by Ali Javan, William Bennet Jr and Harriot at Bell Telephone Laboratories. Refer to
Figure 8.13, which shows a long, narrow discharge tube; it is only a few millimetres in
diameter. A mixture of helium and neon gases (in the ratio of 10:1), which acts as active
medium, is kept inside this discharge tube. The pressure inside the tube is maintained at about
300 Pa. The gaseous system is enclosed between a pair of plane or convex mirrors so as to
form a resonator system. The length of the cavity is between 0.15 m and 0.5 m. As in solid
state lasers, one of the mirrors has very high reflectivity while the other one is partially

Figure 8.13 Schematics of a He–Ne gas laser.


330 Wave Optics

transparent (1% transmission). The pumping is done by a stationary glow discharge fired by
a direct current. When the potential difference between the electrodes is about 1 kV, a glow
discharge is initiated. Lasing levels in this laser are provided by the excited states of the Ne
atoms and He atoms help in pumping neon atoms to the excited states.
To discuss the pumping scheme of He–Ne laser, refer to Figure 8.14, which shows the first
few energy levels of He and Ne atoms. When free electrons produced in the discharge move
through the He–Ne mixture, they are absorbed by He atoms and excite them to higher energy
levels. These levels are metastable and He atoms excited to these states continue to stay there
for fairly long before losing energy through collisions with ground state neon atoms. This
results in transfer of energy to the neon atoms. An interesting feature of He–Ne energy level
diagram is that the energies of E3 and E5 levels of Ne are nearly the same as E1¢ and E2¢ of He.
So when He atoms in these states collide with Ne atoms in ground state, the He atoms transfer
their energy to Ne atoms and raise them to E3 and E5 levels. As a result of this so-called
resonant energy transfer, He atoms fall back to the ground state. This increases population of
Ne atoms in energy levels E3 and E5 as compared to that in states E2 and E4. In this way, the
condition of population inversion is achieved between levels E5 and E4 as well as E3 and E2.
Then any spontaneously emitted photon can trigger lasing action between these levels. The Ne
atoms then drop down from the lower lasing levels (E2 and E4) to level E1 through spontaneous
emission and to the ground state through collisions with the walls.

Figure 8.14 Lasing levels in a He–Ne laser.

The wavelengths emitted in transitions in a He–Ne laser between levels E5 ® E4, E5 ® E2,
and E3 ® E2, are 3390 nm, 632.8 nm and 1150 nm, respectively. Note that laser transitions
corresponding to 3390 nm and 1150 nm fall in the infrared region. However, the light of
wavelength 632.8 nm falls in the visible region and corresponds to the characteristic red light
of He–Ne laser. It may be pointed out here that with appropriate selection of cavity mirrors,
Lasers and Their Applications 331
it is possible to manipulate transitions so as to obtain many other colours in the visible region
(543.5 nm, 594 nm, 612 nm corresponding to green, yellow and orange colour, respectively.)
It may be mentioned here that He–Ne laser gives out monochromatic and highly directional
light of wavelength 632.8 nm due to the absence of crystalline imperfections, thermal distortions
and scattering, which are present in solid state lasers. Moreover, these can operate continuously
without requiring cooling mechanism. In fact, because of excellent beam quality, He–Ne laser
is very useful for holography and spectroscopy. Prior to fabrication of diode lasers, He–Ne
lasers were also used in barcode scanners.
Another important gaseous laser is the CO2-laser. We now discuss it in brief.

8.6.3 CO2 Laser


Carbon dioxide (CO2) laser was developed by C.K.N. Patel in 1964. You may recall that in
ruby and He–Ne lasers, electronic transitions are responsible for lasing action. The active
medium in a CO2 laser is a mixture of CO2 (10%), N2 (10%) and the rest is helium. Note that
carbon dioxide laser is fundamentally different from He–Ne gas lasers in that here molecular
vibrations of CO2 provide for the lasing action. (Nitrogen molecules help in the process of the
excitation of the CO2 molecules.) The CO2 laser can operate continuously at very high power
(» billions of watts) and emits light of wavelengths 10.6 µm and 9.6 µm, which lie in the far
infrared region. Unlike most other gas lasers, the CO2 laser has an appreciably high efficiency
(10–15%). And to attain the high power required from these lasers, cavity lengths can go from
2 to 3 m or even more.
An excellent example of CO2 laser is “Transversely Excited at Atmospheric pressure”, the
so-called TEA laser. In this laser, the gas mixture is made to flow perpendicular to the axis
of a rectangular shaped laser tube at atmospheric pressure. So, the direction of electrical
discharge is normal to the optical axis. In the laser cavity, the gas flows in and out rapidly and
allows excess heat to be removed quickly. The TEA CO2 laser produces intense, short pulses
of microsecond (µs) duration. So, this laser finds applications in industry for heat treatment as
well as medical treatment for dermatology.

8.7 OTHER LASERS

So far we have considered solid state and gaseous lasers. Let us now consider a few other types
of lasers which have been developed for different purposes.

8.7.1 Semiconductor Laser


Gallium arsenide (Ga–As) laser is the most familiar example of a semiconductor laser.
This laser is now available in very compact form. It is unique in terms of its opto-electronic
applications. It consists of a single crystal of Ga–As and emits light in near infrared
(830–850 nm) region (It is also called homo-junction laser.) Semiconductor lasers are used for
information collection and processing, i.e. as image scanning and measurement systems in
bar-code scan, for information storage as in optical disk data storage and for information
transmission as in optical fibre communication.
332 Wave Optics

8.7.2 Chemical Laser


As the name suggests, a chemical laser is one in which chemical reactions stimulate lasing action.
In such lasers, no external source of energy is used/required. Often, these lasers use gases as
active medium and the resultant products of the reaction are excited to higher energy states that
are capable of emitting photons. They produce pulses of energy of the order of a billion watt
(» 200 giga-watt). Majority of chemical lasers utilise Hydrogen Fluoride (HF) and Deuterium
Fluoride (DF) as the active laser molecules because of their attractive features including CW
and pulsed operation, fairly large output power and shorter IR lasing wavelengths (3–4 mm).

8.7.3 Free Electron Laser

Refer to Figure 8.15. You will note that the kinetic energy of the relativistic electron beam
emerging from the source is transformed into a laser radiation in a free electron laser (FEL).
This transformation in laser radiation occurs as the electron beam produced by a particle
accelerator passes through an alternating magnetic field produced by a specially designed
magnet. Because of this, the electron moves in an oscillatory path and the resultant radiation
moves along the axis of the system.

Figure 8.15 Schematic diagram of a free electron laser.

Free electron laser produces waves at wavelength in the range 30 µm £ l £ 0.01 m. FEL
can produce tuneable radiation at infrared, optical wavelengths and even up to X-ray region.
They are suitable for satellite communications, precision work as in air navigation and heating
plasma in thermonuclear fusion.

8.7.4 Dye Lasers

Dye lasers are of comparatively recent origin. These use active fluorescent material (organic
dyes) in a solvent such as alcohol or ethylene glycol. Some of the dyes are Rhodamine 6G,
fluorescein, coumarin, stilbene, tetracene, and malachite green. The lasing action in a liquid
dye laser is usually provided by a flash lamp and therefore these use optical pumping for
population inversion. They emit light waves of continuously varying wavelengths (fluorescent
bands ranging from 50 to 100 nm) and have a great advantage of being tuneable. With several
dyes, the entire visible range can be tuned. The CW power of dye lasers varies from a few mW
to 1W. In pulse operation, peak power of about 100 W is obtained and pulses are in pico-second
Lasers and Their Applications 333
range. Prisms and diffraction gratings are used to select radiation of desired wavelengths from
the emitted fluorescent band of the dyes.
Organic dyes tend to degrade under the influence of light. For this reason, the dye solution
is normally circulated from a large reservoir.

8.7.5 Ion Laser


An ion laser is essentially a gas laser, wherein an ionized gas acts as the active medium. Since
the energy required to excite ionic transitions in ion lasers is invariably large, the current
required for initiating the lasing action is several amperes. As a result, the ion laser tube
produces a large amount of waste heat and therefore, requires active (air/water) cooling.
The most familiar examples of ion laser are Krypton ion laser and Argon ion laser. Krypton
lasers are used for scientific work. When Krypton is mixed with Argon, it gives out white light,
which is used in laser light shows. Krypton lasers are also used in medicine for coagulation
of retina, for manufacturing security holograms and several such other purposes. The light
emitted by Krypton laser corresponding to wavelengths 406.7 nm, 413.1 nm, 415.4 nm, 468
nm, 476.2 nm, 482.5 nm, 520.8 nm, 530.9 nm, 568.2 nm, 647.1 nm, 676.4 nm, which lie in
the visible region.
An Argon ion laser gives out light corresponding to transitions at 488 nm and 514.5 nm,
which fall in the visible region, just like Krypton ion laser. However, it also gives out light
corresponding to wavelengths 351.1 nm and 363.8 nm, which fall in the ultraviolet region.
Common Argon and Krypton lasers can emit waves of several mW for tens of hundreds
of seconds continuously. Their tubes are usually made of beryllium oxide ceramics or copper.
So far we have confined ourselves to understand the physical principle, types and operation
of a laser. You must have realised that the basic principle, insofar as lasing action is concerned,
is essentially the same in all cases. We also discussed how their unique features such as
spectral purity (high coherence) and high directionality arise. In fact, it is due to these
characteristics that lasers find so many and so varied applications—in scientific research,
medical surgery, genetic engineering, energy production, military and industrial processes such
as welding, cutting, heat treatment, isotope separation, among others. Nowadays, lasers are also
used for data transmission and processing (communication), optical fibres, holography, aviation,
laser printers, guidance systems, scanners and space physics with very high degree of reliability.
The most common lasers routinely used in everyday life are weak diode lasers used in CD
players and in barcode scanners, security systems, pointers and movies. The pyrotechnics that
we witnessed in the inauguration and closing of Commonwealth Games from Jawaharlal Nehru
Stadium in Delhi or ICC cricket World Cup in Bangbandhu Stadium in Dhaka and such other
functions worldwide use lasers to produce wonderful effects. It would be no exaggeration to
say that advent of the laser in 1960 heralded the beginning of spectacular developments for our
good. We will now discuss a few typical applications of lasers.

8.8 APPLICATIONS OF LASERS


Communications In the electronics course we learnt that capacity of a communication channel
depends on the frequency of the carrier wave. Since a laser emits highly coherent, monochromatic
334 Wave Optics

light of very high frequency—638.8 nm wavelength emitted by He–Ne laser corresponds to


4.738 ´ 1014 Hz—their use provides scope for multiplexing, i.e. the same pathway can be used
to simultaneously transmit several messages. And when laser beams are used as carrier waves,
the information carrying capacity increases tremendously. In fact, the days are not far away when
messages such as ‘All lines in this route are busy, please dial after sometime’ will be gone forever.
The signal carrying laser beams can be transmitted through free space as well as by light
guides. In the form of optical fibres, light guides have found wide use in optical communication.
Semiconductor lasers with optical fibres are widely used in telecommunication. Wireless
communication using lasers is also being used in Local Area Network (LAN). In wireless
LAN, use of infrared technology is quite widespread, as it saves on cost. (This is analogous
to the one used in the remote control of a TV. The transmitter uses IR LEDs and the receiver
uses a photodiode.) Remote control of rockets and missiles to desired destination is also
facilitated by a laser.
Medicine A laser beam is an excellent diagnostic as well as surgical tool, since it can be
focussed on an extremely small area (spot). Laser beams are being used to remove gall bladder
and kidney stones by breaking them into small pieces, operate hernia, vocal chord surgery, tonsils
removal, removing stomach ulcers, plastic surgery, dermatology, dentistry, urology, neurosurgery,
oncology and performing microsurgery on cells and chromosomes, gynaecology, stopping of
gastric bleeding and such other surgical interventions. An important advantage of laser surgery
is that it is bloodless and painless as the laser beam attacks fewer cells and ‘welds’ blood vessels,
after initial cutting. Moreover, it is very fast. Laser beams are also used in drilling and cutting
bone tissues and destroying specific malignant areas within tissues (treatment of cancers).
Nowadays, bloodless surgeries of heart, abdomen and other body parts are performed as
a routine. (In the decades gone by, surgeons harnessed this potential of lasers maximally to
create riches for themselves.) These involve use of laser beams, which can propagate through
optical fibres. (Fibres are introduced into arteries using catheters.) This facilitates removal of
plaque (a fatty material that accumulates on the arterial wall and blocks the blood flow) from
coronary arteries. Even in root canal therapy, the dentist inserts the laser fibre into the root
canals, removes the infected tissue by vaporizing it and destroys the infection causing bacteria
completely and effectively. Similarly, high precision of a laser facilitates change in the shape
and hence refractive power of the cornea to a desired level, without causing any (thermal)
damage to the surrounding tissues. In laser acupuncture, silver and gold needles are replaced
by fine, micro manipulator oriented laser beams. Lasers also find extensive use in medical
R&D. It is used to induce changes in cells and lends itself very naturally to genetic engineering.
Industry The use of lasers in industry is very far reaching and widespread. Using lasers, it
is possible to weld, cut and drill very fine holes in metals perfectly and develop extremely
sophisticated tools. Due to its sharp focussing characteristic, a highly energetic beam of a laser
can be confined within a diameter of 10–100 µm. (This is smaller than the dot that you can put
with the sharpest pen.) As a result, the use of laser beams leads to higher yields and superior
product quality. Lasers are very effective in heating discrete areas very rapidly. Using CO2 and
Nd: YAG lasers (of power 500 W or more), continuous wave and pulsed welding has become
possible.
For fabrication of the parts of a spacecraft that are made of titanium, steel and aluminium,
Nd: YAG laser is used to cut with high repetition rate. In most of the industrial applications,
Lasers and Their Applications 335
carbon-dioxide (CO2) laser is used for cutting. Lasers are also used for laser markings, i.e.
marking on vernier callipers, screw gauges, reactor and aircraft components, electronic
components, typewriter keyboard, turbine blades, etc. Laser soldering, which leads to high and
reliable quality joints, surface alloying and cladding, annealing and photolithography for the
production of integrated circuit chips are some other uses.
Major advantages of using lasers in industry include no tool breakage or wear, precise
location and drilling in very hard materials like diamond without much stress.
Applications in market goods and library books as bar code In your day-to-day life,
you must have observed a patch of black and white strip on packets of consumer goods,
on books, medicines and so on. The bar code is read by a scanner. This process is very fast
and information is both accurate and reliable as compared to keying in the information in the
computer/register by looking at the label on the article. In libraries, use of lasers has made
book lending and returning faster.
A laser bar code scanner has two parts: laser (LED chip) and detector (a photodiode +
transistor chip). As bar code is held in front of the laser beam, the black lines of the bar code
absorb laser light and the white lines of the bar code reflect it. The reflected light reaches the
detector, which is translated into digital signal. The digital signal is fed into the computer and
the information is displayed on the monitor screen.
Holography In conventional photography, an illuminated (three dimensional) object is
recorded as a 2-D image on a photographic film. But use of a laser beam in recording has
introduced a qualitatively superior method of photography. This is known as holography.
(In Greek, hologram means “whole”.) It is a special three-dimensional photograph of an object
that retains information about the phases of waves coming from the object. During the recording
process, two waves—one coming from the object and another coherent reference wave—are
made to superimpose. (Note that both waves originate from the same laser.) These produce
interference fringes in the plane of the photographic film. Holography records the reflected
light waves themselves. This photographic record carrying the information of amplitude and
phase of the object is called hologram. As such, a hologram looks like a hodgepodge of specks,
blobs, and whorls; it seems to bear no resemblance to the original object. That is, a hologram
does not carry the image of the object but it carries a permanent signature of the object in the
form of an intricate interference pattern.
Holography was discovered by Denis Gabor in 1948 while attempting to improve the
resolution of electron microscopes through wavefront reconstruction. And once he succeeded
in recording a hologram, he realised the significance of his work. This technique became a
practical proposition only after the advent of lasers. We will learn about it in detail in Chapter 9.
Digital storage of information You may recall that a digital system operates on binary
system. Devices such as optical disks, magnetic tapes, CDs, CD-ROMs, DVDs are now widely
used to store information, including music, digitally with advantage of quality. Digital storage
allows continuous updating and dynamic retrieval of information. Laser light is used in reading
information from and recording on a CD/DVD. It is now possible to include special sound
effects and enrich the quality of music. Laser is also used for printing text and its quality is
better than any other printing.
Other applications of lasers Apart from the above-mentioned applications of lasers,
there are several other uses from energy generation to production of low temperatures to
environmental studies and technological applications. These include:
336 Wave Optics

• in automobile industry for wheel alignment of vehicles;


• extremely accurate measurement of distance between heavenly bodies;
• isotope separation for uranium enrichment to produce nuclear fission energy;
• plasma confinement in nuclear fusion reactors,
• production of low temperatures (up to about 1 micro-kelvin);
• environmental study—Lidar (light detection and ranging), which is similar to radar, is
used for weather forecasting and environmental pollution;
• law enforcement—detection of traffic speed violators using laser gun, finger print
detection, etc;
• battle-field device as in star war programme, wherein laser light beam is directed at a
satellite or missile to destroy them; and
• manufacturing decorative pieces through engraving rubber stamp, plaque and award
badges, leather and glass pieces.
We now sum up what you have learnt in this chapter.

