Download as pdf or txt
Download as pdf or txt
You are on page 1of 297

Demonstrational Optics

Demonstrational Optics
Part 1: Wave and
Geometrical Optics

Oleg Marchenko
St. Petersburg State University
St. Petersburg, Russia

Sergei Kazantsev
Paris Observatory
Paris, France

and

Laurentius Windholz
Technical University of Graz
Graz. Austria

Springer Science+Business Media, LLC


Library of Congress Cataloging-in-Publication Data

Marchenko, Oleg.
Demonstrational optics / Oleg Marchenko, Sergei Kazantsev, Laurentius Windholz.
p. cm.
Includes bibliographical references and index.
Contents: pt. 1. Wave and geometrical optics.

1. Optics. 1. Kazantsev, S. A. (Sergei Anatol'evich) II. Windholz, L. III. Title.

QC355.3.M372004
535--dc22
2003061896

ISBN 978-1-4613-4723-1 ISBN 978-1-4419-8925-3 (eBook)


DOI 10.1007/978-1-4419-8925-3
©2003 Springer Science+Business Media New York
OriginalIy published by Kluwer Academic / Plenum Publishers, New York in 2003
Softcover reprint of the hardcover 1st edition 2003
http://www.wkap.nV
10987654321
A C.I.P. record for this book is available from the Library of Congress
AlI rights reserved
No part of this book may be reproduced, stored in a retrieval system, or transmitted in any form
or by any means, electronic, mechanical, photocopying, microfilming, recording, or otherwise,
without written permission from the Publisher, with the exception of any material supplied
specificalIy for the purpose of being entered and executed on a computer system, for exclusive
use by the purchaser of the work.
Permissions for books published in Europe: permissions@wkap.nl
Permissions for books published in the United States of America: permissions@wkap.com
Contents

Preface ix
1. HISTORICAL INTRODUCTION 1
1 Wave nature of light 1
2 Electromagnetic theory 6
3 Optical Spectroscopy 7
4 Quantum Optics 8
5 Atom model and relativity 10
6 Coherence and precision optical instrumentation 13
7 Light correlation and statistics 15
2. PROPERTIES OF LIGHT WAVES 17
1 Laws of electromagnetism 17
1.1 Maxwell's equations 17
1.2 Electric dipole radiation 20
2 Properties of electromagnetic waves 23
2.1 Wave equation 23
2.2 Plane waves 27
2.3 Plane monochromatic wave 28
2.3.1 Frequency, wavelength, light velocity 28
2.3.2 Transversality of electromagnetic waves 30
2.4 Spherical waves 33
2.5 Polarization states 35
3 Quasi-monochromatic waves 37
3.1 Envelope and carrier frequency 37
3.2 Spectrum of quasi-monochromatic oscillations 38
4 Energy and momentum of light 42

v
vi

4.1 Intensity of light 42


4.2 The Doppler effect 47
4.3 Monochromaticity 50
4.4 Natural light 53
4.5 Light pressure 55
4.5.1 Radiation of a harmonic oscillator 55
4.5.2 LORENTZ force 57
4.5.3 Evidence of light pressure 60
4.6 Photon representation 61
3. LIGHT POLARIZATION 73
1 Basic types of light wave polarization 74
1.1 Linearly and circularly polarized waves 76
1.2 Elliptically polarized waves 78
1.2.1 Polarizer and analyzer 80
2 Polarization of quasi-monochromatic light 82
2.1 Polarization degree 82
2.2 The STOKES parameters and J ONES vectors 84
2.2.1 Description of a monochromatic wave 84
2.2.2 Measurements of the STOKES parameters 86
3 Optical dipole radiation and polarizing effects 89
3.1 Polarization under scattering 89
3.2 Polarization due to reflection 90
3.3 Dipole radiation at 3 em wavelength 92
4. LIGHT WAVES IN MEDIA 103
1 MAXWELL 's equations in isotropic media 104
1.1 Wave equations, the POYNTING vector 104
1.2 Boundary conditions 106
2 Reflection and refraction 109
2.1 Snell's law 109
2.1.1 Refraction by a prism 110
2.2 The total reflection 111
2.2.1 Surface wave (evanescent wave) 112
2.2.2 Total reflection of radio waves 114
2.2.3 Optical fibers 115
2.2.4 Rainbow 118
3 The Fresnel formulae 119
3.1 Reflectivity and transparency 121
Contents vii

3.1.1 Normal incidence 123


3.2 The Brewster angle 125
3.2.1 Polarizing devices 125
4 Dispersion 128
4.1 Classical theory of dispersion 128
4.2 Observation of dispersion 134
4.2.1 Crosses prisms 134
4.2.2 The Wood experiment 137
4.3 A wave train in a medium 138
4.3.1 Group velocity 138
4.3.2 Energy transfer 141
4.3.3 The Rayleigh formula 141
4.3.4 A modeling computation experiment 144
5 Radiation under uniform charge motion 147
5.1 VAVILOV-CHERENKOV radiation 147
5.2 Transient radiation 152
5.3 SMITH-PARSELL effect 153
5. OPTICAL ANISOTROPY 165
1 Double refraction in calcite 165
2 The structure of calcite 169
3 A monochromatic wave in an anisotropic crystal 173
3.1 Propogation of the energy and the phase 173
3.2 Nicol and Wollaston prisms 177
3.3 A model of an uniaxial crystal 178
3.3.1 Vector E parallel to the faces of the plates 179
3.3.2 Propagation across optical axis 180
4 Natural optical anisotropy of materials 181
4.1 Retardation plates 181
4.1.1 Experiment with radio waves 182
4.1.2 Quarter-wave plates in optics 183
4.2 Liquid crystals 186
4.2.1 Optical anisotropy of a thin film of a liquid crystal 187
4.2.2 A liquid crystal prism 188
5 Artificial birefringence 189
5.1 Photo-elastic effect 189
5.2 The Kerr effect 191
5.3 An experiment to determine the light velocity 194
viii

6 Optical activity 198


6.1 Natural quartz 198
6.2 The Fresnel experiment 200
6.3 Radio - wave rotation 202
6.4 The Faraday effect 204
6.5 Magnetic domains 205
6. GEOMETRICAL OPTICS 223
1 Homocentric and astigmatic beams 223
2 One refracting surface 227
2.1 Positive and negative optical force 230
2.2 Real and imaginary image 233
2.3 Magnification of a spherical refracting surface 234
2.4 A spherical reflecting surface 235
3 Centered optical systems 237
4 Lenses 240
4.1 General relations 240
4.2 Thick lenses 243
4.3 Thin lenses 246
4.4 Images formed by parallel rays 247
4.5 Two thin lenses 248
5 Errors of optical systems 252
6 Formation of optical images 257
6.1 Human eye 257
6.2 Magnifier 259
6.3 Telescopic system 260
6.3.1 Refracting telescope 260
6.3.2 Reflecting telescope 262
6.4 Microscope 264
6.5 Limitation of light beams in optical systems 265
7 Light propagation inside inhomogeneous media 269
7.1 Concept of eikonal 269
7.2 Light beam bending 275
7.2.1 Mirages 275
7.3 FERMAT'S principle 277

Index
289
Preface

In recent years, Optics has reemerged to be one of the most promising


sciences. Optical applications are found in widely differing scientific
fields - from physics and chemistry to biology, medicine and technical
sciences. Technical applications like laser printers, compact disk players
and data transfer in optical fibers are presently state of the art, but will
certainly undergo remarkable progress in future.
This book is the first of a two-volume textbook which emphasizes the
experimental demonstration of optical effects and properties and - more
importantly - the physical background of optical phenomena in order to
ensure a more didactical approach to the entire field of optics. Besides
its use as a textbook for students of natural or technical sciences, it
may become a valuable compendium of optical experiments for univer-
sity lecturers and motivated teachers in secondary schools. All necessary
parameters required to conduct the described experiments without prob-
lems in a standardly equipped lecture hall are included. Moreover, the
methodological approach to the field of optics via experiments may make
the book well-suited for practicing engineers who wish to increase their
understanding.
The first volume (Wave and Geometrical Optics) discusses wave prop-
erties of light such as its character as an electromagnetic wave and po-
larization and includes a chapter on geometrical optics. Phenomena due
to coherence properties of the waves, like interference and diffraction as
well as the statistical properties of light seen as a flux of photons will be
treated in the second volume (Coherent and Statistical Optics).

ix
Chapter 1

HISTORICAL INTRODUCTION

Optics, treated as a field of science, is concerned with the study of


t he properties of electromagnetic waves with short wavelengths (within
the spectral range of 200 - 1000 nm ), its propagation t hrough vacuum
and matter, and with the int eraction between light and mat ter. Such
elect romagnet ic waves have special significance for gener al hum an de-
velopment and for pract ical applications. ' Human eyes are sensitive to
wavelengths between 400 and 700 nm.
Nearly all basic scientific ideas and general concepts of nature have
been developed using visible radiation and optical studies. The his-
tory and evolution of the basic natural sciences, like astronomy, physics,
chemistry and biology et c., has always been strongly connect ed to op-
tics. Much progress in all these fields has been made by developin g
increasingl y sophis ticated opt ical instruments. But the development of
important physical ideas , principles of various measurements and exp er-
imental methods is also strongly connected to optical concept s. It is safe
to claim that optical methods still remain basic among all the inst ru-
mental techniques of contemporary experimental physics. Even today
many revolutionary changes and achievements are taking place in scien-
tific instrumentation, the creation of new measurement approaches, and
basic research. These developments , based on a deeper understanding
of light as a physical entity and its interaction with matter, cont inue to
expand our knowledge about nature considerably.

1. Wave nature of light


Long before the wave concept of light had been introduced, th e prin-
ciple of linear light propagation was the basic assumption in optics. The
everyday experience of the shadow cast by bodies illuminated . by the

1
O. Marchenko et al., Demonstrational Optics
© Kluwer Academic/Plenum Publishers, New York 2003
2 DEMONSTRATIONAL OPTICS

Sun , as well as Sun rays breaking through clouds, led to the idea that
light propagates from a source as a beam along a straight line trajectory.
For a long time , light was thought to be a stream of particles, emitted
from all visible bodies, penetrating into the eye. This mod el held un-
til the beginning of the 18th cent ury, and even LNEWTON (1643-1727)
considered light to have corpuscular nature. This model of light is in
agreement with the laws of geometrical optics.
But observations showed, that light propagation is a more complex
process. Thorough examination of the boundary between shadow and
light showed, that light also penetrates into the region of the geometrical
shadow, violating slightly the idea of linear light propagation. Obser-
vation of the light and dark bands near the boundary of a geometrical
shadow, and the similar effect of alternating sequences of bright and dark
bands, when light passes through a narrow slit in an opaque screen, lead
to ideas that light may be deflected slightly by geometrical objects.
In order to explain this diffraction phenomenon the Dutch scien-
tist CH.HUYGENS (1629-1695) introduced the following concept of light
propagation [1]. Light can be represented as a kind of oscillation, or
light wave, while propagating. Every point of the light wave originating
from a source may act as the origin of secondary waves propagating in
all directions. The resultant wave front at a particular moment may be
regarded as a superposition of all the secondary waves. This principle,
known as the HUYGENS principle, allows the determination of the posi-
tion of the current wave front , if the position of the wave front is known
for an earlier moment. A geometrical scheme, illustrating this principle,
is shown in Fig.!.!. Based on HUYGENS' geometrical representation the
phenomenon of light diffraction is easy to understand when considerat-
ing secondary sources located outside the obstacle. Light waves arising
from these sources and propagating in all directions are the reason for
the bright bands within the region of the geometrical shadow.
Within the framework of HUYGENS' principle the phenomena of light
refraction and reflection have been explained as well. HUYGENS pro-
vided an explanation for the birefringence phenomenon, discovered by
E .BARTOLINUS (1625-1698) in 1665. It was supposed, that within the
birefringent crystal there exists a second extraordinary wave apart from
the normal ordinary one. Its velocity is dependent on the propagation
direction within the crystal. The shape of the wave front of the extra-
ordinary wave is an ellipse, and not a circle as for the ordinary wave.
Studying the birefringence in Iceland Spar, HUYGENS discovered the po-
larization of the ordinary and the extraordinary beams. An explanation
of this phenomenon was offered much later.
Historical Introduct ion 3

Figure 1.1 Secondary


sources located on the
primary wave front F1
(denoted by dark points)
give rise to a set of elemen-
tary spherical waves. The
envelope of these spherical
surfaces at some moment
gives the new position of
the wave front F2 . The
direction of propagation is
specified by the arrow .

Introduced by HUYGENS, the concept of light as propagating like a


wave (of special, but unknown type) originated in the middle of the
seventeenth century. At that time, the only known wave motion within
material bodies was the sound wave. In analogy to sound waves, the
wave concept of light was based on the general idea, that the light is the
propagation of mechanical oscillations within a light carrying medium,
called aether, which fills all space and penetrates all material bodies .
The next important principle of the wave concept of light was intro-
duced by T .YOUNG (1773-1829). It says, that the superposition of light
waves arriving at a point of observation may increase or decrease the am-
plitude of the oscillations of the resultant wave [2] . This principle, known
as the interference principle, was used by YOUNG to explain the colors
of thin films and to explain the observation of rings in the interference
experiment of I.NEWTON. YOUNG conducted a classical experiment on
light interference from two light sources as is shown in Fig.1.2. These
two sources were realized by illuminating an opaque plate with two small
holes SI and S2 , located close to each other, with a Sun beam that has
passed through a third small hole before. Light beams from these two
sources are superimposed in the region of a screen, forming the interfer-
ence pattern. At points P, where the path difference I SIP - S2P I is
equal to an integer multiplied by the wavelength>' of the light wave, the
amplitude of the oscillations of the resultant wave is larger compared to
the oscillations of each of the separate waves.
If the path difference I SIP - S2P I is equal to an odd integer mul-
tiplied by half the wavelength, >'/2, a decrease in the amplitude of the
oscillations in the resultant wave takes place. These points P correspond
to the dark bands of the interference pattern (Fig.1.3). Analyzing re-
lations between the geometrical parameters of this experiment, YOUNG
4 DEMONSTRATIONAL OPTICS

Figure 1.2. Interference of light waves from sunlight S in the double-slit scheme of
T .Young.

estimated the wavelength of the light wave and found it to be very small ,
on the order of magnitude of >. rv 500 nm.
At the beginning of the nineteenth century, a significant contribution
to the field of optics was made by A.FRESNEL (1788-1827). The main
idea of FRESNEL's work on the quantitative explanation of the diffrac-
tion phenomenon was the unification of HUYGENS' principle together
with the concept of secondary wave interference [3]. This new principle,
called later the HUYGENS-FRESNEL principle, allowed the description
of experimentally well-known diffraction phenomenona: The diffraction
from the edge of an opaque screen and the diffraction from a small hole
in the screen . Later on G .KIRCHHOFF (1832-1887) int roduced a rigorous
mathematical interpretation of the HUYGENS-FRESNEL principle. Even
today the KIRCHHOFF integrands are used for approximate computa-
tions of diffraction patterns.
In 1816 FRESNEL conducted several important optical experiments.
He observed the interference of waves reflected from two mirrors posi-
tioned at an angle close to 1800 • He also noted an interference pattern
created by two refracted light beams intersecting two glass prisms. Sig-
nificant improvements were achieved in the determination of the na-
ture of light by studies of interference phenomena by FRESNEL and
D.F.ARAGO (1786-1853) (Figs.1.4, 1.5). They also examined the in-
terference of polarized light beams . The experimental fact that no in-
terference pattern can be observed when superimposing two light waves
polarized in two mutually orthogonal planes was in contradiction to the
dominant concept of a longitudinally oscillating light wave. Based on

Figure 1.3 Typical int er-


ference pattern produced
by light waves in Young 's
double slit experiment .
Bright bands are the
maxima ofinterference.
Historical Introduction 5
Figure 1.4 The bright
spot in the center of the
geometrical shadow is
resulting from Fresnel
diffraction on a small disk.
The existence of such a
phenomenon was predicted
by Fresnel and confirmed
by Arago . Such kind of
diffraction phenomenon is
called Poisson 's spot .

this crucial result, YOUNG proposed that the oscillations of light waves
are strictly transversal.
The idea of transversal oscillations in light waves enabled FRESNEL
to work out a new approach to the phenomena of polarization and the
propagation of light inside matter. According to FRESNEL light emit-
ted by a radiating body is not polarized. Every molecule of a body
emits a linearly polarized wavelet, but, due to the chaotic motion of the
molecules, the resultant light wave emerging from a radiating body after
superposition of the elementary wavelets results in a wave with oscilla-
tions that chaotically vary their direction. Such light is called natural
light.
Based on these considerations, FRESNEL obtained important results
concerning light wave propagation within anisotropic crystals. Waves
having mutually orthogonal oscillation directions can propagate in every
crystal, but they will propagate with different velocities if the crystal is
birefringent. Studies of the optical activity of quartz became the basis of
FRESNEL'S hypotheses on how the propagation velocity of the circularly
polarized waves depends on the type of polarization. Waves with left and
right circular polarizations have different velocities when propagating
along the optical axis of a crystal. FRESNEL showed experimentally,
that an incoming linearly polarized wave is split into two waves with left

-:0--------
:~
--_. --;,----- Screen
~

Figure 1.5. Light rays, falling on a rather small disk, are deflected from straight lines
due to diffraction on the disk, causing Poisson's spot.
6 DEMONSTRATIONAL OPTICS

and right circular polarizations after refraction by a system of quartz


prisms .
FRESNEL'S theory of light wave polarization has led to analytical re-
lations between the intensity of incident, reflected and refracted waves
as a function of the angle of incidence and of the orientation of the
polarization of the wave with respect to the incident plane. These rela-
tions are known as FRESNEL 'S formulae . Theoretical works by FRESNEL
on the polarization of light were examined by his contemporaries very
carefully, since the initial assumption behind his work is the transversal
nature of light waves. As mentioned above, the predominant concept at
that time was the idea of an aether as an elastic medium, in which only
longitudinal waves may propagate.

2. Electromagnetic theory
The concept of the electromagnetic field, which was completed by
G.K.MAXWELL (1831-1879) was proved to be totally consistent with the
theoretical assumptions and notations of FRESNEL. The concept of the
force field was introduced by M.FARADAY (1791-1867) into the science
of electricity and magnetism in the first half of the nineteenth century
[4] . Initially the lines of force of electric and magnetic fields were treated
as an aid for describing electric and magnetic phenomena in a clear and
pictorial way. But during the following studies of electromagnetic fields,
especially after the discovery of electromagnetic induction by FARADAY,
this concept was filled with a new physical meaning and enabled the
formulation of the laws of electromagnetism in a concise and clear form.
The strict mathematical representation of the concept of the electro-
magnetic field was accomplished by MAXWELL [5]. He expressed the
phenomenological electromagnetic laws in terms of consistent mathe-
matical equations. MAXWELL'S hypotheses of the displacement current
connected the alternating electric and magnetic fields existing within
a limited space region, even when no electric conductivity exists in the
material bodies transporting the electromagnetic field. This fact became
the principle point for the physical introduction of free electromagnetic
waves. MAXWELL came to the conclusion, that light is an electromag-
netic wave with a frequency within a particular range. The velocity
of the electromagnetic waves propagating in a medium should depend
on the mediums electric permittivity and magnetic permeability (c, J.L) .
The measurements of these constants were performed independently as
electrical quantities, and from these figures the velocity of light could
be determined to a value consistent with optical measurements. At this
historical moment electromagnetism and optics led to the same conclu-
sions, namely that (i) a light wave is a transversal wave and (ii) that the
Historical Introduction 7

velocity of light is equal to the velocity of electromagnetic wave prop-


agation. Since then light has been considered to be a special type of
electromagnetic wave with a very short wavelength .
The next principle phenomenon, studied by MAXWELL, was the light
pressure. He showed that light should exert pressure on material bodies .
This pressure turned out to be dependent on the electromagnetic energy
density.
The development of MAXWELL'S electromagnetic theory of gave im-
petus to studies of the effects of electromagnetic waves on matter con-
taining elementary electric charges. Around 1870 G.A.LoRENTZ (1853-
1928) began to devise his theory of displacements of electric charges,
incorporating the basic ideas of electromagnetic and optical phenom-
ena. The first important success of the electron theory was an explana-
tion of the splitting of spectral lines in an magnetic field, discovered by
P .ZEEMAN (1865-1943) in 1896. At the end of the nineteenth century
it was believed, that the electromagnetic theory of MAXWELL , together
with the electron theory were capable to explain all known optical effects.

3. Optical Spectroscopy
At the middle of the eighteenth century it was discovered, that the
radiation of different flames are not only comprised of a continuous-
intensity spectrum. Some parts of the spectrum were observed to be
more intense than the background spectrum, and showed the well-ex-
pressed character of discrete spectral lines. Systematic studies of optical
effects within absorption spectra were started after the discovery of the
dark lines in the solar spectrum by J .F RAUNHOFER (1787-1826), named
Fraunhofer lines in his honor. The chemical spectral analysis was based
on the idea that all the gaseous chemical elements possess a specific line
spectrum. Since then studies of line spectra have become a part of op-
tics, and large efforts were made to create and to improve the required
optical devices and instruments like prism and grating spectrographs.
During the second half of the nineteenth century, atomic spectroscopy
has become a widely developed experimental science, gathering phe-
nomenological data of principle significance. Nevertheless the classical
electrodynamics of MAXWELL did not explain the emission of these dis-
crete line spectra. An explanation of the spectral peculiarities of optical
radiation was not given until the beginning of the twenteeth century,
when the fundamentals of quantum theory were put forward.
The requirements of experimental atomic spectroscopy stimulated the
creation of more sophisticated spectral devices and new generations of
optical instruments. The first diffraction gratings were used by FRAUN-
HOFER to study the spectrum of the Sun at the beginning of the nine-
8 DEMONSTRATIONAL OPTICS

a b

Figure 1.6. (a) Two plane parallel plates, forming a so-called Fabry-Perot interfer-
ometer , give rise to a set of reflections of an incident ray between their inner surfaces.
The interference of these rays is the reason for appearence of distinct and sharp in-
terference fringes. (b) A fragment of a typical interference pattern after a Fabry -
Perot interferometer.

teenth century. Later on, techniques for the production of reflection


gratings were improved by H.A.RoWLAND (1848-1901), A.MICHELSON
(1852-1931) and other researchers. Double-beam interferometers of
JAMIN-, MACH-ZENDER-, and TWYMAN-GREEN-types were created for
studies of the refraction index behavior close to an absorption line of the
studied vapors.
The multi-beam FABRy-PEROT interferometer has earned a special
place among optical spectral devices (Fig1.6) . This optical instrument
is widely used in spectroscopy due to its simplicity as a tool to study
the structure of separated spectral lines. For a typical construction of
the FABRy-PEROT interferometer two well polished glass or quartz plates
with highly reflecting layers put at the inner surfaces are mounted strictly
parallel to each other. The invention of this interferometer made it possi-
ble to determine very accurately the wavelength of atomic emission lines
and to investigate new effects in atomic physics like hyperfine splitting
and isotope shifts of spectral lines in fine details. Later on, the spheri-
cal type of FABRY-PEROT interferometers became the first resonator of a
laser . Still the FABRY-PEROT interferometer is the key to understanding
the physics of lasers.

4. Quantum Optics
Quantum theory was founded mainly on the pioneering results of
M.PLANCK (1858-1947), who resolved the black body problem [6] . Using
a statistical approach to the thermodynamical problem of the equilib-
rium between radiation and matter, he introduced discrete portions of
light energy (quanta) hu (1900).
Historical Introduction 9
radiation substance light quanta substance

0---
0- hv
hv
a b

Figure 1.7. Treating the black-body problem by M.Plank (a ) and A.Einstein (b ).


According to Plank's theory radiation has to be recognised as a continuous field, by
Einstein's hypotheses the radiation consists of light quanta, each carrying the energy
quantum hu ,

The majority of the known optical phenomena and the interaction of


light with matter had been explained at the end of the nineteenth cen-
tury based on the electromagnetic nature of light in association with the
LORENTZ electron theory. Nevertheless some optical phenomena con-
cerning light absorption and emission still remained unexplained. For
instance, the energy distribution within the equilibrium radiation spec-
trum of a black body was not explained.
At a certain temperature a heated solid body emits continuous radi-
ation with an intensity distribution characteristic for this temperature.
A part of this radiation corresponds to the visible range of the spectrum
and causes well-known visible effects. Studies of the spectral energy dis-
tribution of the radiation of heated bodies were made by G .KIRCHHOFF
(1824-1887), J.W.RAYLEIGH (1842-1919) and J.H .JEANS (1877-1946),
L.BoLTZMANN (1844-1906) and J .STEFAN (1835-1893) following the ther-
modynamic approach. The first theory, explaining a part of the spectral
dependency of the emission intensity of the heated bodies was published
by M .K.WIEN (1866-1936). His equation provided a satisfactory de-
scription of the experimental spectral dependency within the short wave
region, but not for large wavelengths . RAYLEIGH and JEANS gave an
equation which describes only the long wavelength region , but gives un-
acceptable results for short wavelengths.
This discrepancy was removed by PLANCK, who made an attempt to
deduce the law of the equilibrium radiation based strictly on thermody-
namics. This approach did not deliver the desired result and PLANCK
had to dismiss the thermodynamic definition of entropy and use the sta-
tistical definition of this quantity, which was introduced by BOLTZMANN .
A consequence of the statistical approach was the assumption of the dis-
crete character of absorption and emission of radiation by matter. The
famous Planck constant h is the proportionality coefficient between the
10 DEMONSTRATIONAL OPTICS

minimal portion of energy (or the energy quantum) and the frequency v
of a monochromatic wave within the continuous spectrum of radiation.
For a number of years the idea of the energy quantization was doubted
and PLANCK 'S method for the computation of the entropy was regarded
as a special mathematical trick. A new interpretation of the idea of
quanta was provided in the work of EINSTEIN (1879-1955)[7] who ex-
plained in 1905 the light-electric effect (or photoeffect ; the emission of
electrons from a metal surface by interaction with ultraviolet light), as-
suming that such a discrete energy portion lu/ is handled to one electron.
His treatment of radiation based on a statistical approach introduced the
concept of the light quanta. These light-quanta much later were named
as photons. In a way this was a step back from the wave concept to
a particle concept of light and did lead later to a dualistic picture of
light where wave and corpuscular properties are necessary to describe
its nature.
While in PLANCK'S theory the entropy of matter was treated statisti-
cally, EINSTEIN considered the statistical nature of the entropy of radia-
tion itself. Such an approach to electromagnetic radiation led EINSTEIN
to the hypothesis of the quanta of the electromagnetic field. Using the
energy conservation law in the process of the absorption of a photon by
an electron, EINSTEIN has explained a number of photochemical effects,
including the phenomenological features of the photoeffect. Therefore,
the field of quantum optics is built upon the quantum nature of in-
teraction of electromagnetic fields with matter and originates from the
explanation of the peculiarities of black body radiation and the empirical
laws of the photoeffect (Fig.1.7).

5. Atom model and relativity


Until the first years of the twenteenth century there enough exper-
imental evidence had been collected to treat electrons as one of the
principle elementary parts of every material body. It was known, that
the typical size of an atom is several Angstrom (1 A= 10- 10 m) and an
estimate for the size of an electron gave r e rv 10- 15 m. The important
interdisciplinary question of a model for the atom arose.
A number of brilliant experiments performed by E.RuTHERFORD (1871-
1937) on the scattering of Q: particles on gold atoms provided an ex-
perimental proof for the existence of dense atomic nuclei, concentrated
within a small volume having a size around 10- 14 m [13] . These ex-
perimental results led to a planetary model of the atom, proposed by
RUTHERFORD in 1911. The negatively charged electrons were kept at
elliptic orbits by the Coulomb attraction forces to the positively charged
nuclei. This model was in contradiction to the conclusions of electromag-
Historical Introduct ion 11

/
/----(t---', hv

,,
/

8 ~
/

o o
I
I
I
I I
\
\ ,
I
\
,, /

,
/
/

a b

Figure 1.8. The planetary model of an atom by Rutherford and Bohr (a) : according
with Bohr 's principles, but in contrast to Maxwell 's electrodynamics, every electron
moves along a stationary orbit . An energy quantum hv is emitted during the transi-
tion of the electron from a higher orbit to a lower one .

netic theory, since an electron moving along a closed orbit should emit
electromagnetic waves. The loss of energy needed for radiation should
lead to the electron falling into the positively charged nucleus. In other
words, such a system of charged particles was treated to be unstable.
In 1913 N.BOHR (1885-1962) made an at tempt to explain th e stability
of the hydrogen atom [8] which turned out to be very fruitful. BOHR
accepted the planetary model of atom by RUTHERFORD , but added the
assumption that the electron can run around the nucleus in stable orbits
when it has certain amounts of angular momentum. Being in one of
these orbits the electron should not emit radiation and therefore not
lose energy. The electron emits radiation only when it is undergoing a
transition between two such orbits (Fig.1.8). The energy of the emitted
quantum should be equal to the difference of energies En of these orbits
(n is an integer, numerating the stable electron orbits) : E n+ 1 - En = lu/
(again in contradiction to electromagnetic theory, where the frequency
of the emitted wave should be determined by the rotation frequency
of the orbiting electron) . When absorbing energy a transition of the
electron from a lower orbit to a higher one takes place. Using this
semi-classical model BOHR obtained a formula for the BALMER spectral
series of atomic hydrogen , which was in excellent agreement with the
spectroscopic observations.
In several respects BOHR'S theory caused discomfort. For example,
this model could not explain the intensity relations between the spectral
lines of the BALMER series, and it could not be used for atoms heavier
than hydrogen. Also the extended model of A.SOMMERFELD (1868-
1951) failed in these respects. Nevertheless, the first works of BOHR
12 DEMONSTRATIONAL OPTICS

caused a large number of experimental and theoretical studies, which


finally did lead to the development of a quantum theory of atoms. Moti-
vated mainly by works of W.HEISENBERG (1901-1975), E.SCHRODINGER
(1887-1961), L.DE BROGLIE (1892-1981), within a very short time (1924-
1928) it was possible to confirm that quantum theory provided a correct
description of atomic processes. Not more than ten years had passed be-
fore the introduction of the new quantum mechanical picture allowed an
explaination of the remarkable results of BOHR at a rigorous theoretical
level.

According to the principles of quantum mechanics objects like micro


particles and radiation quanta exhibit wave characteristics and particle
characteristics at the same time. The wavelength >. of a 'matter wave'
(the de Broglie-wavelength) depends on the momentum p of a uniformly
moving particle through the famous relation discovered by DE BROGLIE:
>. = hlp (h is PLANCK'S constant). This relation holds even for light-
quanta or photons. It depends on the experiment wether electrons and
photons as well as other microparticles manifest their wave or their cor-
puscular properties. It is also impossible to predict the properties of
a single photon or a single microparticle. Only the averaged behavior
of a large number of photons or microparticles may be predicted. For
observing specific states of quantum objects only a probability can be
calculated.

Let us return to the question of a transport medium for light , or, more
general , electromagnetic waves. To prove the existence of an absolute
aether within which the Earth is moving, MICHELSON constructed a
special type of interferometer and performed his famous experiment in
1887 (together with E.MoRLEY (1838-1923)) in which he showed that
the light velocity does not depend on the direction relative to the ve-
locity of Earth and that it is impossible to detect the "aether wind"
(Fig.1.9). This experiment - together with the transformation equa-
tions of LORENTZ - was the basis on which EINSTEIN could later build
his theory of special relativity. This theory removed completely the con-
cept of aether and the electromagnetic field has received the status of
an independent physical entity.

Let us emphasize once more that all the progress in quantum theory
and relativity theory is based on the development of very precise and so-
phisticated optical instruments, like spectrographs, interferometers and
large telescopes (to test relati vity theory in space) .
Historical Introduction 13

stationary aether

c
K K

-v -v

a b

Figure 1.g. With the assumption of a "stationary aether", the reference K, vibrations
of the aether, caused by the light source S in the unifomly moving laboratory K',
would propagate with different velocities cr and C2 . Therefore the velocity of light
propagation should be dependent on the orientation of velocity v relative to the aether
(a) . According to the special theory of relativity the aether doest not exist : the light
velocity c neither depends on the orientation of the unifomly moving laboratory frame
with respect to another "stationary" reference K nor on a choice between uniformly
moving references .

6. Coherence and precision optical


instrumentation
Simultaneously with the development of optical instruments, the in-
fluence of the size and the spectral composition of the source of radiation
on the formation of optical images was studied. The quality of an in-
terference pattern is mostly affected by these quantities for every inter-
ference scheme. Such investigations were stimulated by the specialized
field of optical physics that appeared at that time , which is mainly con-
cerned with metrology and standardization of basic physical quantities.
Metrology as a scientific field emerged under the influence of the works
of MICHELSON on measurements of the coherent length of spectral lines.
MICHELSON found that among all studied spectral lines, the red cad-
mium line (>' = 643.8 nm) is closest to ideal monochromatic radiation
[10]. The wavelength of this line was used for a long time to standardize
the unit of length.
In astronomy, an essential increase in the resolution power, compared
to normal telescopes, was achieved with the help of the stellar interfer-
ometer constructed by MICHELSON [11] (Fig .1.lO). It became possible
to measure the angular dimensions of gigantic stars such as Q Orion.
Two light beams incoming from a star create a system of interference
fringes in the focal plane of the telescope objective. When increasing the
14 DEMONSTRATIONAL OPTICS

Figure 1.10 Operating


principle of Michelson's
stellar interferometer.
Parallel rays from a star
pass through two holes
and are then focused by a
telescope objective. The
contrast of the interference
pattern caused by these
rays is dependent on the
separation between the
holes a and b.

distance between the beam axes (the interferometer base) the contrast of
the interference pattern decreases (Fig. 1.11). A complete disappearance
of the interference takes place at a certain base length. The angular
dimension of the star can be determined from the base length and the
known light wavelength. MICHELSON introduced a quantitative charac-
teristic of the interference contrast (called the visibility function), which
allows analysis of the form of the radiating surface. Similar visibility
functions were obtained by MICHELSON when studying the effect of the
shape of the emission spectral lines on the interference pattern contrast
in the MICHELSON interferometer. In particular he obtained, that the
spectral profile of the red cadmium line has a Gaussian shape.
Later on, the studies of MICHELSON were the background for theoreti-
cal models of the partial coherence theory. At that time the development
of this theory was provided by the works of P.H.VAN ZITTERT [12] .

a b c

Figure 1.11. When Michelson's stellar interferometer is illuminated by a round bright


source of rather small size, the increase of the separation between two holes in sequence
gives rise to patterns (a), (b) and (c) . If the interference fringes just disappears, like
in pattern (c) , the angular size of the source can be calculated.
Historical Introduction 15

Figure 1.12 A sim-


ple schematic of the
Brown-Twiss stellar inter-
ferometer. Parallel rays
Detector from a star are focused
by two mirrors on the
photocathodes of two
~ .----'-- Mirror
identical photodetectors.
The currents of the detec-
tors are multiplied by the
electronic mixer and the
resulting signal is counted.
At a critical distance
between the mirrors the
signal goes down to noise
level, allowing an estima-
tion of the angular size of
the star.