8.9 SUMMARY
• The average time for which the electric field exhibits sinusoidal behaviour is known as
coherence time. It can arise due to finite lifetime of an excited atom in the higher energy
level from which it radiates and goes to a lower energy state or random motion of atoms
or collision of the radiating atom with another atom.
• The path length corresponding to tc defines coherence length and is given by L = ctc,
where c is the speed of light.
• Monochromaticity is strongly related to temporal coherence.
• The degree of visibility or contrast of fringes produced by two light waves equals the
degree of coherence between them.
• The stellar interferometer magnifies the resolving power of the telescope in the ratio
D/d, where d is the diameter of the objective of the telescope and D denotes the separation
of the mirrors M1 and M2.
• When an atom in higher energy level emits energy to attain lower energy level,
the transition is accompanied by spontaneous emission of a photon of frequency
E2  E1
Q
h
• When an atom in an excited state is irradiated by a photon of energy hn, instead of being
absorbed, this photon may trigger the atom to de-excite and emit radiation. If a transition
is accompanied by simultaneous emission of two photons, each of energy hn, it is said
to be stimulated or induced emission.
• Stimulated emission generates an intense, highly coherent and directional beam, whereas
spontaneous emission produces incoherent and broad spectrum.
• If the number of atoms in a higher energy level is more than the number of atoms in a
lower energy level, the system is said to be in the state of population inversion. This is
central for lasing action.
Lasers and Their Applications 337
• For hn = kBT, the ratio of the number of spontaneous emissions to stimulated emissions
for a system in thermal equilibrium is given by

A21 hQ
B21u (Q ) kBT
and stimulated emission will dominate spontaneous emission.
• For constructing a laser, the prerequisites are an active laser medium, laser pumping
mechanism and laser cavity.
• The most general requirement for an active medium is that it should have atoms whose
electrons can be excited to a metastable energy level. In other words, an atom should
have at least three energy states; an upper energy state in which atoms can be ‘pumped’,
a lasing state and a lower energy state to which they can return. Also, it must support
population inversion between any two states.
• The mechanism used for achieving population inversion is known as pumping.
The excitation mechanisms used to activate the lasing medium include optical pumping
or excitation by photons using light source, electrical pumping using gas discharge,
inelastic atomic collisions (semiconductor laser), and chemical pumping using energy
from chemical reactions.
• Optical pumping is a resonant phenomenon.
• In electrical pumping, electrons are usually excited by high voltage electrical discharge.
It is non-resonant pumping.
• In inelastic atomic collision method, an electric discharge through a gaseous medium
such as He–Ne excites He atoms to higher energy states. These atoms make inelastic
collisions with Ne atoms and raise them to their excited states inducing population
inversion.
• In semi-conductor lasers, light emitting diodes are used for pumping by direct conversion
of electrical energy into radiation. In chemical pumping, energy generated in an exothermic
chemical reaction is used to produce an atom/molecule in an excited state.
• Three and four-level pumping schemes are used to attain population inversion but four-
level scheme is more efficient.
• An optical resonant cavity helps to attain amplification of coherent light through positive
feedback. It consists of two plane or spherical mirrors and the active medium is placed
between the mirrors.
• Lasers can be classified in many ways depending on the state of the active medium—
solid, liquid, gas—pumping technique used—optical, electrical, chemical—the purpose—
research, industrial, commercial, spectral range of the laser wavelength—visible, infrared
and so on.
• Lasers find varied applications in scientific research, medical surgery, genetic
engineering, energy production, military, industry, data transmission and processing
(communication), optical fibres, holography, laser printers, guidance systems, scanners
and space physics.
338 Wave Optics

REVIEW EXERCISES

1. The sun subtends an angle of 32¢ on the surface of the earth. Assume that the sun acts
as an extended source. Before reaching the double slit arrangement, sunlight is made to
pass through a filter, which allows light of wavelength 600 nm only. Calculate the
separation between the slits for observing good contrast fringes on the screen.
[Ans. 6 ´ 10–5 m]
2. An atom has two atomic levels separated by 2.26 eV. At 300 K, calculate the ratio of
population of the upper level to that of the lower level. Take kB = 1.38 ´ 10–23 J K–1.
[Ans. 1.1 ´ 10–38]
–30
3. The ratio of populations in two energy levels at 300 K is 1 ´ 10 . If the upper energy
level corresponds to a metastable state, calculate the wavelength of light emitted.
[Ans. 687 nm]
4. Calculate the ratio of stimulated emission and spontaneous emission at an operating
temperature of 1100 K, if the wavelength of emitted light is 550 nm. Do these conditions
correspond to a laser or an ordinary lamp? [Ans. 4.6 ´ 10–11; ordinary lamp]
5. A ruby laser emits light of wavelength 694.3 nm. Assume that it is very nearly a plane
wave. Calculate
(a) the wavelength of this light as it passes through water (m = 1.33)
(b) the angle of refraction. What fraction of each component of polarised light is reflected
as the beam enters water at an angle of 45°?
(c) the amplitudes of the electric and magnetic field vectors of the plane wave propagating
through water, if the time averaged power of the beam in water is 1 kWm–2, and
(d) coherence length of laser in vacuum if the bandwidth of the laser is 30 MHz.
[Ans. (a) 520.8 nm; (b) 32°, R|| = 0.0028, R^ = 0.05;
(c) E0 = 7.5 V m–1, B0 = 3.3 ´ 10–8 T; (d) 10 m]
6. A dye laser consists of a gain medium of bandwidth Dn centred at n0 and two nearly
perfectly reflecting mirrors. Calculate
(a) the allowed frequencies for laser operation in this optical cavity. Express the result
in terms of the time t0 which light takes to make one round trip in the cavity,
(b) the bandwidth required to produce a pulse of duration 10–12 s at a wavelength of
600 nm. The length of the cavity l = 1.5 m.
Assume that the laser operates on all possible cavity modes within the gain bandwidth,
these modes are stable in phase and at t = 0, all modes are in phase.
Ë n Û
Ì Ans. D , n an integer; (b) 104 Ü
Í W0 Ý
7. A He–Ne laser operating at 632 nm has a resonant cavity with plane mirrors placed
0.5 m apart. Calculate the frequency separation between the axial modes of this laser.
It is given that the line width of the laser observed during the spontaneous emission is
1.5 ´ 10–3 nm. [Ans. 3.7]
CH A P T E R 9
HOLOGRAPHY

EXPECTED LEARNING OUTCOMES


After reading this chapter, the student will acquire the capability to:
• Differentiate between conventional photography and holography.
• Describe how holograms are produced, i.e., recorded and reconstructed.
• Discuss important applications of holography.

9.1 INTRODUCTION

We are now advancing into the 21st century, which is designated as the knowledge era. It has
been powered by innovations and revolutionary developments in Information and Communication
Technologies (ICTs) fuelled by the development of lasers. In Chapter 8, we discussed various
applications of lasers. In particular, mention was made about holography. In this chapter,
you will learn about it in detail. Though our emphasis will be on physical principles, references
will also be made to exciting new possibilities.
All of us cherish sweet memories/magnificent works of architecture/natural beauty recorded in
different phases of and times in our life. Many a time, some transient moments leave permanent
impact on our psyche. Such memorable moments are normally captured for posterity through a
regular photograph, which records a (2-D) real image projected on a photographic film or a
photosensitive surface. But in the process of photographic recording, the 3-D character of the
object is almost entirely lost. However, holography enables us to record an object in its complete
3-D grandeur.
In the language of physics, we say that holography is a technique of photography where
both amplitude (intensity) and phase distributions are recorded and the pictures obtained by this
technique have three-dimensional form. (In photography, the phase information is lost, which
results in the loss of depth perception of a 3-D object. This is essentially because the light
sensitive medium responds only to intensity variations. That is why we cannot change the
perspective of the image in the photograph by seeing it at different angles.) As such, a hologram
looks like a hodgepodge of specks, blobs, and whorls; it seems to bear no resemblance to the
339
340 Wave Optics

original object and therefore appears quite unintelligible. That is, a hologram does not carry the
image of the object but it carries a permanent signature of the object in the form of an intricate
interference pattern. And as with an object, we can see a different perspective of the image.
Holography was discovered by Denis Gabor in 1948, though he was not initially interested
in 3-D imaging. (His original aim—to improve the resolution of electron microscope through
wavefront reconstruction—remained unrealised due to non-availability of good coherent sources
at that time. Gabor worked with a mercury arc lamp as the source of light.) But once he
succeeded in recording a hologram, he realised the significance of his work. However, this
technique became a versatile practical proposition only after the advent of lasers and opened
fertile channels for major advances; it now finds important applications in diverse fields of
human activity. Gabor received Nobel Prize in Physics in 1971 for this work.
Nowadays, holograms are so common that even a senior school student is aware of it. These
seldom fail to surprise and mystify those who have not seen holograms earlier. But in 1971,
a jeweller in New York decided to exploit 3-D nature of image for material gain and displayed
a life-like 3-D image of a diamond necklace in his store window. The hologram caused quite
a sensation, rumours and myths. So much so that one day an elderly lady, who was convinced
that the image was the work of the devil, cracked the window.
Holography is a two-step process. The first step is called hologram recording. The second
step is holographic reconstruction. In Section 9.2, we have discussed the basic concepts involved
in the holographic technique using monochromatic light from a laser. Here you will also learn
about hologram recording and wavefront reconstruction. In Section 9.3, we have analysed the
process analytically. In Section 9.4, you will learn to obtain white light holograms. The applications
of holography in holographic interferometry, information processing, production of optical
equipment, data storage, security and other confidential work are discussed in Section 9.5.

9.2 BASIC PRINCIPLE OF HOLOGRAPHY

As mentioned earlier, holography is a two-step process. The first step—hologram recording—


consists of recording the amplitude and phase distributions of the object field on a high-contrast
photographic film as a stable interference pattern using an object wave and a reference wave.
(An object wave is the light diffracted/reflected by the object when it is illuminated by a
coherent source such as laser. The object wave is therefore signal bearing wave. The reference
wave is usually a plane or spherical wave driven from the same coherent source and falling
directly on the photosensitive plate bypassing the object.) The amplitude distribution in a hologram
is recorded in the same manner as in a regular photograph, i.e. in the form of density or opacity
variation across the photographic emulsion. However, this emulsion is completely insensitive to
phase variations. That is, the fundamental problem in holography is the recording of the phase
associated with the object wave. Through interferometry techniques, the reference wave is used
as a vehicle to convert phase variations into intensity variations, which can be recorded.
Note that these intensity variations invariably have periods less than one micrometre and cannot
be seen with unaided eye.
The second step—holographic reconstruction—comprises illumination of hologram with
the reconstruction wave. Usually, it is the same as the reference wave used in step 1. This leads
to reconstruction of a wavefront resembling (in all respects) the diffracted wavefront originating
Holography 341
from the object during the recording of the hologram. The reconstructed wavefront forms a
virtual image at the original position of the object, provided the geometries of the recording and
reconstruction processes are the same. (If you look through the hologram back at the original
position of the object, you will see a 3-D image of the object. Moreover, on moving the position
of the eye, you will detect parallax, i.e. observe different views of the object.)
You will recall that superposition of two coherent light waves having a constant phase
difference results in stable interference pattern. It means that if information about the phase is
to be recorded on a film, we must obtain stable interference. Conversely, to record a hologram,
we should have two coherent light waves. You may now ask: How do we record the phase and
how do we obtain coherent waves? To get insight of the basic physics, we first consider a simple
hologram of a point source.

9.2.1 Transmission Hologram of a Point Source

Holograms can be produced in different ways depending on the relative orientation of the object
and reference waves. When such waves travelling in the same direction superpose, we obtain
a series of bright and dark concentric rings. Such an interference pattern is called Gabor zone
plate, which is similar to the Fresnel zone plate (Chapter 5). It is called Gabor hologram or in-
line hologram. This is obtained with objects such as a wire mesh that have enough open spaces
between them.
Now refer to Figure 9.1(a). It shows two plane waves originating from two distant coherent
point sources. The object wave arrives at the photographic film at an angle q and the reference
wave is perpendicular to the film. (We can also imagine the situation where reference wave
arrives at the photographic film at an angle q and the object wave is perpendicular to the film.)
The superposition of these waves leads to formation of alternate regions of constructive and
destructive interference. The exposure is maximum in the region of constructive interference
and when the film is developed, these regions appear as black silver grains. On the other hand,
the regions of destructive interference appear transparent and the resultant photonegative film
will comprise several parallel, closely spaced opaque lines (perpendicular to the plane of the
figure) separated by transparent lines. Note that a photographic recording of interference fringes
results in a grating-like structure and the photonegative so obtained is transmission hologram
of the distant point source. (Such a hologram is also known as off-axis hologram.) If we look
at the transmission hologram, we would not see a 3-D image of the source because what we
obtain on the photographic plate is a pattern of interfering wavefronts.
To gain information contained in the hologram, we illuminate it by the reconstruction wave,
which is identical to the reference wave. Now view the transmission hologram from the side
opposite to the source of illumination [Figure 9.1(b)], where various orders of interference
fringes are produced. We can imagine the zeroth order wave as continuation of the reconstruction
wave through the grating. If you put your eye along it, you will see the reconstruction light
source shining in your eye. From Chapter 4, you will recall that in addition to the zeroth order
wave, the grating also produces two first-order waves, which leave the hologram at an angle on
the opposite side of the normal to the hologram. One of these appears to be continuation of the
original object wave with the same wavelength and travelling along the same direction.
For this reason, we say that the hologram reconstructs the object wave.
342 Wave Optics

Figure 9.1 Hologram of a distant point source: (a) superposition of object and reference waves produces
interference fringes on the film, which acts like a diffraction grating on development;
(b) On illumination by the reconstruction wave, the hologram emits a replica of the object wave.

You may now ask: Can we say with certainty that the reconstruction wave indeed leaves
the hologram at the same angle as that at which the object wave arrived? To discover answer
to this question, we note that during reconstruction, the first order diffraction wave leaves
the hologram at an angle that depends on fringe spacing. Since the wavelength and fringe
spacing d remain unchanged for the exposure and reconstruction, the angle also remains the
same. It means that spacing of the fringes depends solely on the angle between the impinging
waves; the fringes will be more closely spaced when the object wave reaches the film at a
larger angle.
Note that spacing of the fringes in the hologram contains the phase information of the object
wave, whereas the contrast of the fringes gives a measure of its intensity. That is, more intense
the object wave, greater will be the contrast and vice versa.

Variations in the amplitude of the object wave manifest as the variation in contrast of the recorded
fringe pattern. Larger the angle between the two interfering waves, more closely spaced will be the
fringes and vice versa. Therefore, variations in the phase of the object wave manifest as the variations
in the spacing of the fringes on the photographic record (the hologram). The photosensitive material
used in recording the hologram has high resolution-recording capability (~ more than 10 times that
used in regular photography) since a hologram is actually the recording of a high frequency interference
pattern.
An interesting property of a hologram of a diffuse scatterer is that even if it is broken up in different
pieces, each fragment contains complete information of the object. This has genesis in the fact that
the wave from each scattering point of the object illuminates the complete hologram.
Holography 343
Let us now consider a slightly more interesting case of transmission hologram of a point
source close to the film. Refer to Figure 9.2. It shows a plane reference wave and a spherical
object wave (diverging from the source S). In this case, the interference fringes on the film will
have unequal width. Can you explain it physically? This is because the object and reference
waves impinge on different parts of the film at different angles between them. As may be noted,
the fringe spacing is less at the top and more at the bottom of the film.

Figure 9.2 Transmission hologram of a point source in the vicinity of the film.

Once the film is developed, we obtain a non-uniform diffraction grating signifying unequally
spaced opaque and transparent regions. When this hologram is illuminated by the reconstruction
wave, each small patch acts as a diffraction grating. As before, the zeroth-order wave passes
straight through all parts of the hologram. Moreover, the hologram reconstructs the diverging
object wave: the first-order wave is seen to diverge at a larger angle at the top of the film than
at the bottom [Figure 9.3(a)]. The emitted wavefront is continuous across the face of the
hologram and a diverging wave leaves the hologram. The reconstructed wave forms a replica
of the original diverging object wave just as if the point source S were present and there were
no hologram [Figure 9.3(b)]. If you view this hologram from the far side, you will see a virtual
image of the point source of light at the original position of S. On moving the eye up and down,
you will get different views of the point source. This is essentially because the hologram is
non-uniform and you see the point source through different parts of the hologram. In short,
we perceive the original point source of light being in position, even though no point source
is really present there.
344 Wave Optics

Figure 9.3 Reconstruction of a hologram of a point source.

You may recall that in case of a distant point source, we obtained two first-order waves
when transmission hologram was illuminated by the reconstruction wave. One of these waves
was in the direction of the object wave beyond the plane of the film and the other was downward
at an angle on the opposite side of the normal to the hologram [Figure 9.1(b)].
So you may now ask: What about the second wave for a point source in the vicinity of the film
(screen)? For the hologram of Figure 9.2, this beam converges to point S ¢ on the observer’s side
of the hologram, as shown on Figure 9.4(b). It may be mentioned here that the point S ¢ is as
far in front of the hologram as the point source S is behind the film. Moreover, S ¢ is a real image
of S and appears on reconstruction. (It was not present during exposure). So we can say that,
on illumination, the hologram of a near point source acts simultaneously as a diverging lens
(producing the virtual image at S) and a converging lens (producing the real image at S ¢´ ).

Figure 9.4 Reconstruction using the hologram generated in Figure 9.2. (a) Second first-order wave,
which leaves the top of the hologram at a large angle downwards and the bottom of the
hologram at a smaller angle. (b) Formation of real image at S¢: Wavefronts converge to a
point in front of the hologram.
Holography 345
Sofar we have confined ourselves to the physical features of a hologram, which possesses
complete information about the object in the form of an interference pattern. Proceeding further,
we consider an object beam originating from two points sources. The wave from each source
will interfere with the reference wave. As a result, two interference patterns will be recorded
simultaneously and the final hologram reconstructs both sources on illumination by the
reconstruction wave (Figure 9.5).

Figure 9.5 Hologram of two point sources.

We can likewise extend this argument to three, four or more point sources. We now consider
an extended source, which can be thought of as consisting of several points.

9.2.2 Transmission Holograms of Extended Objects

As mentioned earlier, we can think of a 3-D object as made up of innumerable point sources.
When the source is illuminated with coherent light, Huygens wavelets originating from each
point interfere with the reference wave on the film and give rise to a very complex and non-
uniform interference pattern. (By Huygens’ Principle, these waves can be imagined to be
equivalent to a complex wavefront.) Upon reconstruction, each source point is reconstructed
and thus the entire object. The entire complex wavefront from the source that struck the film
is reproduced when reconstruction wave strikes the hologram. Note that the interference fringes
on the film and the resulting pattern of opaque and clear regions will also be quite complicated.
Nevertheless, when hologram is illuminated properly, order arises and we will see a realistic
image of the object.
We hope that now you will be able to state the basic differences between photography and
holography. The virtual image produced by a hologram appears in complete 3-D form and it
facilitated seeing something very much like the object did in the process of hologram recording.
In fact, the image manifests vivid realism. A photographic image lies in the plane of the film,
whereas a holographic image lies some distance behind the plane. A small part of the hologram can
reconstruct the entire scene, whereas a part of the photograph portrays only its part of the image.
Therefore, unlike in the case of a photograph, even if a part of the hologram is somehow destroyed,
entire information about the object is still available. That is why a hologram is a reliable method
of data storage. A hologram has immense information capacity. This has genesis in the fact that
a number of holograms can be recorded independently on the same film by changing the angle
of the reference beam. A single 100 cm2 hologram may contain at least one volume of
Encyclopaedia Britannica. (This is unlike a photograph.) To sum up, we can think of a hologram
as some sort of magic window in which the object beam has been frozen or locked.
The reference wave acts as the key in the reconstruction process and unlocks the (false) image.
346 Wave Optics

So far we have confined ourselves to the physical features of a hologram, which possesses
complete information about the object in the form of an interference pattern. We now discuss
its theoretical basis.

9.3 THEORY

We now know that superposition of the object wave and the reference wave results in the
production of a hologram. When viewed with unaided eye, this hologram does not even remotely
resemble the object so captured. But every portion of the hologram possesses complete information
on the amplitude and phase distribution of the object in the form of an interference pattern. The
process of obtaining image of the object is known as reconstruction. We now consider the
recording and reconstruction processes and discuss these analytically. For simplicity, we first
consider an in-line hologram, where object wave and the reference wave are assumed to move
in the same direction. The case of oblique incidence is slightly complex but more realistic and
will be considered subsequently.

(a) In-line hologram


Suppose that the photographic plate is in the x-y plane at z = 0. We can respectively represent
the object wave and the reference wave as
y1(x, y) = A1(x, y) cos[wt + f1(x, y)] (9.1)
and
y2(x, y) = A2(x, y) cos[wt + f2(x, y)] (9.2)
Note that the object wave arises due to superposition of waves from individual point sources on
the object. Further, the amplitude of the object wave A1(x, y), is a function of x and y
coordinates as it varies from point-to-point on the photographic plate. On the other hand, the
amplitude of the reference wave, A2(x, y), is almost independent of x and y indicating that it
is constant at all points on photographic plate. Moreover, the epoch of the reference wave
f2 (x, y) will be constant, if the reference wave falls normally on the photographic plate and a
function of x, y, if it is incident at an angle. However, the epoch of the object wave f1(x, y) will
be a function of x and y.
When these two waves reach the photographic plate, the total field distribution obtained by
their superposition is given by
y total(x, y) = y 1(x, y) + y 2(x, y)
= A1(x, y) cos[wt + f1(x, y)] + A2(x, y) cos[wt + f2(x, y)] (9.3)
As mentioned earlier, the photographic plate is sensitive only to intensity variations. Therefore,
to obtain the intensity distribution on the photographic plate, we calculate the time average of
(y total)2 and write
I(x, y) = á(y total)2 ñ
= á{A1(x, y) cos[wt + f1(x, y)] + A2(x, y) cos[wt + f2(x, y)]}2ñ
Holography 347
where the angular brackets denote time averaging. In expanded form, we can write it as
I(x, y) = A12 (x, y) ácos2(wt + f1)ñ + A22 (x, y) ácos2(wt + f2)ñ
+ 2A1(x, y) A2(x, y) ácos(wt + f1) cos(wt + f2)ñ
You will recall that the average value of cos2q over a cycle is 1/2. If we now use the
trigonometric relation 2 cosA cosB = cos(A + B) + cos(A – B), we can rewrite this result as

A12 A22 1
I ( x, y )   2 A1 A2 ¹ Ãcos(2Z t  I1  I2 )  cos(I2  I1 ) Ó
2 2 2
Note that the average value of cos (2w t + f1 + f2) is zero over a cycle. Thus, intensity of the
in-line hologram is given by
A12 A22
I ( x, y )   A1 A2 cos(I2  I1 ) (9.4)
2 2
Let us pause for a while and reflect on the implications of Eq. (9.4). This result shows that the
phase information of the object wave is also recorded in the intensity pattern on the photographic
plate. When this plate is developed, we obtain the hologram. That is how a hologram freezes
complete information about an object in the interference pattern.
Let us now consider the second step: the reconstruction process. You will recall that in this
process, the hologram is illuminated by a reconstruction wave. Let us denote the reconstruction
wave at the hologram plane by y3(x, y) and assume that it is in phase with the reference wave.
Therefore, we can write
y3(x, y) = A3(x, y) cos[wt + f2(x, y)] (9.5)
Now we wish to know the nature of the transmitted wave when the reconstruction wave is made
to fall on the hologram. The hologram is exposed in such a manner that the amplitude
transmittance, defined as the ratio of the transmitted field to the incident field, depends on the
intensity I(x, y) at the time of recording the hologram. By a suitable developing process, it is
possible to obtain a condition under which the amplitude transmittance will be linearly proportional
to I(x, y). Thus we can express the transmitted wave as
y4(x, y) µ y3(x, y) I(x, y)
= Ky3(x, y) I(x, y) (9.6)
On combining this equation with Eq. (9.4), obtained for incident intensity I(x, y) at the time of
recording the hologram, we get

Ë A12 A22 Û
\ 4 ( x, y ) K \ 3 ( x, y ) Ì   A1 A2 cos(I2 I
 1 )Ü
Í 2 2 Ý

On rearranging terms, we can write

( A12  A22 )
\ 4 ( x, y ) K \ 3  K\ 3[ A1 A2 cos(I2  I1 )]
2
For simplicity, we choose the constant of proportionality as equal to one and substitute the value
of y3 from Eq. (9.5) in the second term of the above expression to obtain
348 Wave Optics

( A12  A22 )
y4(x, y) = \ 3  [ A3 cos(Z t  I2 )] [ A1 A2 cos(I2  I1 )]
2
( A12  A22 )
= \ 3  A1 A2 A3 [cos(Zt  I2 ) cos(I2  I1 )] (9.7)
2
As before, using the trigonometric relation 2 cosA cosB = cos(A + B) + cos(A – B), we can
rewrite the expression for the wave transmitted through the hologram as

( A12  A22 )
y4(x, y) = \ 3  A1 A2 A3 cos(Zt  I1 )  A1 A2 A3 cos(Z t  2I2  I1 )
2 2 2
( A12  A22 ) AA A AA A
= A3 cos[Z t  I2 ( x, y )]  1 2 3 cos(Z t  I1 )  1 2 3 cos(Z t  2I2  I1 )
2 2 2
(9.8)
Let us pause here and interpret this expression for the transmitted wave. Note that Eq. (9.8),
which consists of three terms, gives the transmitted field in the plane z = 0. Let us consider each
of the three terms separately. The first term signifies a modulated reconstruction wave;
its amplitude has been modulated by the presence of the object wave. (Whereas A1 is a function
of x and y, the amplitude of the reference wave A2 is constant. As a result, this part of the
transmitted wave will travel, with slight attenuation, in the direction of reconstruction wave.)
But for the constant term (A2A3/2), the second term is analogous to the object wave (y1). It is
important to appreciate that illumination of the hologram with the reconstruction wave has led
to generation of a wave that is almost identical to the wave which originated from the object
itself during the recording process. This part of the transmitted wave forms a virtual image of
the object so that the effect of viewing this wave is the same as viewing the object itself.
The third term is similar to the object wave but the phase carries a negative sign, which denotes
the fact that the wave has a curvature opposite to the object wave. Since the object wave is a
diverging spherical wave, the third term represents a converging spherical wave; it leads to the
formation of a real image, which can be photographed by placing a camera there.
The reconstruction process along with various parts of the transmitted wave is shown in
Figure 9.6.

Figure 9.6 (a) Recording an in-line hologram; and (b) Reconstruction: The reconstruction wave diffracted
by the hologram results in a transmitted wave, which produces a real and a virtual image.
Holography 349
Note that the object is not present when image is being reconstructed. However, the transmitted
wave is identical to the reference wave.
We now consider the more realistic case of off-axis hologram.

(b) Off-axis hologram


As before, we assume that the photographic plate is in the x-y plane at z = 0 and the object wave
is given by Eq. (9.1). Now suppose that a plane reference wave propagating in the
x-z plane is inclined at an angle q with the z-axis. The reference wave at z = 0 is given by
\ 2 ( x, y ) A2 ( x, y ) cos[Z t  kx sin T ] A2 ( x, y ) cos[Z t  2SV x] (9.9)
where s = sinq/l denotes spatial frequency. This equation signifies that the phase varies linearly
with x. Note that there is no dependence on the y-coordinate because the reference wave has
been assumed to have its propagation vector in the x-z plane. Hence, the total field at the
photographic plate obtained by superposition of object and reference waves is given by
ytotal(x, y) = y1(x, y) + y2(x, y)
= A1(x, y) cos[wt + f1(x, y)] + A2(x, y) cos[wt + 2psx] (9.10)
As before, to obtain the intensity distribution on the photographic plate, we calculate the time
average of (ytotal)2 and write
I(x, y) = á(ytotal)2 ñ
= á{A1(x, y) cos[wt + f1(x, y)] + A2(x, y) cos[wt + 2psx]}2ñ
= A12 (x, y) ácos2(wt + f1)ñ + A22 (x, y) ácos2(wt + 2psx)ñ
+ 2A1(x, y) A2(x, y) ácos(wt + f1) cos(wt + 2psx)ñ (9.11)
Following the steps outlined above, it readily follows that
A12 A22

I ( x, y )  A1 A2 cos(I1  2SV x) (9.12)
2 2
Note the similarity between Eqs. (9.4) and (9.12).
Proceeding further in the same manner, you can easily convince yourself that the expression
for the transmitted wave in the plane z = 0 is given by
y4(x, y) = Ky3(x, y) I(x, y)
È A12  A22 Ø A22
= KÉ Ù A2 cos(Z t  2SV x)  K
A1 cos(Z t  2SV x)
Ê 2 Ú 2
A2 A
 K 2 1 cos(Z t  I1  4SV x ) (9.13)
2
On comparing Eqs. (9.8) and (9.13), we conclude that the first and the second terms respectively
signify the reconstruction wave and the object wave. And the third term results in the formation of a
real image of the object.
To investigate the effect of the term containing spatial frequency, let us suppose that the object
wave is also travelling along the z-axis so that f1(x, y) = 0. Then, the third term in Eq. (9.13) will
represent a plane wave propagating along the direction b = sin–1(2 sinq). It means that the effect of
the term 4psx is to rotate the direction of the converging wave. Hence, the third term on the RHS
350 Wave Optics

of Eq. (9.13) represents a conjugate wave, which propagates along a direction different from the
directions of the object and the reconstruction waves.
Having discussed the theoretical basis of a hologram, we now proceed to know the practical
requirements necessary for obtaining a hologram.

9.3.1 Practical Requirements

So far we have confined our discussion to the physics of holography, which is essentially an
interference phenomenon. Therefore, in a practical situation, a few requirements related to
equipment and coherence must be satisfied to be able to obtain a hologram. Some of these
include the following:
• The interference pattern should be stable. To do so, we have to ensure
– spatial coherence so that waves scattered from different regions of the object interfere
with the reference wave; and
– that the maximum path difference between the object wave and the reference wave
does not exceed coherence length.
• The reconstruction wave should have the same wavelength as the reference wave.
• The recording arrangement—the film, the object and the mirrors, if used, should be
stable, in fact, completely free from vibrations with respect to one another during the
recording process. Otherwise, the interference pattern will be hazy.
• The photographic plate must have high resolution.
To appreciate the importance of the last requirement, we recall that when the object and/
or reference wave fall on the photographic plate at an angle, the interference fringes are closely
spaced. In fact, for large angle incidence, these fringes are so closely spaced (»10–6 m) that even
the best emulsion fails to record resolvable fringes. This necessitates use of very special kinds
of fine grain films. But such films increase exposure time, which makes the stability requirement
more stringent. A possible way out of this problem is to use high power laser beam and/or
holographic materials 649F Kodak or 10E 75/8E 75 Agfa-Gaevert films and plates.
In the holograms we have considered so far, the reconstruction wave was assumed to
emanate from a monochromatic source, usually a laser, and have the same wavelength as the
object wave used for exposing the film. You may now logically ask: Can we not construct a
hologram that derives reconstruction wave from white light? We will now discuss production
of white light holograms in brief.

9.4 WHITE LIGHT HOLOGRAMS


(a) Reflection holograms
Refer to Figure 9.7. Two plane waves incident on a thick-emulsion film from opposite directions
produce a standing wave. When this film is developed, silver grains will lie in the layers
throughout the emulsion. On bleach processing such a film, we obtain layers of varying index
of refraction. If one of these waves originated from a distant point source and the other as
reference wave, the developed film can be thought of as a hologram.
Holography 351

Figure 9.7 A reflection hologram of a distant point source. (a) Exposure of thick-emulsion film by
object and reference waves. The hatched lines signify standing waves. (b) Reconstruction
of object wave.

If this hologram is illuminated by a reconstruction wave of the same wavelength as the


reference wave used for exposure, the waves reflected by successive layers interfere constructively
to form a replica of the original object. In this way, we obtain reflection hologram. Note that
it has to be viewed from the side of incidence of reference beam.
Now let us suppose that this hologram is illuminated by white light. We observe that for the
waves corresponding to the wavelength used for exposure, all other waves reflected by successive
layers interfere destructively and annihilate the effect of each other. It means that we will obtain
only the desired wavelength (same as the reference wave) after reflection by silver layers.
Thus by illuminating the hologram by white light, we can reconstruct the object wave.
Note that in white light reflection holograms, the reference wave is quite often made to fall
obliquely, rather than normally, on the photographic film. Similarly, the object waves from extended
sources strike the film at different angles. To understand how reconstruction takes place in this
more realistic case, refer to Figure 9.8(a), which depicts the object and reference waves striking
the film at some arbitrary acute angles. The solid lines represent the positions of instantaneous
crests of the object and reference waves and the dots signify constructive interference due to
superposition of some of these crests. At these points, the silver emulsion will be exposed.
You may now ask: Does constructive interference takes place only at the points denoted by
the dots? The answer to this question is: There are many other points which receive crests of
the object and reference waves simultaneously. To determine these, take two pieces of paper and
place one of these along a wavefront of the reference wave and the other along a wavefront of
the object wave shown in Figure 9.8(a). What do you observe? The edges will cross at a dot
and emulsion will be exposed there. Now slowly advance each piece of paper forward (as if the
wavefronts were propagating in opposite directions) to successive wavefronts. Note that the
object and reference wavefronts intersect at several points along the slanted dashed lines.
The silver emulsion will be exposed along these lines [Figure 9.8(b)].
Next, we illuminate the hologram (the developed film) at a suitable angle with the
reconstruction wave of white light. We will observe that this light is reflected by the silver layers
352 Wave Optics

to form the reconstruction wave, as shown in Figure 9.8(c). The law of reflection ensures that
the reconstructed wave travels in the proper direction and the effect of all but one wavelength
is annihilated by the interference between waves reflected from silver layers.