The next important contribution to the theory of coherence was made


by F. ZERNIKE [131, who introduced a concept of the degree of coherence ,
which is released to the visibility function. By these means , the degree of
coherence became an experimentally measurable quantity. Of particular
importance is that coherence theory, through not self-consistent, leads
to measurable quantities, such as the time-averaged light intensity.

7. Light correlation and statistics


A new stage in the development of coherence theory was initiated
by a number of experiments significanct to the evolution of further op-
tical ideas. These experiments, performed in the microwave or radio-
frequency range, were characterized by a special technique, where the
measured quantities were proportional to the square of the high-frequency
photocurrent and consequently to the microwave intensity. In 1954
R.HANBARRY-BROWN and R .TwISS discussed a new type of radio-
interferometer for measurements of stellar emission in the radio-frequency
range [141 (Fig.1.12). Radio-waves from a star reache two antennae. De-
tectors and signal processing was chosen to give finally signals propor-
tional to the intensities of the electromagnetic radiation at the locations
of the antennas. After multiplication and averaging of the signals , the
intensity correlation coefficient could be calculated as a function of the
of the distance between both antennas. An increase of the distance leads
to a decrease of the correlation signal down to a certain constant level,
which enables the estimation of the angular dimension of the radiation
source.
16 DEMONSTRATIONAL OPTICS

After several successful measurements in the radio-wave region HAN-


BARRY-BROWN and TWISS carried out a laboratory optical experiment
similar to the radio-wave experiment described above [15] . As averaging
detectors they used two photomultipliers. The objective of this exper-
iment was to detect the photocurrent correlations arising between two
registration channels (photodetectors) by partially coherent light beams
created from the same light source . For the first time in this experi-
ment correlations between light beam intensities were detected. Later
on, a stellar interferometer for visible light were designed , based on the
laboratory model. It demonstrated the principle of the correlation mea-
surements, proposed by HANBARRY-BROWN and TWISS . With such
methods the angular dimensions of bright stars like Sirius (or a Canis
Majoris) have been determinedjlfl] . At the same time the theory of light
coherence were significantly enhanced by E .WOLF [17].
Another type of intensity correlation experiment was realized by A.
FORRESTER, R .GUDMUNDSEN and P.JOHNSON [18]. The superposition
of two oscillations having nearly the same frequencies leads to a beat pat-
tern in the resultant oscillation with a frequency equal to the difference
in frequency between the two primary oscillations. The FORRESTER et
al. experiment was designed to measure the signal created by optical
beats of two closely neighboring spectral lines. For observation of the
beats, special experimental conditions were maintained. The width of
each of the lines had to be approximately ten times smaller than the fre-
quency difference between the line centers, and the frequency difference
could not be larger than 1 GHz. The expected beats in the sum radi-
ation of two independent sources (represented by the separate spectral
lines) were observed successfully. Despite the differences in the experi-
mental techniques, the experiments of HANBARRY-BROWN and TWISS
and FORRESTER et al. both demonstrate methods of treating intensity
correlations (and the corresponding photo current correlations) between
partially coherent light beams.
Chapter 2

PRINCIPLE PROPERTIES
OF LIGHT WAVES

1. Laws of electromagnetism
1.1 Maxwell's equations
The transversal wave nature of light was confirmed by studies of elec-
tric and magnetic phenomena carried out during the late in nineteenth
cent ury. Furthermore, light was found to be an electromagnetic wave.
These studies led to laws which established the link between the sources
of electric and magnetic fields and these fields itself. J .K .MAXWELL was
able to find his famous set of equations for the properties of the electric
field strength E and the magnetic induction B. Two of these equations
may be written (for propagation of the electromagnetic field in vacuum)
in the following form:

f
s
E dS = ~q
co
(2.1)

fs B dS = 0 (2.2)

where cO is the permittivity of the free space. Its magnitude in 81 units


is:
cO = 0.8854 X 1O-1l F· m- 1_.
The first equation represents GAUSS'S law, which states that a charge q,
surrounded by a closed surface S (with the surface element dS) is the
source of an electric field E . Let us draw a unit vector n at a certain point
perpendicular to this surface and form an elementary surface vector dS
(Fig.2.1.a) . The scalar quantity EdS is the elementary flux of the vector
E through the element of the surface dS. GAUSS'S law states that the

17
O. Marchenko et al., Demonstrational Optics
© Kluwer Academic/Plenum Publishers, New York 2003
18 DEMONSTRATIONAL OPTICS

a b

Figure 2.1. The electric field flow through an element dS is the scalar product E ·dS,
where S = oS. 0 is the un it vector of the element dS (a) . The total electric field
flow through the closed surface S surrounding the positive charge +q is propotional
to th e quantity of q (b) .

total E vector flowing through a closed surfac e is proportional to the


electric charge q located inside this surface (Fig.2 .l.b) .
Contrary to the electri c field, the flux of the magnetic induction B
through a closed surface is always zero, as follows from equation (2.2).
This result corresponds to the non-existence of magnetic charges and the
fact that the lines of magn etic indu ction are always closed. The cause of
magnetic fields are moving charges (given often as electric currents). The
law of BIQT-SAVART-LAPLACE est ablishes a link between the magnetic
induction B and the total current density j:

c
2
f
L
B dl = c~ J
S
j dS

where c = 299,792.458 km/s is the electrodynamic constant (and the ve-


locity of light in free space) associated with the electric cO and magnetic
/-Lo const ants (/-LO is the permeability of free space) as follows:

c= 1/JcO/-Lo
where f.LO = 471" . 10- 7 N·A-2.
The flux of vector j on the right hand side of the last equation is
calculated through the open surface S, determined by the closed contour
L. The orientation of the elementary vector dl indicates the direction to
pro ceed along the contour L , as shown in Fig.2.2.
Like moving charges and currents , an alternating electri c field also
causes a magnetic field. This physical result may be generalized by the
PROPERTIES of LIGHT WAVES 19

Be?
a b

Figure 2.2. The magnetic induction B caused by current I (a) illustrates the law of
Biot-Savart-Laplace. The line of the vector B around the closed contour L is caused
by the current distribution j, where the magnitude of B is propotional to the flux of
t he vector j through the area surrounded by contour L (b) .

following equation from the MAXWELL system:

c
2
f B dl = ~ JE(t) dS + e10 J j dS (2.3)
L S S
In accordance with the electromagnetic induction law of M .FARADAY,
a variation of the magnetic field flux through a certain surface creates
a curling electric field. The following MAXWELL equation is another
mathematical expression of FARADAY 'S law:

f
L
E dl =- ~ J
S
B(t) dS . (2.4)

The circulation of the vector E around a contour L is equal to a variation


of the magnetic induction vector B through the surface S (Fig.2.4) .
We see that spatial distributions of charges and currents are sources of
electromagnetic fields in vacuum. Additionally, an alternating magnetic
field is caused by an alternating electric field. In turn an alternating elec-
tric field causes an alternating magnetic field and so on. The first term
on the right-hand side of eq, (2.3) is proportional to the displacement

J
current (see Fig .2.3)
eo~ E dS (2.5)
S
Integral equations (2.1-2.4) are called MAXWELL'S equations for electric
and magnetic fields in vacuum.

1.2 Electric dipole radiation


The basic property of electri c and magnetic fields is that they obey
the MAXWELL equations in two specific cases - (i) when the reference
20 DEMONSTRATIONAL OPTICS

Figure 2.3. The magnetic field is Figure 2.4. Illustration of the electro-
caused by changes in the electric field magnetic induction: The electric field
E(t) and the current distribution j around the closed contour L is caused
by changes in the magnetic induction
flow through the contour L.

frame is at rest and (ii) when the reference is in uniform motion. This
statement follows from the special theory of relativity: the laws of elec-
tromagnetism must be similarly represented in a system of reference at
rest and one moving with uniform velocity. In both cases, one has mov-
ing, but not accelerated charges. In the special case of a single charge,
this property leads to the conclusion that a charge at rest or in uniform
motion does not emit electromagnetic radiation. It should be noted that
this statement is completely correct in case of a charge or a system of
charges in vacuum, whereas uniform motion of charges inside a medium
may cause electromagnetic radiation. Such radiation caused by charged
particles moving inside a medium does not contradict MAXWELL 'S elec-
trodynamics. The peculiarities of this kind of radiation will be discussed
in Chapter 4.
The existence of electromagnetic radiation, or electromagnetic waves,
is one of the fundamental consequences of MAXWELL'S electrodynamics.
An electromagnetic wave is considered to be free oscillations of elec-
tric and magnetic fields propagating through space. Let us treat now
the particular case of electromagnetic waves caused by a charge moving
accelerated in vacuum.
We know the electric field of an electric point charge q is radially
symmetric. The electric field strength decreases as the square of the
distance r between the charge position and the observation point:

E(r) - _1_1.. (2.6)


- 47l'eo r 2
At an initial moment t = to let the charge be at rest at the origin of
a Cartesian reference system (point A in Fig.2.5), and then let it move
PROPERTIES of LIGHT WAVES 21

y t
t
I
~- - - - - -

:--f - - - - - ;
. i
"
,-

" r = c (t-to)
i
,, ,,
I ,
. lq
AB .' .. I ..----...

Figure 2.5. Electric field of a moving Figure 2.6. The tangential compo-
positive charge. nent of the electric field line

with an acceleration a during the time interval r along the x-axis. At


the time to + r the charge has the velocity

v = ar (2.7)

and then moves further with constant velocity v. It is possible to dis-


tinguish three phases of charge motion in this case: the first is rest at
the origin of the reference frame, the second one is a motion with steady
acceleration and the third one is a linear motion with constant velocity.
An electromagnetic field is not radiated during the first and third phases
of charge motion. Radiation takes place due to accelerated motion only
during the second phase of motion, thus, during the time interval r .
At the moment t one finds the charge at a point B :

x = v(t - to) + vr /2 . (2.8)


For t » r it is possible to neglect the charge shift during the acceleration
phase vr /2 compared to the distance v(t-to) and one gets approximately

x~v(t-to)

Restricting ourselves to the case v « c it is possible to consider the


charge field at point B to be radially symmetric, as for a charge at rest.
In order to describe the field lines of the accelerated charge we use the
following method. The electric field perturbation caused by the charge
acceleration spreads with the velocity of light c. Therefore the electric
field outside a sphere with radius c(t - to) centered on the origin of the
reference frame consists only as the field created by the charge at rest
at the moment t = to. Inside a sphere with radius c(t - to - r) centered
22 DEMONSTRATIONAL OPTICS

on point B, we have an electric field corresponding to a steady moving


charge. We depict the field of the charge at rest as radial lines originating
from the external spherical surface. Extensions of these lines inside the
sphere intersect at the origin A of the reference frame (where the charge
was at rest before its accelerated motion). Let us add radial lines of the
electric field from the point B to the internal sphere. Using arrows we
join these lines with the outer field lines to maintain the continuity of
the electric field lines. The arrows of the field lines located between the
inner and the outer spheres represent the field due to the accelerated
charge. As it is shown in Fig.2.6, the field of the accelerated charge
has a radial Ell and a tangential El.. component. The radial component
Ell is the usual Coulomb field for which the relation (2.6) is valid. The
tangential component El.. represents the variable radiation field. Let us
find the component El.. for the field line originating at an angle () from
the direction of charge motion (the x-axis). From Fig.2.6 we find:

El.. v(t - to) sin ()


= (2.9)
Ell CT

According to (2.6) we have:

E _ _ 1_!L __1_ q
I - 411"co r 2 - 411"co c2 (t - to)2

Substituting the expressions for Ell and T = via into formula (2.9), we
find the following equation for the tangential component of the field El..

El.. = _1_ qa sin () = 1 qa sin () (2.10)


411"co c3(t - to) 411"co c2r
where r satisfies the simple equation

r = c(t - to) (2.11)

Therefore the radiation field is proportional to the amount of charge,


the acceleration and inversely proportional to the distance between the
charge and the point of observation. This field is fundamentally differ-
ent from a Coulomb field, which is proportional to the inverse square
of the distance. The finite time of the variable field propagation is ac-
counted for (2.10) through of the temporal dependency of expression
(2.11). So, the field variation beginning at the time to reaches the point
of observation after a finite time interval
r
t - to =-.
c
PROPERTIES of LIGHT WAVES 23

An important feature of this type of radiation, called the dipole radiation,


is that there is no field emitted in the direction of charge motion. It
strictly follows from formula (2.10) for 0 = 0 and pictorially it can be
seen in Figs.2.5 and 2.6. Actually electric field line bending does not exist
in the direction of charge motion. At the same time maximal curvature
of the electric field lines and the maximum value of El.. is obtained in
the direction normal to the charge motion (0 = 7C' /2).

2. Properties of electromagnetic waves


2.1 Wave equation
We have discussed above the system of the integral MAXWELL equa-
tions (2.1-2.4). Principle empirical laws of electromagnetism, like the
laws of GAUSS, BIOT-SAVART-LAPLACE, and FARADAY'S law of electro-
magnetic induction are contained in these equations. In addition to the
integral form there is the well known differential form of MAXWELL'S
equations.
In Cartesian coordinates the divergence of a vector function E( x, y, z)
takes the from:
dtv·
E_
-
oEx oE oEz
y
ox + oy + oz .
The curl of a vector function E( x , y, z) is given by:

rotE = i (OEz _ OEy ) + j (OEx _ OEz) + k (OEy _ OEx)


8y OZ oz ox ox 8y
where i,j, k are unit vectors in the positive direction of each axis, re-
spectively [41].
The whole system of MAXWELL'S equations in differential form de-
scribes an electromagnetic field existing in vacuum:

divE = ~p (2.12)
eo
divB = 0 (2.13)
2
c rot
B
= -oE 1.
ot +-J
eo
(2.14)

rotE = -oB
ot- (2.15)

where p is the electric charge density in space and j is the current density
in space.
The elementary flow of an electric or magnetic vector from an in-
finitesimal volume is given by the scalar magnitude of the divergence
(Fig.2.7,a,b). If electric charges are distributed with the density p(x, y, z)
24 DEMONSTRATIONAL OPTICS

s -
//
---
div E
<,
"
B
( -k i
I ~I

--- -
\ /
-, P (X,y,Z) //
.......
b
a

Figure 2.7. Divergence of electric and Figure 2.8. The integration over a
magnetic field lines. A spatial charge finite volume V enclosed by surface
distribution p(x, y, z), existing inside S results in an electric vector flow
a closed surface S, results in a diver- through S . This flow may be consid-
gence of field E from every infinites- ered as caused by the superposition of
imal volume element (a}; closed mag- elementary contributions of divE in-
netic lines result in zero divergence (b) . side the volume V .

within a finite space , every point in this space is an elementary point


source of an electric field (eq.(2.12)). In its integral form (equation (2.1))
GAUSS 's law, which describes the total flow of the electric vector through
the surface surrounding this volume , will superimpose the divergence of
th e elements of the volume (Fig.2.8). In the case of a magnetic field,
where every line of magnetic induction is closed, the elementary flow of
th e magnetic induction through every element of space will be equal to
zero (eq.2.2 and 2.13).
The infinitesimal curl element of an electric or magnetic induction
field line of an alternating electromagnetic field is expressed by the vec-
tor operator rot applied to the appropriate component of the electro-
magnetic field. The resulting vector is always directed normal to the
plane of the curl (Fig.2.9,a). Summing up all the elementary contribu-
tions of the curls one may calculate the field over a closed contour which
surrounds a finite space containing alternating components of the elec-
tromagnetic field (Fig.2.9,b) [42]. Such a process is realized by equations
(2.3-2.4). This shows that there is a complete equivalence between the
integral equations (2.1 - 2.4) and MAXWELL'S equations in differential
form (2.12 - 2.15).
We shall next derive the equations of the wave motion of an elec-
tromagnetic field in free space , which contains no charges and currents:
p = 0, j = O. Under these conditions, MAXWELL'S equations take the
PROPERTIES of LIGHT WAVES 25

•I aE/at it aE/at
I I
I
I "
I
• rot B

~dS I
I I I I
X I I I Sil I u IL
I I I I I
," I
Y I I
I
I
I
I
I
,I
I I

a b

Figure 2.9. The infinitesimal curl element of the magnetic induction B represented
by the vector rotB is caused by an alternating electric field 8E/Ot. The vector rotB
is normal to the element dB (a) . The superposition of all vectors rotB of elementary
curls results in the alternating flow of the electric field throught a finite contour L
(b).

following form:
divE = 0 (2.16)
divB = 0 (2.17)
8E
2rotB
c = at (2.18)

rotE = -8B
-
8t
(2.19)

Applying the vector operator rot to equation (2.19) we get


8B 8
rot [rotE] = -(rot[jt) = - 8t (rotB)
Substituting rotB from (2.18) we obtain
182E
rot[rotE] = - 2- - (2.20)
2 c 8t
In order to simplify the left-hand side of the last equation we use a
formal presentation of the differential operators div and rot in terms of
the standard vector operator "nabla" \7, which in Cartesian coordinates
has the following form:
n 8. 8. 8
v =-l+-J+-k
8x 8y 8z
The differential operators div and rot of a vector function E using \7 are
expressed as a scalar and a vector product, respectively:
divE = (\7 . E) rotE = [\7 x \7E]
26 DEMONSTRATIONAL OPTICS

and the left-hand side of (2.20) takes the form

rot [rotE] = [\7x [\7 x E]]


The following formula is valid for a triple vector product:

[\7x [\7 x E]] = \7(\7. E) - (\7 . \7)E . (2.21)

The first term on the right-hand side of previous expression is equal to


zero according to equation (2.16), since

(\7 . E) = divE = 0 .
The second term of the right-hand side of (2.21), expressed in terms of
the (\7 . \7) operator, takes the form

2 8 2E 8 2E 8 2E
-(\7. \7)E = -\7 E =- 8x 2 - 8y2 - 8z 2

\72 is usually called the LAPLACE operator 6. :


2 8 2E 8 2E 8 2E
(\7 . \7)E = \7 E = 6. E = 8x 2 + 8y2 + 8z 2
Substituting this result into the left-hand side of (2.20) we obtain the
following equation:
2E
6.E =! 8 (2.22)
c2 8t 2
which is called the wave equation. Analogically, the vector B satisfies
the same wave equation:

(2.23)

Solutions of the equations (2.22 - 2.23) describe electromagnetic waves


propagating in free space.

2.2 Plane waves


Let us consider a specific case of electromagnetic wave, where the
electromagnetic field is dependent only on one spatial coordinate, for
example the coordinate z , and the time t. Such a wave is called a plane
wave. In this case, wave equations (2.22) and (2.23) will have the same
form, presented as

(2.24)
PROPERTIES of LIGHT WAVES 27

where / is assumed to be an arbitrary component of the vectors E, B.


To solve equation (2.24) one can re-write it in the form

( ~&t - c~)
8z
(~+
&t
c~)
8z
/= 0 '
and introduce new variables ~, TJ:
~=t-z/c TJ=t+z/c
so that

Then

and

hence the equation for / takes the following form:


82 /
8~8TJ = 0 .

The solution / of this equation is the sum of two arbitrary functions:


h(~) and h(TJ):

/ = h(~) + h(TJ) = h (t - ~) + h (t + ~)
For example, let us discuss the case of h = 0, so that / = h(t - z/c).
In this case the field component f varies with time in a plane z = const ;
and for a fixed time t the field component varies depending with z.
The field component / = h(t - z/c) will have the same magnitude if
coordinate z and time t satisfy the condition t - z/c = const, i.e,
z = const + ct .
This equation implies, that if the field has a certain magnitude at time
t = 0 and point z , the field will have the same magnitude at time t over
the whole (x , y)-plane at a distance ct from point z. This plane, always
normal to the direction of propagation, is called the phase plane (see
Fig.2.1O). So, we can say that all components of the electromagnetic
field of the plane wave will have the same magnitude over the entire
phase plane . This plane propagates along the z-axis with velocity of
light c.
In other words, h(t-z/c) is a plane wave, traveling in the positive z-
direction. It is obvious, that h(t+z/c) is another plane wave, traveling
in the negative z-direction with the same velocity c.
28 DEMONSTRATIONAL OPTICS

c
x

az=ctd
Phase plane

Figure 2.10. T wo positions of t he phase plane separated by a time interval tlt for a
plan e wave propagating in th e z- directio n,

2.3 Plane monochromatic wave


2.3.1 Frequency, wavelength, velocity of light
There is a specific and import ant class of plane waves, in which the
components of the electromagnetic field f are represented by sine or
cosine functions. In this case one considers the propagation of a plan e
monochromatic wave. For example, one may consider the plane mono-
chromat ic wave, traveling in the positive direction of th e z-axis. In this
case, at every fixed z , the electromagnet ic field will carry out harmonic
oscillations within the plane z = cans t . Let T be t he period of t he os-
cillat ions, then the electric vector E and the magnetic induction vector
B may be expressed by the following cosine functions:

E=EO COS[~ (t-~)+ ~]


B = Bo cos [~ (t - ~) + ~] (2.25)

where the phase of the cosine function contains a variable term (27f IT)( t-
zl c) and a constant term - the initial phase sp, Two constant vectors
Eo and B o determine the amplitude of the oscillations.
It is customary to use t he frequency v = liT and th e circular fre-
quency w = 27fv = 27f IT. The const ant factor Tc becomes a length
A = T c = cf» that is called the wavelength of the wave. Apart from the
wavelength, the constant k = 27f I A, or k = wi c may also be used. This
magnitude is called the propagation number. Taking into account th e
definitions of frequency , wavelength and propagation number, the phas e
PROPERTIES of LIGHT WAVES 29

Rotating Light
toothed wheel source
I ~

I:_~====;.:
Mirror
-- 7 km - .

y •

Telescope

Figure 2.11. Schematic setup of FIZEAU 'S experiment to measure the velocity of
light .

of the plane monochromatic wave may be represented by the following


expression:

(27r /T)(t - Z/c) + <p = 27r (lit - ~) + <p = wt - kz + sp (2.26)

The velocity of light c in a vacuum is a fundamental constant; its value


is the same as the electrodynamic constant: c = 299,792.458 km/s,
Let us discuss Ftznau'sexperiment (1849) to measure the velocity
of light, schematically presented in Fig.2.11. Light rays from a bright
light source pass through a rotating toothed wheel to a very distant
mirror and are reflected by the mirror back to the light source . Incoming
rays passing between the teeth of the wheel may be observed by means
of a telescope. The idea of the experiments was to observe periodic
disappearances of the light beam by changing the speed of rotation of the
wheel. These dissappearances are due to the finite time the light would
need to travel to the mirror and back to the observer. The distance
L between the mirror and the rotating toothed wheel was 7 km and
the wheel had n = 720 teeth. The light dissappeared sequently at the
angular velocities 283 and 313 turns per second. The time T needed for
light traveling from the source to the mirror and back to the telescope
is equal to T = 2L/c. Let the wheel turn k teeth during the period T.
At nl = 283 turns per second, corresponding to the first disappearance,
the value of k may be calculated as k = Tnln . For the same time T at
n2 = 313 turns per second, the wheel will turn to k + 1 teeth, therefore
k+ 1 = Tn2n. Now one gets: Tn2n- Tnl n = T( n2 -nr)n = 1. Finally, for
the speed of light we get the estimate: c = 2L( n2 - nl)n = 3.0672 x 105
km/s.
We should note that in a vacuum the phase plane propagates with the
velocity of light. In the general case, the phase velocity is not an univer-
sal constant and can vary when the electromagnetic wave is propagating
30 DEMONSTRATIONAL OPTICS

Table 2.1. The visible region of electromagnetic


waves associated with colors, wavelenths >.
and frequencies IJ .

x IJ
Red 760 - 630 nm (4 - 4.8)·W 4 Hz
Orange 630 - 600 nm (4.0 - 5) .10 14 Hz
Yellow 600 - 570 nm (5 - 5.3).10 14 Hz
Green 570 - 500 nm (5.3 - 6) .10 14 Hz
Light blue 500 - 450 nm (6 - 6.7).10 14 Hz
Blue 450 - 430 nm (6.7 - 7) .10 14 Hz
Violet 430 - 400 nm (7 - 7.5).10 14 Hz

in dense media ; however, in most cases the frequency v is invariable.


This means that the number of complete oscillations of the field per
unit time is the same in vacuum and in a medium. Due to this we can
regard the frequency as the basic quantity of an electromagnetic wave,
represented by the relations (2.26). Because c is a fundamental constant,
for a wave propagating in vacuum the knowledge of v is just enough to
determine such quantities as the period, the propagation number k, and
the wavelength. In practice the wavelength or the frequency is used for
denoting the properties of electromagnetic waves.
For the spectrum of electromagnetic waves associated with visible
light and colors, the wavelengths, measured in nm or 10-9 m, and the
frequencies are given in Table 2.1. The values of frequencies are given
in Hz as the number of oscillations per second.

2.3 .2 Transversality of electromagnetic waves


Let us deduce from MAXWELL 's equations an important property of
electromagnetic waves: electric and magnetic field vectors are located in
plane normal to the direction of propagation of the wave, as we have seen
in a particular case of the plane waves. This fundamental property is
often called the transversality of electromagnetic waves. The transver-
sality of electromagnetic waves results from MAXWELL'S equations in
free space and therefore is also valid in general. Nevertheless, with-
out loss of generality, we describe electromagnetic fields as plane waves.
MAXWELL'S equations in differential form hold for the infinitesimal el-
ements of space , where the wave front of an arbitrary electromagnetic
wave may be considered to be plane.
As we have seen, at every point z there is a plane x, y normal to the
z-axis, where the oscillations of the electromagnetic field occur . We may
PROPERTIES of LIGHT WAVES 31

establish the properties of the oscillations by means of two vectors Eo,


Bo of a plane wave in form (2.25), using the equation (2.18), (2.19):

2 8E 8B
c rotB = tit ' rotE = -tit (2.27)
From the first equation one can write the z-coordinate of vector rotB
as follows:
8By 8Bx
(rotB )z = 8x - By = 0 ,

because all components of the electromagnetic field in the plane wave


depend only on the z-coordinate, hence

The solution of this equation is E; = const. Due to the propagation of


electromagnetic waves, the constant must be equal to zero, otherwise it
would lead to a static electric field. We conclude that possible variations
of the electric field occur only in the (x, y)-plane. The same conclusion
is valid for the magnetic induction of the plane electromagnetic wave.
This means that, while propagating in free space , both vectors of the
electromagnetic wave occur transversal with respect to the direction of
wave propagation.
Now we may determine the mutual position of vectors E and B in the
(x, y)-plane. We use the plane monochromatic wave given by equation
(2.25) to calculate the magnitudes in equations (2.27). Firstly, for 8E/Ot
we write:

~~ = -Eowsin(wt - kz + tp) (2.28)

Now, we calculate the components of vector rotB in the (x, y)-plane :

(rotB)x = 8Bz _ 8B y = _ 8By


8x 8z 8z
Substitution of By with (Bo)y cos(wt - kz + tp) gives

(rotB)x = -k(Bo)y sin(wt - kz + tp)

where the x-component of the vector rotB is expressed by the y-component


of vector Bo. Analogously, we determine the expression

(rotB)y = - 8~x = k(Bo)x sin(wt - kz + tp)


32 DEMONSTRATIONAL OPTICS

Figure 2.12 For a given


k direction of vector k and
radius vector R directed to
a point of the phase plane ,
the distance from the ori-
gin is equal to k . R.

We introduce a unit vector n normal to the (x,y)-plane, which points


in the direction 'of wave propagation. Then the components of rotB and
Bo may be expressed in the vector form:

rotB = -B o x nksin(wt - kz + cp) (2.29)


Now in equation (2.29), we substitute the right-hand side of the first
equation in (2.27) for r otB and the right-hand side of (2.28) for aE/at.
Omitting the factor sin(wt - kz + cp) we can write the following expres-
sions:
w
cBo x (kn) = -Eo, or cBo = n x Eo (2.30)
c
where k = w/ c. A vector k = kn , called the wave vector, is often
used to indicate the propagation of the plane wave (instead of vector
n). Now we have established that the three vectors Eo, Bo and k of
an electromagnetic wave propagating in free space have to be mutually
orthogonal. The plane of equal phase , containing vectors Eo and Bn,
propagates with light velocity; this plane is often called the wave front , or
phase surface which is a plane only in the case of a plane monochromatic
wave.
Due to the properties of the field vectors, the wave front of the plane
wave is normal to the direction of propagation, also to vector k. Let
R be a vector from origin to an arbitrary point of the wave front at
moment t (Fig.2.12) The projection of vector R on the direction of wave
propagation is equal to k . R; in turn this distance is ct. Hence, the
phase of the plane monochromatic wave may be presented in the form:

wt - k R+cp .

With such a representation the phase no longer depends on the choice of


reference, so the expressions for the field components of a plan e mono-
chromatic wave take their most common form:

E = Eo cos(wt - k .R + cp) , B = Bo cos(wt - k .R + cp) (2.31)


PROPERTIES of LIGHT WAVES 33

Figure 2.1S The spherical


wave emitted by a point
source S located at the ori-
gin; the three vectors k , E
and B on the wave front
are mutually orthogonal.

2.4 Spherical waves


Another solution to the wave equation is one where the surface of
equal phase has a spherical shape. This type of wave is called a spherical
wave. A natural source of spherical waves is a point source located at
the center of the sphere. All points of the spherical wave front spreading
from the source with light velocity are located on the sphere of radius
R= ct .
Hence, the phase of th e spheri cal monochromatic wave of wavelength
.x and frequency1/ may be represented by the following expression:

21rl/t - 21rR/.x + cp = wt - kR + cp . (2.32)


As before, cp is the initial phase. In this case, all components of the
electromagnetic field vary with the same phase dependence given by
equation (2.32). For example, the electric component of a spherically
expanding electromagnetic wave may be writ ten as
a
E(R, t) = E a R cos(21rl/t - kR + cp) (2.33)
where E a is the value of the field amplitude at distance a (Fig.2.13).
The spatial dependence of E is given by 1/R, which is characteristic for
a spherical wave. As with a plane wave, the magnetic induction vector
B and vector k satisfy expression (2.30) for all times and at every point.
(Fig.2.13)
In optics there are several ways of generating almost plan e waves and
almost spherical waves. A light beam with a plane wave front may be
formed by means of a small bright light source .located in the focus of
a lens or an objective (Fig.2.14). The light beam on the other side of
the lens may be recognized as a plane wave. The small source of light
is usually achieved with a pinhole or a small round diaphragm, which is
illuminated by means of convergent rays formed by a condenser lens from
an ordinary light source , e.g. a high pressure mercury lamp . In order to
test the quality of the light beam a white screen is placed perpendicular
34 DEMONSTRATIONAL OPTICS

Objective

~
Condenser lens

Hg lamp
:= ~-~ --~
Pinhole
__- _

-
- -~
- ------ ... :
V

Figure J2.14. Plane wave production using a bright light source . A pinhole illumi-
nated by the light of a H9 lamp through a condenser lens acts as a point source .
The lens is used to form a light beam with a plane wave front . A beam with nearly
constant diameter is produced .

to the beam and displaced along the axis of the beam. If the image or the
white spot size on the screen does not change while moving the screen ,
we say that the light beam is parallel and represents a plane wave. But
upon closer inspection of the spot on the screen produced by the mercury
lamp light , one would notice that the white central spot is surrounded
by colored rings. Such an effect implies that different monochromatic
components are present in the beam , which produce colored neighboring
rings of t he image. So one can assume that the white light from t he
mercury lamp consists of a set of monochromatic components.

~\
.> \
~~ \
~ \
Condenser lens

Hg lamp
Pinhole

Figure 2.15. A light wave with a spherical wave front from a bright source. Here a
pinhole plays the role of the source .

The source of exactly spherical waves is a point source. But a point in


its mathematical definition cannot be used as a source, so we say "point
source" when the distance between the source and the point of observa-
tion of the electromagnetic oscillations is much larger than the dimension
of the source. The wavelength of the electromagnetic radi ati on should
also be taken into consideration for an approximation of the distance.
In practice, a source of finite dimensions can be considered as a point
source if R » >. and R » d, where R is the distance between a source
with size d and the point of observation. For example, in many cases
PROPERTIES of LIGHT WAVES 35
x
E

Figure 2.16. Spatial distribution of the electromagnetic field for a plane monochro-
matic wave propagating in the positive direct ion of the z-axis at a fixed moment t .
The wave is linearly polarized in the direction of the x-axis.

the focus of a lens can be regarded as a point source of an expanding


spherical wave (Fig.2.15).
A more appropriate source of a parallel light beam is the light gen-
erated by a gas laser, which radiates an almost parallel bright beam
without using any additional optical instruments. As we shall verify
with examples, such a laser light beam has some advantages over ordi-
nary light sources. The purity of color of a (single mode) laser beam is
the best possible approximation to an ideal monochromatic wave. The
laser light can be described with only one optical carrier frequency to a
higher degree than any other type of optical radiation.

2.5 Polarization states


As shown, the vectors of the electric and magnetic field are directed
perpendicular to the propagation direction, and therefore oscillations of
the electromagnetic field are always transversal. This transversal nature
is a fundamental property of light waves. Long before its theoretical
confirmation, the concept of polarization of light waves had been suc-
cessfully exploited. We are now going to analyze the oscillations of the
electromagnetic field in a phase plane for a plane monochromatic wave.
Let the (x, y)-plane be placed normal to the direction of propagation
z of a plane monochromatic wave (Fig.2.16). For a given vector k ,
normal to (x,y), both vectors E and B will be parallel to the plane at
any time; their projections will be mutually orthogonal. Hence, a single
vector , usually the vector E , is enough to characterize the polarization
state.
There exist only two basic states of polarization: one is linear polar-
ization and the other is circular polarization, which will be considered
in more detail in Chapter 3. Linear polarization is achieved if the pro-
jection of vector E on the (x,y)-plane oscillates on a straight line for
all time (Fig.2.17,a) . As the wave propagates the vector E always stays
within a plane , usually called the plane of polarization, determined by
36 DEMONSTRATIONAL OPTICS

y y
E

x x Figure 2.17 P rojections of


the elect ric vector E onto
the (x , y)-plane in t he case
of a linea rly polar ized wave
(a), and a circularly polar-
a b ized wave (b) .

Figu re 2.18. Oscillations of the electric field of a wave composed by two monochro-
matic waves of equal amplitude and wit h a small difference between their frequenci es.
The slowly varying function A(t ), shown by the dashed line, is the envelope of th e
resul ting oscillat ion .

the projection of vector E and the z axis. For this reason t he linear
polarization is sometimes called plane polarization.
In the case of circular polarization of a plane monochromat ic wave, t he
elect ric field vector can be characterized by two orthogonal projections
on t he (x, y)-plane, both having the same ampli tude of oscillation , but a
phase difference of tt /2. Th e to p of th e vector E rotates with a const ant
velocity on a circle in the project ion plane. Its trace in space is on a
screwed line. An example is shown in Fig.2.17,b.