Figure 9.8 Reflection hologram of a distant point source such that the light strikes the film at an angle
q with the horizontal axis. (a) Exposure of thick-emulsion film. The reference wave arrives
at an angle qref. (b) The fringes where silver emulsion will lie in the developed hologram.
(c) Reconstruction of the object wave.

Figure 9.9 shows an experimental arrangement for producing white light reflection holograms.
Coherent, monochromatic light from the laser is reflected by a mirror M1 on to a beam splitter
(BS). Part of the beam reflects off the beam splitter and travels to another mirror, M2 and then
through a diverging lens, L1 falls on the film. The lens spreads the wave so that the entire film
is exposed. This beam is the reference beam. The other part of the original beam is transmitted
directly through the beam splitter to a mirror M3 and another diverging lens, L2 spreads the
Holography 353

Figure 9.9 Schematics of experimental arrangement for producing white light reflection hologram.

beam to illuminate the entire object. Finally this light strikes the film. This wave is the object
beam. The beam splitter also controls the intensity of the reconstruction wave. Note that the
reference and the object waves have been derived from the same monochromatic source so that
these are coherent and produce interference pattern on the film.
All the optical components are placed on a heavy isolation table, which ensures that the
arrangement is free from the impact of outside vibrations. What do you think will happen if
the optical components were shaken relative to each other? In such as eventuality, it is possible that
the regions of constructive and destructive interference could smear over each other and sharp
fringes may not form in the final hologram. Similarly, the object must remain stationary; the
motion of the object can also lead to smeared fringes in the final hologram and ruin the image.
Note that the object beam and the reference beam strike the thick emulsion film from
opposite directions. And during reconstruction, white light, say from an ordinary spot-light, is
made to fall on the hologram at the same angle as the reference beam.
Another arrangement for producing white light reflection holograms is shown in Figure 9.10.
Here no beam splitter is used; the beam is made to pass through the emulsion and is reflected
back by the object—the object beam. This beam interferes with part of the original beam that
did not pass through the film—the reference beam. Thus these two beams reach the film from
the opposite sides.

Figure 9.10 Schematics of a simple single beam set-up for production of white light
reflection holograms.
354 Wave Optics

(b) Transmission holograms


In Section 9.2, we discussed transmission holograms and discovered that monochromatic light
was necessary for viewing it. You may now ask: Can we use white light to view a transmission
hologram just as in case of reflection holograms? The answer to this question is that we can
do so, provided a thick-emulsion film is exposed with object and reference beams from the same
side. Refer to Figure 9.11. It illustrates the basic principle of a white light transmission hologram.
Just as with the white light reflection hologram, layers of silver are exposed in thick emulsion.
When the hologram is illuminated by white light, different colours are reflected by successive
layers and interfere to produce reconstruction beam of the desired wavelength.

Figure 9.11 Exposure of thick-emulsion film to form a white-light transmission hologram. When
white light is used for reconstruction of image, the desired wavelength is obtained by
interference of waves by successive layers of silver.

It may be mentioned here that most white light transmission holograms are bleach processed
so as to improve the brightness of the final image. Since each wavelength in white light will
interfere to form its own image at slightly different positions, it is quite possible that white light
transmission hologram may exhibit chromatic aberration. Such a colour blurring can be drastically
reduced by the so-called rainbow holography. We will not go into these details here for limitations
of the scope and space.
So far we have discussed the basic principles of physics governing the production of
reflection and transmission white light holograms. Now you may be eager to know the applications
of holography, which are so many and so varied that these cover practically every area of human
activity. In the following paragraphs, we will discuss some of these.

9.5 SOME PRACTICAL APPLICATIONS OF HOLOGRAPHY

The principle of holography has been put to a wide range of applications other than recording
images. The most important areas are interferometry, metrology, astronomy, image processing,
Holography 355
data storage, biomedical techniques and pattern recognition, among others. You will now learn
about some of these.

1. Holographic interferometry
One of the most outstanding applications of holography is in the field of interferometry, which
is used for precise measurements (and comparison) of wavelength of light. Using the techniques
of interferometry, we can also test for stresses, strains, deformations and vibrations in engineering
structures. One area of holographic interferometry is known as non-destructive testing. In this
technique, we first obtain hologram of an object and process it. Thereafter, it is put back in
exactly the same location it occupied during exposure. Suppose that this hologram is illuminated
with the original reference wave. What do you expect to observe? On the basis of our previous
knowledge, we can safely conclude that we will ‘see’ an image of the original object. And if the
original object is not removed, we will see it also. In fact, the object and its image will appear
identical. But if the object has been deformed somehow, it will generate a new wavefront.
Interference fringes will appear due to superposition of the frozen wavefront and the new
wavefront. These give us information to optical interferometric precision, i.e. to an accuracy of
0.1 mm or less.
It may be mentioned here that this technique can be used to detect optical path length
variations in transparent media and facilitates visualization and analysis of fluid flow. Moreover,
the results obtained through this technique for determining changes in the shape of a surface
over large areas cannot be attained as conveniently and efficiently by any other method.

2. Double exposure holographic interferometry technique


The technique of double exposure holographic interferometry is very similar to the real-time
holographic technique described in the above paragraph. In this technique, two superimposed
holographic exposures of an object—one before it is stressed and the other after it has been
deformed—are obtained using the same reference wave on one photographic plate. (It is essentially
based on the ability of the holographic process to release the object wave which can be obtained
in the reconstruction stage and made to interfere with waves which exist at other times.)
What will you observe when this hologram is reconstructed by illuminating it with a reference
wave? Obviously, you expect two object waves, one each corresponding to the unstressed and
the stressed object, to emerge from the hologram. These reconstructed object waves interfere
resulting in a fringe pattern characteristic of the strain on and deformation of the object.
To get a quantitative idea of the fringe pattern, we assume that the deformation of the object
alters only the phase distribution. Therefore, let us represent the object wave in the hologram
plane when the object is in its natural shape as
y1(x, y, t) = A(x, y) cos[wt + f1(x, y)] (9.14)
Similarly, let the object wave corresponding to a deformed/stressed object be
y2(x, y, t) = A(x, y) cos[wt + f2(x, y)] (9.15)
On reconstruction, both these object waves emerge from the hologram and the intensity due to
interference of the two waves will be obtained by time-averaging the square of the resultant amplitude:
356 Wave Optics

I(x, y) = á{A(x, y) cos[w t + f1(x, y)] + A(x, y) cos[w t + f2(x, y)]}2ñ


We can easily simplify this expression using simple properties of the cosine function and obtain
I(x, y) = A2(x, y) + A2(x, y) cos(f2 – f1) (9.16)
Hence, depending on the phase difference between the two object waves, we will obtain
constructive and destructive interference characterized by bright and dark fringes superimposed
on the object:
f2 – f1 = 2mp m = 0, 1, 2, ... (Constructive interference)
S
= (2m  1) m = 0, 1, 2, ... (Destructive interference) (9.17)
2
The holographic interferometry technique describes changes in the shape of a surface but now
techniques are also available which can reveal the shape of the surface itself. One such technique
is called two-wavelength holography. In this technique, we record two superimposed holograms
of an object using light of two different wavelengths for two exposures. When the recorded and
the processed hologram are illuminated with a reference wave, an image superimposed with
interference fringes or contour lines of equal depth is obtained. It is worth pointing out here that
the contour interval between adjacent contour lines is inversely proportional to the difference
in the wavelength of the light used for the two exposures.

3. Information storage
Information can be stored and retrieved more efficiently in the form of holograms than in the
form of real images. This follows from two characteristics of holograms. One, light from each
point object spreads over the entire hologram so that dust or scratches on it do not degrade the
quality of image. Two, it is possible to record several holograms on the same holographic plate
by changing the angle between the object and the reference beams slightly for each hologram.
(A hologram reconstructs the holographic image if the reconstruction beam is incident on it at
the correct angle.) This is particularly true if the recording material is thick. (Is it not analogous
to the manner the information is stored in human brain?) On reconstruction, depending upon
the angle of incidence of the illuminating wave, a particular holographic image will become
visible. Earlier, this technique had limited use because of limitations imposed by the quality of
recording materials but owing to improvements with time, holography has now become an
extremely useful technique for information storage.

4. Holographic microscope
You may recall that Gabor discovered holography while attempting to improve the resolving
power of an electron microscope. Since holographic microscope records information about the
depth of a field (object), holography was expected to find extremely valuable applications in the
study of microscopic events, transient or stationary. To understand how a holographic
microscope works, refer to Figure 9.12. A monochromatic light from a laser is made to fall on
a beam splitter. One part of the beam is made to fall on the object (specimen) and reaches the
film through a microscope. The other part is allowed to reach the film using an arrangement of
plane mirrors. These beams generate an interference pattern on the film, producing a
hologram. The reconstructed image can be seen in any desired cross-section, moving back and
Holography 357

Figure 9.12 Schematic diagram of a holographic microscope.

forth throughout the depth of the reconstructed image. Note that no focussing of microscope is
needed in this case.
A very innovative use of this technique is in the study of transient phenomena. It is not
possible to first locate the position and subsequently record such events using ordinary microscopic
techniques. However, a hologram freezes the event for posterity and we can study the phenomenon
asynchronously at our convenience by analysing the reconstructed image throughout its depth.

5. Optical elements
Another exciting application of holography is in making optical elements such as diffraction
grating. In Chapter 5, we discussed how diffraction gratings were produced by ruling lines with
a diamond point. But now it is possible to create gratings holographically. Holograms generated
by interference of two plane waves or a plane wave and a spherical wave are of particular
interest. It may be mentioned here that holographic gratings find special applications.
For example, holographic gratings made on substrate shapes scatter light minimally and help
avoid the need of additional lenses and mirrors while being used with a spectrometer.

6. Dynamic holography
We now know that in static holography, a permanent hologram is produced by sequentially
following the steps such as recording, development and reconstruction. We now also have
holographic materials which can record a hologram in a short time and we do not have to resort
to developing process. This allows us to use holography to perform simple operations in an all-
optical way. Such real-time holograms include phase-conjugate mirrors (‘time-reversal’ of light),
optical cache memories, image processing (pattern recognition of time-varying images) and
optical computing.
358 Wave Optics

The amount of processed information can be very high, of the order of terabit per second,
since the operation is performed in parallel on the entire image. This compensates for the fact
that the recording time (» ms) is very long as compared to the processing time of an electronic
computer. However, optical processing performed by a dynamic hologram is not as flexible as
electronic processing. For dynamic holography, we need non-linear materials such as
photorefractive crystals.

7. Pattern recognition
Pattern recognition, also known as character recognition, is a promising application of holography.
The hologram of the letter to be recognized is generated by superposition of signal (object) and
reference wavefronts. The hologram of the letter to be read is illuminated with light from
another identical letter and the plane wavefronts that arise can be focused into a bright spot,
which can be recognised photoelectrically. If these wavefronts arise from any other character,
we will not obtain a diffused patch of light rather than perfectly plane wavefronts. In this way,
we can scan a given matrix of characters and determine whether or not a particular character
is present in it.

8. Holographic techniques in video disk systems


In recent years, a lot of work has been reported in using holographic techniques in video disk
systems. In this technique, we record an image of a photograph instead of a 3-D image. This
helps to reduce sensitivity of the hologram to scratches and eliminate the need for a hologram
projector to have the stop-go action of a regular movie projector. In fact, it has already brought
holographic cinema in the domain of reality and holographic television is likely to make its
appearance in the near future.

9. Other applications of holography


Other promising applications of hologram include acoustic imaging. It seeks to replace the
use of X-rays for some situations with acoustic waves. The acoustic waves are recorded
as a hologram, and in the reconstruction step, visible light is used for forming an image.
The measuring techniques included within acoustic holography are becoming increasingly popular
in transportation as well as vehicle and aircraft design.
An innovative application of holography is in security holograms. These are very difficult
to forge because these are replicated from a master hologram, which requires expensive,
specialized and technologically advanced equipment. Nowadays, these are used in national
currencies, credit cards, passports, ID cards, books, DVDs, artworks and sports equipment.
Their use helps to check plagiarism and infringement of intellectual property rights.
Let us now summarise what you have learnt in this chapter.

9.6 SUMMARY

• Holography is a novel technique to obtain 3-D record of an object/ scene/building. It is


also known as wavefront reconstruction photography.
Holography 359
• Holography is a two-step process. The first step is called hologram recording. The second
step is holographic reconstruction.
• A hologram is an interference pattern obtained by superposition of an object wave and a
reference wave derived from the same coherent source of light such as laser.
• The 3-D picture of the object frozen in the interference pattern is obtained by illuminating
the hologram by a reconstruction wave, which is invariably identical to the reference
wave.
• Suppose we represent the object wave and the reference wave as

\ 1 ( x, y ) A1 ( x, y ) cos[Z t  I1 ( x, y )]
and
\ 2 ( x, y ) A2 ( x, y ) cos[Z t  I2 ( x, y )]
Then, intensity distribution of the in-line hologram is given by

A12 A22

I ( x, y )  A1 A2 cos(I2  I1 )
2 2
• When in-line hologram is reconstructed, the expression for the wave transmitted through
the hologram is given by

( A12  A22 ) AA A
y4(x, y) = A3 cos[Z t  I2 ( x, y )]  1 2 3 cos(Z t  I1 )
2 2
A1 A2 A3
 cos(Z t  2I2  I1 )
2
where reconstruction wave at the hologram plane is in phase with the reference wave and
is given by
y3(x, y) = A3(x, y) cos[wt + f2(x, y)]
• For an off-axis hologram, if the reference wave at z = 0 is given by

\ 2 ( x, y ) A2 ( x, y ) cos(Z t  kx sin T ) A2 ( x, y ) cos(Z t  2SV x )


the intensity distribution is given by

A12 A22

I ( x, y )  A1 A2 cos(I1  2SV x )
2 2
and wave transmitted through the hologram is given by

È A12  A22 Ø A22


\ 4 ( x, y ) É
Ê
Ù A2 cos(Z t  2SV x ) 
Ú
A1 cos(Z t  2SV x )
2 2
A22 A1
 cos(Z t  I1  4SV x )
2
where s = sinq/l denotes spatial frequency.
360 Wave Optics

• To obtain a hologram, we must ensure that the interference pattern is stable, the
reconstruction source has the same wavelength as the reference source, the recording
arrangement is completely free from vibrations with respect to one another during the
recording process and the photographic plate must have high resolution.
• Holography finds many and varied applications. Some of these include holographic
interferometry, holographic microscopy, information storage, and confidential work.

REVIEW EXERCISES

1. Suppose that a hologram is formed with a point object and a plane reference wave, which
is incident normally on the photographic plate. Show that the resultant fringe pattern is
circular and centered at the origin.
2. Suppose that the incident plane wave in Problem 9.1 is incident at an angle q with the
z-axis. Calculate the interference pattern.
3. Discuss the differences between the intensity distribution in a Fresnel zone plate and the
one obtained in Problem 9.1.
4. The amplitude of object and reference waves used in double exposure holographic
interferometry technique is 10–4 m. If the initial phases are p/6 and p/3, respectively,
express the time-averaged intensity of the reconstructed interference pattern in SI units.
[Ans. 1.87 ´ 10–8]
5. How will the result of Problem 9.4 be modified, if the amplitude of the reference wave
changes to 1.5 × 10–4 m? [Ans. 2.92 ´ 10–8]
CH A P T E R 10
FIBRE OPTICS

EXPECTED LEARNING OUTCOMES


After reading this chapter, the student will acquire the capability to:
• Explain the process of light transmission through optical fibres.
• Describe characteristic features of different types of fibres.
• Distinguish between pulse dispersion and material dispersion.
• Derive expressions for pulse dispersion in optical fibres.
• Discuss important applications of optical fibres.
• Solve numerical problems.

10.1 INTRODUCTION

You may have witnessed light displays following the opening and closing ceremonies of the
Commonwealth Games held in New Delhi in 2010 or Beijing Olympics in 2008. Similarly,
you may have spoken to your near and dear ones across oceans and continents using Skype or
messenger search engines. The use of Internet for e-mailing, chatting and/or looking for useful
academic information are routine activities nowadays in the life of a college/university students.
Have you ever thought how such communication came into the realm of possibilities? What
helped us to communicate at optical frequencies (» 1015 Hz) and how? You will discover
answers to these and such other related questions in this chapter.
We all know that communication stands for transfer of information in the form of speech,
image or data from one point/place/person to another point/place/person. Optical communication
has been the most preferred vehicle for transmitting information over long distances. In 800
B.C., the Greeks used fire to signal victory in a war, call for help when in distress or raise an
alert against the enemy. In fact, the news of the downfall of Troy towards the end of sixth
century B.C. was transmitted over a distance of about 500 km, from Asia Minor to Argos,
through a chain of fire signal relay stations. During the second century B.C., optical signals
were encoded to send messages. However, there was no worthwhile development in optical
communication systems till the end of 18th century. This was primarily due to non-availability
361
362 Wave Optics

of a reliable light source and the restrictions imposed by the line of sight transmission paths.
(These were affected adversely by the terrain and the atmospheric conditions such as dust
particles, rain and fog.) Nevertheless, interest in optical communication was revived with the
discovery of the laser in 1960. And optical communication in modern sense implies use of
techniques to convert electrical signals into light signals at one end, transmission of these
signals through thin glass guides called fibres and conversion of light signals back into electrical
signals at the other end.
As we know, Samuel Morse invented the telegraph in 1835. This heralded the beginning of
the era of electrical communications. In 1876, Graham Bell invented telephone, which facilitated
use of electrical signals for transmission of speech. (Some people believe that telephone was
invented by Antonio Meucci in 1860.) The idea of using light waves for communication can be
traced to Graham Bell when he invented photophone in 1876. In this experiment, speech was
transmitted by modulating light waves, which propagated in air to the receiver. Generation of
radio waves by Hertz in 1887–88 and wireless transmission of radio signals by Marconi in 1895
proved defining milestones in the development of telecommunication. (In 1895, Indian scientist
J.C. Bose working at Calcutta (now Kolkata), is also reported to have given public demonstration
of radio transmission, using electromagnetic waves of wavelengths 25 mm to
5 m to ring a bell remotely and to explode some gunpowder.) Further refinements made long
distance transmission possible using modulation techniques with radio waves as carriers of
information (speech). The shift from amplitude modulation to phase modulation facilitated to
extend reach and improve quality of transmission. In fact, the telecommunication traffic based
on co-axial cables, radio and microwave links and wire-pair cable increased very rapidly and
in the second-half of the 20th century, it was felt that it could soon saturate. This realization
re-ignited interest in optical communication. Since a carrier wave of higher frequency carries
more information, communication at optical frequencies has obvious advantages over
communication in the radio and microwave frequency ranges.
The first modern optical communication was based on transmission of laser beam—coherent
and spectrally pure light—through open atmosphere. However, its inadequacy for long distance
transmission was exposed as scattering and absorption of light caused severe attenuation and
distortion (by dust particles, water vapours and fog). Thus, to minimize losses due to the
vagaries of terrestrial atmosphere and protect the signal carrying light beam as it propagated
from one place to another, need was felt to provide a guiding medium. Initially, metallic and
non-metallic waveguides were fabricated but these were not suitable because of enormous
losses. Tyndall discovered that light could be transmitted through optical fibres by the phenomenon
of total internal reflection—an optical waveguide in the form of an optical fibre. However,
heavy losses in fibres posed a serious problem.
In 1968, the typical fibre losses were about 1000 dB per kilometre. But it was immediately
realized that these high losses arose due to impurities in the fibre material. A dramatic breakthrough
came in 1970. Using a pure silica fibre, fibre losses were reportedly reduced below 20 dB per
kilometre. Thus optical fibre communication became an engineering reality. Since then,
there has been considerable technological progress in this field. Recent advances have made
fibre optic communication a reliable, versatile and viable proposition for local, trunk and submarine
under-sea inter-continental applications. Optical fibres are hair-thin thread-like structures made
of transparent glass or plastic and surrounded by a transparent cladding. The optical phenomenon
Fibre Optics 363
responsible for propagation of light along an optical fibre is total internal reflection. You will
learn about it in Section 10.2. In Section 10.3, we have discussed the types of optical fibres.
Communication through fibres, which is post-1970 development, is discussed in Section 10.4.
You will also learn the advantages of optical fibres and what efficient optical communication
demands as far as optical fibres are concerned. Dispersion and losses in fibres are discussed in
Section 10.5.

10.2 OPTICAL FIBRES: WORKING PRINCIPLE

Refer to Figure 10.1(a). It shows a typical optical fibre, which consists of a central cylindrical
glass or plastic core surrounded by a cladding of the same material. However, the refractive
index of cladding material is chosen to be lower (» 1%) than that of the core. This is achieved
by doping the silica core with germanium while the cladding is of pure silica. Germanium-
doped core has higher refractive index. The variation of refractive index for a typical optical
fibre is shown in Figure 10.1(b). This assembly is further covered by an outer plastic coating
to guard against loss of light and protect the fibre against damages arising from chemical,
mechanical, environmental or manual handling. The diameter of the core is usually in the range
of 5 µm to 125 µm while the diameter of the cladding is in the range of 100 µm to 150 µm.
The diameter of the plastic coating is around 250 mm.

Figure 10.1 (a) Schematics of an optical fibre; and (b) variation of refractive index with distance from
the axis of the core.

In your earlier classes, you have learnt that if we use ray propagation geometry to depict
propagation of light from a denser medium to a rarer medium, it bends away from the normal
at the interface as it undergoes refraction at the interface [Figure 10.2(a)]. At a particular angle
of incidence, called critical angle (qc), light grazes along the boundary of the interface
[Figure 10.2(b)]. When the angle of incidence is greater than the critical angle (qi > qc ), the light
is reflected back into the denser medium. This optical phenomenon is known as total
internal reflection [Figure 10.2(c)] and forms the working principle of optic-fibre based
communication.
364 Wave Optics

In an optical fibre, when light from the germanium-doped silica core of refractive index n1
reaches the core–cladding interface, it is refracted away from the normal, since refractive index
of the (silica) cladding (n2) is slightly less than the refractive index of the core. The angles of

Figure 10.2 Total internal reflection.

incidence and refraction are related to the refracted indices of the core and the cladding by
Snell’s law:
n1 sinqi = n2 sinqr
or
sin Ti n2
= (10.1)
sin T r n1
As we increase the angle of incidence, the refracted ray moves further away from the normal
and for a particular angle of incidence, the refracted ray grazes along the interface, i.e. it is
perpendicular to the normal. This is the limiting case of refraction and the angle of incidence
is known as the critical angle, which is denoted as qc.
Using Eq. (10.1) for this case, it readily follows that

n2 È n2 Ø
sin T c À Tc sin 1 É Ù
(10.2)
n1 Ê n1 Ú

For the glass–air interface, n1 = 1.5 and n2 = 1.0, so that qc = sin–1(1.0/1.5) = 41.8º. On the
other hand, for glass–water interface (n1 = 1.5 and n2 = 1.33) so that qc = sin–1(1.33/1.5) =
62.7º. Similarly we can calculate the critical angle for any other interface.
Fibre Optics 365
We now advise you to answer a Practice Exercise.

Practice Exercise 10.1 The refractive indices of diamond, glass and air are 4.52, 1.5 and 1.0
respectively. Calculate the critical angles for diamond–glass and diamond–air interfaces.
[Ans. 19.4º; 12.8º]

When a light wave is made to enter the core of an optical fibre such that it hits the core–
cladding interface at an angle qi > qc, it will undergo total internal reflection at that surface.
(The angle of incidence at which light entering the fibre undergoes repeated total internal
reflections is known as acceptance angle. We will denote it by qa.) Further, because of cylindrical
symmetry in the fibre structure, light will be guided through the core by repeated total internal
reflections at the upper and lower core–cladding interfaces. This is depicted in Figure 10.3.
It may be mentioned here that due to this ‘guiding’ property, optical fibres are also known as
optical waveguides. Moreover, discontinuities and/or imperfections at the core–cladding interface
are likely to result in loss of light due to refraction at those centres/points. In other words,
all light entering the fibre core may not continue to propagate along its length.

Figure 10.3 Transmission of light in a perfect optical fibre.

You may now like to ask: Will light undergo multiple total internal reflections and guided
similarly even in a bent optical fibre? The answer to this question is in the affirmative, provided
that the angle of incidence is greater than the critical angle at the curved portion. This is
illustrated in Figure 10.4.

Figure 10.4 Propagation of light through a bent fibre by total internal reflection.

The fact that the phenomenon of total internal reflection can be used to guide light
was demonstrated by Colladon in 1841 through a simple experiment. This experiment was
366 Wave Optics

repeated by Babinet in 1842 and is depicted in Figure 10.5. A beam of light was sent in a water
jet. It suffered total internal reflection at the water–air interface and travelled along the curved
path of water coming out of an illuminated jar. It suggested the possibility of using bent fibres
to guide light, provided that even at the curved portion, the angle of incidence exceeded the
critical angle.

Figure 10.5 Schematics of Colladon’s set up used to demonstrate


propagation of light through a water jet.

You may now like to answer a Practice Exercise.

Practice Exercise 10.2 What will happen if the refractive index of the core were less than
that of the cladding in an optical fibre?

10.2.1 Numerical Aperture

We now know that the critical angle, as also the acceptance angle, depends on the refractive
indices of the materials of the core and the cladding. It would, therefore, be desirable to know
the exact nature of the relationship between acceptance angle and the refractive indices of
the core, cladding and air. To do so, refer to Figure 10.6, which shows that a ray of light is
incident from air (refractive index n0 » 1) on the fibre core (germanium-doped silica; refractive
index n1) at an angle qi(<qa, the acceptance angle). Using Snell’s law at the air–core interface,
we can write
n0 sinqi = n1 sin qr (10.3)
Here qr denotes the angle of refraction.

Figure 10.6 Ray diagram for light incident on one end of an optical fibre
at an angle less than the acceptance angle.
Fibre Optics 367
This ray is incident at the core–cladding interface at an angle f. From the right-angled
triangle ABC, we note that
S
I  Tr
2
Using this result in Eq. (10.3), we get
n0 sinq i = n1 sin(90º – f) = n1 cosf (10.4)
Now recall the trigonometric relation sin2f + cos2f = 1 so that we can rewrite Eq. (10.4) as

n0 sin Ti n1 1  sin 2 I (10.5)

For the ray incident of the core–cladding interface to undergo total internal reflection, Eq. (10.2)
implies that f must be greater than the critical angle and we can write

n2
sin I !
n1

On substituting this value of sinf in Eq. (10.5), we get

n22
n0 sinqi < n1 1 
n12
1/2
È n2  n 2 Ø
< n1 É 1 2 2 Ù n12 2
 n2 (10.6)
Ê n1 Ú

Note that for the limiting case of incidence, the maximum angle of incidence—the acceptance
angle—defines the light gathering capacity of the optical fibre. Therefore, we can write
qi < qa
and
n0 sin T a n12  n22 (10.7)
Equation (10.7) defines the numerical aperture (N.A.) of the fibre. Numerical aperture is a very
useful measure of the light gathering capacity of an optical fibre. Note that it depends only on
the refractive indices of the core and cladding materials. It may be mentioned here
that the ray treatment of light does not hold when the fibre optic diameter is less than 8 mm.
This is essentially because geometrical optics excludes the possibility of interference effects,
which become important for such thin structures.
Within the framework of geometrical optics we can say that if a cone of light is incident
on one end of an optical fibre, it will be guided through it provided the semi-angle of the cone
is less than qa:
Ë (n12  n22 )1/ 2 Û
Ta sin 1 Ì Ü
Í n0 Ý (10.8)
368 Wave Optics

If the angle of incidence is greater than qa, the light will be refracted into the cladding material.
For a typical silica fibre, the refractive indices of the core and the cladding are 1.50 and 1.47
respectively. If the fibre is immersed in water (n0 = 1.33), its numerical aperture and acceptance
angle are respectively given by

N.A. n12  n22 1.502  1.472 2.25  2.16 0.2985


È 0.2985 Ø
Ta sin 1 É Ù sin 1 (0.2244) 12.97º
Ê 1.33 Ú

It may be mentioned here that if all rays in the acceptance cone defined by q i = 0 and q i = q a
are launched in a short optical fibre, the light emanating from the other end of the fibre will
also appear as a cone of semi-angle qa. If this beam is made to fall normally on a sheet of white
paper and we measure its diameter, we can easily estimate the NA of the fibre.
You should now answer a Practice Exercise.

Practice Exercise 10.3 The refractive indices of the core and the cladding are 1.45 and 1.40
respectively. If light is incident from air, calculate its acceptance angle. [Ans. 22.2º]

In the previous sub-section, you have learnt that interference effects become important for
thin optical fibres. When interference effects are considered, the fibre is known to support a
discrete number of guided modes. (A mode is a mathematical concept to describe the nature of
propagation of electromagnetic waves in a waveguide. It specifies a transverse field distribution
associated with the electromagnetic waves which propagate along the fibre without any change
in its field distribution, except for a change in phase.) Can you guess the factors which determine
the number of modes in an optical fibre? If you think that the number of modes is a function
of the refractive indices of core and cladding, core diameter and wavelength of light used, you
are thinking logically and correctly. It means that if we fix any two of these factors, we can
change the number of modes by varying the third variable. In general, the core, cladding and
wavelength of light are fixed but the core diameter is changed to obtain the single-mode of
operation. Therefore, the optical fibres are put under three broad categories based on the profile
of the refractive indices of the core with respect to the cladding, the number of modes that can
propagate in a fibre and a mix of these. We now discuss these in brief.

10.3 TYPES OF OPTICAL FIBRES

We now know that in its simplest form, an optical fibre consists of a germanium-doped silica
core and a cladding of slightly lower refractive index. Based on the refractive index profile,
optical fibres can be classified as
(i) step-index fibre and
(ii) graded-index fibre.
Similarly, we can categorise fibres depending on the number of modes that can propagate
in them as
Fibre Optics 369
(i) single-mode fibre and
(ii) multimode fibre.
In practice, we will have a mix of these as step-index single-mode fibre, step-index multimode
fibre and graded-index multimode fibre. We now discuss these in brief.
• Step-index fibre In step-index fibre, the refractive index of the core (n1) is uniform
throughout and undergoes an abrupt change at the core–cladding interface. (The refractive
index of the cladding is also uniform but less than that of the core (n2 < n1). The fibre
core is very narrow (» 2–10 µm). In a single mode step-index fibre [Figure 10.7(a)],
the light propagates along the axis of the core. But in a multi-mode step-index fibre
[Figure 10.7(b)], the light propagates in the form of meridional rays as well. These cross
the fibre axis after every reflection at the core–cladding interface and propagate in a zig-
zag fashion.
• Graded-index fibre In a graded-index fibre, the refractive index of the core material
decreases parabolically along the radius such that it is maximum at the centre of the core
and equal to the refractive index of the cladding at the core–cladding interface. The light
rays propagating through it are in the form of helical rays. These do not cross the fibre
axis at any time [Figure 10.7(c)].
• Single-mode fibre In a single-mode step-index fibre, only one mode can propagate
through the fibre [Figure 10.7(a)]. To specify the number of modes propagating through
a step-index optical fibre, we define a dimensionless parameter, called waveguide
parameter through the relation
2S 2S 2S
V n1a 2 ' a n12  n22 a (NA) (10.9)
O O O
Here l is the wavelength of light propagating through the fibre, a is the radius of the fibre core,
n12  n22 n1  n2
n1 is the refractive index of the core material, and D =  defines the relative
2n12 n1
refractive index difference, where n2 is the refractive index of cladding.
For a step-index fibre, if V < 2.4045, the fibre is said to be single-mode fibre. The single-
mode fibre has a smaller core diameter and small relative refractive index difference. As such,
fabrication of single-mode fibre is very difficult and this increases its cost. However, transmission
loss and dispersion/degradation of the signal are very small. That is why single-mode fibres are
very useful in long distance communication.
The wavelength for which V = 2.4045 is known as the cut-off wavelength for propagation
of single-mode and is denoted by lc. This mode is also referred to as the fundamental mode.
You may now like to answer a Practice Exercise.

Practice Exercise 10.4 The optical fibre specified in Practice Exercise 10.3 is operated at
wavelength 1.3 µm. It is characterized by D = 0.003 and a = 4.2 µm.
(i) Calculate V and determine the number of modes that will be excited in it.
(ii) Also calculate the cut-off wavelength at which the fibre will support only a single-
mode. [Ans. (i) 2.284, single-mode; (ii) 1.235 µm]
370 Wave Optics

• Multimode fibre A multimode fibre supports several modes simultaneously [Figure 10.7(b)
and (c)]. The waveguide parameter V for a multimode fibre is greater than or equal to
10. Bergano and Davidson (1996) have shown that the total number of modes propagating
through a multimode step-index fibre is given by
2 2
V2 È n d 2' Ø 1 È 2S a Ø 2
NSIF 4.9 É 1 Ù É Ù (NA) (10.10)
2 Ê O Ú 2Ê O Ú
where d is the diameter of the core of the fibre.

Figure 10.7 (a) A step-index single-mode fibre, (b) step-index multimode fibre, and (c) graded-index
multimode fibre. The modes supported by these fibres are also shown.