3. Quasi-monochromatic waves
In reality, light emitted by a particular light source may be represented
as a mixture of monochromatic waves having different frequen cies, but
it can hardly be treated as a monochromatic wave. Nevertheless, such
light can be regarded as consisting of different spectral components,
each distribut ed within a narrow rang e of wavelengths. Such a type of
light radiation is often called quasi-monochromatic. In practice, we can
generate quasi-monochromatic light by means of different types of optical
filters and other devices which provide a frequenc y selecti on within a
narrow band of wavelengths.
PROPERTIES of LIGHT WAVES 37

a b

Figure 2.19. The quasi-monochromatic oscillation composed by harmonics of a


Gaussian spectrum . The carrier frequency VQ has the largest amplitude in the spec-
trum.

3.1 Envelope and carrier frequency


In a simple case, the superposition of two monochromatic oscillations
results in oscillations of a quasi-monochromatic type:

E(t) = Eo cOS(271'I/lt) + Eo cOS(271'1/2t) = (2.34)

2+1/1]
2Eo cos [ 271' 1/2 -2 1/1t ] cos [1/
271' 2 t

When the difference between 1/2 and 1/1 is small, the resulting oscillations
look like a monochromatic wave with a slowly varying amplitude:

E(t) = 2Eo cos [271'~l/tl ' cos [271'I/ot] = A(t) cos [271'I/ot]

where A(t) = 2Eo cos [271'~l/t] , and I/o = (1/2 + 1/1)/2 is the mean fre-
quency or carrierfrequency. The frequency difference between each orig-
inal wave and the carrier is given by ~I/ = 1(1/2 -l/d/21 . The amplitude
A(t) , a slow function of time, is the envelope of the high frequency func-
tion E(t). The resulting oscillations of the electric field E(t) are shown
in Fig.2.18.
For the common case of a quasi-monochromatic oscillation A(t) is
a time dependent amplitude and «p(t) is a phase, varying slowly in
time with respect to the period T of the fast oscillations of the field
at carrier frequency I/o (T = 1/1/0). The oscillations of such a quasi-
monochromatic wave at any point may be represented by the expression

E(t) = A(t) cos(271'1/0t + «p(t)) (2.35)


where I/o specifies the carrier frequency.
38 DEMONSTRATIONAL OPTICS

3.2 Spectrum of quasi-monochromatic


oscillations
The quasi-monochromatic oscillation expressed by eq. (2.34) is spec-
ified by two spectral harmonics of the same amplitude Eo. The low
frequency harmonic has the frequency Vo - D.v , and the high frequency
harmonic Vo + D.v, where D.v = (V2 - vd/2. In a more general case, the
spectral harmonics of quasi-monochromatic oscillations, represented in
terms of the time-dependent amplitude and phase in (2.35), vary con-
tinuously within a certain range of frequencies. Such a continuous dis-
tribution of spectral harmonics forms a continuous spectrum of quasi-
monochromatic oscillations (2.35). For example , the oscillations with the
Gaussian envelope, shown in Fig.2.19,a, are described by the following
expression:
E(t) = exp[-at 2] cos(27Tvot) .
The spectrum associated with E(t) also has a Gaussian shape and is
given by

rv exp [- 4~2 (27TV - 27TVO)2]

This function has its maximum at the carrier frequency Vo (Fig .2.19,b).
In principle the carrier frequency Vo and the functions describing the
amplitude A(t) and the phase ~(t) can have an arbitrary form. But in
all cases, a mathematical procedure exists , which restricts the arbitrari-
ness by expanding the dependence E(t) in terms of the monochromatic
oscillations. E(t) can be represented by a FOURIER integral of a complex
exponential function of the frequency, calculated from -00 to 00:

J
00

E(t) = ~ £(v) exp( -i27TVt) du , (2.36)


-00

where £(v) is a complex function called the spectrum of E(t) [43] . The
spectrum £(v) results from the FOURIER spectrum of the function E(t):

J
00

£(v) = ~~ E(t) exp(i27Tvt) dt (2.37)


-00

where £(v) exp( -i27Tvt) is the contribution to E(t) by the spectral com-
ponent at frequency t/ ,
Mathematically, the left-hand part of eq. (2.36) is the FOURIER int e-
gral of the function £(v). But if v is a frequency, what about negative
PROPERTIES of LIGHT WA VES 39

IE(v)1 IE(v)1

~_1_11
-v -v
2 I
o o

a b

Figure 2.20. The total spectrum of oscillations caused by the superposition of two
monochromatic oscillations of frequencies Ill, 112 (112 - III « (Ill + 112)/2) (a) . The
spectrum of positive frequencies associated with cosine expansion of the original func-
tion of time (b) . Spectral components of the spectrum used in the cosine expansion
are twice as high as that of the full spectrum .

frequencies? Negative frequencies may be seen as a mathematical rep-


resentation to avoid dealing with two functions, A(t) and <]}(t), instead
of one, which is called the complex amplitude:

E(t) = A(t) exp(i<]}(t)) (2.38)

However, the electric and magnetic fields of the electromagnetic wave


are real functions in time and space. We shall denote the electric field by
E(r) and call E(r) the physical field. The fact that we deal with real func-
tions of electromagnetic fields allows a simplification of the description
of spectral functions associated with quasi-monochromatic oscillations.
We may verify our assumptions using the example of two monochro-
matic components . Let us consider the sum of two such terms, defined
at the same absolute value of frequency Ivl:

E(r)(t) = ~£(_)(v)exP(-i21l'vt) + ~£(+)(v)exP(-i21l'vt) (2.39)

Here the first value £(_)(v) is defined for v ~ 0 and the second £(+)(v)
is defined for v 2: O. Now we shall reduce the sum (2.39) to a function
of the positive frequency v only.
If E(r)(t) is a real function, the sum in (2.39) will be real under the
condition
(2.40)
where both values £(+) (v) and £(-) ( -v) are defined for positive t/, Now,
defining £(+)(v) = 1£(v)1 exp(icp) and using (2.40) for £(_)(v) in (2.39),
we obtain £(_)(v) = 1£(v)1 exp(-icp).
40 DEMONSTRATIONAL OPTICS

One may note that complex exponential factors in (2.39) correspond-


ing to £(_)(v) and £(+)(v) for positive frequencies v 2 0 will take the
forms exp(i27fvt) and exp( -i27fvt) respectively. Hence, the sum on the
right-hand side of (2.39) may be written as:

~£(_)(v) exp(i27fvt) + ~£(+)(v) exp(-i27fvt) =

= ~1£(v)1 ex~[i(27fvt - cp(v))] + ~1£(v)1 exp[-i(27fvt - cp(v))] =

= 1£(v)1 cos(27fvt - cp(v)) . (2.41)


where cp(v) is the phase of the complex amplitude £(v) represented as a
function of the frequency u.
The condition (2.40), therefore, allows us to avoid operations with
negative frequencies when dealing with physical fields [37]. In a similar
way let us represent the integral in the right-hand part of (2.36) as a
sum of integrals:

J
00

~ £(v) exp( -i27fvt) di/ =


-00

J J
o 00

~ £(_) (v) exp( -i27fvt) du + ~ £(+) (v) exp( -i27fvt) du ,


-00 0

where £(_)(v) is defined for v:::; 0, and £(+)(v) for v 2 O. Then, for the
given lvi, we always find two values £(_)(v) and £(+)(v), which satisfy
condition (2.40) that the original function E(r)(t) is real. Summation
of each pair of similar values results in expression (2.41), so that a su-
perposition of monochromatic waves given by the integral (2.36) may be
reduced to a summation of the positive frequencies only. Therefore, for
a physical field E(r) we may write:

J
00

E(r)(t) = A(t) cos(27fvot + <p(t)) = 1£(v)1 cos(27fvt + cp(v)) dv


o
(2.42)
Hence, a real function E(r)(t), describing quasi-monochromatic oscilla-
tions, may be written in terms of its spectrum, which contains only
positive frequencies in a cosine expansion. The total spectrum of os-
cillations described by (2.34) is presented in Fig.2.20,a. There are two
regions of the spectrum, one lies in the positive region of frequency and
the other in the negative . Both regions contain two components of the
PROPERTIES of LIGHT WAVES 41
same magnitude. A spectrum associated with the same oscillation, which
is obtained by applying a cosine expansion using positive frequencies , is
presented in Fig.2.20,b. Here, one also gets two components like the pos-
itive part of the total spectrum, but in this case the magnitude of the
spectral components is exactly twice as high as in the total spectrum.
Apart from this real function the complex function E(t) can be of
interest. It is not difficult to prove that this function can be represented
by an integral over all positive frequencies:

J
00

E(t) = £(11) exp( -i21rvt) du (2.43)


o
This function is called the analytical signal which is associated with
the real function E(r)(t) [37] . Taking into account the definition of the
complex amplitude (2.38) the complex signal may be expressed as

E(t) = A(t) exp(i<l>(t)) exp(i21rllot) (2.44)

It is obvious that
E(r)(t) = Re{E(t)} ,
and
E(t)E*(t) = A2(t) (2.45)
The complex signal provides simpler mathematical transformations than
the physical field.

4. Energy and momentum of light


4.1 Intensity of light
Electromagnetic waves transfer energy, a fact which everybody has
experienced in daily life. Life on the Earth itself is the result of light
energy flow from the Sun to the Earth. The simple heating of bodies by
sunlight, associated with a transfer of radiation energy, demonstrates the
fundamental idea that light radiation transports energy. The exploration
of basic features of light are based on measurements of light energy.
According to MAXWELL's electromagnetic theory we associate an en-
ergy density with an electromagnetic field. This energy density is ex-
pressed in terms of the electric and the magnetic fields:

coE 2 + coc2 B 2
w=----- (2.46)
2
where w is the quantity of the electromagnetic energy contained within
a unit of volume.
42 DEMONSTRATIONAL OPTICS

~f
Figure 2.21 Assuming
a plane monochromatic

s ~. z
wave, the energy density
inside the volume La is
constant.

Let w be the energy density of the plane monochromatic wave rep-


resented by equation (2.31) propagating along the z-axis as shown in
Fig.2.21. We choose a cylindrical volume of cross section a normal to
the propagation direction and with the length l. The total electromag-
netic energy contained within this volume is w .i, a. The time T, needed
for the light wave which passes through the cylinder with velocity c, to
transport energy on the path l is T = l/ c. We define the flow of elec-
tromagnetic energy as the quantity of the energy passing through the
perpendicular cross section a per unit time :
wla eoE 2 + eoc2 B 2
- =cwa=c a
T 2
The quantity
2 2 2
s = cw = /oE + eoc B (2.47)
2
is called the energy density flux, and it defines the value of the electro-
magnetic energy passing through a unit area normal to the propagation
direction per unit time. The energy density flux is completely described
by vector S, which is called the POYNTING vector and is defined as
follows :
S=ExB (2.48)
Among all the possible ways of transforming light energy into other
energy forms (generally heat) , we consider the mechanism of freeing an
electron from the surface of a material under the influence of a light
wave (photo effect). Normally the electrons of an electrically neutral
substance are kept inside the surface by electrostatic forces within the
neutral atoms, in the case of dielectrics, or within the limits of the crys-
talline lattice for metals. Some additional external energy is needed for
an electron to leave the surface of a substance. A free electron caused
by absorption of the energy from light flow is called a photoelectron.
We consider the process of photoelectron creation to occur within
some interval !:i.t, which is longer than a few periods of oscillations of the
electromagnetic field (Fig.2.22). This is a principle point, since in this
case the energy of the light flow needed for creation of a photoelectron
can be estimated by means of simple calculations.
PROPERTIES of LIGHT WAVES 43

bodyof photodetector

ITI 0 photoelectron 0
00 --0 0
1\fV\fv~ 0 0
0 0
0
coupled electrons
I. ~t .1
Figure 2.22. The release of a photoelectron takes a finite time interval l:1t, which is
much longer than the period T of the oscillations of the incident light wave. Each
coupled electron in the matter of the photodetector could receive the energy portion
hu, where h is Plank's constant and II is the light frequency. Photoelectrons released
from the photodetector's body by incident light give rise to the photocurrent of the
detector.

We write S from (2.47) as


2 2B2
S = C coE + coc = ccoE2
2
since c2 B2 = E2.
For the electric field of the plane monochromatic wave in the form
(2.31) the choice of the initial phase is arbitrary. We define the coordi-
nate of the electron from the condition: kR - sp = O. The oscillations of
the E vector at the electron position are given by the harmonic depen-
dency
E(t) = Eo cos(271"vt)
We substitute Eo cos(271"vt) for E in the expression for the energy flow
S:
S = ccoE5 cos2(271"vt)
The choice of the initial time is also arbitrary, so it can be taken as
to = O. Hence, the net light energy passing through the physically small
space of the electron's location during the interval tlt is proportional to
an integral

J
Dot
S dt = ccoEo
2 J
Dot
2
cos (271"vt) dt
E5 (
= ccoTtlt Sin(471"Vtlt))
1 + 471"vtlt
o 0

According to our assumption tlt > T = 1/v, or tltv > 1, the value of
the phase in the second term inside the parenthesis is much larger than
271". Hence,
Isin(471"vtlt)I j(471"vtlt) « 1
44 DEMONSTRATIONAL OPTICS

--
A(l)
pholodeleclor
° 0 °0 'J 1(1)-41)

""
o
o 0 o 0
00
o 0
0 00
o
0
~~.

-
0 0 0 00 I
E(l) 0° 0° o 0
o
o 0
pholocurrenl

Figure 2.23. Photodetection process caused by an incident quasi-monochromatic


wave. Oscillations in the electric field of the wave occuring at the carrier frequency
are modulated slowly by the envelope A(t) . Light energy is taken up by the coupled
electrons of the photo detector and ·t hen released as the energy of photoelectrons,
creating a photocurrent. The time dependency of the photocurrent i(t) is proportional
to A 2 (t ).

and can be omitted. Finally, for the energy we obtain c£0D.tEZ/2 =


D.tS, where
- 2
S = c£oEo/2 (2.49)
is the mean value of energy flow density, produced by smoothing the fast
oscillations of frequency of v during the interval D.t. We shall call the
value S the intensity of light. Usually this value is denoted by I. The
intensity of a monochromatic plane light wave is constant and directly
proportional to the square of the amplitude Eo. The example considered
here shows a typical situation dealing with a measurement procedure.
Now the intensity (2.49) can be represented in terms of the complex
field of the electromagnetic wave in the following way:

1= c£oEE*/2 (2.50)

Since we shall often use the intensity as a measurable quantity in different


experimental situations, we shall omit the constant factor c£0/2, using
the simpler expression I '" EE* .
The representation of the intensity in terms of the complex field func-
tions allows the calculation of the average value of the intensity during
the smoothing interval D.t with ease even for more complicated cases,
in which the amplitude is a function of time and coordinates too. For
example, following from (2.50) we can write

I(t) = E(t)E*(t) = A2(t) (2.51)

Therefore, the intensity is the square of the envelope of the rapid oscil-
lations at the optical carrier frequency vo.
Let us discuss a photodetection process caused by a light wave, where
oscillations of the electric vector occur according to expression (2.34)
PROPERTIES of LIGHT WAVES 45

(Fig.2.23). The light wave penetrating into the material of the pho-
todetector gives rise to an absorption of light energy by the electrons of
photosensitive atoms. A number of these electrons is no longer coupled
to the atoms and give rise to a photocurrent.
As we have discussed, the creation of a single photoelectron takes the
finite time interval !:i.t. The same duration may be taken as a time inter-
val which provides a smoothing of the photocurrent, consisting of a huge
amount of photoelectrons. It means that variations of the photocurrent
in time i(t) follow the time dependency of the light wave intensity. In
turn, according to the expressions (2.34) and (2.51) we may write
i(t) A 2(t ) coS2[7r(Vl - V2)t]
roJ roJ

It is clear, this time dependence oscillates slower than the electric field
of the incident wave, which oscillates at the carrier frequency Vo = (VI +
v2)/2. No electronic device is able to resolve the carrier frequency itself.

Example 1.
It is well known that the mean flux energy density of the Sun radiation
near the Earth's surface is equal to (8) = 1.4 kW/m 2 . Let us estimate
the amplitude of the electric field strength of a light wave arriving the
Earth surface, assuming a quasi-monochromatic radiation.
The mean flux energy density (8) is given by (8) = !ceoE6. Substi-
tution of numerical quantities leads to Eo ~ 1 kV/m.
Example 2.
Let two waves of identical frequency propagate along the same direc-
tion. If these waves have mutually orthogonal linear polarization states
the flux energy density will be equal to the sum of the flux energy den-
sities 8 1 + 82 for every wave.
Actually, the flux energy density may be represented in terms of
the scalar product of electric field strengths 8 = !ceo(E. E). The
electric field vector E of the resulting wave is the sum of two mutu-
ally orthogonal vectors Ex and E y : E = Ex + E y , therefore one gets:
(E· E) = (Ex + Ey)·(E x + Ey) = (Ex'Ex) + (Ey·E y) for the scalar prod-
uct (E · E). Hence, the flux energy density for the resulting field should
be equal to the sum of the individual flux energy densities.
Example 3.
Let us find the spectral distribution for a wave train given for a short
time interval as a cosine function Eo cos(27rvot) . This wave train exists
for a limited time interval 7, so that t satisfies the following inequalities:
-7/2 < t < 7/2. At times t < -7/2 and t > 7/2 the field strength is
zero (Eo = 0).
46 DEMONSTRATIONAL OPTICS

In order to find the spectral distribution of the field corresponding to


Eocos(271"vot) it is useful to represent the cosine function in terms of two
exponential factors : Eocos(271"vot) = Eo[exp(i271"vot) + exp( -i271"vot)1I2.
Now, calculation of a FOURIER spectrum of the last expression gives rise
to

-f
E jr/2
-r/2
[exp(i271"vot ) + exp( - i271"vot)] exp(i271"vt)dt =

Eo [sin[7I"(v - vo)r] + sin[7I"(v + vo)r]]


2 71"(v - vo) 71"(v + vo)
For the important case of a quasi-monochromatic wave (for a relatively
long time r » l/vo), the value of the second term, containing the sum
of frequencies v + vo, will be much smaller than the first one, and can be
omitted. In good approximation, this enables us to describe the spectral
intensity distribution of such a wave train by the following expression:

For a given duration r the spectral distribution of intensity depends


more on the difference of frequencies v - Vo than on t/,

4.2 The Doppler effect


When a moving source radiates, a change in the frequency of the
emitted wave will be noted in a reference frame being at rest. This phe-
nomenon is common for many different types of radiation and is called
DOPPLER effect. In general, formulae from EINSTEIN'S special theory of
relativity are "necessary to express the change in frequency caused by the
motion of a radiating source . Nevertheless, in the particular case of an
optical source with a velocity much less than the velocity of light , the
approximation of classical theory is valid.
We will consider a simple example of the DOPPLER effect in its classi-
cal treatment. Let a source S move uniformly with the velocity v (v « c)
in a rest reference K (Fig.2.24). In another reference K', moving par-
allel to K with the velocity v, the source S is at rest . Let the source
emit N complete oscillations during some period t1t'. This number N
of complete oscillations will be the same in both references. In the mov-
ing reference K' the length of the emitted wave train is equal to AG,
whereas in the rest reference K the wave train is shorter and equal to
BG. The difference in the lengths AB is caused by the motion of the
source and it may be expressed via the ratio v I c: AB I AG = v I c. Thus
PROPERTIES of LIGHT WAVES 47

K'

I
I
I
I
I
I

1 i
A B c

Figure 2.24- A source S emits N complete oscillations in the moving reference frame
K' over the distance AG. The same number of complete oscillations is emitted in the
rest reference frame K . The length of this wave train in the rest reference frame is
BG .

for the lengths BC and AC we have

BC AC-AB v
== =1-- (2.52)
AC AC c
According to the main principle of the special theory of relativity the
wave train runs with the velocity c in both references, hence the dura-
tions of the wave train are different, and we express their ratio (2.52)
as
At _ BC -1 V
At' - AC - --;;
If the frequencies of the emitted light are denoted by Vo and VI in K'
and K references respectively, where Vo = N I At' and VI = N I At, then
a relation between Vo and VI will occur:

VI = vo/(l- vic)
With the assumptions v « c and v > 0, the previous expression results
in a positive difference between the frequencies Av = VI - Vo in the form:
v
Av = VI - Vo = Vo-c (2.53)

If a source moving towards an observer (at rest) emits a monochromatic


wave of frequency Vo , the observer will receive the radiation at the larger
frequency VI, where the change of the frequency is given by formula
(2.53). In a similar way it may be shown that if a source moves away from
the observer, the observer will receive a lower frequency than emitted
by the source.
48 DEMONSTRATIONAL OPTICS

He-Ne
laser

/ J:.
Photo Telescopic Retro-rellecting

Dr-
detector sytem prism

Diaphraqrn
Mirror
:--==t=
eam splitter -
v

Figure 2.25. A setup for observing the Doppler effeet by means of a moving retrore-
fleeting prism.

The DOPPLER effect may be directly demonstrated with the setup


schematically presented in Fig.2.25. The light beam from a laser is split
by a semi-transparent plate into two beams. One beam falls normally
at a mirror, then it is reflected back to the beam splitter and reaches a
diaphragm placed in front of a photodetector. The other beam, reflected
by the beam splitter, passes through a telescopic system and then falls
on a retroreflecting prism.
At small angles of incidence , a ray, falling on the retroreflecting prism
will be reflected by the inner surfaces of the prism in the direction exactly
opposite to that of incidence. This property of the prism permits all re-
flected rays, passing back through the telescopic system, to be collimated
within traces of the incident rays. For this reason small deviations of the
prism caused by its motion will have little or no effect on the position
of the beam reflected by the prism within the plane of the diaphragm.
Therefore the two beams from the mirror and from the moving prism
give rise to a stable superposition on the photodetector. If the prism is
at rest oscillations of the resulting electric vector on the photocathode
will be represented by
E = E l exp[i27l'vt] + E2exp[i27l'v(t + Lie)]
where E l , E 2 are the amplitudes of the electric vectors of the waves re-
flected by the mirror and prism, respectively, and L is the difference
between the geometrical pathes of these waves. (Here we use the com-
plex representation of the resulting field to simplify calculations of the
intensity) . The distance L spans 5 - 8 meters, whereas a lateral dis-
placement of the moving prism of about 1 em is allowed without dis-
turbing the desired effect. Because the intensity of both superimposed
beams may differ, one can provide approximate equality in the intensi-
PROPERTIES of LIGHT WAVES 49
ties by means of a polarizer inserted near the mirror and adjust EI = E2.
Uniform motion of the prism with velocity v toward the beam splitter
will cause the beam reflected by the prism to have the new frequency
VI = v(l + 2v/e) , where the factor 2 occurs due to the back reflection of
the incident beam by the retroreflecting prism. Therefore, the resulting
field may be represented as

E = Ed exp[i27fvt] + exp[i27fVI (t + L/e)]} =


= E I exp[i27fvt] (1 + exp [i27f(2v~t + .6.</»])
where .6.</> = i/L]« + (2v/e)vL/e is a constant phase , provided that v is
constant. By using the formula (2.50) it is easy to calculate the intensity
of the resulting oscillations :

where II = E? Since the photocurrent of the photodetector is pro-


portional to the intensity, one can observe harmonic oscillations of the
output signal , which occur at the frequency f = v(2v/e) , by means of
an oscilloscope . Let us estimate f for a prism velocity of v = 0.1 cta]« .
For a laser wavelength>. = 632.8 nm and the well known magnitude of
c we get f ::::: 3 kHz.
The superposition of the two laser beams, resulting in the DOPPLER
effect, holds if the requirements of mutual coherence are maintained.
Mutual coherence means the polarisation states of both superimposed
beams are the same, and the path difference between them (about 2L) is
less than the so called coherence length of laser radiation. Additionally,
this experiment demonstrates the large value of the coherence length of
laser radiation.

4.3 Monochromaticity
The model of a plane monochromatic wave is the first approximation
of a real light wave. Nevertheless, all the features of light waves, such
as frequency, wavelength, polarization state and the shape of the wave
front, may be understood through this model. In turn, the concept of a
quasi-monochromatic wave is closer to the radiation from real sources ,
since it deals with wave trains of finite durations. One may treat such
wave trains being composed of monochromatic waves, the frequencies of
which are distributed within a finite interval.
Another reason why the quasi-monochromatic approximation is often
used is that optics, as any other physical science, must be based on
50 DEMONSTRATIONAL OPTICS

Figure 2.26 The function


I(Av) = sin 2(Avr )/ (Av)2
specifies the spectral dis-
tribution of intensity of a
wave-train of a fixed time-
period r. The dependency
of the electric field vec-
tor of the wave on time
is given by the expression:
cos(27l"vot) with conditions
-r/2::; t::; r/2.
!iv

measurements. Such measurements always determine the light energy


averaged over a certain time interval. In other words, we will be generally
interested in two main questions: what is the spectral composition of the
light wave, and how large is the intensity of the wave? In this connection,
it is of interest to analyze the results obtained in Example 3 (calculations
of the spectral distribution of the intensity of a wave train).
Fig.2.26 shows a plot of the following function :
2(D.vT)
f(D.v) = sin (2.54)
(D.vT)2
where D.v = v - VQ , and VQ is the carrier frequency of the wave train.
The function f(D.v) is even with respect to the argument D.v and has a
maximum f(O) = T2 at D.v = 0, or at v = VQ .
The fact that the maximum value of the spectral distribution of inten-
sity is equal to T 2 is caused by the superposition of monochromatic waves
with frequencies concentrated around the carrier frequency VQ. Such a
spectral distribution is typical for all wave trains or quasi-monochromatic
waves. The superposition of monochromatic waves mentioned above
leads to a narrow spectrum of radiation. It is typical that most of the
energy is grouped within a finite frequency interval D.vQ. In this example
D.vQ denotes the frequency region between the two first minima of the
function f(D.v). Since the first minima of the sine function occur for
D.YT = ±7f, the required frequency interval D.vQ is given by
(2.55)
It is important to note that it follows from this relation that the greater
the time interval T, the smaller is the spectral interval D.vQ , and vice
versa.
PROPERTIES of LIGHT WAVES 51
In turn, with an increase of the time interval T, the wave train becomes
increasingly monochromatic, and the light energy becomes increasingly
concentrated in a narrow region around the carrier frequency. In the
limit T - 00, the whole energy of this wave groups in a single peak at
frequency Yo. For this case a sharp and infinitely high peak would be
assigned to the function f(!:i.v) . However, in reality the value of T never
reaches such limit , since we always deal with wave trains of limited time
periods in optics. The fact that intensity always has a finite magnitude
also confirms the requirements mentioned above.
Usually thermal light sources emit natural light in a wide region of
frequencies or wavelengths. But the selection of a narrow spectral in-
terval is often required when measurements of the intensity are carried
out. For this reason different devices are used to increase the degree of
monochromaticity of light radiation. The degree of monochromaticity
may be defined by the ratid' of the carrier frequency Vo to the width of
the spectral interval !:i.voi

(2.56)

The simplest optical device of this kind is the human eye. The ability
of the human eye to distinguish colors is based on the difference of the
eye's sensitivity to different wavelengths of visible light. Colored glasses
and optical color filters for a camera are also examples of such simple
devices.
We should note that the function f(!:i.v) was discussed based on the
argument !:i.v, rather than on the frequency v (Fig.2.27). Since we deal
with quasi-monochromatic waves, it is more convenient to represent the
spectral dependency of the intensity in terms of the frequency difference
to emphasize the fact that the time-dependency of an envelope of the
original oscillations of an electromagnetic field has to be a slow function
of time.
In practice such spectral dependencies are often used, for example,
in optical spectroscopy when studying shapes of spectral lines. A lot
of methods for selecting very narrow regions of the visible spectrum are
available. The frequency value corresponding to the maximum intensity
in the spectral distribution is usually recognized as the carrier frequency.
For a given value of Vo the dependency I(!:i.v) can be investigated by
measuring the line shape. There are a large number of spectral lines,
which can be described by special functions known as the Lorentzian
and Gaussian curves. These functions are also even with respect to the
central frequency, so the concept of the carrier frequency may be used
in these cases as well.
52 DEMONSTRATIONAL OPTICS

1(L\v)

o L\v

Figure 2.27. Illustrating the principle of optical filtration . The dependency of the
light intensity I(~/I) of a thermal source (the curve presented by the thin line) and
the spectral characteristic of an optical filter, presented by the thick line. The total
intensity selected by the optical filter is proportional to the area below its transparency
function . /10 specifies the carrier frequency of a quasi-monochromatic wave selected by
the filter . The spectral region, which may be associated with such a wave, is denoted
by ~/lo .

Another method for selecting a required spectral region with well know
spectral characteristic is the use of so called interference filters . An
optical interference filter is characterized by a spectral curve of a well
known shape. This symmetric curve has a sharp maximum at a certain
frequency that may be taken as the carrier frequency. We should note
that applying the concept of quasi-monochromatic waves is more correct
than the monochromatic concept, even when the spectral distribution
does not have a symmetrical shape.

4.4 Natural light


Radiation from thermal light sources is sometimes called natural light
according to the specification given by FRESNEL. Features of natural
light may be deduced from the specific conditions of radiation by a ther-
mal source, where radiating particles undergo chaotic disturbances due
to the thermal motion of the atoms and molecules composing the heated
body.
One can treat such radiating particles as oscillators which emit elec-
tromagnetic waves (visible light), assuming that every oscillator emits
independently from the others. An important point is that there is no
correlation between the moments of initiation and termination of the
radiating process for each individual oscillator. From this observation
natural light may be regarded as chaotic composition of a huge amount
PROPERTIES of LIGHT WAVES 53

=k; .-~ - - - --
fA,
'II
\ X/~---'"
:-'

a b

Figure 2.28. The states of two beams of natural light are identical until a certain
polarizing device is inserted into the beam . In t he first state, the original beam it is
assumed to be composed of two mutually orthogonal linearl y polar ized beams of the
same intensity (a) . In the second state the same original beam is assumed to be a
composition of two beams of circularly polarized light : one is left and another is right
cicularly polarized (b) .

of wave trains where each train is emitted randomly with respect to the
others. Therefore no certain time interval and no polarization state can
be assigned to natural light.
Nevertheless , it is possible to introduce some order or correlation into
the chaotic radiation of a thermal source by using some passive methods.
For example, if an optical filter is inserted into a beam of natural light
t hen most of the wave trains passing t hrough the optical filter will have
approximately the same duration, which can be roughly estimated by
(see (2.55)):
T rv l/llv
where tlv is the effective spectral region of the optical filter. The symbol
for the duration T is used to point out the fact that we are dealing with
a huge amount of practically identical wave trains, instead of a single
train considered in the previous example. In other words, this symbol
is usually used to specify a mean or average magnitude, since, in fact ,
T plays the role of the mean value of the wave train duration obtained
after inserting the optical filter.
It is typical for natural light that the observer is designing the polar-
ization state. A beam of natural light does not have a certain polariza-
tion state due to the chaotic nature of the radiating processes. Neverthe-
less, we can prepare a light beam with a desired polarization, using an
appropriate polarization device. Here we can assume that the original
beam of natural light consists of two beams with mutually orthogonal
polarization. Every beam has the same magnitude of the intensity, so
that the total intensity is equal to their sum. In 'other words, for linear
polarization state we assume two beams of intensity of 10/2, polarized in
two orthogonal directions, propagating along the same direction, where
54 DEMONSTRATIONAL OPTICS

10 is the intensity of the original beam. This assumption is based on


the fact that two waves with mutually orthogonal linear polarization
give rise to a resulting wave, where the flux of light energy is equal to
the sum of individual fluxes, corresponding to the polarized beams. We
proved this rule in Example 2 considered above. In the same way, one
can consider the original beam of natural light to be composed of two
beams of circular polarizations, one having left circular, the other right
circular polarization (Fig.2.26). As in the prior case both beams have
the same intensity 10/2. Devices to generate a desired polarization state
are treated in Chapter 4.

4.5 Light pressure


4.5.1 Radiation of a harmonic oscillator
Dipole radiation plays a very important role in optics, because it is
possible to explain the majority of optical phenomena arising from in-
teraction of the optical field with matter in terms of the linear harmonic
oscillator model. The key point of the simplest model is that a free
electron may undergo a constrained oscillating motion around its equi-
librium position under the action of the external electric field of the
light wave. These oscillations will be harmonic under the influence of a
monochromatic wave with frequency w. It is clear, that the electron dis-
placement from its equilibrium position may be described by a harmonic
function of the form
x(t') = Xo cos(wt') (2.57)
With the assumtion that the velocity of the electron has to be always
much less than the light velocity we may apply here the model of dipole
radiation. Let us derive an expression for the radiation field E(t - t') of
a harmonic oscillator, where the time interval is determined by
, r
t - t =-
c
analogous to (2.11), where t' is the time when the electron is located at
a particular point and t is the time corresponding to the coordinate of a
point of observation specified by the radius-vector r with respect to the
electron location.
The electric dipole moment vector is

p = -qx , (2.58)

where -q is the electron charge and the radius vector x is drawn from the
origin of the reference frame to the negative charge. The electric dipole
moment of the displaced charge is subjected to harmonic oscillations as
PROPERTIES of LIGHT WAVES 55

x r

Figure 2.29. Stimulation of electric dipole radiation by the electric field E of a light
wave. The dipole moment p of the electron -q is along the direction of E. The
radiation intensity is dependent on the azimuthal angle (J and on the distance r .

a function of time t' occuring with light frequency w. Let us find the
second derivative of the dipole moment vector with respect to t', using
(2.57) and (2.58):
2
d p(t' ) = _q~x(t') = -qa(t') = qx w2 cos(wt') (2.59)
dt,2 dt,2 0

where a(t') is the electron acceleration and Ixol is the amplitude of elec-
tron's oscillations.
Substituting a(t') from eq.(2.59) to (2.10), we obtain the following
relationship for the radiation field of the harmonic oscillator:

') _ qw2 cos(wt') . () _ w2 cos(wt') . ()


E( t - t - xo 4 2 sm - -po 4 2 sin , (2.60)
'lTcOC r 'lTcOC r

where IPol = Iqxol is the amplitude of electric dipole moment oscillations


and () is the angle between the radius vector passed from the electron
equilibrium position to the point of observation and the dipole moment
vector (Fig.2.29).
The electric dipole moment vector is directed parallel to the electric
vector of the incident light wave. The electron displacement in this case
always occurs against the external electric field direction (because of the
negative charge) and perpendicular to the path of the incoming light
beam. As the vector P points from the negative charge to the equilibrium
state of the oscillator, it will be parallel to the electric field vector E of
the incident wave field (Fig.2.29) . Therefore, the electromagnetic wave,
emitted by the harmonic oscillator possesses the same polarization state
as the incident light wave. For example, a linearly polarized light wave
causes oscillations of the dipole moment along the direction transversal
to the light wave propagation. A circularly polarized light wave causes
circular motion of the dipole moment vector in the plane orthogonal to
the direction of the incident wave propagation.
56 DEMONSTRATIONAL OPTICS

Figure 2.30 The radiation


characteriatics of a dipole .
Intensity of the radiation is
represented by brightness
of white pixels. The ver-
tical white arrow indicates
the direction of the dipole
oscillations .