For a multimode graded-index fibre having parabolic refractive index profile core, the total
number of modes is given by
V2 NSIF
N GIF (10.11)
4 2
Fibre Optics 371
Note that the total number of modes propagating through a multimode graded-index fibre
is half of that supported by a multimode step-index fibre.
In a multimode fibre, different modes travel with different group velocities leading to what
is known as inter-modal dispersion. (In the language of geometrical optics, this is known as ray
dispersion and arises because different rays take different time in propagating through the fibre.)
It may be mentioned here that in multimode fibres, the core diameter and the relative difference
in the refractive indices are larger than in single-mode fibres. In the case of multimode graded-
index fibre, signal distortion is very low. The light rays travel at different speeds in different
paths of the fibre because of the parabolic variation of refractive index of the core. As a result,
light rays near the outer edge travel faster than the light rays near the centre of the core.
However, launching of light into the fibre and fabrication of the fibre are relatively easy. These
fibres are used in LAN.
If the refractive index of the cladding is greater than the refractive index of core, incoming
light will not undergo total internal reflection. In such a scenario, the incoming light will be
refracted and no wave propagation will take place.
So far, we have discussed the basic principle involved in transmission of light in optical
fibres. Let us now learn how an optical fibre works as a component of optical communication
system.

10.4 OPTICAL COMMUNICATION THROUGH OPTICAL FIBRES

We now know that optical communication refers to transmission of audio and/or visual information
in the form of speech or picture by light waves. Since optical frequencies (» 1015 Hz) are several
orders of magnitude higher than radio and microwaves, transfer of information through optical
communication is comparatively more rapid, secure and can be used over longer distances.
Moreover, the volume of information transacted is significantly higher. Other added advantages
of optical communication are that it offers the possibility of complete electrical isolation and
immunity to electromagnetic interference so that signals are almost free from transmission loss/
leakage.
Refer to Figure 10.8, which shows the basic configuration of a typical optical fibre
communication system. It uses light waves as a carrier, optic fibres as transmission media and
light sources and detectors as opto-electronic transducers. The electrical to optic transducer
converts input electrical (communication) signals into light signals and modulates the light from

Figure 10.8 Basic configuration of typical optical fibre system.


372 Wave Optics

the optical source. The optical carrier—laser light—can be modulated internally or externally
using an electro-optic modulator or acousto-optic modulator. Nowadays electro-optic modulators
are widely used as external modulators. These modulate light by changing its refractive index
through the given input electrical signal. Light travels through the optical fibre cable, which are
available in lengths of 2 km or so and are joined together by fusion of glass called splicing.
For long distance optical transmission, a repeater is used to amplify the signal.
In digital optical fibre communication systems, the input electrical signal is encoded as
digital pulses from the encoder and these pulses modulate the light from the laser diode or LED
and convert them into optical pulses. At the other end, a photo-detector converts optical pulses
into electrical pulses. A decoder then converts the electrical pulses into their original electrical
signal. So retrieval of the signal at some distance down the line depends only on the recognition
of either the presence or the absence of a pulse representing binary (0 or 1) digit.
As of now, optical fibre communication has evolved through four generations.
The characteristic features defining each generation are given in Table 10.1. The first generation
optical fibre communication systems used GaAs based LEDs and laser diodes, which emitted
light of wavelength 0.8 µm. In the period 1974–78, graded-index multimode fibres were
used but thereafter, only single-mode fibres have been put to use in long distance communication.
In the 2nd generation, InGaAsP hetero-junction laser diodes served as optical sources.
These helped to shift the operating wavelength upward to 1.3 µm and overcome loss/dispersion
of information. In the 3rd generation systems, the operating wavelength was further shifted to
1.55 µm and dispersion-shifted fibres were used. These allowed considerable decrease in
transmission losses. In the 4th generation, erbium-doped optical fibre amplifiers are used and
the entire transmission and reception are performed in the optical domain. To increase the bit
rate, Wavelength Division Multiplexing was introduced. The 4th generation fibres are characterized
by a smaller core and larger relative refractive index difference. It is hoped that soliton based
lossless and dispersionless optical fibre communication will become a reality soon, increasing
the data rate beyond 1,000 Tb s–1.
Table 10.1 Parameters defining different generations of optical fibre communication systems
Generation Wavelength Bit rate Loss Repeater spacing
(µm) (Mb s–1) (dB km–1) (km)
I 0.8 4.5 1 10
II 1.3 1.7 ´ 102 < 1 50
III 1.55 1.0 ´ 104 < 0.2 70
IV 1.55 1.0 ´ 105 < 0.002 100
Soliton based 1.55 >1.0 ´ 109 < 0.0002 > 100

You should now answer a Practice Exercise.


Practice Exercise 10.5 A fourth generation optical fibre is characterized by D = 0.008 and
a = 2.4 µm. If n1 = 1.47, calculate
(i) V and determine the number of modes that will be excited in it.
(ii) Also calculate the cut-off wavelength at which the fibre will support only a single-
mode. [Ans. (i) 1.81, single-mode; (ii) 1.16 µm]
Fibre Optics 373
For telecommunication to be an enjoyable experience, it is necessary that the quality of
audio and/or video is distortion-free. In the language of physics, we say that communication
should be dispersion-free and carry large volume of information. We can consider an analogy
with a congested city such as Mumbai, Kolkata, Bengaluru, Chennai and Delhi, where density
of population is high and there is dearth of space. To make better utilisation of space, the
concept of high rise buildings evolved. Similarly, very efficient use of duct space becomes
possible due to extremely thin dimensions of the fibre. The optical fibre based communication
offers several such advantages. We now discuss these.

10.4.1 Advantages of Optical Fibre Communication

Optical fibre based communication offers several advantages. The important ones include the
following:
(i) Wider bandwidth The information carrying capacity of a transmission system is directly
proportional to the frequency of carrier wave. You will recall that radio and microwave
carrier frequencies are about 106 Hz and 1010 Hz respectively, whereas the optical
carrier frequency is in the range of 1013–1015 Hz. Thus an optical fibre offers significantly
greater transmission bandwidth and wavelength division multiplexing increases its
information carrying capacity by several orders of magnitude. Bit rates up to 2.5 Mbps
are already in use in India.
(ii) Low transmission loss With the use of ultra-low loss fibres and erbium-doped silica
fibres as optical fibres, now it is possible to achieve almost lossless transmission.
Moreover, it is possible to achieve attenuation of about 0.002 dB km–1. Also, appropriate
optical amplification can be achieved using erbium-doped silica filters over a short
length in the transmission path at selective points. Thus the repeaters can be spaced
at intervals of about 100 km and the distortion produced during strengthening of the
signal is almost negligible since amplification is done in the optical domain.
(iii) Low cross-talk In a well-designed optical fibre cable, optical interference between
the fibres is almost absent. As a result, there is no cross-talk between the fibres though
the number of fibres in a cable is very large. Moreover, optical fibres are not affected
by any interference originating from power cables, railway power lines and radio
waves. This is essentially because optical fibres are made from silica, which is an
electrical insulator.
(iv) High signal security Optical fibre communication provides almost foolproof security
because signals transmitted through the fibres neither radiate nor can be tapped easily.
(v) Small size Fibre optic cables have small radii, are flexible, compact and lightweight.
A fibre cable can be bent or twisted without any damage.
(vi) Low cost The major advantage of optical fibre communication systems is that these
are low cost. This has genesis in wide spacing between the repeaters, abundance of
silica along the sea line/coast and easy maintenance of the networks.
We now know that communication between two satellite stations is carried out through open
space, whereas communication between a space satellite and a terrestrial station is influenced
by earth’s atmosphere. Similarly, in fibre-optic communication losses may be caused by
374 Wave Optics

microscopic defects. These lead to absorption and scattering of light by the building blocks of
the supporting medium. The signal travelling along the fibre can also lose strength due to
quantum tunnelling, in addition to geometrical/structural imperfections. Attenuation and pulse
dispersion are the two most vital parameters of an optical fibre that determine its quality,
efficiency, information carrying capacity and cost of a communication system. As mentioned
earlier, lower attenuation and dispersion help to reduce cost and improve efficiency of the
system, thereby improving the quality of output. Let us now discuss these.

From the classical point of view, a particle having energy E can not penetrate a potential
barrier of height V > E. But quantum-mechanically speaking, there exists a small but finite
probability of such a particle penetrating the barrier. This is observed in emission of a-rays
from the nucleus. This effect is known as quantum tunnelling.

10.5 ATTENUATION AND LOSSES IN FIBRES

The transmission loss or attenuation of the signal in an optical fibre is a measure of the loss
of optical power as it propagates. The attenuation of an optical beam is measured in decibel per
unit length (dB km–1). Attenuation essentially determines the maximum length over which no
repeater is required for signal strengthening (amplification). Signal attenuation is defined in the
logarithmic unit of decibel, which is used to compare two power levels. For a particular optical
wavelength, we take logarithm of the ratio of the transmitted (input) optical power Pin to the
received (output) optical power Pout from the fibre:

È Pin Ø
D 10 log10 É Ù
(10.12)
P
Ê out Ú

Thus, if the output power is 50% of the input power, the loss will be
a = 10 log10 2 = 3.01 dB
That is, a loss of half of the initial power is nearly equal to three decibel loss. It is clear that
if output power is one-hundredth of the input power, the loss will be 20 dB. On the other hand,
if 99% of the light is transmitted through the fibre, the loss will be about 0.2 dB. This corresponds
to the third generation optical fibre.
Now go through the following example.

EXAMPLE 10.1 The power of a 4 mW laser beam decreases to 40 µW after traversing


40 km in an optical fibre. Calculate the attenuation.
Solution The attenuation of the fibre is given by Eq. (10.12). Therefore, we can write

È 4 – 103 W Ø
D 10 log10 É 6 Ù š (40 km) 0.5 dB km 1
Ê 40 – 10 WÚ
Fibre Optics 375
You should now answer a Practice Exercise.

Practice Exercise 10.6 In fourth generation optical fibre, a < 0.002 dB km–1. What per cent
of light is transferred? [Ans. Almost completely]

Attenuation can be classified into two broad categories: intrinsic and extrinsic. The intrinsic
losses include molecular absorption, Rayleigh scattering due to random molecular orientations,
and structural imperfections. On the other hand, extrinsic losses are caused by bending of fibres
and imperfect connections. The spectral dependence of intrinsic losses for a typical silica optical
fibre is shown in Figure 10.9. Note that two low loss windows occur around 1.33 mm and
1.55 mm. This explains why most fibre-optic systems operate around these wavelengths in the
infrared region.

Figure 10.9 Wavelength dependence of various intrinsic losses in a typical silica optical fibre.
(Source: Arumugam, M., PRAMANA Journal of Physics, p. 858, vol. 57, Nos. 5 and 6 (2001).

The mechanisms generating intrinsic losses include:


1. The ultraviolet absorption due to electron transition at wavelengths around 0.8 µm.
Absorption by molecular vibrations of hydroxyl (OH) – ions dissolved in glass occurs at
2.8 µm but its harmonics are present at wavelengths 1.38 µm and 0.95 µm. However,
this absorption is eliminated almost completely by reducing the water content in the
fibre to below 10 parts in a billion. The infrared absorption by Si-O coupling is present
at wavelengths from about 1.4 µm to 1.6 µm. Absorption by transition metal impurities
such as chromium, iron, manganese, nickel, etc., causes absorption losses at wavelengths
greater than 0.8 µm. This absorption is almost negligible in ultra low pass fibres.
2. Due to imperfections in the structure introduced during the manufacturing process,
the refractive index inside the glass fluctuates spatially. Moreover, these lead to random
molecular dislocations, which scatter light passing through the fibre and divert out of
376 Wave Optics

the core. This loss, known as Rayleigh scattering, is inversely proportional to the fourth
power of operating wavelength:

D 1.7 ÈÉ 0.85 Ø
4 Ù
(10.13)
Ê O Ú

The wavelength is expressed in micrometre.


The maximum loss due to Rayleigh scattering occurs in the ultraviolet region. In the region
from 0.8 µm to 1 µm, it introduces a loss of about 0.6 dB km–1.
A vivid demonstration of spectral dependence of Rayleigh scattering can be made using a
long optical fibre. Take a halogen lamp and inject the light in a long multimode optical fibre.
Observe the colour of light emerging from the fibre. It will appear reddish. Next, take only
about one metre long piece of the same optical fibre and again allow light to fall on it. Do you
expect any change in the colour of the light coming out of the fibre? The light will now appear
whitish. This difference is essentially due to Rayleigh scattering.
The mechanisms generating extrinsic losses are:
• Geometrical non-uniformity at the core–cladding boundary.
• Imperfect connections or alignments between fibres.
• Micro-bending.
It may be mentioned here that extrinsic losses are very small compared to intrinsic losses.
Moreover, these can be minimized by proper care during manufacturing and installation of fibres.
Yet some distortions arise in the quality of information as it is transmitted from the input-end
to the receiver-end due to pulse dispersion. We now discuss the possible reasons of pulses
dispersion.

10.5.1 Pulse Dispersion

In the preceding sections, you have learnt that in digital communication systems, information
is transmitted through the optical fibre in the form of coded pulses, which are then decoded at
the receiver end. While propagating through the fibre, the width of these pulses broadens/
spreads in time. For instance, if the digitised signal pulses are sent in the form of square pulses,
they are converted into broadened Gaussian pulses (Figure 10.10). This broadening of signal
width is known as dispersion.

Figure 10.10 Dispersion of a pulse in a fibre.

From Chapter 7, you may recall that dispersion arises due to dependence of the refractive
index of the material medium (fibre in this case) on the wavelength of the carrier wave.
The dispersion induced overlapping of pulses leads to loss of resolution. Such distortion causes
Fibre Optics 377
errors in decoding and degrades the quality of information at the output end. In fact, in such
a situation, the data transfer becomes unreliable and the transmission capacity is severely restricted.
You will therefore agree that pulse dispersion and attenuation determine the efficiency of the
system as well as the distance between two consecutive repeaters in a fibre-optic link.
There are two types of pulse dispersion mechanisms in a fibre:
(i) intra-modal dispersion, which arises basically in single-mode fibres because of
wavelength dependence of the fundamental mode (spot size) and
(ii) inter-modal dispersion, which arises because different modes travel with different group
velocities in a multimode fibre, as shown in Figure 10.11. This is more important in
step-index multimode fibres. We now discuss these in detail.

Figure 10.11 Modal dispersion.

(i) Intra-modal dispersion


To understand the effects of intra-modal dispersion analytically, we analyse the behaviour of
group velocities of the guided modes in the optical fibre. From your knowledge of waves and
oscillations, you may recall that group velocity is the velocity at which the energy in a particular
mode travels along the fibre. Mathematically, we can write

dZ dZ d O
vg (10.14)
dE dO dE

where propagation constant b is given by

2S Z
E n1 n1
O c
We know that refractive index in a dispersive medium depends on the wavelength of light.
Using this fact, we can write
dE 2S dn1 2S
n (10.15)
dO O dO 1 O2
Similarly, using the relation between angular frequency and wavelength (w = 2pc/l), we get
dZ 2S c (10.16)

dO O2
378 Wave Optics

On combining Eqs. (10.14), (10.15) and (10.16), we obtain

1 È 2S c Ø
vg = – É
È 2S dn n 2S Ø Ê O 2 ÙÚ
É ¹ 1  1 2 Ù
Ê O dO O Ú
2S c c c
= (10.17)
È n1 1 dn1 Ø È dn Ø ng
2SO 2 É  ÉÊ n1  O 1 ÙÚ
Ê O 2 O d O ÙÚ dO
where c is the velocity of light and
dn1
ng n1  O (10.18)
dO
is the group index of the fibre.
This result shows that the group velocity (vg) and the phase velocity (vp = c/n1) vary with
wavelength and have different values in an optical fibre.
The intra-modal dispersion arises due to the dependence of the group velocity on the
wavelength, i.e. dispersive properties of the fibre material and increases as the spectral width
of the optical source increases. This spectral width defines the range of wavelengths emitted by
the optical source. For instance, the spectral width of LED is about 25 nm with peak emission
wavelength at 8.50 nm. In case of a typical laser diode, the spectral width is about 2 nm or so.
It means that in an optical fibre, the intra-modal dispersion can be reduced by using single-mode
laser diode as an optical source.
Before proceeding further, go through the following example.

EXAMPLE 10.2 In the range of 0.5 µm–1.6 µm, the refractive index of pure silica core
varies with wavelength and can be represented by the empirical relation
a
n1 (O ) A0  aO 2  (i)
O2
where A0 = 1.451, a = 0.003 and l is measured in µm. Calculate the group velocity for
l = 0.80 µm.
Solution From Eq. (10.18), we recall that group index is given by
dn1
ng n1  O (ii)
dO
It readily follows from (i) that
dn1 2a
 2aO 
dO O3
Using this result in (ii), we can write
a 2a
ng = A0  aO 2  2
 2a O 2 
O O2
Fibre Optics 379

3a
= A0  aO 2  (iii)
O2
Hence, for l = 0.80 µm, the value of group index is

0.003 – 3
ng = 1.451  0.003 – (0.80) 2 
(0.80) 2
= 1.467
And group velocity is given by

c 3 – 108 m s 1
vg 2.045 ms 1
ng 1.467

Since every source of light has a finite spectral width and each wavelength component of
a pulse travels with a slightly different group velocity in a medium, broadening of the pulse
occurs as it propagates through the fibre. It means that information carried by a fibre optic
system will be reliable only when pulse propagation is free from pulse broadening. Let us now
study the impact of variation of refractive index of the core with wavelength on inter-modal
dispersion.

(a) Material dispersion


Material dispersion arises due to variation of refractive index of the core material with the
wavelength of light. It generates intra-modal dispersion. It is also known as chromatic dispersion.
A material is said to exhibit material dispersion only when second differential of the refractive
index with respect to wavelength is non-zero. Mathematically, we write

d 2 n1
›0
d O2
where n1 is the refractive index of the core material.
We now know that the group index as well as group velocity vary with the operating
wavelength of light. To calculate the expression for pulse broadening due to material dispersion,
we note that the time taken by a pulse to propagate through a length L of the fibre is given by
L
tm = (10.19a)
vg
Using Eq. (10.17), we can rewrite it as

LË dn1 Û
tm = Ì n1 (O )  O (10.19b)
cÍ d O ÜÝ
This equation implies that different wavelengths will travel with different group velocities in a
dispersive medium.
380 Wave Optics

For a source with root mean square (rms) spectral width Dl and mean wavelength l,
the rms broadening due to material dispersion, Dt m may be obtained from the Taylor series
expansion about l:

'W m
dW m
dO
'O 
d 2W m
dO2
( 'O ) 2  "
For sources operating between wavelengths 0.8 and 0.9 µm, it is sufficient to retain only the linear
term in the Taylor series expansion. In this approximation, we obtain the expression for material
dispersion using Eq. (10.19) as

L 2 d 2 n1
'W m 'O O (10.20)
cO dO2
Note that the broadening of the transmitted pulse is directly proportional to the spectral width
of the source and the distance traversed by the pulse in the medium. It means that by using a
source with narrow spectral width, the material dispersion can be reduced. For pure silica, the
material dispersion tends to be zero at l = 1.3 µm and rises almost exponentially for wavelengths
around 0.6 µm to 0.8 µm.
We define material dispersion coefficient through the relation

1 'W m O d n1
2
(10.21a)
Dm
L 'O c dO2
If wavelength is measured in µm and c = 3 ´ 10–4 m ps–1, the material dispersion coefficient
is expressed in picoseconds per kilometre length of the fibre per nanometre spectral width of
the source:

1 È 2 d 2 n1 Ø
ÉO
–1 –1
Ù – 10 ps km nm (10.21b)
4
Dm 
3O Ê d O 2 Ú
A negative value of the dispersion coefficient implies that the longer wavelengths travel faster
and vice versa. In 4th generation optical systems, laser diodes operate at a wavelength of
1550 nm with a spectral width of 2 nm. At this wavelength, the material dispersion coefficient
is 21.5 ps km–1 nm–1. Therefore, in traversing one kilometre length of the fibre, the material
dispersion comes to about 43 ps. Since every light source, including a laser, exhibits finite
spectral spread howsoever small, we can say that material dispersion is an intrinsic property
of a fibre-optic communication system.
You may now like to answer a Practice Exercise.

Practice Exercise 10.7 In the first generation optical fibre, the spectral width of the light
source was about 20 nm around l = 0.80 µm. For the material dispersion coefficient of
84 ps km–1 nm–1, calculate the pulse broadening. Compare it with 3rd generation optical
communication system characterised by l = 1.3 µm, for material dispersion coefficient of
2.4 ps km–1 nm–1. [Ans. 1680 ps; 48 ps]
Fibre Optics 381
On answering this Practice Exercise, you will note that pulse broadening is extremely small
in the 3rd generation optical communication system. This explains why optical communication
is more efficient when operated around 1.3 µm. Physically it is due to the fact that group
velocity vg is approximately constant around 1.3 µm. In fact, the wavelength 1.27 µm is referred
to as the zero material dispersion wavelength.
You may now like to ask: Is there any mechanism other than material dispersion for
intra-modal dispersion that causes pulse broadening? You will soon discover that waveguide
dispersion is another important source of pulse broadening. We will learn about it now.

(b) Waveguide dispersion


We know that material dispersion gives rise to intra-modal dispersion. For the sake of argument
assume that the fibre is free from material dispersion. You may then ask: Will the pulse propagate
without any broadening? The answer to this question is in the negative; there is another source
of intra-modal pulse broadening. This is known as waveguide dispersion, which arises due to
the finite frequency bandwidth and the dependence of group velocity of each mode on the
wavelength of light. The waveguide dispersion will be more for higher frequency bandwidth of
the transmitted pulse. A detailed analysis of waveguide dispersion is quite complicated and we
will refrain from going into the details. However, it may suffice to mention that for a single mode
of propagation constant b, a fibre is said to exhibit waveguide dispersion when d2b/dl2 ¹ 0.
For 1.4 < V < 2.6, an empirical formula for a step-index single-mode fibre is given by
L 'O
'W w  n2 '[0.080  0.549(2.834  V )2 ] (10.22)
c O
For L = 1 km, Dl = 10–9 m and c = 3 ´ 10–4 m ps–1, we can write the expression for waveguide
dispersion coefficient as
n2 '
Dw  – 107 [0.080  0.549(2.834  V )2 ] ps km–1 nm–1 (10.23)
3O
The total dispersion is obtained by adding the material and waveguide dispersions.
For the step-index fibre operating at wavelength 1.3 µm with n2 = 1.450, D = 0.003 and
a = 4.2 mm,
2S
V – 4.2 – 106 – 1.45 0.006 2.280
1.3 – 10 6
Since V < 2.405, this fibre will support only one mode at the operating frequency. And waveguide
dispersion coefficient in this case is given by

1.450 – 0.003
Dw =  – 107 [0.080  0.549(2.834  2.280)2 ] ps km 1 nm 1
3 – 1300
= 2.772 ps km–1 nm–1

The typical variation of Dm, Dw and Dtotal (=Dm + Dw ) with operating wavelength is shown in
Figure 10.12. Note that the material dispersion as well as total dispersion passes through zero
around 1.3 µm. This is known as zero dispersion wavelength. If we operate 1.3 µm single-mode
382 Wave Optics

fibre at l = 1.55 µm, the material dispersion in single-mode fibre is positive and large
while waveguide dispersion is negative and small. (This leads to residual dispersion of about
15 ps km–1 nm–1) Therefore, to achieve zero dispersion, one can increase waveguide dispersion
by adding more GeO2 in the core. (This increases D and the refractive index of the core.)
Alternatively, one can add some fluorine in the cladding, which decreases its refractive index.
With such dispersion-shifted fibres, we get minimum loss and zero dispersion at l = 1.55 µm.
This is what characterises fourth generation communication system.

Figure 10.12 Variation of material dispersion coefficient, waveguide dispersion coefficient and total
dispersion with wavelength.

For fourth generation optic fibre communication system, the material dispersion coefficient
is nearly equal and opposite to waveguide dispersion coefficient so that they cancel mutually
and total dispersion is zero. Physically, we can say that in contrast to material dispersion,
waveguide dispersion causes longer wavelengths to travel faster than shorter wavelengths and
their effects annihilate mutually.

(ii) Inter-modal dispersion


Pulse broadening due to inter-modal dispersion arises because different modes within a multimode
fibre travel with different group velocities. As a result, the pulse width at the receiver-end
depends on the transmission times of the slowest and the fastest modes. The inter-modal dispersion
is very prominent in step-index multimode fibre and causes significant pulse broadening. However,
inter-modal dispersion in multimode fibres can be minimised substantially by using graded
(parabolic) index fibre. We discuss these now.

(a) Multimode step-index fibre


Refer to Figure 10.11 again, which depicts the paths taken by three rays incident on the axis
of the fibre at different angles. (Each ray corresponds to a mode.) Note that each ray enters the
fibre at the same time but they cover different distances in the core; the ray incident on the core–
cladding interface at the largest angle travels the longest distance. As a result, there will be a
Fibre Optics 383
spread in time as these rays reach the receiver-end. This causes broadening of the information
carrying signal and restricts the transmission capacity of the optical fibre. Can you explain it
by arguing on physical ground? It is so because pulse broadening may cause the pulses to
overlap as they propagate. To avoid overlapping, we resort to increasing the time delay between
two consecutive pulses. As a result, the number of pulses that can be transmitted through the
fibre per unit time goes down reducing transmission capacity.
To quantify these ideas about pulse dispersion, let us consider the fastest and the slowest
modes propagating in a perfectly structured step-index fibre (Figure 10.13). These modes are
respectively represented by the axial ray and the extreme meridional ray. Suppose that the
meridional ray is incident at the fibre axis at an angle q i and is refracted at an angle qr before
it reaches the core–cladding interface and undergoes total internal reflection. Since both rays
travel with the same velocity within the core, the delay in reaching the receiver-end is governed
by the difference in their respective path lengths in the optical fibre, which, in turn, depends
on the angle a ray makes with the axis of the fibre. Let us suppose that the meridional ray takes
time t in going from P to R along PQR. If the refractive index of the core is n1, it will propagate
at a speed c/n1. Hence, we can write

PQ  QR n1 ( PS  SR ) n1PR
t (10.24)
c c cos T r c cos T r
n1

Figure 10.13 An axial and a meridional ray passing through a step-index fibre.

Hence, the time taken by the meridional ray to traverse a length L of the fibre will be

n1 L
tL (10.25)
c cos T r
This result shows that the time taken by a ray depends on the angle it makes with the axis of
the fibre. So a ray travelling along the fibre axis (q r = 0; cosqr = 1) will take the minimum time:

n1 L
tmin (10.26a)
c
On the other hand, the maximum time will be taken by a ray for which qr = 90º – qc, where
qc is the critical angle at the core–cladding interface. Then using Snell’s law at the core–
cladding interface, we can write sinqc = n2/n1 = cosq r so that qr = cos–1(n2/n1), where n2 is the
384 Wave Optics

refractive index of the cladding. Using this result in Eq. (10.25), we find that the maximum time
taken by the extreme meridional ray is
n12 L
tmax (10.26b)
n2 c
Hence, if several rays travel in the fibre simultaneously, the spread in time at the receiver-end
will be between the axial ray and the meridional rays. On combining Eqs. [10.26(a) and (b)],
we obtain the desired expression for inter-modal dispersion:
n1 L È n1 Ø
( 'W i )SIF tmax  tmin 1
c ÉÊ n2 Ù
Ú

Since relative refractive index difference

n12  n22 n1  n2
' 
2n12 n1
and numerical aperture
NA = 2 n1 (n1  n2 ),
we can express this result as
n1 L L
( 'W i )SIF  ' (NA) 2 (10.27)
c 2n1c
Note that inter-modal dispersion is directly proportional to the length of the fibre and square of
NA. It means that inter-modal dispersion will be small if we have smaller NA. This, however,
reduces the acceptance angle and hence the light gathering capacity of the fibre.
In a typical multimode step-index fibre, n1 = 1.5 and D = 0.01. Then in a two kilometre long
fibre, the inter-modal dispersion is
1.5 – 2000 m
( 'W i )SIF – 0.01 107 s 100 ns
(3 – 108 m s 1 )
This shows that after traversing a fibre of length one kilometre, the pulse will be broadened by
about 50 ns. It means that if two consecutive pulses are separated by less than 50 ns at the
transmission-end, they cannot be resolved at the receiver-end and no useful information can be
retrieved. Hence, in a 1 Mbit/s fibre-optic communication system, dispersion of such a magnitude
will require repeaters to be placed every three to four kilometre. However, in 1 Gbit/s fibre-optic
system, 50 ns dispersion will cause havoc even within 50 m. Thus smaller pulse dispersion leads
to more efficient and increased information carrying capacity of fibre-optic communication
system. It is therefore important to know the methods that can be employed to reduce inter-
modal pulse dispersion in multimodal fibres. As mentioned earlier, a graded (parabolic) index
fibre helps us to achieve the desired objective. Let us learn about it now.

(b) Multimode graded-index fibre


From Section 10.3, you may recall that the refractive index of the fibre core remains uniform
throughout in a step-index fibre. On the other hand, in a graded-index fibre, the refractive index
Fibre Optics 385
of the core material decreases parabolically along the radius such that it is maximum at the
centre of the core and equal to the refractive index of the cladding at the core–cladding interface.
The refractive index profile of the core and cladding in such a fibre is respectively given by
2 1/ 2
Ë ÈrØ Û
n(r) = n1 Ì1  2 ' É Ù Ü r<a
Í Ê aÚ Ý
1/ 2
= n1 >1  2' @ r>a (10.28)
For a typical multimode graded-index silica fibre, the relative refractive index difference
D = 0.01, n1 = 1.45 and a = 25 µm.
Figure 10.14 shows axial as well as meridional ray paths within a index grading fibre.
As may be noted, the meridional rays follow sinusoidal trajectories of different path lengths,
which arise due to index grading. However, the ray traversing a larger path length does so in
a region of lower refractive index and hence moves at a greater speed (since it is inversely
proportional to the local refractive index). As a result, the longer sinusoidal paths are almost
compensated by the higher speeds and there is an equalization of the transmission time even
with the axial ray, which travels in the region of the highest refractive index along the axis of
the fibre (core) with the slowest speed. So we can say that in a multimode graded-index fibre,
all rays (modes) take approximately the same time in traversing the length of the fibre from the
input-end to receiver-end and the disparity in mode transit times is almost eliminated and the
transmitted information is more or less free from pulse broadening.

Figure 10.14 Trajectories of rays within a multimode graded fibre.

Detailed mathematical calculations for the inter-modal dispersion are somewhat involved.
Therefore, we just quote the result without going into details:

2
n2 L È n1  n2 Ø n L L
('W i )GRIN É Ù  2 '2  (NA) 4 (10.29)
2c Ê n2 Ú 2c 8cn13

Note that pulse dispersion in case of graded-index fibre is proportional to the fourth power of
numerical aperture. For a typical multimode parabolic index fibre characterized by n2 = 1.45,
and D = 0.01, inter-modal dispersion is about 0.25 ns km–1, which is 200 times less than that
obtained for multimode graded-index fibre. For this reason, the first and the second generation
386 Wave Optics

optical communication systems used graded-index fibres. The use of single-mode fibres completely
eliminated inter-modal dispersion. We may now conclude that
• Single-mode step-index fibres exhibit only intra-modal dispersion, which is made up of
material dispersion and waveguide dispersion.
• The multimode step-index fibres cause significant pulse broadening. However, inter-
modal dispersion due to multimode graded-index fibre is lesser by a factor of about 100
or more.
• Single-mode fibres are used in modern long distance communication systems because of
negligible dispersion.
We now know that attenuation and pulse broadening limit the quality of information received
at the output-end and limit the distance between two repeaters in a fibre-optic link. The maximum
permissible bit rate for a given pulse dispersion is inversely proportional to pulse dispersion.
Recall that pulse dispersion is induced by inter-modal dispersion, material dispersion and
waveguide dispersion. However, the waveguide dispersion is not important for multimode fibres.
Then the total dispersion can be written as

'W ( 'W m ) 2  ('W i )2 (10.30)


You may now like to answer a Practice Exercise.

Practice Exercise 10.8 The pulse dispersion due to inter-modal and material dispersion are
0.24 ns km–1 and 0.05 ns km–1 respectively. Calculate the maximum bit rate.
[Ans. 2.8 Gbit-km s–1]
It may be mentioned here that in an actual link, the source and detectors can also influence
pulse broadening and may have to be accounted for in rigorous calculations. However, we will
not go into these details here.
(c) Dispersion compensating fibres
Many countries have installed capacity of millions of kilometres of single-mode fibre-optic
links in underground ducts operating at 1.3 µm. Such fibres give residual dispersion of about
15 ps km–1 nm–1 at 1.55 µm. This causes significant transmission losses and drop in information
carrying capacity of the communication system. In order to overcome these shortcomings, an
obvious choice is to shift the operating wavelength of single-mode optical fibre to 1.55 µm.
However, this will necessitate replacement of the entire network with new dispersion-shifted
fibres and entail enormous expenditure, apart from huge effort. To avoid this and yet upgrade
the old fibre-optic links, dispersion compensating fibres have been evolved. These have very
large negative dispersion at 1.55 µm. A short length (a few hundred metres to a kilometre) of
dispersion compensating fibre used in conjunction with (tens of kilometres of) 1.3 µm fibre link
helps to obtain minimum loss and zero dispersion at the end of the link.