Let us find an expression for the radiation intensity of the harmonic os-
cillator as a flux of the mean radiation energy density. Following (2.50),
the density of the energy flux is expressed by the quantity

S = CEoE2
Substituting expression (2.60) for the electric dipole field, we find

S= 1 IPol2 w4 cos 2 (wt' ) sin2 () (2.61)


(471" )2c:O c3r 2
The intensity of radiation as an averaged value over time t' may be
obtained by averaging the function cos 2 (wt' ):

J
00

cos 2 (wt' ) = cos 2 (wt' )dt' = ~


-00

As a result we obtain, for the intensity of radiation of the harmonic


oscillator,
4 2
IPol
2
10 = w sin () (2.62)
3271"2c: oc3r 2
The dependency of the harmonic oscillator radiation intensity on the
polar angle () is shown in Fig.2.30.

4.5.2 Lorentz force


Light absorption and light reflection and refraction at the boundary
of two media are accompanied by mechanical forces acting upon a sub-
stance. These forces are usually referred to as light pressure.
Consider the simplest case of orthogonally incident light on the surface
of an absorbing substance. This substance will feel mechanical pressure
in the direction of the incident wave. Indeed, the incident light wave
PROPERTIES of LIGHT WAVES 57
x
v
E -qo---

z
B

Figure 2.31. The Lorentz force initiated by a linearly polarized plane monochromatic
wave. The velocity v of the negative elementary charge -q is anti parallel to the
electric vector E of the wave; the Lorentz force is in the same direction as the unit
vector n of wave propagation.

interacts with the electrical charges of the medium and gives rise to a
force applied to the elementary free electron charge -q:

F = -qE-q[v x B] (2.63)

where v is charge velocity. The second term is the LORENTZ force that
may be transformed by expressing the magnetic field B in terms of
the electric field strength E. We introduce a unit vector n along the
direction of propagation of the light wave (Fig.2.31). Then we may
write B = [n x EJle and the expression for the LORENTZ force F L takes
the form
FL= -q"!.. x [n x E] = - ~ n(v. E) (2.64)
e e
In transforming the triple vector product we have assumed that vector
v is anti-parallel to vector E of the electric field, causing motion of the
elementary charges. The product of the force -qE and the velocity v
is equal to the power N delivered by the electromagnetic wave to the
charge. Hence, it is possible to write

(2.65)

While interacting with a charge, an electromagnetic wave transmits en-


ergy. The period averaged power N is positive. Therefore it follows
from expression (2.65), that the LORENTZ force is directed along the
wave propagation vector n . The period averaged force -qE turns out to
be zero. Therefore time averaging of the total force (2.64) acting upon a
charge leads to a macroscopic light pressure effect, created only by the
LORENTZ force (2.65)
We note that the force in (2.65) is not directly dependent on the
elementary charge q. This fact allows us to generalize expression (2.65)
for the case of a light-absorbing macroscopic object. Let the object have
58 DEMONSTRATIONAL OPTICS

= = I

a b

Figure 2.32. Light pressure force caused by reflection by a mirror (a), and through
scattering by a small sphere (b) .

a cross section (J on a plane normal to the propagation direction n. For


the force we again get
N
Flp=n- (2.66)
c
where Fl p is light pressure force and where the power of the incident
radiation N may be represented in terms of the POYNTING vector S.
For the case of a completely absorbing surface the expression for the
power then takes the form: N = a(n · S).
The existence of the light pressure force allows the conclusion that the
electromagnetic field may be characterized by a certain momentum. The
light pressure force changes the mechanical momentum of the material
body on which it acts. Based on the general law of momentum conser-
vation, one might assume that a body should acquire momentum due to
a change in the light wave momentum. A similar momentum change of
light occurs not only by absorption but also by reflection and even by
light scattering by an obstacle. An expression for the light pressure force
in the last two cases may differ from (2.66). This difference is caused
by an effective energy balance between incident, reflected and absorbed
light fluxes. It may be expressed by a certain numerical dimensionless
factor K, :

N
Fl p = K,n-
c
(2.67)

Let us make an estimation for the hypothetical case of a plane mir-


ror. The ideal plane mirror totally reflects the incident light flow in the
opposite direction and the coefficient K, has its maximum value K, = 2
(Fig.2 .32,a). In other cases 0 ::; K, < 2. For example, for uniform scatter-
ing of the incident light by a small sphere, one gets K, = 1 (Fig.2.32 ,b).
PROPERTIES of LIGHT WAVES 59

Cell 1-- - - - ---1


II • "
• .J • : ..: ' ~'
.: ..''a:
~
(
• ~ . . !

/
Laser beam \
Lens Water /
Lens Small scattenng Trace of laser
sphere beam

a b

Figure 2.33. A glass cell to demonstrate the light pressure effect (a); a small sphere
moving when the laser beam is on (b).

4.5.3 Evidence of light pressure


As the light of a laser beam is a highly monochromatic wave with
a high degree of monochromaticity, it can be focused to an extremely
small area having a linear dimension close to its wavelength. Therefore,
with a laser beam it is possible to produce a high degree of light energy
concentration even with moderate laser output power. These properties
of laser radiation can be effectively exploited for the demonstration of
the light pressure effect.
The laser beam in this experiment is focused by a lens onto a small
sphere of practically transparent material which is placed in the focal
plane of the lens. We estimate the acceleration of the sphere , assuming
that its density is p ~ 1 g . cm- 3 , its diameter d = 10.x is equal to the
diameter of the light beam in the focus of the lens, the power of laser
beam is N = 1 Wand /'i, = 1. Using (2.66) it is not difficult to obtain
N N
a=-=--
me Vpe

where V is the volume of the sphere , which for the above parameters
may be estimated as V = 7rd3/6 ~ 500.x 3 . Now, for X = 550 nm one gets
a ~ 4 .10 4 m/s 2 • If the acceleration of gravity 9 ~ 10 mis, so a ~ 4 ·10 3g.
In the demonstrational experiment that we shall consider now, very
small spheres of transparent plastic material, suspended in water, which
is used to overcome the difficulties associated with the heating of the
spheres (by absorbing laser light to a small extent) and thermal forces.
The preferable mean radius of these spheres is about if = 50 nm and
the radiation of an argon ion laser in the green range of the spectrum
(.x = 514 nm with a power of N ~ 1 W) is used. The experimental
setup for observation of the light pressure effect is shown in Fig.2.33,a.
The laser beam passing along the horizontal direction falls on a short-
60 DEMONSTRATIONAL OPTICS

focus lens (f = 0.5 cm) mounted on the vertical side of a glass cell.
The trajectory of the light beam can be seen due to light scattering in
the water . As there are only a small number of the particles at the
scattering centers , the water will appear dark. However, if a scattering
particle passes close to the focus of the lens, a bright green asterisk is
observed which moves along the horizontal direction.(Fig.2.33,b) The
speed and the brightness decrease when it moves away from the focus.
Of course a friction force works against the accelleration due to the light
force, so the particles are moving slowly in the water.
The kinetics of these suspended particles is quite complicated since
there are many contributing physical effects. For example , the linearly
polarized laser beam induces an electric dipole on a sphere . For a con-
dition if :::::; O.lA, the angular distribution of the emitted intensity of
this dipole will be symmetric with respect to the axis of the dipole.
(Fig.2.32,b). Such symmetrical angular distribution also implies that
the net momentum of the scattering light is equal to zero, hence the
change of light momentum caused by scattering by a particle is equal to
the light momentum falling on the particle per unit time . Therefore the
factor K, in (2.67) is equal to unity.
Owing to the small difference between the optical density of the
spheres and the water , the quantity of pressure force is sufficiently small .
Apart from light pressure, the moving particle undergoes a flow resis-
t ance by water, which is described by STOKES ' law. Therefore the mo-
tion of the particle is slow enough for observation. Here we shall not
give the calculations of these forces, but it should noted that a thorough
examination shows that the two forces - the light pressure and the
STOKES force - give a sufficient explanation of the observed velocity of
the particles without including forces induced by thermal gradients. A
particle experiencing light pressure is shown in Fig.2.33,b .

4.6 Photon representation


In Chapter 1, we saw that a monochromatic wave may be treated in
terms of light quanta, or photons. Each photon possesses the energy:

Eph = hu .

From the expression for the total energy E of a moving particle,

deduced in the theory of relativity, where p is the momentum and mo the


mass at rest, one gets Eph = hv = pc for photons (mo = 0). Therefore
PROPERTIES of LIGHT WAVES 61

Figure 2.34. A monochromatic wave with a certain frequency is treated as photons.


A photon carries the momentum likj the projection of the angular momentum of a
photon on the propagation direction is equal to +Ii for left - circular, -Ii for right -
circular and 0 for linear polarization.

each photon carries a momentum


h 21T
IpPh! = tu/]c = hj>" = - - = lik
21T >..
and the vector of the momentum is given by

PPh = lik ,

where h is PLANCK'S constant (h = 6.6260755 x 10-34 J. s and Ii =


hj(21T) ~ 1.053 x 10-34 J . s).
From the energy density w we get the number of photons n per unit
volume

and the photon flux


N 2
z = hvA [1j(sm )]

if a light beam of power N fills and is normal to the area A.


The total energy per unit volume is nhu, whereas the absolute value
of the total moment is nhkj(21T) = nhvje. This means , that the energy
density and the absolute value of momentum density Ipi for light are
related as follows:
Iple=w (2.68)
where pe is the momentum flux density. The light pressure force F
acting upon a unit area results from changes in p by an obstacle:

F = -bope . (2.69)

For example, a completely absorbing material (a perfect absorber) will


give the following change of incident light momentum: bop = - p. For
the light pressure force we get: F = pe, or F = w. Additionally, the light
power N incident on a unit area per second is N = we, that also gives
F = N j e = w. A similar expression is valid for the case of scattering
light by an oscillator, when the resulting angular distribution of light
62 DEMONSTRATIONAL OPTICS

momentum is equal to zero, as it was regarded in the light pressure


experiment. So we see that wave theory and photon representation are
equivalent.
A monochromatic wave with circular polarization carries an angular
momentum. This is also valid for a single photon. For a given direction
of propagation of the wave, specified by a unit vector s, the projection
of the angular momentum of a photon on this direction may possess the
following magnitudes: +Ii, 0, -Ii, where photons with zero projection
are associated with a linear polarized monochromatic wave. Photons of
angular momenta projection +Ii belong to a monochromatic field of left-
handed, and -Ii of right-handed polarization. If a circular polarized light
beam is represented by photons, possessing the projections of angular
momentum ±Ii, then the angular momentum per unit volume is L = ±
slin . Transforming the last equation by means of w = rdu» we get

L=±wjw (2.70)

For example, the change ~M in the mechanical angular momentum of a


body, which completely absorbs a circularly polarized wave, results from
changes of the light angular momentum: ~M = L Taking into account
(2.70) for (e.g, with sign "+" for left - polarized light) one can write
the following expression :
w N
~M=s- =s-
w cw
where N is the light beam power.

SUMMARY

The principle properties of light are explained by treating it as an


electromagnetic wave, which can be described by MAXWELL'S equations,
which incorporate basic characteristics such as velocity of propagation,
polarization, capacity for transmission of energy and momentum. A
resultant electromagnetic field due to two or more sources is obtained
by adding the fields of each source at the point of observation. Thus
electromagnetic waves obey the principle of superposition. Using this
principle, electromagnetic theory gives a simple explanation of polariza-
tion and quasi-monochromatization.
PROPERTIES of LIGHT WAVES 63

k z

Figure 2.35.

Electromagnetic waves, generated by a certain initial source, propa-


gate as oscillations of electric and magnetic fields. According to classical
MAXWELL electrodynamics, an accelerating charge emits an electromag-
netic field. Electromagnetic waves, and light waves in particular, trans-
port energy and momentum. Apart from energy and momentum, light
waves transport angular momentum. The value of the angular momen-
tum is connected to the circular polarization of an electromagnetic wave.

PROBLEMS

2.1. A circularly polarized plane monochromatic wave of frequency


W, propagating in vacuum, may be regarded as a flux of photons with
a constant mean density n. (Fig.2.35). Derive the expressions for the
mean values of following quantities:
a) the energy density,
b) the density of energy flux,
c) the momentum density,
d) the density of momentum flux,
e) the angular momentum,
f) the density of angular momentum flux.
2.2. A laser emits ultra-violet pulses with energy U = 1 J, duration
T = 8 ns and an effective beam diameter d of 1.5 mm. Calculate the
length 1 of each wave group and the energy density w immediately after
the laser output.

2.3. A continuous wave He-Ne laser having an effective beam diameter


of d = 2 mm radiates red light (A = 633 nm) with power N = 1 mW.
Calculate the energy density of the output of the laser.

2.4 It is known that our eyes are sensitive to a very weak flux of visible
light under conditions of full darkness near the peak of the spectral
64 DEMONSTRATIONAL OPTICS

V .~· S

Figur e 2.36.

sensitivity of the eye, which lies in the blue-green region of the spectrum
(.A = 550 nm). Normal observers require only approximately 50 quanta
to see a light flash, provided that such weak flashes from a light source
follow periodically in intervals of TO :::::: 5 s needed to reach threshold
visibility. It should be taken into account that only about 20% of the
quanta incident upon the sensitive cells of the eye will be absorbed, on
other words the effective quantum efficiency q of an eye is about q = 0.2.
In the experimental setup shown in Fig.2.36 a weak flux from a source
S passes through the lens L , the optical filter F and then the parallel
beam of blue-green rays falls on a slit of dimensions a = 2 mm times
b = 5 mm. A rotating disc, placed behind the slit , having an orifice of
th e same size, opens periodically the hole, and periodical flushes may be
observed by the eye. The distance R between the centers of the wheel
and the second slit is equal to R = 10 em. Estimate the intensity of
light necessary to permit threshold visibility when penetrating into the
eye.

2.5. Let a parallel laser beam of power N be represented by a flux


of light quanta of density n. Every quantum carries the energy lU..J and
the momentum lik. The light beam falls on a totally reflecting disk of
cross-section a. The effective cross section of the beam is S. (Fig.2.37).
Derive the formula for the light pressure force.
2.6 A short light pulse from a laser is reflected by the mirror M and
propagates through a thin parallel plate of mica as shown in Fig.2.38.
The light beam is linearly polarized, whereas the mica plate provides
circular polarization for the transmitted light wave.
Estimate the net angular momentum obtained by the mica plate, pro-
vided that the duration of the light pulse ' T = 1 ns, its wavelength
PROPERTIES of LIGHT WAVES 65

n ~
_·_·_·-6~:'~~·_·_·~ -. ~.

Figure 2.37.

Figure 2.38.

t ~F:=
1 =~=====J Ph 011
D

Figure 2.39.

A = 480 nm, its power N = 10 kW, and the effective cross-section of the
light beam S = 5 mm 2 .

2.7. Today measurements of weak light intensities are generally car-


ried out by means a photomultiplier in the so-called photon counting
mode (each detected photon creates a current pulse which is counted) .
A simple scheme is shown in Fig.2.39. Light rays from a sodium spec-
trallamp L working at low vapor pressure (A = 590 nm) are collimated
by an objective 0 into a parallel beam which passes through an optical
filter F and circular diaphragm D.
This parallel beam arrives at a photomultiplier operating in the pho-
ton counting mode. The light penetrating the cathode of the photo-
multiplier causes photoelectrons. Every photoelectron occurs due to the
absorption of a light quantum tu». After amplification within the mul-
tiplier the original photoelectron sequence is detected as a sequence of
66 DEMONSTRATIONAL OPTICS

1-4-----C------.l

Figure 2.40.

photo-counts. The total photo current is usually expressed as the number


of photo-counts appearing per second in the counting circuit .
Estimate the mean number of quanta in space between the diaphragm
D and the diaphragm d placed in front of the photocathode and the
mean intensity behind the diaphragm D if the mean frequency of the
photo-counts is 7 = 100 kHz and the distance between the diaphragms
l = 50 em and both diaphragms have the radius 0.2 mm. The quantum
efficiency of the photocathode is q = 0.2.

SOLUTIONS

2.1. We choose a plane normal to the propagation of the wave at


a fixed point in space. For the unit area a on this plane we draw a
cylindrical volume in the direction of propagation with the length l = c-
1 s, where c is the velocity of light (Fig.2.40). Since every photon carries
the energy lu», the momentum lik , and the angular momentum lik/k (in
the case of left circular polarization), the energy density is w = nn», the
momentum density is p = nlik and the angular momentum density is
m = nlik/k. The net number of light quanta within the chosen volume is
N = nul. Since all the quanta cross the unit area a during 1 second, we
get for the mean density of energy flux we = new, for the mean density
of momentum flux pe = nelik and for the mean density of angular
momentum flux me = nelik/k. The quantity we is the intensity of the
wave I = new. The power falling onto the area 8 is given by N = 1.8.
2.2. The length of every wave group is estimated to be l = CT ~ 240
ern. Immediately after the laser exit the energy 1 J is contained in
PROPERTIES of LIGHT WAVES 67

the volume 1rid?- /4, thus the energy density may be estimated to be
w ~ E/(1rld?- /4) ~ 2.5.10 5 J/m 3 .

2.3 Using the results of Problem 2.1 one gets I = N / S = we and


w = N / eS, which gives, for the numerical values used here , w ~ 10- 6
J/m 3 .
2.4 Let I be the intensity required for threshold vision of the eye.
Within one complete revolution of the wheel the illuminating slit is open
during the period T ~ alu, where v is the linear velocity of second slit 's
center v = OR (Fig.2.36). The magnitude of the angular velocity 0
may be estimated from the fact that eye's sensitivity should be restored
during the dark period.
Hence, the duration of one revolution of the wheel may be estimated
to be TO ~ 5 s, then 0 ~ 21r /TO. Therefore the duration of a single flash
T may be estimated by the following equation:

a a aTo
T > - = - = --
- v OR 21rR
For the given interval T and slit's area ab, light energy U = I abr passes
through the opened slit:
a2bTo
U = Iobr = I 21rR
Because only 20% of quanta are absorbed by the sensitive cells, the
threshold amount of energy 50hv is approximately equal to the quantity
qU:
a2bTo
50hv=qI--
21rR
Finally, for the intensity we can write the following:

1= iu/ 501rR
qa2bTo
Substitution of the numerical values gives the following estimation for
the intensity: 1= 10-10 W/m 2 . As we have seen in Example 1, the the
mean flux energy density of the Sun radiation near the Earth's surface
is equal to 1.4 . 103 W/m 2 . This means that our eyes are sensitive to
light intensity over an enormous range of intensities (about 1013 ) .

2.5 We fix a cylindrical volume within the light beam of length e as


shown in Fig.2.41. Under conditions of stationary flux of the light beam,
the mean number of quanta within this volume is equal to: ncB and the
68 DEMONSTRATIONAL OPTICS

I c I

Figure 2.41.

net momentum is (see Problem 2.1):


P =ncS/ik
Since light crosses through this volume in 1 second, the magnitude P
is also the net momentum of light, which passes through the fixed cross
section S per unit time . Because the cross section of the disk 17 < S ,
only a part of the net momentum falls on disk 's face:
17
~P = P s = nCl7/ik .

Due to total reflection by disk 's surface , the change in t he light mo-
mentum per unit time is equal to -2~P. In turn, the change in the
momentum of the disk per unit time is equal to 2~P . Hence, the light
pressure force acting upon the disk is:
F = 2~P = 2nCl7/ik .
Since this force is directed normal to the disk surface we can introduce
a unit vector I directed along k then write the following expression for
the light force:
F = 2l7cn/ikl .
We substitute tu»/ c for /ik in the expression for light force:
F = 217nfiw I .
The magnitude nfiw is th e energy density of the light beam that , with
w = ntu», gives the formula:
F = 217wl .
Now, since the power of the beam N = Swc, then substituting the last
expression in the formula obtained above gives:
PROPERTIES of LIGHT WAVES 69

+L

-L

Figure 2.42.

2.6. Since the cross section of the laser beam is less than that of the
mica plate the total energy flux passes through the plate. The polariza-
tion state of the light pulse becomes circular as it propagates through
the plate.Let the direction of circular polarization of the pulse become
that shown in Fig.2.42 and the total angular momentum of the pulse be
L. After passing of the light pulse through the mica plate, the plate has
got the angular momentum -L.
In order to calculate the magnitude of L we represent the light pulse
in terms of a flux of light quanta, every quantum carries the angular
momentum h. It is our task to calculate the total number of quanta Q
contained within the light pulse. We use the expression for the power
of the light beam that was obtained in the previous problem 2.5: N =
Srduac, where the power N represents the light energy of the beam,
passing through the fixed area S per unit time. Hence, the total light
energy of a light pulse of duration T is equal to the following magnitude:
NT, whereas the total amount of light quanta is Q = NT/IUv. Therefore,
the net angular momentum of the transmitted light pulse is lNT /w, and
the mica plate receives the following angular momentum:

L= _INT
W

where 1 is a unit vector specifying the direction of L. Substitution of


numerical data in last expression gives the magnitude of the angular
momentum: L ~ 5 .10- 21 J·s.

2.7 Under the conditions of photo counting regime provided by the


photomultiplier, a single photo-count should be associated to every ab-
sorbed quanta. Hence the mean number of quanta absorbed by the cath-
ode during 1 second is 1 , whereas the mean number of quanta falling on
the cathode per second is l/q. We assume that all the light flux passing
70 DEMONSTRATIONAL OPTICS

through the diaphragm D gets to the photocathode, hence 1j q is also


the mean amount of quanta, passing through an arbitrary cross section
of the beam between the diaphragms D and d per unit time. Since it
takes 'T = L j c for light to travel the distance between the diaphragms,
the mean number of quanta within the space between the diaphragms
may be estimated by the following value:

_ 1T fL
n=-=-
q qc
The intensity I behind the diaphragm D, represented in terms of flux of
light quanta, is
1= 1tiJ..J
q1rr2
Substitution of the numerical magnitudes gives the intensity I :::::: 8.5 .
1O-7 Jj (m2s) and the mean number of quanta is only n:::::: 8 .10- 4 •
Chapter 3

LIGHT POLARIZATION

Before the concept of electromagnetic waves, and before the famous


experiments of MICHELSON and MORLEY and EINSTEIN 'S conception
of relativity, the wave theory of light was based on the assumption,
th at light waves propagate through a specific medium , the aether, filling
every empty space and all transparent bodies . The propagation of light
waves was regarded as mechanical oscillations of the aether as an elastic
medium (in a similar way as sound propagates in air) . As different
oscillation processes in elastic media had already been well studied that
time, the propagation was thought to happen as longitudinal mechanical
waves.
As mentioned in Chapter 1, one of the first phenomena indicating the
transversal character of the oscillations of light waves was the observation
of birefringence in natural calk-spat cryst als. Transversality of light
waves was comprehended and discussed for the first time in the works of
FRESNEL and YOUNG. This hypothesis arose as a result of experimental
studies of interference of polarized beams conducted by FRESNEL. It was
discovered that no interference exists between two light beams polarized
in two mutually orthogonal directions. To explain this phenomenon
YOUNG put forward the idea of the transversality of oscillations in the
light wave. Despite the fact, that this contradicted to the elastic nature
of the aether, FRESNEL used this concept to derive a number of principle
relations of the wave theory of light, including the well known FRESNEL
formulae, which depict the laws of light refraction and reflection at the
border of two homogeneous dielectric media. Therefore in the middle
of the nineteenth century a controversial picture of light propagation
arose, since in an elastic medium (aether) only longitudinal oscillations

71
O. Marchenko et al., Demonstrational Optics
© Kluwer Academic/Plenum Publishers, New York 2003
72 DEMONSTRATIONAL OPTICS

are possible, but some optical experiments showed that light waves have
a transversal character.
MAXWELL 'S electromagnetic theory provided an adequate explana-
tion of the transversal nature of light waves. Despite discrepancies with
the mechanic model of aether, physicists began to treat light as electro-
magnetic waves in the visible region of the spectrum. Later, based on
the principles of EINSTEIN'S theory of relativity, which had been per-
fectly corroborated by a number of basic optical experiments such as
MICHELSON'S experiments and observation of star aberration, the prob-
lem of the aether was finally solved. The fundamental ability of a varying
electromagnetic field to reproduce itself, which produces a running elec-
tromagnetic wave, propagating even in vacuum, allowed a dismissal of
the concept of aether as a special light transport medium.
Adoption of the electromagnetic nature of light initiated new studies
concerning interaction between light and matter. Practically all optical
phenomena known at the beginning of the XIX century, had been ex-
plained by LORENTZ 's electron theory, based on the electron structure
of matter. Interaction of a light wave with matter was explained as an
effect of the electromagnetic field on the bounded atomic electrons. All
approaches of the classical theory of optical phenomena were founded
on the concept of a harmonic oscillator affected by the field of a mono-
chromatic alternating electromagnetic field.
Basic types of light wave polarization, ways of polarizing light from
a natural light beam and a number of polarization phenomena arising
when the light waves interacts with dielectric bodies will be discussed in
this chapter.

1. Basic types of light wave polarization


As we have seen in Chapter 2, one of the solutions of the wave equa-
tions (2.22,2.23) of the electromagnetic theory is a plane monochromatic
wave, propagating through empty space. We also have briefly consid-
ered polarization as a fundamental property of electromagnetic waves
and mentioned basic types of a plane monochromatic wave such as lin-
early and circularly polarized. Now we shall discuss these phenomena in
more detail.
Let a plane monochromatic wave propagate along the positive direc-
tion z of a right-handed Cartesian system. Hence, the electric vector E
and the magnetic induction vector B oscillate within a plane parallel to
the plane x, Y , whereas the wave vector k is directed along the z--axis
(Fig.3.1). The fundamental relation between the.triple of vectors E, B,
and k, being mutually perpendicular, allows the study of all types of
polarization state by describing only the behavior of the vector E.
Light Polarization 73

E
y

Figure 3.1 For a plane


monochromatic wave prop-
agating along the z- axis,
the motion of vector E
within the x, y-plane to be
analysed for determining of
the polarization state. A
possible orientation of E
within this plane is shown
by the arrow.

According to the superposition principle, the polarization state of a


plane monochromatic wave may be seen as a result of the superposition
of two waves propagating along the same direction and having mutually
orthogonal linear polarizations. Let the first wave be polarized along the
x-direction and other one be polarized along the y-direction:

Ex(z, t) = Eox exp [i(kz - wt) ] = al exp( - i!pI) exp[i(kz - wt)]


Ey( z , t) = Eoy exp[i(kz - wt)] = a2 exp( - i!p2) exp[i(kz - wt)] , (3.1)

where Eox = al exp( -i!pI ) and Eoy = a2 exp ( -i!p2 ) are the complex
amplitudes; al and a2 are the real amplitudes; !PI and !P2 are the initial
phases.
The ratio of the complex components Eox, EOy is given by the following
expression:
EOy = a2 exp( -ic5) (3.2)
Eox ci

where 15 = !P2- !PI is the phase difference between the orthogonal os-
cillations. Depending on the value of 15, different polarization states
may be realized. For example, electric field oscillations at point z = 0,
represented in terms of real functions, may by written as follows:

Ex(O, t) = al cos[271'vt + !PI]


Ey(O, t) = a2 cos[271'vt + !P2] (3.3)
Different cases of the resulting polarization state for the special case
al = a2 are shown in Fig.3.2 depending on the phase 15 = !P2 -!Pl ·
If the ratio EOy/Eo x is real (15 = 71'n,·n = 0,±1 ,±2, ...), we see that
the resulting electric field oscillations occur along a line corresponding
to a linear polarization state. The right circularly polarized wave results
from 15 = 71'/2, whereas left circular polarization arises at 15 = -71'/2.
74 DEMONSTRATIONAL OPTICS

/00 Cj",Cj 00 /
/) = 0 0< /) < ItI2 Ii = 7tI:! 7tl2 < /) < It /) = -It -It < 0 < -It/2 0 = -na -7tl2 < 0 < 0 0=0

Figure 3.2. The polarization states depend on the phas e difference 6 for al = a2 .

We specify right-handed polarization by clockwise rotation of the elec-


tric vector as the wave is travelling toward you, whereas left-handed
polarization by counter-clockwise rotation. If the phase difference does
not equal a multiple of 1r /2, elliptical polarization will arise. In these
cases, Eoy/ Eox = exp( -i8) has a complex value, thus the electric vector
E(x,y) rotates around the z-axis, and its head moves in an ellipse or a
circle.

1.1 Linearly and circularly polarized waves


Let the real amplitudes al and a2 be arbitrary but the ratio (3.2)
be a real value. The resulting oscillation from two orthogonal linearly
polarized waves given by (3.3) will still take place on straight lines,
possibly positioned as shown in Fig.3.3 . The direction of these resulting
oscillations depends on the ratio ada2 as well on the phase difference
8, provided that 8 = 1rn (n = 0,1,2, ..). Therefore, even if the phases of
the mutually orthogonal polarization states, <PI, <P2, differing by 8 = 1rn,
remain the same, a change in one of their amplitudes al or a2 will give
rise to a resulting wave with a new direction of linear polarization.
The possibility to use two orthogonal linear polarizations in all cases
discussed above is evident, but it is also possible to represent any po-
larization state of the electric field in terms of right and left circularly
polarized waves as well. Such polarization state can also be treated as
mutually orthogonal. We will next illustrate the ability of circularly
polarized waves to create a linearly polarized state.

L x

Figure 3.3. Different cases of the superposition of two waves of orthogonal linear
polarizations, depending on the wave amplitudes (8 0). =
Light Polarization 75

Figure 3.4 Projections


of the electric vectors
E on (x , y)-plamis for
righ t-handed (a), and for
a b left-handed polarization
(b) .

By using the definition of left-handed circular polarization, one may


write for two orthogonal components (in analogy to eq. (3.3)) (see
Fig.3.2):

EiL ) (0, t) = al cos(wt)


E~L)(O , t) = al cos(wt - 1f /2) = al sin(wt) (3.4)
where the superscript" L" specifies the left circular polarization state.
Let another wave, having right circular polarization of the same ampli-
tude, propagating along the same direction have a phase shift sp with
respect to the first wave. Two orthogonal components of this wave are
expressed as:

Ei R ) (0, t) = al cos(wt + <p)


E~R) (O , t) = al cos(wt + <p + 1f /2) = -al sin (wt + <p) (3.5)

where the superscript " R" specifies right circular polarization . Due to
th e superposition of th ese four waves a linear polarization st ate will
occur. Resulting oscillations of such a superposition have the following
two orthogonal components Ex(O, t ), Ey(O , t ) :

Ex(O , t) = EiL) (0, t) + EiR) (0, t) = 2al cos ~ cos(wt + <p/2)

Ey(O , t) = E~L) (0, t) + E~R) (0, t) = - 2a l sin ~ cos(wt + <p/2) , (3.6)

L 'P/2 = 0
x

'P/2 = rtl2 'P/2 = -11/2 'P/2 = 11

Figure 3.5. Oscillations of a linearl y pol arized wave cau sed by superpositio n of two
circular orthogonal polariza tion stat es of equa l amplit udes a nd th e ph ase differen ce
<p .
76 DEMONSTRATIONAL OPTICS

Figure 3.6. For an elliptically polari zed wave, the ellipse may be described by two
pro jections of the electric vect or a l , a 2 at a given magnitude of the phase difference 8.
Another repr esent ation of the ellipse may be perfor med in t erm s of its major semiax is
a and minor semiaxis b and t he an gle Xi tan X = =r- bja.

as follows from (3.4) and (3.5). The fact th at th e resulting oscillation rep-
resents a linear polariz ation state follows from t he link between Ex(O, t)
and E y(O, t) :

(3.7)

which is valid for any inst ant in t ime and for any point of observatio n.
Examples of resulting oscillat ions are shown in Fig.3.5 (in case of <p = 7f,
when cos(<p / 2) = 0, t hen Ex = 0).

1.2 Elliptically polarized waves


Based on results obtained above and on th e general relation (3.2), one
can derive condit ions for observing ellipt ically polarized waves. Treat ed
in te rms of sup erp osition of two mutually orthogonal linearly polarized
waves, ellipt ical polarization states will occur if the phase shift 8 f:-
a, ± 7l', ±27l', ... for an a rb it rary ratio of real amp lit udes a l an d a 2 of the
original waves. If al = a2, elliptical polarization is observed for 8 f:-
0,±7f , ±27f, ... and for 8 f:- ± 7f/2 , ± 37f/2 , ... (see Fig.3.2).
We note that for given values of 8, a l and a2 t he ellipse, corres ponding
to an ellipt ically polarized wave, may be characterized by an angle ex
Light Polarization 77

determined by cj and a2 :

(3.8)

as shown in Fig.3.6. The ellipse is also completely determined by its


major semiaxis a and minor semiaxes b and the angle X (Fig.3.4) [39]:
b
tan X = =f- (3.9)
a
where the signs =f permit two possible directions of the minor semiaxis
for one given a. Between the angles X and a the following relation is
valid:
sin2X _ . I:
. 2 - smu (3.10)
sm a
where as before 8 is the phase difference between the two orthogonal
components of the electric vector.
We will see in Chapter 4 that electromagnetic theory can satisfactorily
explain the transmission of light in transparent media and the reflection
and refraction of light at the boundaries of such media. However, no
medium, except vacuum , is perfectly transparent for any region of the
electromagnetic spectrum. That means all material media show ab-
sorption in some region of the electromagnetic spectrum. For example,
metals are an important class of absorbing media. When the incident
light falling on the boundary of a glancing metal, is linearly polarized,
th e light reflected from the surface is elliptically polarized. Such a po-
larization transformation takes place for every absorbing medium.

Example. A point electrical charge q moves uniformly on a circular


trajectory with radius a, see Fig.3.7. Let us describe the polarization
state of the electromagnetic wave emitted by this rotator.
Circular motion with constant angular velocity may be represented
in terms of two linear oscillations occurring in two mutually orthogo-
nal directions, provided the phase difference between these oscillations
is equal to 7r /2. According to the superposition principle the resultant
electromagnetic wave will be a superposition of two waves emitted by
two harmonic dipole oscillators. Let the origin of a Cartesian system be
at the center of the rotator and the z-axis be normal to the rotator's
plane, x, y. Let the charge move counter-clockwise if looking from the
point P at the pole of a sphere with radius R »a. For an arbitrary
point on the sphere specified by the vector S, the directions of the electric
field vectors Ex and E y emitted by these linear oscillators are mutually
orthogonal. The oscillations of Ex and E y occur with a relative phase
shift 7r /2. If e is an angle between z and the radius vector S drawn from
78 DEMONSTRATIONAL OPTICS

Figure 3.7. Spherical waves emitted by a circularly moving charge. These waves
have different polarization states, depending on the direction of observation.

rotator's center, so that the vector S lies in the plane x , Z, then the am-
plitude of E y will not depend on B, whereas for Ex one gets Ex rv cos B.
Hence, the superimposed wave will have an elliptical polarization.
In the upper semi-sphere one will observe left elliptical polarization
due to the given motion direction of the charge. This polarization state
is reduced to circular polarization at the pole point P where both am-
plitudes of the waves Ex, E y have the same value. At the opposite pole
of the sphere right circular polarization will be observed. For equatorial
points the amplitude of one of the waves emitted by the linear oscilla-
tors is always equal to zero which leads to linear 'polarization within the
plane x, y.

1.2.1 Polarizer and analyzer


As real light sources never radiate light with a pure polarization state,
it is necessary to use special optical elements in order to prepare light
of a certain, e.g. a linear , polarization state. Such elements are called
polarizers. Polarizers can operate based on different optical phenomena,
which will be speciall y considered in Chapter 4.
One of the most convenient methods for producing linearly polar-
ized light is the use of Polaroid films, based on the effect of dichroism.
Light Polarization 79

P,

He - Ne laser
Analyzer Polarizer Analyzer

Figure 3.8. The given direction of the Figure 3.9. Linearly polarized light
polarization plane of a He-Ne laser Pi caused by the polarizer Pi is tested by
may be used to determine the principle the analyzer P2 . The intensity of the
dire ction of an analyzer P2 . transmitted light is dependent on the
prin ciple directions Pi an d P2.

Dichroism means that certain materials have different absorption coeffi-


cients for light polarized in different directions. Polyvinyl alcohol films
impregnated by iodine, for example, transmit nearly 80 percent of light
polarized in one plane, and less then 1 percent of light polarized at a
right angle with respect to this plane .
If a polarizer is inserted into the beam of a laser emitting linearly
polarized light (for instance the beam of a He-Ne gas laser, generating
bright linearly polarized red radiation of >. = 632.8 nm) , one can find
that the maximum of transmitted intensity is obtained when the po-
larization plane of the polarizer is parallel to the polarization plane of
the laser light . In this way one determines the principle direction of the
polarizer. When turning the polarizer by an angle of 90° with respect
to this direction one can find that practi cally no light is transmitted
(Fig.3.8).
Light originating from a mercury lamp is close to natural light with
respect to polarization behavior and can be changed to linearly polar-
ized light by inserting a polarizer into the beam (Fig .3.9). The polar-
ization plane is determined by the polarizer Pl, Introducing a second
polarizer P2 into the beam , which is called the analyzer, and rotating
it independently, one can easily find that the maximum amount of light
transmitted thought both polarizers is observed when the principle di-
rections of the polarizer and analyzer are parallel, and a minimum of the
brightness corresponds to transversal orientation of these directions. If
this happens, we say that the two polarizers are crossed .
80 DEMONSTRATIONAL OPTICS

Figure 3.10 . Natural light passed Figure 3.11. When natural light
through an ideal polarizer and an ana- passes through a real (imperfect)
lyzer , the latter rotated by an angle a polarizer, one can find two orthogonal
with respect to the polarizer. polarized components of intensities
III and Is: in the outgoing beam . l s:
is much smaller, depending on the
quality of the polarizer.

2. Polarization of quasi-monochromatic light


2.1 Polarization degree
We can assume that a beam of natural light is composed of wavelets
having two mutually orthogonal linear polarizations, each of them being
equally presented (50% and 50% in average). The same is true when
natural light is regarded as composed from two orthogonally circular
polarizations.
The dependency of the intensity of the light passing through the sys-
tem of the polarizer and the analyzer as a function of the angle a be-
tween their principle directions may be found exploiting simple geomet-
rical considerations (Fig.3.1O). The electric vector E of the light passing
through the polarizer undergoes oscillations along its principle direction.
The projection of this vector on the principle direction of the analyzer
is given by: Eo = E cos a. The intensity of the light beam passing the
analyzer is proportional to the squared electric field strength E~ or:
10 = I cos2 a (3.11)
where I is the intensity of light beam leaving the polarizer (absorption
is neglected).
As the polarization state of natural light is undetermined, it is possi-
ble to assume that the polarizer selects one of all possible polarization
directions within the incident beam. This direction is not determined
until the polarizer is inserted into the light beam.
But only one half of the incoming light intensity passes through the
polarizer (this case corresponds to an ideal polarizer without losses due
to light reflection and absorption). Therefore the intensity of the linearly
Light Polarization 81

polarized light can not exceed one half of the intensity of the natural light
beam incident on the polarizer. In real cases the intensity of the linearly
polarized beam is less that one half of the incident beam intensity due
to various light energy losses due to the polarizer. Moreover, a small
component of the light beam polarized orthogonally to the principle
direction might also exist in the outgoing light beam (Fig .3.11). So,
in a light beam of natural light passing through a real polarizer two
orthogonal components will always exist . One says the light beam is
characterized by a non-zero value of its degree of polarization P, and
this beam has a partial polarization, in contrast to totally polarized
monochromatic light.
The degree of polarization P mentioned above results from measure-
ments of the transformation of natural light by a polarizer. It is deter-
mined in terms of light intensities polarized in two orthogonal directions
t, and III :
P = III - h (3.12)
III+h
where III corresponds to the principle direction of the polarizer. For an
ideal polarizer this quantity is equal to unity but for real devices it is
always less than one.

Example. A natural light beam with the intensity 10 passes through


a polarizer and analyzer with their principle directions rotated at an
angle a = 30°. The polarizer and the analyzer transmit 95% of the light
flux along its principle directions and 1% at the orthogonal ones. Let
us find intensity and the degree of polarization for a light beam passing
such a device.
Designating Ilr) and if) as the intensities of light after the polar-
izer with the polarization directions parallel and perpendicular to its
principle direction we have:

Ilr) = 0.5 x 0.9510 = 0.475 10


if) = 0.5 x 0.01 10 = 0.005 10

because the intensities of these polarization components incident to the


polarizer both were equal to 0.5 10 . Designating II~a) and It) as the light
intensities after the analyzer we get:

II~a) = 0.95 Ilr) cos2 30° + 0.95 if) sin 230°::::::! 0.34 10

a
Ii ) = 0.01 It) sin 2 30° + 0.01 i:t> cos2 30° : : : ! 0.0012 10
82 DEMONSTRATIONAL OPTICS

The total intensity of light leaving the analyzer is equal to:

I(a) = II~a) + It) = 0.3412 10


Finally for the degree of polarization we get:
/a) _ I(a)
p = II .1 = 0.34 - 0.0012 ~ 0 993
/a) + I(a) 0.3412 .
II .1

2.2 The Stokes parameters and Jones vectors


2.2.1 Description of a monochromatic wave
The use of the amplitudes at, a2, which we introduced above (see
(3.1)) and which characterize a polarization state, is not possible in
practice since these magnitudes can not be measured directly. Because
intensity is a primary measurable quantity, values proportional to the
squares of amplitudes are therefore generally used to characterize the
polarization state. Four STOKES parameters are often used:

= a 1 + a2
2 2
80
2 2
81 = a 1 - a2 (3.13)
82 = 2aIa2 cos {j ,
83 = 2a1a2 sin {j ,

where 8 = <PI - <P2 is the phase difference between two orthogonal com-
ponents of the electric vector having amplitudes a1 and a2. It can be seen
that three independent magnitudes a1, a2 and 8 are needed to express
the four STOKES parameters. Only three among them are independent.
It is easy to verify the following equality:

(3.14)

Using an analyzer the intensities of two orthogonal components of polar-


ization, at and a~, are measured. The first parameter 80 is proportional
to the intensity of light. The second parameter 81 may be found by
measuring of intensities by means of an analyzer. Since a1 and a2 are as-
sociated with amplitudes of two orthogonal components of polarization,
either at or a~, depending on the ratio between them, is proportional to
the maximum intensity of the transmitted light through the analyzer.
In this case another parameter will be expressed by the intensity at the
minimum of transmitted light. Two other parameters 82 and 83 may be
established via measurements using the analyzer and a so-called retar-
dation device. The latter, when placed into the beam, provides a desired
Light Polarization 83

phase difference ¢ between the two orthogonal electric field vectors. For
example, a quarter-wave plate is one frequently used retardation device ,
the operating principle of which will be examined in Chapter 5.
Another representation of the polarization state based on a vector
and matrix formalism was introduced by R .C.JONES. In this method
an electric amplitude vector E is written as:

E = [ Ex ] or E = [ . Eox exp(icpx) ] (3.15)


.e, EOy exp(icpy)

where Eox , Eoy are the real amplitudes and CPx , cpy the phases of the x-
and y-components of the electric field vector. With such notation, for
example, right-handed polarization may be represented by the vector

E = [ Eox exp(icpx) ]
R Eox exp( icpx - in/2)

where EOy = Eox , and cpy = CPx - n /2. It is convenient to normalize this
vector by dividing each element by th e value of the length of the vector
ER . The amplitudes then have t he value Eox = Eoy = ER/V2, hence
after factoring this electric vector will have the form:

An arbitrary choice of the phase CPx is possible so we set CPx = 0 which


leads to a simple vector not ation of right-hand ed polarization:

where the factor ER is omitted and exp( -in /2) = -i. In a similar way
one can write an expression for left-handed polarization:

The sum of these left- and right-handed polarization states is given by


a sum of the vectors: ER + EL :

which represents linear polarization in x-direction. The intensity of the


resulting wave is equal to 2E6x' To show t hat left- and right-handed
84 DEMONSTRATIONAL OPTICS

Figure 3.12 A pa r-
tially polar ized quasi -
monochromatic beam
passes throu gh a re-
z tardat ion plate , which
ca uses a retardat ion 1J for
orthogonal components of
th e elect ric vect or, and
then through a polarizer.
The pr inciple direction of
Retardation Polarizer the polar izer P makes an
plate
an gle (J wit h t he z--axis.

polarization are mutually orthogonal we build the scalar product of the


vecto rs E R and EL in t he following form : E RE'i :

Similar expressions may be found for all pairs of ort hogonal polariz ation
states.

2.2.2 Measurements of the Stokes parameters


A pure polarizati on state does not exist in realit y, just as a truly
monochromat ic wave does not exist. At t he ot her ext reme, totally un-
polarized or natural light does not in practice exist since a ret ardation
and a parti al polariz ation of light emitted by each arbit rary source is
always pr esent unde r experiment al conditions, for example, via scatter-
ing or reflection of light . It implies that the STOKES paramet ers should
be repres ented in te rms of mean intensit ies, which are measur ed under
real condit ions. In ot her words , t he math ematical descripti on of po-
lar ization must be performed in terms of a quasi- monochromatic wave,
having a polarization st at e with a given polariz ation degree, or in terms
of partially polarized light .
We assume that two ort hogonal components of the quas i-monochro matic
wave are expressed by the physical fields E~r) and E5 ):
r

E5r) = Ay cos(27fvt + <I>y)

where Ax, <I>x and A y, <I>y are slowly varying functions of t ime with
respect to t he carrier frequency v (see.(2.35)) . With the phase difference
Light Polarization 85

8 = <P x - <P y , the STOKES parameters take the form:

80 = (A;) + (A~ ,) , .
81 = (A;) - (A~)
82 = 2 (AxA y cos 8) (3.16)
83 = 2 (AxA y sin 8)
where 8 is a slowly varying function of time, too . Here we assume
that the interval for the measurement (for the averaging procedure) is
much longer than the effective time for any change of the functions A;
and A~, so that the mean magnitudes (A;) , (A~,), (AxA y cos 8) and
(AxA y sin 8) can be treated as constants.
Let us consider how the STOKES parameters may be determined by
means of measurements of intensity. In order to simplify the calcula-
tions of intensities we introduce complex functions associated with the
components of the electric vector (see (2.44)):

Ex = Ax exp[i(27rl/t + <P x )] Ey=Ayexp[i(27rl/t+<p y)] . (3.17)


We assume now that a retardation device, producing a retardation </J,
and a polarizer behind the retardation device, are inserted into the beam .
Let the principle direction of the polarizer have an angle 0 with respect
to the x-axis. A projection of the electric field on this principle direction
is thus expressed as

E(B, </J) = Ex cos B + Ey exp(i</J) sin B . (3.18)


Hence, for the intensity of light transmitted through the retardation
device and the polarizer as function of the angle 0 and the phase </J one
can write:
1(0, </J) = (ExE;) cos2 0 + (EyE; ) sin 2 0+
+ (ExE;) exp( -i</J) + (EyE;) exp(i</J)) sin Ocos 0 (3.19)
Substitution for Ex and E y from (3.17) in (3.19) gives:

1(0, </J) = Ix cos2 0 + Iy sin 2 0+


+ (AxA y exp[i(8 - </J)] + AyAx exp[-i(8 - </J)]) sin Ocos 0 ,
or

1(0, </J) = Ix cos2 0 + Iy sin 2 0 + (AxA y cos(8 - </J)) sin 20 , (3.20)


where Ix = (A;), Iy = (A~). It can be seen that to determine the
STOKES parameters it is ,necessary to obtain a set of measurements of
86 DEMONSTRATION~L OPTICS

the intensity 1(0, ¢) at several values of the angle 0 and the retardation
¢. For example, the following set 1(0,0), I(rr/2,0), I(rr/4,0), I(3rr/4,0),
I(rr/4,rr/2), I(3rr/4,rr/2) results in the following STOKES parameters:

80 = 1(0,0) + I(rr/2 ,0)


81 = 1(0,0) - I(rr/2,0)
82 = I(rr/4, 0) - I(3rr/4,0) (3.21)
83 = I( rr/ 4, n /2) - I(3rr / 4, rr /2)

The polarization properties of quasi-monochromatic light are deter-


mined by the term (AxA y cos(8 - ¢)) in (3.20), where the time depen-
dency of the phase 8(t) is crucial. If we assume that random changes
in the phase 8(t) occur within a very narrow phase interval, then the
wave will show a high degree of polarization. In contrast, with random
variations of the phase 8(t) between -rr and n , the averaged magnitude
(AxA y cos(8 - ¢)) is equal to zero, even if the retardation ¢ takes a fixed
value. But even in this case (incomplete) linear polarization of the wave
can be obtained, provided that (A; ) =1= (A;) . For example, such con-
ditions are realized when a beam of natural light passes through a real
polarizer, where there is no retardation (¢ = 0) and only a negligible
intensity is associated with the polarization direction orthogonal to the
principle direction of the polarizer. Then only two parameters 80 and
81 are not equal to zero, whereas 8 2 = 83 = O. In the opposite case, for
(A;) = (A;) and with 8(t), changing randomly and uniformly within
the interval -rr, n , three parameters 81 , 82, 83 are equal to zero, which in
terms of the STOKES parameters specifies an unpolarized beam. When
any of the parameters 81, 82, 83 has a non- zero value the light beam
J
shows a polarization state and the positive magnitude 8i + 8~ + 8~ will
describe the intensity of the "polarized portion" of this beam, in con-
trast to an "unpolarized portion" of intensity, 80- J8i + 8~ + 8~. The
degree of polarization of the beam is therefore equal to the ratio:

P = V./8 21 + 822 + 8 32 (3.22)


80

Since, for a quasi-monochromatic beam , P is always less than one,

(3.23)

and only in an ideal case of monochromatic light is the equality (3.14)


valid.
Light Polarization 87
Polarizer Glass pipe

H91~--~~~--_ ·-!- ·-~1


. Screen
I
I
I
I
=f= Analyzer

+
'A

Figure 3.13. a) Polarizer inserted, analyzer removed : Generation of linearly polarized


light through scattering by microparticles; turning th e polarizer results in varying
light intensity at point A. b) Polarizer removed , analyzer inserted : The induced
dip ole radiation caused by natural light has its maximum intensity in the direction
normal to light beam propagation .

3. Optical dipole radiation and polarizing effects


The peculiarities of dip ole radi ation examined in Chapter 2 manifest
themselves in many optical phenomena concerning t he generation of po-
larized light from natural light . The state of polarization always changes
when light is reflected by the boundary of a transparent dielectric, when
light is refracted, and when light is scattered by microparticles in dull
media. Let us describe these effects quantitatively by considering the
principles of dipole radiation emitted by particles of a medium which
have been excited by natural light .

3.1 Polarization under scattering


In order to demonstrate this phenomenon let us regard the following
experiment . A parallel light beam formed by a condenser lens passes
through a long glass pipe filled with water with glass windows fixed at
its ends (Fig.3.13). Microparticles contained in the water scatter the
light and cause a visual effect in the beam.
Rotation of an analyzer mounted after the exit window does not affect
the brightness of the light spot on the final screen. It indicates the
absence of linear polarization of the light beam passing the pipe filled
with water. Now let linearly polarized light, generated by a polarizer,
pass through the pipe and let an observer look on the tube from the side.
Rotation of the polarizer will cause the scattered light to periodically
appear and disappear. The light beam practically disappears if the main
88 DEMONSTRATIONAL OPTICS

direction of the polarizer is horizon tal, that is parallel to the line of sight .
The scattered light brightness is maximal for vertical orientation of the
polarizer.
Rot ating an analyzer in front of th e eye of th e observer shows that the
scattered light is linearly polarized (Fig.3.13), even when natural light
passes through th e tube (polarizer removed) . Maximal scattered inten-
sity also corresponds to a vertical orientation of the anal yzer , whereas
its minimum corr esponds to the horizontal orient ation.
An explanation of these observations is easily possible using the basic
prop erties of scattered dipole radiation which are described by eq. (2.62)
(see Chapter 2, 2.4). Let Ix and I y (Ix = I y) be the intensities of two
polarization states of the incoming light along th e x- and y-axes of a
Cartesian system, as shown in Fig.3.14. The cha rges of th e micropar-
tides begin to oscillat e in both vertical and horizontal directions, but
only th e vertical oscillations (in th e direction of the y-axis) lead to the
emission of linearly polarized light in the direction of the observer at
point A. For the int ensity of light scattered wit hin the (x, z)-plane the
dependency I(B) rv (Iy + Ix sin 2 B) is valid, where B is th e angle between
t he x- axis and the direction of observat ion (Fig.3.14). In particular ,
B = 0 corresponds to zero and B = 7r / 2 to maximal scattered dipole
radiat ion. Since th e radi ation ente ring th e eye of the observer at point
A is polarized in th e y-direction, a polarizer in front of th e eye causes
also a dependence of the intensity on rotation of th e polarizer angle. At
point A th e stray light is verti cally polarized. At point B, observing at
an angle B, horizontally polarized light is also observed .


Figure 3.14. The polarization of the scat tered light is dependent on th e angle of
obs ervation. At point A, the stray light is vertically polarized. At point B, obser ving
at angl e 8, horizontally polarized light is also observed.
Light Polarization 89

Figure 3.15. Generation of a linearly polarized beam by the reflection of natural light
from the flat boundary of a transparent dielectric medium (nl > n2). If the angle
of incidence is equal to the angle of polarization Bp , the reflected beam is completely
linearly polarized. The electric vector of the reflected beam is directed perpendicularly
to the plane of incidence.

Polarization of the scattered sunlight by the Earth's atmosphere when


the weather is clear can be observed with the help of an ordinary com-
mercial polarizing filter used in photography or with polarizing sun-
glasses. Light beams originating from different locations in the sky pos-
sess different degrees of polarization. If the angle between the line of
sight and the visible position of the sun is not too large, the degree of
polarization of the scattered light is low. The maximal degree of linear
polarization is observed for locations in the sky which are orthogonal
with respect to the position of the Sun.

3.2 Polarization due to reflection


A light wave arriving at the surface of a transparent dielectric medium
is partly reflected and partly penetrates into the medium. The electric
field of the light 'wave causes oscillations of the bounded electrons of
the medium around their equilibrium positions. Wave packages emitted
by the oscillators inside the medium are - after superposition with the
original wave - responsible for the light wave propagating through the
medium. Atomic oscillators located close to the boundary of the dielec-
tric medium are responsible for the reflected light wave. Let us now
consider, in detail, a particular case of reflection of natural light, falling
on the flat boundary of a transparent dielectric medium at an angle Bp ,
as shown in Fig.3.15 .
Bp is chosen in such a way that the directions of propagation of the
reflected and the refracted beams are perpendicular. For such a geo-
metrical arrangement of the propagation directions, the axes of the os-
90 DEMONSTRATIONAL OPTICS

__ LE t " - - - l - + - - f ".

Figure 3.16. Principle setup for generation and detection of linearly polarized super
high frequency radio waves (wavelength 3 cm).

cillat ions of the induced electric dipoles, denoted by D , coincide with


the direction of the reflected wave. Let us represent the incident light
beam as a composition of two linearly polarized waves, one with its
polarization plane inside the (x, z)-plane (the figure plane) and having
the amplitude Ell and the other orthogonal to the figure plan e in the
(y, z)-plane, having the amplitude E J.. The oscillations of the atomic
oscillators take place within the same planes. Wave trains emitted by
these oscillators passing through the dielectric medium have the same
linear polarizations: one in the (x, z)- plane and the other in the (y,z)-
plane. At the same time th e oscillators vibrating in the (x, z)-plane do
not contribute to the reflected wave, therefore in this wave only linear
polarization in the direction perpendicular to the (x , z)-plane, denoted
by EJ., is present. In this way the reflected light beam is totally linearly
polarized for angle the ()p, called th e BREWSTER angle.

3.3 Dipole radiation with 3 em wavelength


It is important to note that wave optics, being part of the more gen-
eral theory of electromagnetic waves, allows analysis of the basic features
inherent to all electromagnetic waves. On the other hand, model exper-
iments and demonstrations in another wavelength region , in particular
the range of radio-frequency waves, may help to deep en the readers un-
derstanding of optical waves. A wavelength of rv 3 em (corresponding
to a frequency f = 10 G H z = 1010Hz) exceedes the optical wavelength
considerably, and may efficiently be used to demonstrate and to model
optical phenomena, allowing the display of principle features of optical
phenomena at a macroscopic scale. The operation principle of a ba-
sic apparatus for experiments with electromagnetic waves of super-high
radio frequen cies is shown in Fig.3.16.
A high-frequency electric current (10 GHz) flowing through a thin wire
located at the closed end of a rectangular waveguide causes radiation of
electromagnetic waves. The current is produced by a clystron generator.
Inside the waveguide, a propagating electromagnetic field is established.
Light Polarization 91

Detecting horn

Emitting horn

Figure 3.17. The emitting and detecting horns of electromagnetic waves with 3 ern
wavelength . If the detecting horn is rotated on the axis of the propagation direction,
the output signal of the detector is changing.

The polarization plane of the waves is given by the orientation of the


emitting wire and the axis of the waveguide. Due to the physical features
of the waveguide, a linearly polarized electromagnetic wave of >. = 3 em
will be emitted by the open end , which works like an emitting horn.
The receiver is another waveguide with an entrance horn and a high-
frequency semiconductor diode at the other end. The diode axis is per-
pendicular to the waveguide axis. When the axis of the diode is paral-
lel to the polarization plane of the incoming electromagnetic wave, the
maximal current is induced by the electromagnetic field. If the reception
horn is rotated on its longitudinal axis, the current value will decrease
(Fig .3.17) until, at the perpendicular position, the diode current will be
equal to zero. This simple equipment allows the generation and detection
of a plane electromagnetic wave in the super-high frequency range. In
order to easily detect the output signal of the high-frequency diode , the
power of the clystron generator is modulated by a low-frequency voltage
(e.g. in the acoustic frequency range) . The power of the low-frequency
component of the output signal of the diode is a measure of the detected
signal.

SUMMARY
The polarization state of a wave can be measured for a certain po-
sition, but may change for different points of observation, e.g. from
elliptical to linear polarization. Only in the case of a plane wave is the
polarization state always the same.
The superposition principle provides a simple solution for most typical
problems by representing a certain polarization in terms of two orthogo-
nal polarization states. Such orthogonal states are given by two mutually
orthogonal linear polarizations, or by a pair of left and right circularly
polarized waves.
92 DEMONSTRATIONAL OPTICS

.
A J'

\ . :

a \."

Figure 3.18.

Dipole radiation is the most common model for describing the radi-
ation of elementary oscillators within a medium under the influence of
a light wave. It is characterized by a simple angular dependency of the
emitted intensity and of the direction of the linear polarization.

PROBLEMS

3.1. Deduce the STOKES parameters and J ONES vectors for the fol-
lowing polarization states:
x - and y-linear polarization;
linear polarization at 45° and at -45°;
right-handed and left-handed circular polarization.

3.2. As we have seen, two orthogonal polarization states show a zero


scalar product of two appropriate JONES vectors . Show that a similar
rule is valid for the STOKES parameters.

3.3. Express the STOKES parameters in terms of the JONES vectors


for a quasi-monochromatic wave.

3.4. Let a beam of natural light propagate in a horizontal direction


and cross with a second beam of quasi-monochromatic light at an angle
Q (Fig .3.18).

The intensity of the first unpolarized beam is h. The quasi-monochromatic


beam is characterized by the intensity /2 and the degree of polarization
P . What is the degree of polarization at a point A ?

3.5. Let us assume that the polarization state of a quasi-monochromatic


wave is characterized by two orthogonal components of the electric vee-
Light Polarization 93

F R P

Figure 3.19.

tor:

Using the STOKES parameters 80, 81, 82, 83, one may introduce an ef-
fective phase difference between these components. Deduce formulae for
such an effective phase difference and discuss the results.

3.6. Derive the formula (3.12) using the definition of the degree of
polarization (3.22) as given by the STOKES parameters.

3.7. Consider the experiment with light scattered by microparticles


within water as described above. Estimate the degree of polarization of
light rays arriving at an observer at the angle () with respect to the di-
rection normal to the optical axis of the light beam (Fig.3.14). Calculate
the polarization degree at () = 30°.

3.8. How mayan elliptically polarized wave be analyzed by means


of a polarizer and a so-called quarter-wave plate? There a direction
exists in the plate, which is called the optical axis. With propagation
normal to the plate surface one of the orthogonal polarised component
must be positioned parallel to the optical axis, the another one therefore
becomes orthogonal to this direction. If this is the case the quater-wave
introduces the retardation <p = 'IT/2,or -'IT /2 depepending on properties
of the plate, between the two orthogonal components of the incident
light beam.

3.9. Let a light source S emit quasi-monochromatic light of partial


polarization. In order to determine the parameters of the polarization
state, the following setup is proposed (Fig .3.19): Light from the source
S is collimated into a parallel beam by means of a pinhole and a lens.
This beam passes an optical filter F, then a retardation device Rand
a polarizer P. The operating principle of the retardation device, which
is schematically shown, is that of a so-called BABINET-compensator:
94 DEMONSTRATIONAL OPTICS

The thi ckness of two wedges at t he point where the light ray passes
may be varied to obtain a desired ret ardation ¢ between two orthog onal
compone nts of th e electric vector. In order to select a certain thickness
a narrow slit is mounted in front of the compensator. Discuss a way to
measure t he phase difference 8, provided that the polarizer is fixed and
¢ is varied by t he t hickness of t he compensator plates. The t ra nsmit ted
int ensity is measured by a photodetector.

SOLUTIONS
3.1. To simplify the not ation we assume that two possible ort hogonal
components of the elect ric vector of a polarized monochromatic wave
have unit amplitudes. Then for x-linear polarization these orthogo-
nal components at a point of observation are given by the expressions:
Ex(t) = cos(27rvt), Ey(t) = 0. These expressions result in JONES vectors
and th e STOKES parameters:

80 = 1, 81 = 1, 82 = 83 = °
In a similar way, in the case of y-linear polarization Ex(t ) = 0, Ey(t) =
cos(27rvt ):

E= [~ ] 80 = 1, 81 = -1 , 82 = 83 = °.
For linear polarizat ion at 450 the ort hogonal components are given by
Ex(t) = cos(27rvt) , Ey(t) = cos(27rvt), hence the appropri at es J ONES
vector and STOKES par ameters take t he form:

E=~[~] 80 = 2, 81 = 0, 82 = 2, 83 = °
For linear polarization at -45 0 the orthogonal components are given
by Ex(t) = cos(27rvt), Ey(t ) = - cos(27rvt), t hen the appropriate Jones
vect or and the STOKES parameters take the form:

E ~ !1 ]
= [ 80 = 2, 81 = 0, 82 = -2, 83 = °.
The JONES vectors associat ed with right- and left-handed circular polar-
ization have been considered ab ove (see 3.2). For th e right-handed po-
larization , real functions associated wit h t he orthogonal component s are
Ex(t) = cos(27rvt), Ey(t ) = cos(27rvt -7r j2), so that 8 = cf>x - cf>y = 7rj2
80 = 2, 81 = 0, 82 = 0, 83 =2
Light Polarization 95
In the case of left-handed polarization 83 = -2:
80 = 2, 81 = a, 82 = a, 83 = -2

3.2. We represent the J ONES vector of a quasi-monochromatic wave


by slowly varying amplitudes Ax, A y and phases <I>x, <I>y connected with
two orthogonal components of a polarization state:

E y = A y exp( -iwt) exp( i<I>y)

For the intensity of these components we get

hence: 85 = (ExE;)+(EyE;) . Now, two complex magnitudes ExE; and


EyE; are expressed by the amplitudes Ax, A y and the phase difference
0= <I>x - <I>y in the following forms:

thus

ExE; + EyE; = AxAy(exp( io) + exp( -io)) = 2A xA y cos 0 ,


EyE; - ExE; = AxAy(exp( -io) - exp( io)) = 2iA xA y sin 0
The STOKES parameters 82 , 83 are represented by the products of the
J ONES vectors as follows:

3.3. We use the results which have been obtained in the previous
problem and rewrite the STOKES parameters for several polarization
states in a vector form:

51 = {I, 1, a, a} 52 = {I, -1, a, a}


53 = {l,a, lD] 54 = {l,a,-l,a}
55 = {2,a, a, 2} 56 = {2, a, a, -2}
where the pair of vectors 51, 52 represents x- and y-linear polarization,
the pair 53, 54 linear polarization at 450 and -450 and the pair 55,56
right - and left- handed circular polarization, respectively. Since we are
96 DEMONSTRATIONAL OPTICS

Ph.D.

/
/
/
a /
/

/
/
/
/
/
/
/
/
/

Figure 3.20.

dealing with the vectors of real components, then the scalar pr odu ct for
the first pair is given by

S182 = 1. 1- 1. 1+0.0+0 .0 = 0

The scalar products of the other pairs can be calculat ed in t he same way.
It can be seen that all th e pairs connecte d to two ort hogonal polarization
states permit a scalar product zero.

3.4. The degree of polarization P of the quasi-monochromati c light


beam is represented by STOKES par ameters of t he following form : P =
vsi + s~ + sVso, where So = 12. Th e STOK ES parameters of th e un-
polar ized beam are: It , 0, 0, O. One may assume that a mixture of
t he two beams contains unp olarized and polarized portions, which are
completely independent. This means t he STOKES parameters of t his
combinat ion are given by the sum :

where the magnitudes Sl, S2,S3 are associat ed only with the second
beam. Since prop erties of a polarization state follow from measurements
of int ensity, we also assume that the sensitive area of a photodetector
used for the measurements is norm al to the propagation of th e second
beam , which is shown by a dashed line in Fig.3 .20. Hence th e t ot al in-
tensity of the combin ation is measur ed as It cos a + h . Fin ally, for th e
degree of polarizat ion we have:

P, . - V
/ Sl2+ s2
2 + s3
2 _ P 12
T

m tx - It cos a + 12 - It cos a + h
Light Polarization 97

3.5. For the given amplitudes and phases of a quasi-monochromatic


wave the polarization state is characterized by the STOKES parameters in
the form (3.16). One may propose the following definition for the cosine
of an effective phase difference DelI between the x- and y-components
of the electric vector:
o (AxA y cos D)
cos ell = V(Ai) (A~)

where Ix = (A;) and Iy = (A~) are the intensities of these orthogonal


components. Since the sign of DelI is not defined by the cosine function,
the second relation should be exploited:
. 0 (AxA y sin D)
sin ell = V(Ai) (A~)

In accordance with the definition of the STOKES parameters we sub-


stitute (AxA y cos D) by (1/2)82 , (AxA y cos D) with (1/2)83, (A;) with
(1/2)(80 + 81) and (A~) with (1/2)(80 - si) in both relations, which
gives:
82 . 83
cos oeff = sm Dell =
J 2 - 82
8o 1 8J
2 2
0 - 81
The effective phase difference is determined by these formulae. Never-
theless, for such a definition the following relation must be valid:
2 2
2 . 2
cos Dell +sm Dell = 82
2
8 -
+ 83
2
81
=1
0

Indeed we have stated before that 86 = 8i+ 8~ + 8~ , therefore this re-


lation is fulfilled. We may interpret this result in the following way: If
we assign a specific phase difference DelI to the polarization state of a
monochromatic wave, we assume that this wave does not show any ran-
dom changes in phase or amplitude. Therefore the degree of polarization
is equal to one and the relationship 86 = 8i+ 8~ + 8~ is valid . In practice,
for any quasi-monochromatic wave 86> 8i+ 8~ + 8~ is valid (see (3.23))
due to the fact that random changes in the phase difference 0 and/or in
the amplitudes Ax, A y always occur. The appearance of such random
variations in a quasi-monochromatic wave always causes an unpolarized
portion of intensity.

3.6. The formula (3.12) was obtained under the assumption that two
orthogonal components of the electric vector Ex, E y are completely inde-
pendent, therefore the phases cfI x(t) ,.cfI y(t) are independent , too. Hence,
98 DEMONSTRATIONAL OPTICS

in thi s case, the phase difference 0 = <P x - <P y varies uniformly within
the inte rval -7f , n , and th e mean values (AxA y cos 0) and (AxA y sin 0)
are both equal to zero. With these condit ions the STOK ES par ameters
corresponding to such a polarizat ion state are as follows:

So = Ix + I y Sl = Ix - I y o , o .
According to th e definition of the degree of polarization (3.22) we get
t he expression:

p = J sr + s~ + s5 Sl Ix - I y
So = So = Ix + I y

which is similar to (3.12).

3.7. We assume a natural light beam of the int ensit y 10 to be rep-


resent ed by two beams of the int ensity 10/2 with mutually orth ogonal
linear polarization st at es. Let th e linear polarization of one beam oc-
cur along the vert ical direct ion and the other beam along the horizont al
direct ion as shown in Fig.3.1 4 by the crossed arrows.
Th erefore an induced dip ole at some point inside the water will oscil-
lat e along both directions. The int ensities of scattered light denoted by
I II and Is. give the degree of polarizat ion:
2 2
p = I II - h = 10/2 - (10/2) sin B = 1 - sin B
I II + h 10/ 2 + (10/2 ) sin 2 B 1 + sin 2 B
where I II specifies t he intensity emitted by the dipol e with a vertical
oscillat ion, which is not dependent on e, whereas h is the int ensity
associated with th e horizont al direction of polarization, associated with
I sin 2 e. In the particular case B = 30° t he degree of polarization is equal
to P = 0.6.