10.6 APPLICATIONS OF FIBRES


Optical fibres find widespread applications in different areas of human interest and social
welfare. In fact these have brought about a revolution in healthcare and medicine, engineering,
Fibre Optics 387
education and communication. Initially, optic-fibre was installed for use in high capacity links
between different countries or metropolitan cities within a country. Now it is increasingly being
installed for use in Local Area Networks servicing the business community for transferring
funds almost instantly. The development of new communication systems is taking place very
rapidly. The most promising networks are based on Wavelength Division Multiplexing (WDM)
technology. It facilitates transmission of several signals at different wavelengths independently
in the same fibre. These networks necessitate development of new opto-electronic devices,
which may be electronically controlled. These devices include optical amplifiers, which can
amplify many signals simultaneously.
The use of optical fibre networks by educational institutions has changed the form and
format of teaching-learning tremendously; leading institutions are using computer-assisted
networks for facilitating their on-campus students in a number of ways. In fact, a silent
transformation from chalk and talk based classroom teaching-learning to networks supported
mouse and wire based mobile learning in anytime, anywhere, anyone paradigm is being touted
as the wave of future. However, the most important application of optical fibres is in the field
of telecommunications. Several hundred thousand kilometre of optic fibre links under the ground
as well as in the sea-bed have made it possible to communicate with our dear ones across
continents, oceans and civilizations in real-time, as if they are staying next door. We have got
opportunity to use Skype or Messenger search engines and noticed that simultaneously several
million people are in audio-visual contact one-to-one or one-to-many. We learnt the physics of
optic-fibre communication systems in detail in the preceding sections of this chapter.
The advent of viable optical fibres has led to significant developments in medical technology.
These have paved the way for laproscopic surgery, which is usually used for operating hernia,
gall bladder/kidney stone, appendectomies, or look inside the heart while it beats. It makes use
of two or three bundles of optical fibres. The surgeon makes a number of small incisions in the
target area. One of the bundles of optic fibres is used to illuminate the chosen area and another
one carries the information (image) back to the surgeon. It is usually coupled with laser surgery.
Another important application of optical fibres is in sensors. These offer advantages of
lower cost, compactness, greater accuracy, reliability and flexibility. Unlike electrical sensors,
fibre-optic sensors are not influenced by external electromagnetic fields. That is, fibre-optic
sensors exhibit electromagnetic immunity. In air travel, passengers are warned against the use
of mobile phones as these interfere with the navigation of the aeroplane. However, the use of
sensors in an operation theatre helps to monitor the levels of dissolved oxygen or pH of blood.
If a fibre is squeezed or stretched, heated or cooled, a small but measurable change occurs
in its light transmission characteristics. For this reason, fibre-optic sensors can be put in a
hazardous/explosive environment and the effects can be measured at a central point several
kilometres away. These can be used to measure pressure, temperature, current, rotation, strain,
etc., with greater precision and speed. Such advantages have facilitated integration of fibre-optic
sensors into civil structures such as bridges and tunnels, process industries, medical instruments,
aircrafts, missiles, etc.
Fibre-optic sensors are classified in two broad categories: extrinsic and intrinsic. In extrinsic
sensors, the optical fibre serves as a device to transmit and collect light from a sensing element
external to the fibre. On the other hand, in intrinsic sensors, the light beam does not leave the
fibre. But the physical parameter to be sensed alters the properties of the fibre directly. This in
388 Wave Optics

turn leads to changes in a characteristic such as intensity, polarisation, phase, etc., of the light
traversing the fibre.

10.7 SUMMARY

• An optical fibre consists of a central cylindrical glass or plastic core surrounded by a


cladding of the same material but slightly lower refractive index.
• The optical phenomenon of total internal reflection forms the working principle of optic-
fibre based communication.
• Numerical aperture (NA) is a measure of the light gathering capacity of an optical fibre.
It is given by
NA sin T a n12  n22
where qa is angle of acceptance.
• Optical fibres are classified as step-index fibre, graded-index fibre, single-mode fibre and
multimode fibre.
• In a step-index fibre, the profile of refractive index is given by
n(x) = n1 0<x<a
= n2 x<a
where n1 and n2 are refractive indices of the core and the cladding respectively. In a
graded-index fibre, the refractive index of the core material decreases parabolically along
the radius such that it is maximum at the centre of the core and equal to the refractive
index of the cladding at the core–cladding interface.
• In a single-mode fibre, only one mode can propagate through the fibre, whereas a multimode
fibres support several modes simultaneously. The number of modes propagating through
a step-index optical fibre is expressed in terms of the waveguide parameter through the
relation
2S
V n1a 2 '
O
Here l is the wavelength of light propagating through the fibre, a is the radius of the fibre
core, n1 is the refractive index of the core material, and D is the relative refractive index
difference.
• The total number of modes propagating through a multimode step-index fibre is given by:
2
V2 È n d 2' Ø
NSIF 4.9 É 1 Ù
2 Ê O Ú
where d is the diameter of the core of the fibre.
• Optical communication refers to transmission of audio and/or visual information in the
form of speech or picture by light waves. It is comparatively more rapid, secure and used
over longer distances. Optical fibre communication has graduated through four generations,
which are characterised by a unique operating wavelength.
Fibre Optics 389
• Optical fibres offer unique advantages of wider bandwidth, low transmission loss, low
cross-talk, high signal security, small size and low cost. Above all, it is very reliable and
efficient.
• At a particular wavelength, attenuation of an optical beam in a fibre is defined as the
logarithm of the ratio of input power to output power:

È Pin Ø
D 10 log10 É Ù
Ê Pout Ú

The attenuation is measured in decibel per unit length (dB km–1).


• The intra-modal pulse dispersion arises due to dispersive properties of the fibre material
and increases as the spectral width of the optical source increases.
• Every light source has finite spectral width, i.e. emits light over a range of wavelengths.
Since refractive index is a function of wavelength, different wavelengths travel the same
fibre length in different times. This is known as material dispersion and is present in
single-mode as well as multimode fibres. The material dispersion coefficient is given by

1 dW m O d 2 n1 1 È 2 d 2 n1 Ø
Dm  ÉO Ù – 10
4
ps km 1 nm 1
L dO c dO2 3O Ê d O 2 Ú
where wavelength is measured in µm and c = 3 ´ 10–4 m ps–1. For pure silica, the values
of Dm (in picoseconds per kilometre length of the fibre per nanometre spectral width of
the source) are respectively –84.2, 2.39 and 21.52 at l = 0.85 µm, 1.3 µm and 1.55 µm.
• In single-mode fibres, pulse broadening is also caused by waveguide dispersion.
The waveguide dispersion coefficient is given by

n2 '
Dw  – 107 [0.080  0.549(2.834  V )2 ] ps km 1 nm 1
3O
• Pulse broadening due to inter-modal dispersion arises because different modes within a
multimode fibre travel with different group velocities. For a multimode step-index fibre,
inter-modal dispersion is given by

n1L È n1 Ø n1 L L
('W i )SIF  1Ù # ' (NA)2
c ÉÊ n2 Ú c 2n1c
• For a graded-index fibre, inter-modal dispersion is given by
2
n2 L È n1  n2 Ø n L L
( 'W i )GRIN É Ù  2 '2  (NA)4
2c Ê n2 Ú 2c 8cn13
• Optical fibres find widespread applications in healthcare and medicine, engineering,
education and communication.
390 Wave Optics

REVIEW EXERCISES

1. The refractive indices of the core and the cladding of a silica optical fibre are 1.50 and
1.48 respectively.
(i) If light is incident from air, calculate qa.
(ii) This fibre is replaced by another one with n(core) = 1.45 and n(cladding) = 1.41.
What happens to its light gathering capacity?
[Ans. (i) qa = 14.13º, (ii) qa = 19.77º; increase]
2. The velocity of light in a SIF is 2.0 ´ 108 ms–1 and the critical angle is 75º. Calculate the
acceptance angle for the fibre in air. Take c = 3 ´ 108 ms–1. [Ans. 22.6º]
3. The core of a silica fibre has refractive index 1.5. The cladding is doped to give relative
refractive index difference of 0.0005. Calculate NA and critical angle.
[Ans. NA = 0.0474; qc = 88.19º]
4. The core and cladding of a step-index fibre are characterized by refractive indices 1.47
and 1.45 respectively. It is operated at 1.3 µm. Calculate the number of modes that it will
support. Take the radius of the core as 2.8 µm. [Ans. N = 5]
5. A step-index optical fibre has core diameter of 50 µm and is operating at wavelength of
0.85 µm. If NA = 0.3, calculate the number of modes in a GRIN optical fibre.
[Ans. 767]
–1
6. The attenuation of a pulse in an optical fibre is estimated to be 2 dB km . Calculate the
fraction of initial intensity available after the pulse has travelled a distance of 1 km, 5 km
and 10 km. [Ans. 0.63, 0.1, 0.01]
7. The power of 2.5 mW laser beam drops to 25 µW after traversing 100 km. Calculate a.
[Ans. 0.2 dB km–1]
8. A step-index silica fibre is characterised by n1 = 1.5, a = 25 µm and D = 0.02. It is
operated at wavelength 0.85 µm with a spectral width of 50 nm.
(i) How many modes will it support?
(ii) Calculate the inter-modal pulse dispersion.
[Ans. (i) 1537; (ii) 100 ns km–1]
9. A graded-index fibre is characterised by the same parameters as given in Problem 7.
(i) How many modes will it support?
(ii) Calculate the inter-modal pulse dispersion.
[Ans. (i) 769; (ii) 2 ns km–1]
10. Discuss the remedies to reduce pulse dispersion in an optical fibre.
INDEX

Active laser medium, 316 Destructive interference, 96


Acceptance angle, 365 Dextrorotatory, 74
Airy function, 145 Diffraction, 155
Airy pattern, 242 Fraunhofer, 156
Anomalous dispersion, 262 Fresnel, 155
Anti-Stokes Raman scattering, 286 transition to Fraunhofer region, 156
Arena, 308 Diffraction by a
Attenuation in optical fibres, 374 circular aperture, 172, 189, 210
straight edge, 176
Diffraction grating, plane, 231
Babinet compensator, 71 resolving power of, 250
Biaxial crystal, 61 spectrum, 231
Birefringence, 56
Diffraction pattern, production of, 158
Boundary conditions, 25
Dispersion of light, 257
Brewster’s angle, 57
anomalous, 258, 262
Brewster’s law, 58
normal, 258
rotatory, 75
Cauchy equation, 260 Displacement method, 102
Circular polarisation, 49 Displacement of fringes, 97
Circularly polarised light, 51 Double refraction (see Birefringence), 56
Coherence Double slit interference experiment, 7
length, 299
sources, 88
time, 299 Echelon gratings, 239
Concave reflection grating, 235 Einstein’s A and B coefficients, 314
Constitutive relations, 15 Electromagnetic spectrum, 9
Constructing a laser, the prerequisites, 316 Electromagnetic theory of dispersion, 266
Cornu’s spiral, 180–188 damped system, 272
diffraction by straight edge, 186 undamped system, 268
Corpuscular model, 4 Electromagnetic theory of light, 14

391
392 Index

Electromagnetic waves Hologram, recording of, 340


plane waves in free space, 16 theory, 346–349
reflection and refraction of, 25 transmission holograms, 341–45
reflection and transmission coefficients, 30 white light holograms, 350
Elliptically plane polarised wave, 51 Holography, 4, 339
Elliptical polarisation, 49 applications of, 354–358
Energy density of an electromagnetic wave, 23 basic principle, 340
Extraordinary ray, 60 Human vision, 12
Huygens’ construction, 81–83
Huygens’ explanation of birefringence, 62
Fabry–Perot interferometer, 146–151
fringe pattern, 148
sharpness of spectral lines, 150 Information processing by human brain, 13
width of transmission peaks, 149 Interference by a plane-parallel thin film, 113
Fibre optics, 4, 361 conditions of maxima and minima, 116
Fraunhofer diffraction, 194 reflected light, 114
missing orders, 221 Interference of light, 7
N-identical slits, 224 division of amplitude, 88, 112
single slit, 196 division of wavefront, 87
two vertical slits, 216 Interference of polarised light, 7
Fresnel’s biprism, 100 Interference pattern, production of, 100
Fresnel construction, 160 Interferometers, 113
and rectilinear propagation, 167 Fabry–Perot, 146–151
Fresnel diffraction, 155 Jamin, 139–140
circular aperture, 172, 189 Lummer–Gehrcke, 151
patterns of simple obstacles, 172 Mach–Zehnder, 141
rigorous analysis, 188 Michelson, 133–139
straight edge, 176 Twyman–Green, 140, 141
Fresnel half-period zones, 161–167 Inter-modal dispersion, 377
radii of, 163 Ion laser, 333
Fresnel’s two-mirror arrangement, 105
Fresnel integrals, 183
Fringe shape, 93 Kathavate experiment, 158
Fringe-width, 92
Fringes of equal inclination, 120
Fringes of equal thickness, 121 Laevorotatory, 74
Full-wave plate, 70 Lasers, 297
applications, 333–336
in communication, 333
Glan–Thompson prism, 64 CO2, 331
Graded-index fibre, 369 He–Ne, 329–331
Grating (see Diffraction grating), 230–231 holography, 335
ion, 333
medicine, 334
Haidinger fringes (see Fringes of equal thickness), ruby, 327
121 semiconductor, 331
Half-wave plate, 71 solid state, 326–329
Helium–Neon laser, 329–331 types, 326–333
Hertz experiment, 8 Lasing action, 297
Index 393
Left circular wave, 52 Optical resonant cavity, 324–326
Light waves, 3 Ordinary ray, 60
Limit of resolution, 245
Linearly polarised light, production of, 54
Lloyd’s mirror arrangement, 105, 106 Perception of light, 12
Localised fringes, 117 Phase plates, 70–71
Longitudinal coherence, 306 Photoelectric effect, 8
Photon, 8
Plane polarised light, 36, 46
Plane of polarisation, 46
Malus’ law, 54
Plane of vibration, 46
Material dispersion, 379
Poisson spot, 157
Maxwell’s equations, 14–15
Polarisation of light, 43
Michelson interferometer, 133
applications of, 75–77
adjustment of, 136 by biaxial crystal, 66
applications of, 136–139 double refraction, 59
formation of fringes, 134 reflection, 56
Michelson stellar interferometer, 309 scattering, 69
Microscope, resolving power, 247 selective absorption, 67
Multimode fibres, 370 uniaxial crystal, 61
graded index, 384 Polarisation vector, 50
step index, 382 Polaroid, 44
Multiple beam interferometry, 88, 142–146 Population inversion, 315, 317–320
coefficient of Finesse, 144 Poynting vector, 5, 21
Pulse dispersion, 376
Pumping, 320
Nature of light, 3 schemes, 322
Neuronal cells, 12
Newton’s rings, 125
Nicol prism, 63 Quantum tunnelling, 374
Non-localised fringes, 117 Quarter wave plate, 71
N-slit Fraunhofer diffraction pattern, 224
intensity distribution, 225
Raman effect, 258, 284–289
positions of minima and secondary maxima, 227
Raman scattering, 258, 286
positions of principal maxima, 226
Raman spectroscopy, 289–291
width of principal maximum, 230
Rayleigh criterion of resolution, 243
Numerical aperture, 366 Rayleigh scattering, 278
Reflection of electromagnetic waves, 27
normal incidence, 27–31
Observing diffraction, 156 oblique incidence, 31
Optic axis, 60 Refractive index, 257, 259
Optical activity, 74 Resolving power of optical instruments, 242
Optical communication, 4, 371–374 astronomical telescope, 244
advantages of, 373 diffraction grating, 250
Optical fibres, 363 microscope, 247
applications of, 386 Rayleigh criterion, 243
attenuation and losses in, 374–386 Right circular wave, 49
types of, 368–371 Rochon prism, 64
working principle, 363–368 Rods and cones, 12
394 Index

Scattering of light, 257 Twyman-Green interferometer, 140


Sellmeier equation, 263
Single mode fibres, 369
Single slit diffraction pattern, 196 Ultrasound waves, 3
intensity distribution, 198 Uniaxial crystals, 61
positions of maxima and minima, 204 normal incidence, 81
Snell’s law, 34 oblique incidence, 82
SONAR, 3 propagation of plane waves, 81
Spatial coherence, 304–307
visibility of fringes, 307
Spatial evolution of Fresnel diffraction pattern, Water jet experiment, 366
159–160 Wave equation, 15–16
Specific rotation, 74 Wave plates (see Phase plates), 70–71
Spontaneous emission, 311 Waveguide dispersion, 381
States of polarisation, 48 White light holograms, 350
Step-index fibres, 369 reflection holograms, 350–353
pulse dispersion of, 376 transmission holograms, 354
Stimulated emission, 311 Wire-grid polariser, 68
Stokes Raman scattering, 288 Wollaston prism, 64–65
Superposition of linearly polarised waves, 49 Wood’s experiment, 262

Temporal coherence, 298 Young’s double slit experiment, 88, 89


spectral linewidth, 301 conditions for maxima and minima, 91, 92
Theories of light, 4–9 fringe shape, 93, 94
Total internal reflection, 363–365 intensity distribution, 94–97
Transverse electromagnetic waves, 7 interference pattern, 90
Two beam interference, 87
Two slit Fraunhofer diffraction pattern, 216
intensity distribution, 217 Zone plate, 168–170
missing orders, 221 and convex lens, 171

You might also like