3.8. In order to analyze an elliptical polarization state, the quarter-


wave plate and the polarizer are rotated independently until the intensity
of light tr ansmitted through the polarizer is equal to zero. If this is
the case, th e quarter-wave plate compensates the phase difference of
7f / 2, which exists between two orthogonal polarised components of th e
incident beam. These are precisely the components which are associated
wit h t he directions of the major semiaxis a and minor semiaxes b of
the ellipse assigned to elliptical polarization (see Fig.3.6). We denote
these components by E a and Eb. After the ret ard ation plate, th e electric
vector E a and Eb occur in phase , hence their superposition gives rise to
a linear polarizati on, which is specified by Ex , inclined at some angle X
Light Polarization 99

with respect to the major semi-axis a of the the ellipse. It is the semi-
axis which is known due to a position of the optical axis of the plate.
This linear polarization may be detected by the polarizer. When the
intensity transmitted through the polarizer is zero, its principle direction
is perpendicular to Ex. By rotating the polarizer, the intensity '" E~
can be expressed in terns of the intensities '" E~ and", El

3.9. The two positions of the polarizer corresponding to the maximum


and minimum of transmitted light may be established before the retar-
dation device is inserted into the beam. The directions of the maximal
and minimal transmission may by associated with the x - and y-axis
respectively. The polarizer can then be positioned at some angle 0 with
respect to the x- axis and BABINET 'S compensator can be inserted into
the beam. With the displacement of one plate of the compensator, the
retardation ¢ changes and the intensity of the transmitted light under-
goes periodical oscillations as follows from (3.20):

1(0, ¢) = Ix cos2 0 + I y sin 2 0 + (AxA y cos(8 - ¢)) sin 20 .

The output current of the photodetector varies between the maximum


and minimum values, and one gets

for the maximum of the intensity and

Imin(f}, ¢» = Ix cos2 B + I y sin 2 0 - (AxA y ) sin 2B


for the minimum . One notices different positions of the moved com-
pensator plate corresponding to maxima and minima of the transmitted
light, which belong to the retardations ¢>max and ¢>min. These positions
allow the estimation of the phase difference, because the following rela-
tions hold:

1= Imax(O, ¢» when 8 = ¢max ± 271"m (m = 0,1,2 ...)

1= Imin(O, ¢) when 8 = ¢>min ± (271" + l)m (m = 0,1,2...)


Between two neighboring maxima and minima the compensator plate is
shifted by .6.£0 and the phase difference is given by (271"/ .6.(0).6.f. Since
the value of the refractive index of the compensator plate (and therefore
.6.(0) is dependent on the wavelength of the beam, a narrow band optical
filter F is placed into the beam if the light source emits white light.
Chapter 4

PROPAGATION OF
LIGHT WAVES IN MEDIA

It is well known from daily experience, that a light beam falling onto
the surface of a transparent medium partially passes into the substance
and is partially reflected on the surface. In this chapter we shall consider
the laws governing these phenomena, restricting ourselves to the case of
absolutely transparent media, wherein the absorption and scattering of
light are absent.
This restriction to absolutely transparent media is a simplification,
usually applied in optics. Real media always contain a certain amount
of distributed microparticles. Being centers for scattering, these particles
provide the visibility of light trajectories, known as light rays. It is of
interest to note that even such a clear medium as the Earth's atmosphere
with very few dust impurities possesses scattering properties, which take
place on the air molecules itself. Due to the large path length in the
atmosphere these scattering processes are extremely effective. Blue light
is scattered more effectively then red, yellow or green light. This is the
reason for the blue color of the sky. Sun rays traveling into the Earth's
atmosphere also undergo scattering by random air concentrations (or
fluctuations of the air density) arising due to chaotic molecular motion.
Additionally, the presence of dust and water vapors in the lower layers
of the atmosphere yields a strong forward scattering of light rays when
the sun is near to the horizon and where the thickness of air is large.
Under these conditions the sun appears to be red.
But for absolutely transparent media under laboratory conditions the
effect of molecular scattering is negligible, so the molecules of such a
medium may therefore be considered as immovable. Moreover we assume
that no free electrons or ions exist , and the medium is an ideal dielectric,
wherein the positive and negative charges are joined together and the free

101
O. Marchenko et al., Demonstrational Optics
© Kluwer Academic/Plenum Publishers, New York 2003
102 DEMONSTRATIONAL OPTICS

particles are electrically neutral. These assumptions enable the analysis


of the fundamental properties of dielectric media.

1. Maxwell's equations in isotropic media


1.1 Wave equations and the Poynting vector
We shall discuss properties of elect romagneti c waves with frequencies
corresponding to visible light tr avelling in media. For t his treatment
several simple assumptions are necessary. First , we assume th e mag-
netic permeability J-L of the dielect rics to b e pr actically equal to t hat of
vacuum, which is 'valid for nearly all transpa rent subs tances at optical
frequencies. The mat erial of a substance will be regard ed as isotropic, i.e.
for every point it s properties are not dependent on a direction. Finally,
we shall assume t hat th e elect romag net ic field is not strong enough to
produce nonline ar polarization or ionizati on effects within the material
(ot herwise a nonlin ear dependence of the polarizati on on the electromag-
netic field strength would be observed). Und er these assumptions two
vectors are created, which describ e the behavior of the material under
t he presence of an extern al electromagn etic field. One is the elect ric dis-
placement D and t he oth er the magnet ic field H. For isot ropic dielect ric
media the following linear relati ons with t he electric and magnetic field
st rengt hs are valid:
D = cE , (4.1)
B = J-LH (4.2)
These linear relations introduced by MAXWELL are called t he material
equations. Here e is t he dielectri c permittivi ty and J.L is t he magnetic
permeability of the substance. Under the assumpt ions mentioned abo ve
MAXWELL's equations take the form:
divD = 0 , (4.3)
divB = 0 (4.4)

rot E =
aB
-- (4.5)
at
rotH
aD
= at (4.6)
The wave equations appropriate to thi s syst em of MAXWELL 'S equations
are similar t o (2.22) and (2.23):
a 2E a2 E a2 E a2E
ax 2 + 8y2 + az2 = J.LC at2 and
a 2H a 2H a 2H a 2H
ax 2 + ay 2 + a z2 = J.LC at 2
Light Waves in Media 103

As before, propagating plane and spherical monochromatic waves obey


the ~ave equations presented above. In media the POYNTING vector of
a monchromatic wave is expressed by the vectors E and H :

S=ExH

in contrast to the definition of S for the case of a monochromatic wave


propagating in vacuum (see (2.48)). Vector S points in the propogation
direction of the wave. The direction of the phase velocity in isotropic
media coincides with that of the POYNTING vector S.
In a Cartesian system let a plane monochromatic wave of linear polar-
ization along x-direction propagate in the positive z-direction. D and
H are then given by:

D = Doexp[-i(wt - kz)], H = Hoexp[-i(wt - kz)]


where the vector D oscillates within the (x, z)-plane, and vector H
within the (y, z )-plane. Using equation (4.6) the property of transver-
sality is expressed as follows:
w
Ho = kS x Do , (4.7)

where S is the unit vector in the direction of propogation; the factor co] k
is equal to the magnitude of the phase velocity of light in the material:

V=-=--=-
w
k
1
.jJi£
A
T
(4.8)

where A is the wavelength and T is the period of the electromagnetic


oscillations . In a medium let the number of complete oscillations of a
wave of the wavelength A during an interval T be equal to N . In vacuum,
this number of complete oscillations needs th e same time interval T,
th~refore the frequency v = NIT will be the same for vacuum and for
the material. In contrast, the wavelength changes:

(4.9)

where Ao is the wavelength in vacuum , and n :

n=J;o~~ (4.10)

is the refractive index of the optical material, where /-l ~ /-l0, that is, we
assume the magnetic permeability /-l of the dielectrics to be practically
104 DEMONSTRATIONAL OPTICS

--,
d
...<. "::
/'

---
<,
/'

1.._-
<,
(
I
--- -....
b

Figure 4.1. A small rectangle abed surr ounds an element AS , enclosing two med ia
having £1 and £2 .

equal to that of vacuum. Here a dimensionless magnitude c/co is often


used as relative dielectric permittivity.
Th e relation (4.7), showing the t ra nsversal nature of elect romagnet ic
waves, may be expressed by vectors H and E. We replace t he electric
displacement vector D by t he electri c vector E , using t he ma teri al equa-
tio n (4.1), and get the relation between the electric and the magne tic
vector:
Ho = /f;,s x Eo (4.11)

1.2 Boundary conditions


In optics t he case of a light wave propagating across t he bounda ry
of two media is often of interest. The polarization of the wave may
be expressed by two orthogonal components of the electric (as well as
th e magnetic) vector. M AXWELL'S equations allow us to establish laws
associated with t he cha nges in th ese components (and therefore of th e
polarization of t he wave), when the wave penetrates one medium from
anot her one. In order to get solutions for t he general case it is enough to
t reat two components : one pa rallel and the ot her norm al to the bound-
ary plane to be considered.
Let a small rect angle abed surround an element of area 6.S, enclosing
element s of both media as shown in Fig.4.1. Now, according to (2.5)

f
L
E dl = - ~ I
S
BdS ,

we t ake the first int egral around t he rect angle abed and the second one
over its area 6.s. If the height of th e rect angle h = ad = be is smaller
t han th e lengths ab and ed, then the largest contribution to the first
integr al will be given by the tangenti al components of the field, rather
Light Waves in Media 105

than those normal to the boundary. In turn, if both lengths ab and ed


are rather small these tangential components may be replaced by the
constant magnitudes tl . E 1 and t2 . E2. Hence, integration around the
rectangle abed gives:

f
abed
E dl ~ (t 1 . E 1 + t2 . E2)ab ,

where tl and t2 are the unit tangents along the sides of the rectangle:

where n12 is the unit normal vector to the boundary pointing from the
first medium into the second one; s is the unit normal vector of the plane
of the rectangle (see Fig.4.1). Integration over the area of the rectangle
gives, for a rather small fj.S, the following:

-~
dt
J BdS ~ - (~B)
dt
·sfj.S = - (~B)
dt
·s(ab)(bc)
118

Thus, we get the following relation:

tl . E 1 + t2 . E 2 = - (:t B ) ·s(bc)

If the sides of the rectangle be and ad are decreased, the right-hand


side of this equality will tend towards zero, hence for the tangential
components, for t2 = -tl = S x n12, we get:

This condition is not dependent on the vector s since the orientation of


the rectangle is arbitrary. Hence for the tangential components E t2 =
n12 x E2 and En = n12 x E 1 the following equation is true:

E t 2 = En (4.12)
in other words, the tangential component of the electric vector does not
change across the boundary. An analysis of equation (4.6), which is
similar to (4.5), results tina the boundary condition for the tangential
components of the magnetic vector H,

Ht2 = H n (4.13)

Now let a small cylinder of volume hfj.S surround a space in two


media as shown in Fig.4.2. Due to the fact that equation (4.3) is similar
106 DEMONSTRATIONAL OPTICS

;'

-._-
;'

'-. h Figure 4.2 A small cylin-


der of volume h6.S, enclos-
ing two media having <:1
and <:2 .

to (2.17) , according to (2.2) the surface integral of the vector D over a


closed surface S is equal to zero (since we assume that no charges are
present in the dielectric medium) :

f DdS =0 .

Thus we have to perform the integration of the electric displacement D


through the closed surface enclosing the volume h~S. If the areas of
the top and bottom side of the cylinder are much larger than that of the
walls the calculation of this integral reduces to:
(Dj n, + D2n2)~S = 0 ,
where nl and n2 are external unit vectors of the boundary at an internal
point of the cylinder. When the height h tends to zero, both vectors D 1 ,
D 2 will express the electric displacement at one point, with the vector
D 1 in the first medium (s = cl) and the vector D2 in the second medium
(e = c2) ' In this case ni = -n12 = n2, thus the boundary condition for
vector D is given by:

which implies that the normal component of the electric displacement


does not change across the boundary. Since in most cases we shall
consider the electric vector of waves propagating across boundaries of
isotropic materials, where e does not depend on coordinates and where
the relation D = cE is true, we replace the last condition with the fol-
lowing one:

or
En2c2 = cIEnl (4.14)
Here we use the notations E n 2 , E n 1 for the normal components of the
electric vector. It is seen that equation (4.4) is similar to (4.3) , therefore
for the normal components of the vector B one can derive the condition:
Light Waves in Media 107

Water
surface

Figure 4.3 Refraction and


Mirror
n =1.33 reflection by the surface of
water .

which shows the normal component B n to be continuous across the


boundary.

2. Reflection and refraction


2.1 Snell's law
When a plane wave falls on a plane boundary formed by two homoge-
neous media of different optical properties, it is split into two waves: the
transmitted wave proceeding into the second medium and the reflected
wave propagating back into the first medium.
Let, for instance, a light beam fall from inside water on the boundary
surface water - air . A part of it is reflected, and a part is refracted into air
as shown in FigA.3 (a light beam enters the glass box filled with water ,
falls on the mirror, and after reflection by the mirror travels towards
the water surface). The angle of incidence ()i is measured between the
direction of the light propagation and the surface normal vector . The
reflection angle ()r has the same value as ()i :

(4.15)

a relation which was already known in ancient times .


When increasing the angle of incidence ()i by rotating the mirror M
clockwise, the angle of refraction ()t increases faster and follows the re-
lation
sin ()i n2
- . - = - =n12 (4.16)
sm ()t nl
Here nl and n2 are the refractive indices of water and air (n2 :::::; 1).
Since nl > n2, one can say that the water is optically more dense than
air . The three rays: incident, reflected, and refracted lie in the same
plane , determined by the entrance direction and the normal vector of
the surface. This plane is called the plane of incidence.
The relation (4.16) is called the law of reflection, which, together with
the relation (4.15) and the statement that the refracted, reflected and
incident rays are in the sameplane (see Fig. 4.3), constitute SNELL'S law
of refraction.
108 DEMONSTRATIONAL OPTICS

2.1.1 Refraction by a prism


A prism is one of the mostly used optical devices. Let us consider the
refraction taking place on a triangular prism, as shown in Fig.4.4. If n
is the refractive index of the prism, the angles (3 and 0, corresponding
to refracted rays, are related to the angles of incidence a and "( by the
law of refraction:

nsin(3 = sin o, n sin-y = sino (4.17)


From the triangle ABC one can find:
7f 7f
(2" - (3) + (2" - "() + 8 = 7f , (4.18)
therefore, for the angles (3 and "( a simple relation is valid:

8=(3+"( ,
where 8 is the angle between the two active sides of the prism , called the
refractive angle. So, relations (4.17) and (4.18) provide the solution of
refraction on the prism. In any application the case of minimal horizontal
inclination of the incident ray is of interest. If this inclination is specified
by an angle cp, the angle between the incident ray and the outgoing ray,
the minimal value of ip will be observed at f3 = "(, or when the refracted
ray passes inside the prism parallel to its basic side. In this case

(3 = "( = ()j2, sin c = sino = nsin(()j2) (4.19)

is valid.
The refractive index of the prism material can be determined using a
spectrometer (Fig. 4.5). Light from a source passes through a narrow
vertical slit and is then collimated by an objective into a practically
parallel beam. This beam is refracted by the prism and focused by
the second objective into the output plane. The colored pictures of the
entrance slit are observed by means of an eye-piece while slowly turning
the prism to find the position which gives the minimal inclination cpo
By means of a spectral lamp (e.g. a Hg-Cd discharge lamp) emitting
several well-known wavelengths, the refractive index can be determined
in its dependence on A. For each wavelength the minimal deviation is
observed for a different angle.
In the opposite case, when the refractive properties of the prism are
known, one can examine the spectrum from a light source. For example,
the line spectrum of a Hg-Cd discharge lamp consisting of a set of bright
lines, is shown in Fig.4.6,a. Contrarily, the continous spectrum of a
xenon lamp (presented in Fig.4.6,b) is close to the visible region of the
Light Waves in Media 109

A A

a b

Figure 4.4. Refraction by a triangular prism for a general case of incidence (a) . The
case of minimal inclination for a certain wavelenth is fulfilled for rays which pass
parallel to the prism's basic side (b) .

Vertical
slit Objective
(Collimator)

Spectral
lamp

Eye
piece

Figure 4.5. A spectrometer used for the determination of the angle of minimal decli-
nation of rays . For each spectral line of the discharge lamp the prism has to be turned.
From these angles , the refraction index of the prism material can be determined.

solar spectrum, which has an intensity maximum between green and


yellow.

2.2 Total reflection


Total reflection occurs when light propagates from an optically denser
medium into one which is optically less dense (n2 < nl) and the angle
of incidence ()i exceeds the critical angle Bi given by the relation:

(4.20)
110 DEMONSTRATIONAL OPTICS

Wh en Bi = O i, sinB t = 1, i.e. Bt = 900 , so t he light propagat es in a


dir ection tangential to the boundary. If Bi exceeds the critical angle
Oi , no light enters the second medium. All the incident light is then
reflected back into the first medium. This phenomenon is called to-
tal reflection. This behavior can be demonstrated using the equipment
shown in Fig.4.3.

2.2.1 Surface wave (evanescent wave)


Nevertheless, in the case of total reflection the electromagnetic field
in the second medium does not disappear completely, even if there is
no longer a flow of energy across the boundary. Assuming that a laser
beam falls on the plane side of a 45° glass prism, it is reflected at the
inner side and leaves the prism through the third side (Fig.4.7). The
refractive index of the glass prism is 1.43, so that the critical angle of
incidence (see (4.20)) is:
Oi = arcsin(I/1.45) = 43.50
At Bi = 450 , total reflection takes place.
When a second identical prism is pressed closely to the first one , the
intensity of the reflected beam will decrease and a weak beam transmit-
ted through both prisms will appear. A strong pressure is needed for
this phenomena, because the thickness d of the air layer between the
pressed surfaces of the prisms should be of optical dimension, i.e, d A f'V

(A ~ 630 nm).

Figure 4.6. The spectrum of a spectral lamp , e.g. containing Hg and Cd, shows some
discrete wavelengths (or spectral lines) (a) . The continous spectrum of a xenon lamp
is close to the solar spectrum (b) .
Light Waves in Media 111

Incidentray Transmitted ray

Reflected ray

Figure 4.7. Total reflection of a laser beam, passing through two nearly-touching
glass prisms . The horizontal incident beam is divided into two beams - one is the
horizontal transmitted beam, the other is the vertical reflected beam. Their intensity
ratio is determined by the distance between the touching surfaces. The ext ernal forces
needed to provide a distance small enough are indicated by th e arro ws.

This experiment shows that a certain amount of field strength ex-


ists in the air close to the reflecting surface. It turns out that the field
strength decreases exponentially with increasing distance from the sur-
face. The second prism can pick up energy from this field, when its
distance is about the wavelength, leading then to an electromagnetic
wave propagating inside the second prism .
Based on the theoretical analysis given before it is impossible to an-
swer the question of how the light energy of the incident wave penetrates
into the second medium. This analysi s was based on the simplifying as-
sumption that the boundary surface and the wave itself are extended
infinitely. But a real incident light wave is always limited in space and
in time . The formulae derived above are therefore not completely valid
when an incident light beam possesses a limited cross section near th e
boundary (points 1 and 2 in Fig.4.8 ). A small part of the light energy
penetrates into the second medium and causes an inhomogeneous wave.
The amplitude of this wave, which is called the surface or evanescent
wave, is characterized by an expon ential extinction with respect to th e
distance d along the normal to the boundary. When d rv A its amplitude
is already close to zero.
So light energy may be transported from a more dense medium to a
less dense one even in the case of total reflection. It is that very small
amount of energy which is transported by the inhomogeneous wave inside
the second medium from one side of the beam (point 1) to the other
(point 2).
112 DEMONSTRATIONAL OPTICS

z
Evanescent
field

Figure 4.8. Energy transport in the case of an evanescent wave.

2.2.2 Total reflection of radio waves


This phenomenon can be more impressively demonstrated by means of
ultra high-frequency electromagnetic radiation of>. = 3 cm. A paraffin
prism shows a refractive index n ~ 1.4 for such electromagnetic waves,
which is the same value as the refractive index of ordinary glass for
visible light.
At first the emitting horn is located in front of one short face of the
paraffin prism and the detecting horn (A) is placed behind the other
short face of the prism (Fig.4.9). The electromagnetic wave from the
emitting horn passes into the detecting horn due to the total reflection
on the longer side and is detected as a highly pronounced output signal
of the electrical circuit of the detector. Then a second paraffin prism is
placed near the first one, so that the longer faces of the prisms are parallel
at a distance of more than 3 em from each other. A second detecting
horn (B) is placed behind the short face of the second prism, but no
signal is detected because all the electromagnetic energy passes into the
first detecting horn. Displacing the second prism towards the first one,
so t hat their distance becomes approximately 3 cm or smaller, an output
electric signal will appear from the second detector, whereas the signal
from the first detector will decrease. If both prisms are touching each
ather, the signal of the first detector disappears and the signal of the
second reaches its maximum value: as the boundary between the prisms
disappears, the reflection dissappears.
This demonstrational experiment shows the existence of a wave prop-
agating along the boundary between two media, even in the case of total
reflection. An energy flow from one medium into the other, as men-
tioned above, is possible but decreases with increasing distance from the
boundary to the optically less dense medium.
Light Waves in Media 113

A Detecting horns ~

Paraffin prisms ~ B

Emitting hom

Figure 4.9. The evidence for the surface wave demonstrated with radio-waves.

2.2 .3 Optical fibers


T he phenomenon of total reflection is extensively used in technical
applications for transmitting information and images by means of optical
fibers. An optical fiber is usually a glass or an optically transparent
plastic fiber, usually called a core, inserted into an envelope , called a
cladd ing, characterized by a smaller refractive index than th e refractive
index of the core (Fig. 4.10).
This construction enabl es th e observation of total reflection inside the
fiber, even for substantial bending, so that the incoming light ray un-
dergoes a large number of total reflections on the boundary between
the core and the cladding before leaving the fiber at its end face (Fig.
4.11). For the given refractive indices of core n2 and of cladding nl for
).. = 650 nm the optical fiber is characterized by its numerical ap erture
sin a = J n~ - ni , where a is the acceptance half-angle. An incoming
bundle of rays within the accept ance angle will be guided by total reflec-
tion . For example, typ ical specifications for glass fibers as of n2 = 1.492
and nl = 1.402 will provide an acceptance angle of 2a = 61 0 • Th e qual-
ity of an optical fiber is specified by the minimum radius of bending,
which is usually given in terms of the outer diameter of the fiber, D.
For most commercially fabricated optical fibers the value of the minimal
radius R of bend ing is equal to 20 x D. R may vary from 5 mm , when
D = 0.25 mm, to 5 em, when D = 2.5 mm .
When a narrow laser beam should be coupled into the fiber , the accep-
tance angle may be much smaller and a smaller difference between the
refractive indices of core and cladding is possible . Let us now obtain th e
relation ment ioned above between the numerical aperture sin sp and the
refractive indices nl , n2 . Since two opposite sides of the fiber are paral-
lel there exists the following correlation between "I and {3 : sin "I = cos {3.
114 DEMONSTRATIONAL OPTICS

Figure 4.10 Structure of a


ljl y typical optical fiber. The
\
refractive index of the core
- - - {.~ . _ . n2 is greater than that
ljl \
of the cladding ni . which
Core causes the incident ray ar-
riving at the optical fiber
Cladding '----_ _--c_ _-+-_ _----'
at the angle 'P. to be totally
reflected by the boundary
"core - cladding" .

Under the condition of total reflection by the boundary" core - cladding"


we can write: sin{3 = nIfn2' hence one obtains the following expression
for cos{3:
sin-r = cos {3 = (1/ n 2) Jn~ - n~
The angle of refraction 'Y of the incident ray may be expressed by the
acceptance angle cp and indices n 1, n2: n2 sin 'Y = sin sp, provided that
the light fiber is surrounded by air (n ~ 1). Finally, for the numerical
aperture we get :
sincp = -
Jn~ n~
When coupling light from a laser into an optical fiber , a small ball lens
is usually used (Fig.4.12). In this case the necessary numerical aperture
is determined by the diameter of the laser beam d, the diameter of the
ball D and by the refractive index n of the lens.
Let a ray fall on the ball lens at point A. The half acceptance angle
sp may be found from triangle OBC: ip = a - 'Y, where a is the angle of
incidence of the light beam. Further one gets a + (1r - 2{3) + 'Y = 1r and
'Y = a - 2{3, where {3 is the angle of refraction: sin {3 = (sin a) In. Finally

Figure 4.11 A set of re-


flections of a light ray in-
side an optical fiber .

.~';i)':_>c- -
/ /" /
Figure 4.12 A small ball
lens used to increase the
_l _ numerical aperture of an
optical fiber.
Light Waves in Media 115

we obtain: ip = a - / = 2a - 2f3. Therefore, sin o = sin2(a - (3) . The


incidence angle is determined by the ratio of d]D: a = arcsin(d/ D). For
example, for given the values n = 1.5 , d = 0.5 mm and D = 1 mm we
find: sin a = 0.5, a = 30°, sin f3 = 0.33, f3 :::= 18.5°, a = 60° - 37° = 23°
and sin ep = 0.39.
Apart from wavelength and the optical fiber materials the impor-
tant property of light transmissivity is dependent on the outer diam-
eter of the fiber and, for example, may vary from 0.36 to 0.40 dB/m
for D = 250 J.Lm. These numerical specifications , valid around 1960,
showed that optical grade fibers are capable of transmitting light as far
as 80 meters. Applications of light fibers for data transmission enor-
mously pushed the developments of radically less absorbant fiber ma-
terials. Nowadays values of,..., 0.2 dB/km are achieved. Due to these
improvements in long-distance fiber cables, amplifiers are needed only
at distances larger than 50 km. Apart from lower damping, the data
rate transmitted through fibers could be significantly enhanced by the
use of single-mode fibers. Such fibers, having a light-guiding core of
very small diameter, allow light to propagate only at one certain path
or "mode" , whereas in thicker fibers light is transmitted with quite dif-
ferent numbers of reflections along the fiber (such different numbers of
reflections lead to different signal velocities) . Further spreading of data
bits can be avoided if the wavelength of the transmitted light is close
to a region where no dispersion occurs. For quartz fibers this wave-
length is approximately 1.55 J.Lm. Further developments are devoted to
all-optical micro-switches, data-multiplexers, repeaters and amplifiers ,
in order to avoid the transformation of optical data transmission into
electronic transmission and back.
Sometimes it is necessary to transmit pictures by optical fibers. For
this purpose, optical image fiber bundles are used, which consist of a
set of identical fibers of small diameter, grouped so that the position of
every fiber at the entrance face is reproduced at the output face.

2.2.4 Rainbow
Refraction and reflection of sun rays by water drops contained in the
atmosphere is responsible for the well-known natural phenomenon of the
rainbow. We can see a rainbow as a colored arc in the sky after rain when
a cloud moving away from us is illuminated by the Sun . The description
of this phenomenon is based on the reflection and refraction of solar rays
in spherical water drops . In the simplest case there are one reflection
and two refractions for a ray, falling on the inner surface of a drop as
shown in Fig.4.13. One can easily find a relation between the directions
116 DEMONSTRATIONAL OPTICS

Figure 4-13 Light path


u ' through a drop of water. A
horizontal solar ray S pass-
ing throught the spherical
drop reflects at the point
B and leaves th e drop at
n point C at angle n with re-
spect to the horizon .

of the incident and the reflected rays , assuming a spherical shape for the
water drop.
For example, let a parallel beam SA fallon a drop at the point A and
let a = 60° be the angle of incidence. For a given refractive index for red
rays in water, nred = 1.331, we find, for the angle of refraction (3, sin (3 =
sin a / nand (3 = 40°33'. After total reflection on the inner surface of the
drop this ray will return to the surface, refracting again at point C. The
angle of refraction is also equal to a = 60° as shown in figure Fig.4 .13.
Through geometrical considerations, now one can find the angle between
a ray reaching an observer and a horizontal line to be n = 4(3 - 2a. For
th e red ray in our example this angle is equal to n ::= 42°. Due to
dispersion a blue ray will leave the drop at a smaller angle, as follows
from the value of the refractive index for blue rays nblue = 1.345. In this
case n is approximately 40°, therefore the observer will see the blue arc
of the rainbow below the red arc , i.e. as the inner arc. It is clear , the
angle of 60° assumed above is rather arbitrary, and a pencil of incident
rays with different directions and different incident angles should be
considered, instead of only one ray falling on the drop at the particular
angle a. The problem has to be treated with scattering theory, and is
not so simple . However, the estimation given above gives approximately
correct values of the visible angular dimension of a rainbow.

3. The Fresnel formulae


There is a difference in reflection and refraction for two types of lin-
early polarized electromagnetic waves, one polarized in the plane of in-
cidence, and the other one with a polarization normal to the plane of
incidenc e.
Light Waves in Media 117

Boundary
of medium

Figure 4.14. Reflection and refraction at the boundary of two media. Amplitudes
of the incident, reflected and refracted waves polarized in the plane of incidence are
marked by II, and those polarized normal to this plane by .L.

Let Ey) , Ey) and Ef) be the complex amplitudes of incident, re-
flected, and refracted waves polarized normally to the plane of incidence,
t
respectively, and let EI~i) , EI~r) and Efl ) be the complex amplitudes of
incident, reflected and refracted waves polarized in the plane of inci-
dence , respectively (see FigA.14) . The fact that these three waves are
located in one plane, according to SNELL'S law and the law of reflection
(4.15),( 4.16), enable one to find relations between their amplitudes.
The boundary conditions for the electrical components of the waves
are also used for this purpose. Each component of the electric vector of
the incident, reflected and refracted wave may be presented in a similar
form:
E = Eoexp[-i(27l"lIt - kr)]
For the origin, where x = y = z = 0, the variable parts for all the waves
are reduced to the factor exp( - i 27l"lIt ). This factor, being common to
all three waves, may be omitted, hence one can apply the boundary
conditions in the forms (4.12), (4.13) to the complex amplitudes. Since
the tangential components of the vectors E and H are continuous across
th e boundary we get:
E(i)
y
+ E(r)
y
= E(t)
y
118 DEMONSTRATIONAL OPTICS

H(i)
x
+ H(r)
x
= H(t)
x
H(i)
y
+ H(r)
y
= H(t)
y

Using the relation (4.11): Ho=~(n x Eo) and the angles ()i, ()r and
()twe represent the projections of the elect ric and magnetic vectors in
terms of E.l and Ell. In this way we obtain from the first pair of relations

E(i)
.l
+ E.l(r) = E(t)
.l
and ~l cos ()·(E(i) _ E(r)) = ~e2E(t) cos () .
t . l . l . l t
J.L J.L

for the components of the electric vector polarized normal to the plane of
incidence . The second relation allows the introduction of the refractive
indices nl = Jel/eo and n2 = Je2/eo (see 4.10). Thus this pair of
relations may be presented in the form:

E(i) +E(r)
.l.l.l
= E(t) and nl cos ()·(E(i)
t .l
- E(r))
.l
= n2E(t) cos ().
.l t
. (4• 21)

These relations, connected with the magnetic vector, permit the follow-
ing relationships for the parallel polarized components:

or by introducing the refractive indices:


nl(E(i)
II
+ E(r))
II
= n2 E(t)
II
and cos ()·(E(i)
t II II
cos () .
- E(r)) = E(t)
II t
(4.22)
For a given Ey) and ()i, the first pair of the equations (4.21) contains two
independent variables : EY) and E~). Therefore these variables can be
found independently from E~r) and EI~t) , which confirms that two waves
of orthogonal polarization states are independent of each other. From
this result follows that the problem of refraction and reflection may be
solved for two orthogonal components of polarization independently for
natural light.
When resolving (4.21) and (4.22) with respect to the components of
the electric vector of the reflected and refracted waves one obtains:
E(t) _ 2nl cos ()i E(i)
II n2 cos ()i + nl cos ()t II

E(t) _ 2nl cos ()i E(i)


(4.23)
.1. - nl cos ()i + n2 cos ()t .1
E(r) _ n2 cos ()i - nl cos ()t E(i)
II - n2 cos ()i + nl cos ()t II
Light Waves in Media 119

E(r) _ nl cos ()i - n2 cos ()t E(i)


1.. - n2 cos ()i + nl cos ()t 1..

The relations (4.23) are called the FRESNEL formulae , which directly
provide the simple phenomenological consequences connected with the
reflection and the refraction of polarized light. We shall introduce re-
lationships between the intensity of the incident light beam and the
intensities of the refracted and the reflected beams before considering
applications of the FRESNEL formulae.

3.1 Reflectivity and transparency


According to the definition given in Chapter 2, the intensity of a plane
monochromatic wave is represented by the product I '" EoEo, where Eo
is the complex amplitude of the electric field of the wave. The ratios of
the complex amplitudes from the FRESNEL formulae (4.23) are real. For
instance, after transformations which use the reflection law (4.15, 4.16),
one gets:
E(r)
sin(()i - ()t)
.......d:....=
E(i) sin(()i + ()t)
1..

The real ratio of EY)/ Er) can be used to derive a ratio between the
incident intensity and the reflected one in the following form:

(4.24)

The quantity R1.. defines the portion of the intensity corresponding to the
reflected wave. It is called the power reflectivity for the wave polarized
normally to the plane of incidence. If we calculate the ratio Ef)/ Er)
from (4.23) and call the square of the absolute value T1.. = IEf)/ Er) 12 ,
we obtain
T 1.. = sin 2()t sin 2()i
(4.25)
sin 2 ( ()i + ()t)
This quantity represents the portion of the intensity corresponding to
the transmitted light , provided the polarization is normal to the plane
of incidence. The quantity T 1.. is called transparency. Assuming that ab-
sorption is absent and that a wave polarized in the plane of the incidence
is also absent, we get

in agreement with the law of the conservation of energy. Therefore the


relation
(4.26)
120 DEMONSTRATIONAL OPTICS

I j--" E(~
II
Figure 4.15 Components
O)
E11---11 of the electric vector of an
incident and reflected wave
n,
E II(i ) an d E(r)
II ' bot h In
i t he
plane of incidence, are in
opposite directions.

is valid. It can be verified by substituting R1. (4.24) and T1. (4.25) in


(4.26).
In a similar manner, for a wave polarized in the plane of incidence we
obtain
2
R - tan (Oi - Ot) and (4.27)
II - tan 2 (Oi + Ot)
71 = sin 20t sin 20i
(4.28)
II sin 2 (Oi + Ot) sin 2 (Oi - Ot)
Again we can verify, using the relations (4.27) and (4.28), that RII +111 =
1. Moreover, the net reflectivity R defined as R = R II + R1. and the total
transparency T = 111 + T1. are related by the similar expression:

This fact results from the general concept, that the intensity of two waves
linearly polarized in two mutually orthogonal directions is represented
by the sum of the intensities of both components.

3.1.1 Normal incidence


For normal incidence, 0i = 0, 0t = 0, the complex amplitudes of the
transmitted and reflected waves from the FRESNEL formulae (4.23) are:

E(t) _ 2nl E(i) E(r) _ n2 - nl E(i)


(4.29)
II - n2 + nl II' II - n2 + nl II

E(t) _ 2nl E(i) E(r) _ nl - n2 E(i) (4.30)


1. - n2 +
nl 1.' 1. - n2 + nl 1.

Now we shall discuss the relationships between the transmitted and


the incident complex amplitudes. For reflection at an optically denser
medium, where n2 - nl > 0, we obtain from (4.15) that the ratio
. re al an d POSIitiive,
E II(r ) / E II(i ) IS

Taking into account that EI~r) is oriented in the opposite direction with
respect to EI~i) (see Fig.4.15), we conclude that the phase undergoes a
Light Waves in Media 121

change of 71" in this case. For the same case, n2 - nl > 0, the ratio
Et) I E~) = - E~r) I EI~i) is real and negative (see (4.30)) and again a
phase change of 71" takes place. So the oscillations of the electric field
vector in the wave reflected by an optically thicker medium in the case
of normal incidence undergo a phase change of 71" with respect to the
incident wave. Contrarily, the vibrations of the refracted wave are in
phase with the incident wave.
The distinction between the amount of reflectivity of the parallel and
the perpendicular components disappears in the case of normal inci-
dence, and from the relations (4.29) and (4.30) we find:

RII = R.L = ( nn22 +- nl


n1 ) 2
, rr:
.111
--T.L-- 4n2nl
(n2 + nl)2
(4.31)

The value of R when the wave falls normally on the boundary between
glass and air (n2 ~ 1.5 and nl ~ 1) can be estimated as:

R = (1.5 - 1) 2 ~ 0.04 .
1.5 + 1
This result means, that only 4% of the incident intensity is reflected.
The FRESNEL formulae allow the establishment of phase changes be-
tween incident , refracted and reflected wave. All ratios between par-
allel and normal components of these waves, which can be formed by
means the FRESNEL formulae (4.23), are real. This means that the phase
change between appropriate pairs of the components may be invariant
or changes by 71" . It can be seen from (4.23) that E~t) and E~) have the
same signs as EI~i) and E~), therefore the phase of the transmitted wave
is equal to that of the incident wave.
In the case of the reflected wave, the ratios between the orthogonal
components depend on the relative refractive index or on the relationship
between the incident and refracted angles. With reflection by a denser
medium, when n2 > nl and ()i > ()t, the signs of E~) and Et) are
different, which leads to a phase change of 71". Similar considerations
predict a phase change of 71" for EI~r) and E~) .

3.2 The Brewster angle


If the incident wave is polarized in the incident plane , at a special
inclination angle the reflected intensity can become zero. This special
case results directly from the FRESNEL formula for EI~r) (4.23). From
the condition E~r) = 0 we obtain a straightforward relationship between
122 DEMONSTRATIONAL OPTICS

Boundary
of medium

Figure 4.16. Refraction and reflection at the BREWSTER angle (nl < n2). An ind uced
dipole located on the dielectric boundary may be represented by projections of the
dipole moment PI! and Pl., where the PI! lies within the incident plane and Pl. is
normal to this plane. Oscillations of PI! do not cause radiation in the direction of the
reflected wave, so only Pl. gives rise to th e reflected linear polarized wave.

the angles:
cos Ot n2
= (4.32)
cos Oi nl

This relationship together with the law of reflection (4.15, 4.16) is valid,
when
(4.33)

This implies that the reflected and the refracted rays form an angle of
90° (Fig. 4.16). Now, using (4.32) and (4.33) we find the condition for the
angle of incidence, which is called the BREWSTER'S angle or polarizing
angle, in the form:

(4.34)

As we have seen, the existence of this effect can easily be deduced using
the model of harmonic oscillation of surface charges and the radiation
charact eristics of a dipole (see Chapter 3, sec.3.2).
Light Waves in Media 123

3.2.1 Polarizing devices


As an evident application of the BREWSTER effect, linearly polarized
light may be created by reflection at the polarizing angle. This effect
builds the operating principle of N ORRENBERG 's polariscope, schemati-
cally represented in FigA.17 . We let natural light fall on the glass plate
(M1) at the BREWSTER angle. The light reflected towards mirror M con-
tains only E-vectors oriented perpendicular to the incident plane. This
linear polarized light is reflected by M and passes through M1 towards
the second glass plate M2. The first plate fills the role of a polarizer,
and the second one is the analyzer. If M2 is positioned at the polar-
izing angle with respect to the incident beam, and its normal vector is
parallel to the normal vector of the first plate, the intensity of light re-
flected by M2 reaches its maximal value. This case corresponds to the
linear polarization of light leaving the system of mirrors. The mirror M2
can be rotated around the axis of the propagation of the beam, so that
the plane of incidence can be changed . As a result, the intensity of the
reflected light can be decreased , and two orthogonal linear components
of polarization are present in the outgoing light beam; the polarization
degree is less than unity. When the mirror is rotated by 900 , no light is
reflected.
The main disadvantage of the polarization effect produced by the
N ORRENBERG mirror is the small fraction of light reflected at the po-
larizing angle. In this regard another polarizing device, consisting of
a stack of thin plane parallel plates, is preferably used. The incident
beam passes through the stack without changing its direction, as shown
in Fig. 4.18. Let the incident light be natural, which enables the assump-
tion that the light is composed of two linear polarized components by
the same amount (50%). One component is polarized in the plane of
incidence and the other is polarized normally to this plane .
After passing through one boundary of a plate the ratio of intensities
of the two polarized components (see (4.25),(4.28)) will be:

711 = (nlcosBi +n2sinBt)2


T.!. n2 cos Bi + nl sin Bt
At the polarizing angle, when cos Bt = sin Bi following from (4.25), and
using (4.28), we find

711 =
T.!.
(ni + n~)2
2n2nl
(4.35)

Taking the values of the refractivity indices of the glass as n2 :::::: 1.45 and
of air as nl :::::: 1.0 we estimate this ratio as 711fT.!. :::::: 1.145. This ratio
124 DEMONSTRATIONAL OPTICS

~
I
I

a b

Figure 4.17. NORRENBERG's mirrors .

E, E,

Figure 4.18. A stack of thin plane-parallel plates. The electric vector component Ell
of the incident light beam lies in the plane normal to the flat surfaces of the plates.
Th e transmitted light beam is polarized to a high degree in the plane normal to the
surfaces of the plates.

increases as the power of the number of reflecting surfaces. For instance


if the number of plates is 20, the ratio will be 1.14540 ::::: 225, and when
the number of plates is 40 the value of 1l1/TJ. will be 1.14580 = 50634.
Such a polarizer is mainly used in the far ultra-violet or infra-red spectral
ranges .
Another important application is the use of so-called BREWSTER win-
dows which allow the entrance (or exit) of light of a certain polarization
direction into (out of) a device. For example, such windows are used for
Light Waves in Media 125
the discharge tubes of gas lasers (Fig. 4.19). The windows are fixed at
the polarizing angles with respect to the optical axis of the laser . A light
wave propagating between the two laser mirrors, polarized in the plane of
incidence, will experience no reflection losses when passing through the
windows. Other polarization directions undergo partial reflection at the
BREWSTER windows and are less amplified by the laser medium. Since
the light wave, while running forward and backwards inside the laser
resonator, passes very often the windows, therefore linear polarization
of the emitted laser light is caused.

4. Dispersion
As we already pointed out the propagation properties of electromag-
netic radiation (and of light waves) within an ideal uniform dielectric
can be described in terms of the dielectric constant e and the refrac-
tive index: n 2 = cleo. The simplest model of an ideal uniform dielectric
medium is an ensemble of identical classical oscillators, consisting of elec-
trons which can move around their equilibrium positions. If the initial
displacement of the electron is ~x , it experiences (for small dislocations
~x ) a linearly increasing force with displacement: F = -k~x and will
then oscillate with its resonant frequency wo = y'klrne.
The dependence on frequency of the elementary oscillators in the elec-
tromagnetic field of the light wave is called the dispersion of the medium,
i.e. the dependency of the dielectric permittivity and the refractive in-
dex on the frequency of the wave. It is well known from mechanics and
from the theory of electricity that the properties of an oscillator depend
on the ratio between the frequency of an external force and the resonant
frequency of the oscillator. Generally two different limit cases can be
distinguished. The first is when the frequency of the external periodical
force is much smaller or much larger than the resonant frequency wo , and

Laser mirror

Discharge lube
Brewsterwindow

Figure 4.19. Generation of laser light of high degree of linear polarization state.
Inside the laser resonator, the light passes many times through the BREWSTER window
of the discharge tube. This provides the linear polarization of the laser beam .
126 DEMONSTRATIONAL OPTICS

the second case when the frequency of the external force is close to woo
In optics the second case is commonly associated with light absorption
within a certain frequency band being characteristic for the medium.
The optical equivalent of the first case is the region of dispersion far
from the band of absorption.

4.1 Classical theory of dispersion


We shall now discuss the dispersion of an ideal gas for demonstrating
the principles of classical dispersion theory. When an electromagnetic
wave penetrates into the gas, an electric dipole moment is induced in
each atom (treated as a system of a heavy positively charged ion and
a mobile electrically negative electron) due to the action of the electric
field of the wave. Let p be the electric dipole moment of each atom of
the ensemble, its value being directly proportional to the electric field of
the wave E
p = aE , (4.36)
where a is the mean polarizability of the atom.
According to the representation of an electric field within an ideal di-
electric media, the vector of the electrical displacement D of the medium
is determined by the electric dipole moment of an unit volume P , and
the electric field strength:
D =eoE+P
Let N be the number of atoms per unit volume. Taking into account
(4.36) we obtain
P=Np=aNE ,
and now
D = eoE + aNE = (co + aN)E = eE
where
e=eo+aN=eo (1+ ::) .
We introduce the relative dielectric permittivity E:

E = e/eo = 1 + aN/co. (4.37)

So the problem of the frequency dependency of E as well as that of the


refractive index is reduced to an analysis of the relationship (4.36), which
is the relation between the harmonic oscillations of the electric field E
and the oscillations of the elementary dipole p induced by this field. By
definition p = -ex, where -e is the charge of the electron and x is its
displacement under the action of E .
Light Waves in Media 127

NEWTON'S equation of the motion of the electron (of the elementary


oscillator) takes the form:

d 2x dx 2 e
-dt 2 + 2')'-
dt
+ wox = - -~e'
E (4.38)

where Wo is the resonance frequency. Energy damping of the oscillator


is taken into account by the second term in the left part of the equation,
where')' is the damping constant.
For a given temporal dependency E(t), the equation allows us to find
the time-dependency of the displacement x . To simplify the solution
of (4.38) we calculate the oscillations induced by a linearly polarized
electric field, which is represented by the complex expression

E(t) = Eoe- iwt . (4.39)

As (4.38) is a linear differential equation, its solution for the temporal


dependenc y of an external force like (4.39) can be assumed to have the
form:
x(t) = xoe- iwt . (4.40)
Therefore dxld: = -iwxoe-iwt and ~x/dt2 = -xow 2e-iwt . Using E(t)
in the form (4.39), we obtain

xo(w5 - 2i')'w _w 2) = -~Eo ,


me
and
-e/~e
x(t) = 2 2 2" E(t) , (4.41)
W o- w - 1/YW

which tells us that x(t) has the same temporal behavior as E(t) . Multi-
plying both parts of (4.41) by -e we find the expression for p(t)
e2/~
p(t) = -x(t)e = W 2 - 2
w -
e2 "
t')'w
E(t)
o
Hence, the atomic polarizability a in (4.36) is given by

e2/~e
a = --;;----';:;----::---
w6 -w 2 - 2i')'w

Taking into account equation (4.37) for the relative dielectric permittiv-
ity we get:
e2N 1
E = 1 + -- -."..-----,,,---- (4.42)
2
~co w6 - w - 2i')'w
128 DEMONSTRATIONAL OPTICS

The fact that E in (4.42) is a complex number is caused by our assumption


of damped oscillation of the electron (4.41). In the undamped case
(r = 0) E will be real. Analyzing the denominator in (4.42) we can
assume (for high frequencies) that w2 »w5 and also (far enough away
from resonance) (w5 - w2) »2i'yw . So apart from resonance the
imaginary part of E is very small and nearly no damping (no absorption)
takes place.
The opposite C8."!e is true close to resonance: if (w5 - w2) < 2i'yw
(or even w5 - w2 = 0 for exact resonance), the imaginary part is large
compared to the real one. In this case a reasonable absorption takes
place due to 'Y > O.
As we see, the complex value of E is associated with an absorption of
energy by atomic oscillators when w ~ wo o The refractive index therefore
is also complex

e2 N 1
fi2 =E= 1+---~---- (4.43)
meo w5 - w2 - 2i'yw
We formally represent the complex refractive index in (4.43) by intro-
ducing its real and imaginary parts:

fi = n + ir:
so that
fi2 = (n + iK,)2 = n 2 - K,2 + 2inK (4.44)
Introducing the constant

(4.45)

having the physical dimensional of angular frequency, and substituting


w p into (4.43) we separate the real and the imaginary parts as follows:

w2(w2 - w2) 2iw2w'Y


fi2 - 1 + p 0 + p
- (w5 - w + 4w2'Y2
2)2 (w5 - w2)2 + 4w2'Y2
and we get

2 2 wp2( Wo2 - w 2) w2w'Y


n - K = 1 + (2 W o -w
2)2
+ 4w2'Y 2 and ti« = (2 W o -w
2)2
+ 4w2'Y 2
(4.46)
For light field frequencies close to the resonant frequency Wo the following
approximation is valid

w5 - w2 = (wo - w)(wo + w) ~ 2wo~w ,


Light Waves in Media 129

where ~w = wo - w. Using this approximation we can write (4.30) as:


w2
n« ~ -L (~)~ 2
4wo w +,
. (4.47)

Example
For diluted gases under normal conditions (at 00 C and 1060 mbar)
the frequency wp ~ 3 . 1014 radian per second, whereas for the visible
range of spectrum w ~ 3.6.10 15 radian per second (green line). Referring
to the fact that hard ultraviolet is strongly absorbed by air it is possible
to estimate Wo ~ 1 . 1016 radian per second. (This fact is a fundamental
condition for the existence of biological life of the Earth!)

We also assume that the damping constant , is small compared to


wo, so that the inequality is satisfied:
w~/wo,« 1 (4.48)

For this case we get


w2
,2 «
,
nr: ~ -L - 1 (4.49)
4wo
for ~w = 0, i.e. the maximum of ns: at ~w = 0 has a value much smaller
than the one. Because K « 1 and K2 « 1, K2 on the left hand side of
(4.47) can be omitted, and we get

Finally we obtain (using the approximation formula (1 + x) 1/2 ~ 1 + x /2


for x « 1):
(4.50)

As we see, n is close to one. Hence, the following expression for K is


valid (see (4.47)):
w2 ,
K ~ -L .,....,....--:-:::----~ (4.51)
4wo (~w)2 +,2
The dependencies of the real and the imaginary parts of n are shown in
Fig. 4.20 as functions of the parameter ~w h. The function n(w) is called
the dispersion curve and describes the refractive index as a function of
w (as a result of the theoretical consideration of the elementary theory
of dispersion).
130 DEMONSTRATIONAL OPTICS

Figure 4-20. The function -x/(l + x 2 ) corresponds to the dispersion curve new) at
x = llwh, and the function 1/(1 + x 2 ) of the same argument is associated with the
imaginary part of n.

The second curve K(W) shows the absorption profile, which in this
particular case is given by a LORENTZIAN curve described by the depen-
dency

(4.52)

Two domains of frequencies, W « WQ and W rv WQ, are usually of in-


terest. The region of frequencies lower than the resonant frequency WQ
is called the domain of normal dispersion. In this frequency region the
refractive index increases with an increase of the frequency. The behav-
ior of the refractive index near and within the light absorption region is
more complex. The odd function n(w) has a maximum and a minimum
symmetrically located with respect to the resonance frequency, the low
frequency maximum at WQ-, and the high frequency minimum at wQ+"
In the frequency region between WQ - , and WQ + , the refractive index
decreases with increasing frequency and we say that anomalous disper-
sion is taking place. Light rays having shorter wavelengths are then
refracted less than those of longer wavelengths . The region of anom-
alous dispersion, as we can see in Fig.4.20, coincides with the central
part of the absorption line of the medium .
Normal dispersion beyond the absorption line, can be described by
the formula, derived from (4.46) when (w6 - W 2)2 » 4w2 , 2 ,
Light Waves in Media 131

Horizontal Objective yl
slit
X'
--++-~H-#--:>I't-- 'L.I- ' _ ._ ._ . . _ . _ . - . _. _. - . -=--
. -'Ar - _ . ~

Observer
Vertical y
slit

Figure 4.21. A setup for observing the disperion of glass in Newton 's scheme of
crossed prisms.

Taking into account that w«wo or w/wo « 1 we can estimate:

1 1 w2
----".---=-2 ~ -+-
w6 -w w6 w6
and the expression for the refractive index, represented in terms of the
wavelength, takes the form

(4.53)

where
w2 41T
2C2
w2
A = 1 +...E. and B = p
w6 w6
Relationship (4.53) is called CAUCHY 'S formula. This approximation
provides an accurate description only for low density media, for instance,
gases, within which the refractive index is close to one.

4.2 Observation of dispersion


4.2.1 Crossed prisms
NEWTON suggested a method for observing the dispersion phenom-
enon with two crossed prisms. A setup using the method of crossed
prisms is shown in Fig.4.21. A xenon lamp emits a continuous spectrum
in the visible light region, similar to that of sunlight (see Fig.4.6,b). By
means of the condenser lens the bright light beam of the lamp is con-
centrated on a narrow vertical slit . An image of the slit is projected
by means of the objective in the plane of a second narrow slit placed
horizontally. Then a real image of the second slit is projected by the
second objective onto the plane of observation.
If a glass prism is inserted into the beam behind the vertical slit , the
spectrum of the xenon lamp will appear in the plane of the horizontal
132 DEMONSTRATIONAL OPTICS

slit . If the refractive edge of the prism is vertical and the base of the
prism is located as shown in Fig.4.21, then the spectrum in the (x ,y)-
plane will be horizontal; its" blue" side is associated with larger x than
the "red" side. In (XI ,yl) we have to rotate the coordinate system due
to the real image formed by the second objective. When now inserting
a second glass prism after the horizontal slit, the spectrum within th e
(XI, yl)-plane will bend (Fig.4.22).
The observer behind the (XI, yl)-plane may interpret the bending of
the spectrum as caused by the dependency n(>.), because the displace-
ment of the spectrum's colors is connected to wavelengths.
It is obvious, that the shape of. the spectrum, presented in Fig.4.22 ,
has to depend on the positions of the bases of the crossed prisms , which
may be illustrated by the following exampl e. Let point A lie on the
optical axis of a lens (Fig.4.23) and let the lens produce a real image PI
of the plane P. Without the prism the image AI would be located on
the optical axis too.
Now, let a point source of monochromatic light be set at the point
A, and let us insert a prism. We find the location of the real image of
this source on plane PI caused by refraction in the prism. Let a light

-- x'

Red side Blueside


of spectrum of spectrum

Figure 4.22. The spectrum of a xenon lamp formed by two crossed prisms. The
shape of the spectrum gives the dependency n(A) in the Cartesian system x', y' . The
value of A decreases in the positive direction of the x'-axis.

Figure 4.23 Illustrating


the geometrical path of
a ray emitted by a po int
source at A . After refrac-
tion by the prism and by
the lens a real image BI of
the source appears within
plane PI .
Light Waves in Media 133
Objective Quartz Horizontal
Condenser y'
lens cell slit

V~rtical / R y
slit Sodium U Objective
vapor Heat source

Figure 4.24. A setup for observation of the absorption and dispersion behaviour
around the D- lines of sodium, using a sodium vapor prism .

ray, leaving from A along the optical axis, be refracted by the prism in
the direction CD . The extension of the short line CD intersects plane
P at point B, which is a virtual image of point A formed by the prism .
In turn, a straight line drawn from B through the center of the lens
will intersect PI at a point BI . Therefore BI is the real image of the
monochromatic source located at A.
If a source of quasi-monochromatic light is set at point A, a spectrum
of the source will appear within PI . In the case of a point source , the
spectrum is a colored straight line, which runs from B" to BI, provided
that B" and BI are associated with the red and violet limiting wave-
lengths of visible light provided by this source . Now it is easy to take
into account the effect of the second prism crossed with the first one.

4.2.2 The Wood experiment


The method of crossed prisms was modified by WOOD for investigating
the behavior of the refractive index close to an absorption line. As an
absorbing medium WOOD used sodium vapor, which has two strong

D · lines


y'

t Figure 4-25 The behav-


ior of the continuous spec-
trum around the D-lines
of sodium. Wavelength in-
creases from the right-hand
side of the spectrum to the
left. The resolving power
of the prism does not allow
Yellow side Green side separation of the doublet of
of spectrum of spectrum the sodium D-lines.
134 DEMONSTRATIONAL OPTICS

sharp absorbing spectral lines (called the D-lines of sodium at )'1 = 589.0
nm and A2 = 589.6 nm). Metal vapors are commonly used in atomic
physics experiments since they provide strong absorption lines in the
visible region of the spectrum.
The setup for observation of the dependency of the refractive index
on frequency near the absorption lines of sodium is shown in Fig. 4.24.
A xenon lamp , a bright source of light with a continuous spectrum, is
used . With the help of a condenser lens it illuminats a narrow vertical
slit. By the first objective, an image of this slit is produced in the plane
of a narrow horizontal slit mounted at a considerable distance from the
vertical one. A second objective provides an image of the horizontal slit
on the screen xt, yl. If a glass prism is placed after the vertical slit, a
spectrum of the lamp is formed on the screen. An evacuated tube of
suitable size ('" 70 em length, '" 5 em in diameter) acts like a prism
with a horizontal base when it contains a column of sodium vapor with
vertically varying density. For generating this vapor, a small amount of
metallic sodium is placed in the middle of the tube before its evacuation.
By heating the bottom of the tube and keeping the top of the tube at a
low temperature, a vertical column of sodium vapor of varying density
appears.
The inhomogeneous heating causes the vapor density to be higher at
the bottom of the tube and smaller near the top of the tube. This column
of sodium vapor refracts the light rays like a prism with a horizontal edge,
therefore a curved spectrum just like that in NEWTON's experiment can
be observed. The spectrum around the D-lines (which are not resolved)
is shown in Fig.4.25. In the vicinity of the D-lines all light is absorbed by
the metal vapor (dark band), and the refractive index increases and then
jumps to another value on the other side of the absorption region. The
colored spectral band is strongly bent, but its slope always has the same
sign (normal dispersion). This spectral curve represents the behavior
of the refractive index with respect to the variation of the wavelength.
When using a spectral device with more dispersion (e.g. a set of several
prisms instead of the one shown in Fig. 4.23), a small region of normal
dispersion would then appear between the resolved D-lines.

4.3 A wave train in a medium


4.3.1 Group velocity
As we have seen above, domain the refractive index may be larger or
smaller than unity, for instance n(w) > 1 in the normal dispersion, but
n(w) < 1 for the frequency region with anomalous dispersion. Therefore
in the latter case the phase velocity v can exceed c (v > c). This fact
Light Waves in Media 135
seems to contradict the basic statement of relativistic theory concerning
the upper limit of velocity as c.
Nevertheless when examining this problem more closely this contra-
diction disappears. The concept of the phase velocity as a characteristic
quantity of a monochromatic wave is based on MAXWELL's electromag-
netic theory, where the monochromatic wave is a special case of the
solution of the wave equation. The case of such a wave is never realized
in nature.
However, monochromatic waves are a useful tool to represent a real
wave with a finite net energy by a superposition of monochromatic waves
with a certain frequency distribution. Whereas the phase velocities of
the monochromatic components in a medium satisfy relation (4.9), the
velocity of the wave package formed by the sum of all superimposed
waves is never larger than c. We shall now consider a way of defining
the velocity of waves having a net finite energy in a medium where
dispersion is taking place. This wave is represented by a superposition
of monochromatic waves.
We assume that two monochromatic waves of identical amplitudes
of the same linear polarization and of a small frequency difference are
superimposed in a medium having normal dispersion :
(4.54)
For simplicity, we consider the initial phases of the waves to be equal to
zero.
Let Wo be the mean frequency and 8w be the frequency difference
between Wo and the frequencies of the monochromatic components WI ,
W2:
Wo = (WI + W2)/2, WI = Wo - 8w, w2 = Wo + 8w, 8w «wo
Let >'0, >'1 and >'2 be the wavelengths associated with these frequencies.
Now ko = 2rr/>.o, k1 = 2rr/>'1 and k2 = 2rr/>'2, so that for a small
difference between the propagation numbers one gets:
>'1 - >'2 >'1 - >'2
8k = (k2 - k1)/2 = rr >'1>'2 :::: rr >.~ (4.55)

Taking into account these specifications we rewrite the expression for


the superimposed wave (4.54) as follows:
E(t, z) = 2Eo cos(8wt - 8kz) cos(wot - koz) (4.56)

The z - dependency of the function E(t , z) at t = const is shown in


Fig. 4.26. Fast oscillations take place inside a slowly varying envelope,
136 DEMONSTRATIONAL OPTICS

Figure 4.26. The spatial distribution of the electric vector in a wave composed by
two monocromatic waves with .a small difference in frequencies. With propagation in
a medium , the envelope A(t) propagates with the group velocity u .

denoted by the dotted line. It is clear that the time - spatial dependency
of the envelope is given by the cosine function :

cos(c5wt - c5kz) (4.57)

One can find two neighboring zero points of the envelope from the rela-
tion
c5kfj.z = ±7f/2 , (4.58)
at t = canst, or from
(4.59)
at z = const . If t = const , .6.z is the spatial separation of points having
the same phase , whereas for z = canst, fj.t is their temporal separation.
In both cases (4.57) and (4.58) the net wave, being limited by the two
neighboring zeros, is usually called a wave group, or wave train. The
velocity of the wave group , which is denoted by u , is the velocity of the
envelope , which is called the group velocity. When fixing any point of
the envelope, the group velocity can be found from the condition

c5wt - c5kz = 0 ,
where
z c5w
u= - = - (4.60)
t 15k
In the general case, a wave group may be treated in terms of the su-
perposition of elementary waves within a narrow region of wavenumbers
k around k o associated with the carrier frequency Wo of the wave group.
Let us consider a wave group E(t, z), propagating along the z-axis.
Mathematically such a wave group is represented by summing up the
elementary monochromatic waves:

E(t,z) "" J g(k)exp{i[w(k)t - kz]}dk , (4.61)


Light Waves in Media 137

where the function g(k) describes the amplitude distribution of the ele-
mentary monochromatic waves.
We fix a certain propagation number ko and assume that the concept
of group velocity is valid, because the dependency of w on k is very weak
within a narrow region of k. This means that the expansion of w(k) near
ko,

_ dw(k) d2w(k) 2
w(k) - w(ko) + ~(k - ko) + 2dk 2 (k - ko) + ... ,
may be restricted only to its linear dependency on (k - ko) :

dw(k)
w(k) ~ w(ko) + ~(k - ko) .

For w(k) we substitute the latter expression into the exponential factor
of the wave and expand k to k = k - ko + ko . We then get

exp{i[w(k)t - kz]} = exp{i[w(ko)t - koz]} x

~(k - ko)t - z(k - ko) )]


x exp [z. (dw(k)

Thus the expression for the wave group takes the form:

E(t,z) '-"exp{i[w(ko)t - koz]} x

x Jg(k) exp [i (~~k) (k - ko)t - z(k - ko))] dk ,

where the first factor is associated with a monochromatic wave of carrier


frequency w(ko), travelling in the positive direction of the z-axis with
the phase velocity w(ko)/ko. The second factor is the amplitude of the
wave group, which consists of elementary monochromatic waves:

Every elementary monochromatic wave runs in the same direction with


the group velocity
dw(k)
u=~ .

It is seen that the phases of the elementary monochromatic waves de-


pend on the difference k - ko, rather than on k. This implies that slow
variations of the envelope A( t, z) of the wave group occur with respect
to the fast oscillations of the carrier frequency.
138 DEMONSTRATIONAL OPTICS

4.3.2 Energy transfer


Under dispersion conditions the concept of light phase velocity may
only refer to the propagation of a monochromatic wave. However, a
pure monochromatic wave does not have a beginning and an end, and
its net energy is undefined in the sense of physical measurements. Only
electromagnetic waves with a finite net energy are of interest , because
those can be associated with the real procedure of measurement . In any
method for the measurement of the velocity of light , using an interrupted
light beam, one is always measuring the group light velocity.
So in reality, the energy of a wave group - contrary to the idealized
a
example shown in Fig.4.26 - is finite quantity. This energy propagates
with the group velocity. For light waves, th e group velocity in a real
medium can never exceed the vacuum light velocity c. As we have seen
above, in a medium showing a weak dispersion, the envelope A(t, z)
propagates with the group velocity u. The intensity A2(t, z) of a wave
group is assosiated with energy transfer in the medium. Therefore the
velocity of energy transfer in a medium corresponding to the velocity
of the mean flux energy has to be equal to the group velocity. We
should note that the concept of group velocity is correct as long as the
approximation of a wave group or a quasi-monochromatic wave is valid,
and we may associate the energy transfer of light waves in media with
th e group velocity only within the limits of this approximation.

4.3.3 The Rayleigh formula


When dispersion takes place, the phase velocity will be a function of
the wavelength v = V(A) . An example in which the effects of dispersion
can be studied quite well is the dependency of the velocity of waves
propagating on the surface of water .
Every small surface element of water oscillates along verti cally. Mo-
tion of this element gives rise to an oscillating motion of its neighboring
surface elements due to the coupling of the water molecules. The mo-
tions of these elements, in turn, cause vertical displacements of the other
neighboring surface elements. The oscillations of the last elements have
a phase shift with respect to the oscillations of the first selected element ,
therefore a surface wave propagating along the water's surface is created.
If the depth of the water is greater than the longest possible wave-
length, two principle factors are responsible for the formation of a wave.
The first is the size of the initial deformation of the water surface, since
the energy of the water elements increase with increasing initial ampli-
tude. A detailed analysis shows the energy of a unit mass element of
the water's surface, which is called the surface energy, to be expressed in
terms of the the capillary constant (J, the wave number k and the water
Light Waves in Media 139

Figure 4-27. The dispersion curve v(>.) associated with waves propagating on the
surface of water . The function v(>.) has its minimum at Vo = 23 cmls (under normal
conditions) and >'0 = 1.7 em. The region of wavelenths between >'0 and>. = 0
corresponds to anomalous dispersion, the region between >'0 and >. -+ 00 to normal
dispersion.

density p. This part of the surface energy of the unit mass element is
given by the expression kCJ/ p. Therefore, the larger the deformation is,
the higher the surface energy. The second part of this energy is formed
by gravity. The unit mass element of the water surface possesses a higher
energy depending on the gravitational acceleration 9 and the wavelength
>.. Hence, this part of the surface energy is directly propotional to 9 and
A. Therefore the relationship between the phase velocity v and the prop-
agation number k = 21T/ A is given by the expression

v2 = fL + k~ (4.63)
k P
The function V(A) is shown in Fig. 4.27. Under normal conditions the
minimal velocity is v ~ 23 cui]« when Ao = 1.7 em. If >. < 1.7 em, the
so-called capillary waves are observed . The phase velocity of this type
of wave increases extensively with a decrease in the wavelength. In the
other case, when>. » 1.7 em, gravity is the dominant factor for the
appearance of surface waves, which are called gravity waves. One can
observe gravity waves with a wavelength up to >. "" 3/4 km, a period of
"" 23 s and a phase velocity of "" 120 km per hour.
Now we will obtain a relationship between the phase velocity v and
the group velocity u. From the relationship between the propagation
number and the angular frequency, k- v ="W, we find
kbv + v8k = 8w
140 DEMONSTRATIONAL OPTICS

and using (4.60) we can write:


8w 8v
u = 8k = v + k 8k (4.64)

Now k); = 271"; and therefore


)"8k + k8)" = 0 ,

or 8k = -k8)../)... On substituting the last equation into (4.64) we obtain


8v
u=v-)..- (4.65)
8)"

which is called the RAYLEIGH formula.


We see that the group velocity in the studied case is larger or smaller
than the phase velocity, depending on the sign of the second term in
(4.65). This relationship provides a graphical method to find the group
velocity. Let )..1 < )..0, as shown in Fig.4.27, and VI be the phase velocity
associated with )..1. The tangent to the curve v()..), expressed by dv/d)",
at the wavelength)" = )..1 intersects the ordinate axis at V2, where V2 =
VI - )..ldv/d)" at dv/d)" < O. Therefore the group velocity is given by
u = V2 > VI. In the case of gravity waves, for example at wavelength )..2
with phase velocity V3, the tangent to the curve intersects the ordinate
axis at V4. Here, V4 = V3 - >"ldvjd>" at dvjd>" > 0, therefore the group
velocity associated with the gravity waves is u = V4 < V3 .
Assuming X « 1.7 cm, the contribution of the gravity factor to v 2 in
(4.63) is negligible and it is possible to estimate
2 ak
v::::::-
p
Now applying the RAYLEIGH formula , we find u = (3/2)v. In the case
of large wavelengths, where gravity is dominant, we have

v2 :::::: ft
k
and the RAYLEIGH formula takes the form u = (1/2)v.
4.3.4 Simulation of the motion of wave trains
In practice it is difficult to realize convenient conditions for observ-
ing the motion of wave trains on the surface of water. Nevertheless,
this phenomenon is unique in showing both normal and anomalous dis-
persion. By using the constants g, p and a in (4.63) we may create a
computational experiment describing the propagation of wave trains on
the surface of water.
Light Waves in Media 141

Figure 4.28. Images produced by the computing technique described in the text
illustrates the motion of wave groups on the surface of water. The initial position
of every wave group is marked by the point A on the z- axis, which corresponds to
th e centers of the wave groups at t = O. Points B,a,n correspond to the centers of
th e wave groups after the time interval At = 8.5 s. The wave group positioned at
point a moved without dispersion; the wave group arriving at B moved under normal
dispersion; and the group arriving point D moved under anomalous dispersion. The
white strips indicate maxima of the amplitudes within every wave group.

First let us consider the problem of the propagation of a wave train in


the case of no dispersion. According to the dispersion curve in FigA.27
this case is realized in a small region around>. = >'0 = 1.7 ern where
the tangent to the curve v(>.) is parallel to the abscissa, so that u =
Vo = 23 csa]«. Therefore, one may regard monochromatic waves within
this region to have approximately the same phase velocity, hence a wave
group of these monochromatic waves will move with the same group
velocity u = vo.
The spatial distribution of the amplitude of the wave group a(z, t) at
a fixed moment t may be represented by the following sum of 121 waves
having slightly different wavelengths distributed around >'0 :

a(z, t) rv ~
120
cos [(2
)..: + a m4~060) (tvo - z + 50)] (4.66)

v
'-....2.- 1 .
-- - - -~
I I I Figure 4.29 The region of
I I I
I I I wavelengths between A1
I I I
I I I and A2 is considered to
I I I
I I I form a wave group, propa-
I I I
I I gating without dispersion.
142 DEMONSTRATIONAL OPTICS

Vo -- --~ I I I
J I I Figure 4-90 The region of
I I I
I I I wavelengths between A1
I J I
J I I and A2 for forming awave
I I J
I I J group , which propagates
I I : under conditions of normal
dispersion .

v
2
Vo -- - -~
I I I
I I I Figure 4-91 The region of
I I I
I I I wavelengths between A1
I
I
I
I
I
I and A2 for forming a wave
I
I
I
I
I
I
group, which propagates
I I I under conditions of anom-
alous dispersion.

where Q: = 0.25, 400 is the net number of points z taken for the calcu-
lation and 50 specifies the initial position of the wave group at t = O.
Summing up the monochromatic components is similar to moving from
point 1 to point 2 in Fig.4.29. We convert the distribution a( z , 0) into
a set of values, each providing a level of a gray scale, and create a pic-
t ure showing the spatial distribution of the amplitude. The distribution
a(z , t) at t = 0 is shown in Fig.4.28, where the point A specifies the cen-
ter of the wave group. In a similar way, the distribution a(z, t) for the
time t = 8.5 s is calculated. This distribution is also shown in Fig.4.28,
and the center of the wave group is marked by C. If the wave group
propagates between points A and C with the velocity u = Vo = 23 cui]«
the distance AC = 23 x 8.5 = 195.5 em.
To illustrate the case of normal dispersion we take into account the
dependency of the phase velocity on the wavelength and choose a new
region for variation of the wavelength (Fig.4.30), where the wavelengths
are now distributed around A2 = 1.8 em. The spatial distribution of the
amplitude is given by the expression

120
a(z, t) '" . ; cos
[(2A: - Q:
60) {t(Vl +
m~O f3 * m) - z
]
+ 50}

where A2 = 1.8 em, Q: = 0.25, and the constant f3 = 0.0015 provides


the dependency of the phase velocity on A, v = V(A). When increasing
the wavelength , the propagation number of m-th monochromatic com-
Light Waves in Media 143

ponent, 27f/A - a(m - 60)/400, decreases , whereas the phase velocity of


the same component increases with VQ + f3 * m .
The final position of the wave group is shown in Fig.4.28, where the
center of the wave group is marked by B . During the same time t = 8.5
s, this wave group shows up at a distance from the initial point A of 124
em, which provides for a group velocity of u = 124/8.5 ~ 14.6 ern / s.
In turn, the magnitude of the phase velocity VQ calculated from (4.63) at
A2 = 1.8 em is still approximately equal to VQ = 23 cm/s; hence u < VQ .
In the case of anomalous dispersion, the calculation was carried out
for wavelengths located around Al = 1.6 em, as shown in Fig.4 .31. The
wave group composed of monochromatic components within this region
of wavelengths has a spatial distribution of amplitude a(z, t) described
by the following expression:

a(z,') ~; cas [G: + ;;'0)


Q {'(va + I), m) - z + 50}]
where A = 1.6 cm, a = 0.25, f3 = 0.0015 and VQ ~ 23 cui]«. The positive
sign in the factor 27f/ A+ am/ 400 provides an increase in the propagation
number with decreasing wavelength. After the same time t = 8.5 em the
center of the wave group is at point D in Fig.4.28. The distance AD is
284 em, which gives for the group velocity u = 284/8.5 ~ 33 em, so that
u > VQ .
5. Radiation under uniform charge motion
The principle concept of electrodynamics is that radiation of electro-
magnetic waves always occurs during accelerated or decelerated motion
of electrically charged particles. Nevertheless , a number of optical phe-
nomena of light wave radiation exist under very fast uniform motion of
a charged particle in an optically transparent media. The mechanism of
radiation is principally based on the interaction of the electromagnetic
field of the particle with the medium. A qualitative analysis of such
processes allows a simple interpretation in the frame of classical electro-
dynamics, exploiting the phenomenon of light wave superposition known
as the HUYGENS-FRESNEL principle.

5.1 Vavilov-Cherenkov radiation


Radiation of an unknown nature was discovered by the Russian physi-
cists P .CHERENKOV and C.VAVILOV while studying the luminescence of
liquids under , -ray illumination [26, 27] .
Later on this phenomenon, now called VAVILOV-CHERENKOV radia-
tion , was the subject of extensive studies to discover the acting mecha-
144 DEMONSTRATIONAL OPTICS

nisms since the properties of this radiation did not correspond to any
known optical process of luminescence in liquid or solid bodies and did
not even correspond to the radiation of decelerated charged particles.
One of the most thoroughly studied subjects at the time of this discov-
ery was the phenomenon of luminescence, which was understood as light
emission by the microparticles of material substances, atoms or mole-
cules. Experimental and theoretical investigations, as well as known
laws of classical and quantum mechanics, allowed the classification of
different types of luminescence in terms of the characteristic radiation
time by an atom or a molecule. This effective radiation time depends
on the type of particles and on the excitation conditions. All the known
experimental methods of measurements for the radiation time lead to
the conclusion that VAVILOV-CHERENKOV radiation was of a different
nature. This conclusion was supported by the experimental fact that
th e intensity of VAVILOV-CHERENKOV radiation was weakly dependent
on the kind of liquid. All experiments showed weak blue emission of
VAVILOV-CHERENKOV radiation, which had almost the same intensity
for all liquids , and was dependent only on the incident , - ray intensity.
As a possible mechanism of this phenomenon, VAVILOV proposed the
so-called deceleration radiation. Under the influence of , - rays, fast
free electrons are set free from atomic particles and move through the
liquid. As a result of the electrostatic interaction with the surrounding
at oms or molecules, these fast electrons are quickly decelerated. As it is
known, accelerated or decelerated motion of a charged particle creates
radiat ion. Nevertheless, further studies of the intensity and the spectrum
of the VAVILOV-CHERENKOV radiation showed that the relative intensity
of the deceleration radiation in the optical part of the spectrum is about
100 times lower than the experimentally recorded intensity of this new
typ e of radiation.
Later, a rigorous theory of the VAVILOV-CHERENKOV phenomenon
was created by LFRANCK and LTAMM [28]. According to this theory
t he cause of VAVILOV-CHERENKOV radiation is an electron moving in
th e medium with a very high velocity exceeding the phase velocity of
light in this medium. Such fast electrons leave the atoms of the liq-
uid after absorption of hard "t: radiation in the VAVILOV-CHERENKOV
experiments.
Let us discuss the mechanism of VAVILOV-Chrerenkov radiation at
a qualitative level. Suppose that a charged particle is moving in an
optically transparent medium with the velocity v q exceeding the phase
velocity of light in this medium:
c
vq > v =-
n
Light Waves in Media 145

Figure 4.32. The polarization effect on the particles of a medium caused by a negative
charge moving with relativistic velocity v. The polarized particles located along the
trajectory of the charge become coherent sources of secondary spherical waves. The
superposition of the secondary waves is the reason why coherent Vavilov-Cherenkov
radiation is emitted by the medium.

where n is the refraction index of the medium.


The energy of the moving particle is very high, therefore it has to
be treated as a relativistic particle. Moreover, its energy considerably
exceeds the excitation energy of the atoms of the medium. In other
words , the portion of the kinetic energy lost by the particle to the exci-
tation of atoms of the medium and another portion of energy, which is
transformed into VAVILOV-CHERENKOV radiation, are extremely small
in comparison with its total energy. For this reason , one may assume
that the particle is moving in the medium with constant velocity v q . The
electromagnetic field of the moving charged particle excites atoms of the
medium located along the trajectory of this particle (FigA.32) . These
atoms become coherent sources of secondary waves. A pictorial scheme
of the successive positions of the particle along its trajectory and the
wave fronts of the secondary waves, represented by spheres of decreasing
radii are shown in FigA.33.
The envelope of these wavelets gives the position of the front of the
resulting wave or, effectively, of the VAVILOV-CHERENKOV radiation. As
it can be seen from this figure, the resulting wave front propagates in
a direction oriented at an angle () with respect to the velocity vector
of the particle. The spherical fronts of the secondary waves create a
resulting field in space, the front of which propagates within a cone with
an opening angle equal to (). Let a particle moving along its trajectory
during the time interval ~t start at the moment to from the point xo.
During the same time span the light wave reaches the position of the wave
front emerging from the point xo. Through geometrical considerations
146 DEMONSTRATIONAL OPTICS

Figure 4.99. Initiation of a wave front of Vavilov-Cherenkov radiation by a moving


charge q. According with the Huygens-Fresnel principle, secondary coherent spheri-
cal waves emitted by points along the charge's trajectory form a conical wave front
propagating at the angle 9 with respect to the direction of charge mot ion.

it is easy to find the following relationship between the velocities of light


and of the particle and the angle ():
v c
cos()=-=- (4.67)
vq nV q

This simple result is obtained under the supposition, that the moving
charge creates a secondary wave at every point along its trajectory and
that the resulting field arises by superposition of these secondary waves
while accounting for their phases.
A similar mechanism of sound wave creation takes place when a body
is moving in a medium with a velocity exceeding the velocity of sound.
The well-known effect of shock wave generation, when an airplane ex-
ceeds the sound barrier, has a similar physical nature (secondary wave
creation by the flying airplane) to the generation of an "opt ical shock
wave" by the relativistic charged particle flying through a medium. In
the acoustic case the shock wave front is located at the surface of the
MACH cone and propagates at the angle () with respect to the airplane
trajectory. The opening angle of the MACH cone () may be determined
from the expression analogously to (4.67):
V
cos() = - s
VI

where VI is the velocity of the airplane after exceeding the sound barrier,
V s is the velocity of sound in the air, (VI> Vs).
CHERENKOV radiation is used by a number of devices known as CHE-
RENKOV counters for the registration of fast charged particles. In order
Light Waves in Media 147

Photo
-t--- multipliers

Fluoresc ing glass 4~~~~~-1 Opaque screen


Organic glass - ~

Opaque housing

Figure 4-34. Scheme of a Cherenkov counter for detecting cosmic particles . Two
identical photomultiplies, both having a photocathode diameter of 16 em, are arranged
facing each other vertically. A light flash in the fluorescing glass is detected by the
first photomultipier and the Vavilov-Cherenkov radiation in the organic glass by the
second one.

to demonstrate quantitative parameters of CHERENKOV radiation, let us


discuss the operation of a cosmic radiation counter. A primary cosmic
radiation exists, which falls on the Earth's atmosphere consisting mainly
of protons with an energy £p between 5 and 10 GeV. In passing through
the Earth's atmosphere, this primary cosmic radiation provoke nuclear
reactions resulting in muons or j.L-mesons, which have the charge of an
electron but possess about 200 electron masses. Apart from muons, hard
')'-quant a also appear, however the hard ')'-quanta are affected consider-
ably by absorption in the upper layers of the atmosphere. Therefore
the main component of the secondary cosmic radiation near the Earth
surface is a flux of highly energetic muons. Using the relativistic formula
for the energy of a particle
1
£ = £0 ---:::::==
J1-~
we estimate the muon velocity assuming £ = 10 GeV and £0 = 100 MeV.
From this relativistic formula we obtain, that

and v ~ 0.999 c which means that these muons have to be treated as


relativistic particles.
148 DEMONSTRATIONAL OPTICS

An ordinary CHERENKOV counter is comprised of a block of transpar-


ent material having a thickness of several centimeters (Fig.4.34) . Special
measurements showed that a relativistic charged particle, flying through
a glass block generates about 100 optical photons of CHERENKOV radi-
at ion for a path of 1 em. The counter described in this figure consists of
organic glass of 16 em diameter, 6 em thick; therefore, while a particle
flies through it, a flash of CHERENKOV radiation of about 600 photons is
formed. These light flashes are recorded by the photomultiplier, which
transforms it into an electric pulse. The flux of muons is very weak: one
particle runs through the radiator every 1-2 minutes. For more reliable
detection of such rare events, a fluorescing glass plate together with a
first photomultiplier is located before the organic glass block. A particle
moving through the fluorescing glass provides a light flash detected by
this photomultiplier. This flash can not be detected by the second pho-
to detector, because an opaque screen is mounted between the fluorescing
glass and the organic glass block. Thereby, only a particle passing both
successive detecting units generates two practically synchronized pulses
at the exit of the photomultipliers. This pair of pulses creates the exit
pulse of a radiotechnique coincidence scheme. This registration tech-
nique is used in the majority of optical devices for registration of weak
flows of particles or photons and allows an increase in the reliability of
counts of CHERENKOV radiation.

5.2 Transient radiation


The principle condition for CHERENKOV radiation is that a charged
particle is moving in a medium with a velocity exceeding the phase veloc-
ity of light: v q > cfn , However, under certain circumstances, emission of
radiation is also possible if a charged particle is moving with a constant
velocity lower than the phase velocity of light. The principle condition
for the creation of such a type of radiation is a variation of the refractive
index of the medium or the phase velocity of light along the trajectory
of the particle. This situation is realized when a moving charge crosses
the boundary between two media with different refractive indices. The
possibility of such transient radiation was analyzed for the first time
by by GINZBURG and FRANCK [29]. This radiation is generated most
effectively when crossing of a vacuum-metal boundary. In order to ex-
plain transient radiation let us use the electric reflection method. In
this method we substitute the metal boundary with an image plane and
project an image of the moving charge as shown in Fig.4.35 . In the
region left of this plane a charged particle is moving toward the plane .
In the region right of this plane, the motion of the image of the charged
Light Waves in Media 149

+q -q
---o-.... ......::::;j~-"-""O
-v
Figure 4.35 A positive
charge moving towards
a flat infinite metallic
boundary will give rise to
transient radiation when
penetrating the metal.

+q

Figure 4,36. A set of dielectric flat plates may cause an appreciable amount of tran-
sient radiation.

particle is also directed towards the plane. The sign of the image of the
particle is opposite to the sign of the real particle.
Therefore , the problem of the motion of a charged particle towards the
boundary is reduced to the equivalent problem of an electric dipole with
a variable dipole moment (FigA.35). Once the charged particle crosses
the metal boundary, both the image charge and the dipole disappear.
It is this disappearance of the dipole moment that creates the transient
radiation. Apart from the transient radiation an ordinary deceleration
radiation exists as well. As the transient radiation is characterized by
several distinguishing features, it is possible to separate it from the back-
ground of the deceleration radiation.
Transient radiation arising from the crossing of a transparent dielectric
boundary is rather weak. For this reason a set of immediately adjacent
transparent plates is used (FigA.36). Transient radiation becomes more
concentrated in a specific direction when the energy of the charged parti-
150 DEMONSTRATIONAL OPTICS

e- Ve
11?A~ B
----'+6'..,1 :, 12 ,------=i1
'--_...J
e+ I
6

Figure 4.37. A moving electron e" and its image e+ caused by the flat surface of a
diffraction grid. Electron and image compose an electric dipole of the length h . The
length of the dipole increases to b if the electron moves along a depth of the grid .
This change of the electric dipo le is a reason for production of visible light waves.

de is increased. It is spatially reduced to a small solid angle propagating


out from the boundary of two media.

5.3 Smith-Parsell effect


Special experiments in order to observe radiation created by an elec-
tron moving along the surface of a metallic diffraction grating were car-
ried out by SMITH and PARSELL in 1953. The periodic structure of
the diffraction grating plays the role of a boundary with a periodically
varying profile along the trajectory of electron motion.
In order to explain this effect we imagine two planes : the first one is
tangent to the surface of the grating and the second one, parallel to the
first, is located at the lower depth of the groove (Fig.4 .37).
An electron mirrored by these planes leads to the creation of its charge
image . The image in the first plane creates a dipole moment ell and in
the second plane elz , where iz > ll ' In this manner the electron mo-
tion over the grooves provokes abrupt periodical changes of the induced
electric dipole moment between el2 and eh. These jumps in the dipole
moment lead to secondary waves, propagating out of the surface of the
grating (Fig.4.38) .
We can find a relation, allowing to determine the wavelength of the
radiation propagating at an angle 0 with respect to the surface of the
grating (Fig.4.39). At the moment to the electron is located at point A
at the beginning of a groove. Moving with the velocity V e the electron
will pass point B (beginning of the next groove) after the time span
7 = diVe, where d is the period of the grating, or the distance between
the adjacent grooves. The light perturbation, provoked by the jump of
the dipole moment at point A will propagate along the direction AS
for the short length AG during the time span 71 = d cos 0I c. Light
perturbation emerging at point B at a moment when the electron passes
this point, in turn, reaches point G by the shortest path BG. Light
perturbations originating from points A and B superimpose at point
Light Waves in Media 151

Figure 4.38 Initiation


of radiation emitted by
P (t+xlv)
a moving electron when
passing a rectangular step.

C. As a result, the light field will interfere constructively if the time


difference T - Tl is equal to an integral number m of periods of the light
wave (T = 1/1/ = A/C):

A
T-Tl = mT=m-
C

Substituting in this relation T = dive and Tl = d cos ()/ c, we obtain the


following condition:

m); = d (~ - cos ()) (4.68)

While the electron is moving to the next groove of the grating the
next pair of wave perturbations arise in the same way. The amplifica-
tion condition for these waves in the direction AS is the same. Hence,

5
c

AI--_...J
1---- d

Figure 4.39. Two steps located at the points A and B. Both points emit spherical
waves. A superposition of these waves will give rise to a maximum of radiation of A
at the angle 0, if rnA = d(c/v e - cos 0), where V e is the electron velocity and m is an
even number.
152 DEMONSTRATIONAL OPTICS

while moving over the grooves of the grating the electron excites an in-
creasingly amplified wave, propagating at angle () with respect to the
surface of the grating.
To estimate the wavelength of the radiation we use the parameters of
the PARSELL and SMITH experiment: an electron energy of 300 keY and
a grating with a groove distance of d = 1.67 um. Using the relativistic
formula for the kinetic energy of a moving electron:

£ = £0 _ £0
J1-~
we determine the ratio velc for an electron with the energy 300 keY.
The total energy of an electron at rest is £0 = 511 keY. Substituting
this value into the last formula we find vel c ~ 0.8. Referring to formula
(4.68) we see that radiation in the visible spectral range is possible at
small angles () for m = 1. For example, for () = 20° and cos () ~ 0.94,
cju; - cos () is close to 0.3. So, for a grating of period 1.67 um, the
wavelength of the generated light is estimated to be A ~ 0.5jlm. This
radiation corresponds to the green part of the visible spectrum.
Summing up these phenomena, we note that an occurrence of radia-
tion is possible even for uniform motion of charged particles, if it takes
place either inside a material medium or when the particle is crossing the
boundary between two media with different refraction indices . The qual-
itative interpretation of these phenomena lead to the fact that uniform
motion of a charged particle creates secondary waves inside a medium.
These secondary waves will superimpose according to the equi-phase
condition, because they are generated by the uniform linear motion of a
single charge . It is important to not e, that the emergence of radiation is
strict ly connected to the motion of one charged particle perturbing the
medium in the same way over its trajectory. In cases where an ensemble
of charged particles is moving, with a certain velocity spread, such as an
electron beam, these effects disappear.

SUMMARY

Optical phenomena arising when light propagates inside a homoge-


neous transparent medium may be explained by the interaction of the
initial wave with atomic oscillators of the medium. Superposition of the
initial wave with the wavelets emitted by these atomic oscillators forms
the resulting light wave propagating inside the matter. This wave is
characterized by a constant refractive index far from an absorption line.
Light Waves in Media 153
Excitation of such a light wave inside a material medium caused by
light, penetrating into the medium through a plane boundary, depends
essentially on the incident angle and the polarization direction of the
incoming light wave. FRESNEL'S formulae and its consequences are ap-
plicable in this case.
Propagation of light becomes more complicated when the oscillation
frequency of the incident light wave tends to the resonant frequency of
the atomic oscillators . The energy absorbed by the oscillators increases
and anomalous dispersion takes place. Under strong dispersion of the
medium, a difference between the phase and the group velocity of light
waves becomes noticeable. The group velocity, which is simultaneously
the velocity of energy transport, can not exceed the velocity of light in
vacuum.
The velocity of a macroscopic particle can never exceed the velocity of
light in vacuum, but it may be higher than the phase velocity of light in-
side a medium. Motion of a charged particle in this velocity range causes
VAVILOV - CHERENKOV radiation. Distinct from the radiation from an
accelerated charge, the electromagnetic wave of VAVILOV - CHERENKOV
radiation is the result of the superposition of secondary waves emitted
by atomic oscillators excited by a steadily moving charge.

PROBLEMS

4.1. The definition of a refractive index is based on the assumption


that a light wave penetrating into a transparent dielectric substance
gives rise to oscillations of atomic oscillators. Secondary waves emitted
by these oscillators sum together to form a new wave, which propagates
in the medium with a phase velocity v, thus v = cfn. Let us assume
that a plane wave falls normally on a thin parallel glass plate. Derive
the expression for the wave transmitted through the plate and discuss the
obtained results for the particular case d « A, where d is the thickness
of the plate and A is the wavelength of the incident light wave.

4.2. A plane monochromatic wave, represented by a stream of pho-


tons, falls on a flat surface of a dielectric media. Derive SNELL'S law,
assuming that every photon carries the momentum Iik. Take into con-
sideration the refractive indices of nl and n2 of the two media and () as
the angle of incidence (Fig.4.40).

4.3. A light wave falls on the flat surface of a dielectric material with
nearly grazing incidence (Fig.4.41). Prove that the oscillations of the
154 DEMONSTRATIONAL OPTICS

electric vector of the reflected wave will have a phase shift of 7r with
respect to the incident wave.

4.4 A linearly polarized wave falls on a flat boundary of a dielectric


material (Fig.4.42). The electric vector of the wave E deviates from the
plane of incidence by the angle ai , so that tan( ai) = Ef) / E~i) .
Introduce similar angles for reflected and transmitted waves and show
that the original linear polarization undergoes rotation after refraction
and reflection.

4.5. Crystalline quartz has refractive indices of n1 = 1.547 and n2 =


1.557 at wavelengths Al = 550 nm and A2 = 410 nm, respectively. Use

Figure 4.40.
I
I

:; ) / ) ) .) / ) > ;::) T; ) ) ) 4:;


I

~ 7; )
I
I

Figure 4.41.

Figure 4.42.
Light Waves in Media 155

Figure 4.43.

the CAUCHY formula n = A + B/A 2 and calculate the refractive index


n3 at A = 610 nm.

4.6. A muon of energy E.p. = 0.5 GeV propagates within dense barium
crown glass. Due to dispersion, CHERENKOV radiation is grouped within
a narrow range of conical surfaces. The deflection of rays within the
visible range of wavelengths may be taken into account by a difference
in the angles () of the CHERENKOV radiation. Estimate the difference
fj.(} occurring between blue (AI = 486 nm) and red rays (Ared = 659
nm) provided that the refractive indices are nl = 1.615 and n2 = 1.605,
respectively.

= 5 Gev, moving within glass, gives rise to


4.7. A proton of energy E.p
CHERENKOV radiation. Due to the dispersion of the glass, light waves
of different wavelengths are deflected at different angles with respect to
proton's trajectory.
This implies that the radiation propagates within a solid angle fj.O
(FigAA3). Derive the expression for light power of CHERENKOV radia-
tion within this solid angle. Estimate the power for the case of flint glass.
It may be approximated that all radiation is emitted within the blue -
red region of wavelength (Ablue = 486 nm, nblue = 1.575, Ared = 659 nm,
nred = 1.555). Use the fact that the mean number of quanta TJ emitted
over 1 cm of the trajectory is approximately equal to 100.
156 DEMONSTRATIONAL OPTICS

",

Figure 4.44.

SOLUTIONS

4.1. We write the expression for the complex amplitude of the electric
vector in the incident wave as follows:

E, = Eo exp[i(wt - kz + <p)]
The glass plate introduces the phase difference t::.<p = kd(n - 1) thus
the electric field occurring after the plate, may be presented by the
expression:

Et = Eo exp[i(wt - kz + ip - kd(n - 1))] =

= Eo exp[i(wt - kz + <p)] exp[-ikd(n - 1)]


(reflection on both sides of the plate has been neglected). If this phase
t::.<p is rather small , the phase factor can approximately be written as
(exp[i(wt - kz + <pm ~ (1 - ikd(n - 1)) . Further, using the relation
- i = exp( - i7r /2) we write the field of the transmitted wave:

E t = Eo exp[i(wt - kz + <p)] + Eokd(n - 1) exp[i(wt - kz + <p - 7r /2)]

It can be seen that the first term of the transmitted wave is the same as
the incident one, whereas the second term may be treated as an action
of the atomic oscillators of the glass plate. In other words, the field
Eokd(n - 1) exp[i(wt - kz + sp - 7r/2)] emitted by the glass plate has
a very small amplitude compared with that of the incident wave and a
phase shift of 7r/2.

4.2. Specifying the photon momenta in the incident wave as pi = lik i ,


orthogonal projections of each momentum on the horizontal axis x and
the vertical axis y will be p~ = lik~ and P~ = !ik~ , respectively. We
know from discussing light pressure forces that the y-component of the
photon momentum has to change at the boundary of media and we
assume that the x-component is unchanged. This statement may be
Light Waves in Media 157

Figure 4.45.

expressed as p~ = p;, where p; is the x- component of the momentum


for a transmitted photon (Fig.4.44). With the angles of incidence Oi and
refraction ot we rewrite p~ = p; in the following form:
pi sin Oi = pt sin ot

With pi = tiki and pt = tik t , we get the relations for the transformations
of propagation number k and wavelength A:

l'Oi l'Ot
Ai sm = At sm

Since the frequency of the wave v is unchanged while the photons prop-
agate through the surface of the media, after multiplying both sides of
the last expression by civ, we can write SNELL'S law:

4.3. We use expression (4.23) (FRESNEL'S formulae) for the ampli-


tudes of the components EI~r) and EY) of the electric vector of the trans-
mitted wave to relate them to the magnitudes of the incident wave E~i)
and E~) :

E(r) _ n2 cos Oi - nl cos Ot E(i) E(r) _ nl cos 0i - n2 cos Ot E(i)


II - n2 cos Oi + nl cos Ot II .L - nl cos Oi + n2 cos Ot .L
where n2 is the refractive index of the dielectric material and nl is the
index for air, hence n2 > nl (FigAA5).
Under the condition of grazing incidence the angle of incidence Oi can
have very small values, whereas the angle Ot of the transmitted wave
always has a finite magnitude, because, in this case, the incident wave
is reflected by a denser medium . Therefore, in the first expression, the
158 DEMONSTRATIONAL OPTICS

Figure 4.46.

term n2 cos ()j tends to zero at ()j -+ 7r /2, whereas the term nl cos ()t
remains a finite value. The result for E~r) becomes E~r) -+ -E~j). In
a similar way EY) -+ - Er) at ()j -+ 7r /2. Hence, for grazing incidence
both components of the reflected wave obtain a change in phase of 7r
with respect to the appropriate components of the incident wave.

4.4 We use the FRESNEL formulae (4.23) to represent the angles at


and a r of the electric vectors of the transmitted and reflected waves in
terms of the angle a j (Fig.4.46):

t) =cos(()j-()t) t) =cos(()j-()t}tanaj
E(t) E(j )
tanat=
Ell Ell
It can be seen that for the refracted, transmitted wave, the rotation of
the electric vector occurs towards the plane of incidence with respect
to the position of the electric vector in the incident wave, because the
factor cos(()j - ()t} leads to a decrease of the value of tan at. In turn, in
the case ofthe reflected wave the term cos(()j - ()t)/ cos(()j + ()t} increases
with tan ar , which means the absolute value of the angle a r increases
with respect to the angle aj of th e incident wave. The rotation of the
polarization direction occurs away from the plane of incidence.

4.5. For the given refractive indices at two wavelengths one can esti-
mate the constant B in the Cauchy formula . From n l = A + B/>..i and
Light Waves in Media 159

n2 = A + B/>'~ for B we get:

1 1) -1 >.21 >.2
B = (n2 -nl) ( >.2 - >.2 = (n2 -nl) >.2 ~2
2 1 1- 2

Calculations give B = 0.378 . 10- 14 m 2. In order to calculate term A


we use the relation: nl = A + B / >'I,
which gives A = 1.522. Finally the
refractive index calculated for>. = 610 nm is n3 = 1.532 .

4.6. In order to calculate the difference 1:10 in the angles of CHERENKOV


radiation we use the formula :
v c
cosO=-=- ,
vJ.l nvJ.l
where vJ.l is the velocity of the muon. A small increment of 0 is thus
represented by a small increment of the refractive index I:1n

L.l.
. 0 =cl:1n
"0 SIn --
n 2vJ.l
or using the well known representation sin 0 by cos 0, the expression for
1:10 takes the form
1:10 = I:1n 1
n J(nvJ.l/c)2 - 1

The relativistic formula for energy e= £01 )1- v~/c2 , where £0 = 100
MeV is the muon rest energy, permits the following ratio for (vJ.l/c)2 :
(vJ.l/c)2 = 1- (£0/£)2 = 0.8. One may take the mean value of the refrac-
tive index as n = (nl + n2)/2 = 1.61 for n. Substitution of the numeri-
cal values in the expression for 1:10 gives 1:1() = (I:1n/nh/1.61 · 0.8 - 1 ~
6.10- 3 rad.

4.7. Let the proton run a short distance I:1l between two adjacent
points of its trajectory (Fig.4.47). Radiation emitted by the proton at
every point will group between two conical surfaces enclosing the angles
01 for red rays and ()2 for blue ones. Hence, in order to find the light
power we must calculate the light energy within a volume 1:1 V bounded
by the two ring-shaped areas and by the short line I:1l , which is shown
in Fig.4.47. We assume that these rings are rather small and have the
area (J. We represent (J by the radius R and the angular width 1:10 =
()2 - ()1, provided that 1:10 is rather small:

(J = 27r R 2 1:1() sin "0 ,


160 DEMONSTRATIONAL OPTICS

Al

Figure 4.47.

where B is a mean angle associated with radiation of a wavelength >:


between the blue and red wavelengths. Hence, for the desired volume
we can write the expression:

6.V = 2rr6.IR 2 6.() sin (j

The number of quanta occurring within this volume n is the same as


emit ted while the charge moves along the short path 6.1, therefore

n = 2rr6.IR 2 6.() sin BTJ ,

since TJ is the number of photons emitted per unit length. The .mean
light energy contained within 6.V is given by

Since light moves the distance 6.1 during the time interval T = 6.1/v,
where v is the mean phase velocity of light, the power may by expressed
as
N = E/T = 2rrvR26.()sinBTJh"iJ .
By definition, the solid angle 6.0 containing the radiation is equal to
6.0 = a / R 2 and the power N/),o within the solid angle 6.0 is equal to:

N M! = 2rrv6.() sin -()TJh"iJ = 2rrc6.() _nsin BTJh"iJ


Light Waves in Media 161

where n is the mean refractive index of the glass. Now we use the result
obtained in the previous problem: boO sinO = cbonl(n2~) and substitute
b.OsinO by cbonl(n2~) . Then we get for N/:;.n :

_ 21rc2 b.nryhlJ
N /:;.0. - -3 V.
n p

From the relativistic energy of the proton e; J


= £01 1 - VlI c2 , the
proton velocity can be estimated provided that £0 ~ 1 Gev, so that:
1 - Vp2 Ic 2 ~ 0.2 , thus ~/c ~ 0.9. For numerical calculations we take
n = 1.565, bon = 0.02, ry = 100 quanta per 1 em and IJ = c/>', where
>: = 572 nm. Substitution of these magnitudes in the expression for N /:;.0.
gives Nss. ~ 4.10- 12 st«
Chapter 5

OPTICAL ANISOTROPY

Light propagation in crystals is accompanied by a number of surpris-


ing phenomena which have taken a special place in the history of optics
by enabling fundamental discoveries. When a light wave falls on a plane
boundary of an isotropic medium the wave passes from the boundary
into the medium as discussed in Chapter 4. But when the medium is an
anisotropic crystal, in general, two waves exist inside the crystal, prop-
agating along two different directions with two different velocities. The
oscillations of the electric vectors of these waves are mutually orthogo-
nal and the waves have mutually orthogonal linear polarizations. This
phenomenon is called double refraction or birefringence. It was found by
BARTHOLINUS with calcite in 1670, and later was carefully investigated
by HUYGENS . He explained the phenomenology of double refraction by
assuming that the incident wave gives rise to two waves. One wave cre-
ates wavelets with a spherical wave front. Therefore, as in an isotropic
medium, the velocity of the wave is the same for all directions, and a
beam which behaves in such a way is called the ordinary ray. The other
wave creates wavelets with an elliptical wave front and is called the ex-
traordinary ray. The velocity of the extraordinary ray is dependent on
the direction of propagation.

1. Double refraction in calcite


Natural calcite (or chalc - spar or Iceland spar) is available in the
form of large optically transparent crystals. These large natural crystal
specimens can be cut as a crystal rhombus by cleaving along the natural
planes . Then the surfaces can be polished . The final form of this optical
crystal may be thought of as a cube, slightly pressed along the diagonal
00' (Fig .5.1). The image of an object observed through the Island

163
O. Marchenko et al., Demonstrational Optics
© Kluwer Academic/Plenum Publishers, New York 2003
164 DEMONSTRATIONAL OPTICS

o o

a b

Figure 5.1. The double image of letter D observed throught a calcite crystal (a) ; if
the crystal is rotating the double image rotates too (b) . The dotted line 00' specifies
a spatial diagonal of the crystal.

crystal doubles. From this doubling of the image the name birefringence
was derived .
Fig.5.1 shows an image of the letter D observed through the faces of
the crystal. Let us draw a line 00' through two apexes of the crystal
and through geometrically similar points of the images (88') , as shown
in Fig .5.2,a,b. When the crystal rotates, the double image will rotate
in space in such a way, that the lines 00' and 88' stay parallel. This
simple observation indicates the complicated nature of light refraction
in the calcite crystal.
For observing double refrac tion in a crystal in another way, a narrow
parallel beam is directed normally onto a natural face of the crystal.
Two beams exit from the opposite side, parallel to the incident beam

o -, 0,
..... s ,
,s
-. ,

,
s' ., s'
,
,
0'
0'
a b

Figure 5.2. When rotating the crystal the double image turnes in a way that a line
ss' drawn through two similar points of the images remains parallel to the spatial
diagonal 00' , as shown for the cases (a) and (b).
OPTICAL ANISOTROPY 165

-
Figure 5.4. Testing the polarization
Figure 5.3. Spatial splitting of a nat- states of the ordinary and extraordi-
ural light beam into an ordinary and nary rays . Double arrows indi cate two
an extraordinary ray. orthogonal linear polarization orienta-
tions.

(Fig.5 .3). The ordinary beam passes along the direction of the incident
beam, and the second, extraordinary beam is shifted with respect to the
direction of the original beam. In other words the angle of refraction of
the extraordinary beam is not equal to zero, even in the case of normal
incidence . When rotating the crystal around the axis of the incident
beam the image of the extraordinary beam on a screen behind the crystal
moves in a circular path around the ordinary beam.
For natural, non-polarized incident light the two images of the out -
going beams have the same intensity. It is easy to verify that the output
beams are polarized in two mutually orthogonal directions. If we put an
analyzer in the outgoing beam, its rotation results in a periodic decrease
in the ordinary beam intensity and an increase in the intensity of the
extraordinary beam (and vise versa). The maximum intensity for the
ordinary beam corresponds to complete extinction of the extraordinary
beam (and vise versa) . The arrows in Fig.5.4 . indicate the orientation
of the polarization for both beams.
Now let a broad beam of light fall on the crystal to produce over-
lapping spots on a screen as shown in Fig.5.5. Rotation of the crystal
around the incident beam does not change the position of the ordinary
beam , whereas the image of the extraordinary beam will rotate. Insert-
ing an analyzer in the outgoing beams provides successive extinction and
brightening of images, but for the region where the images overlap the
screen brightness remains invariable (Fig.5 .5,b) .
This procedure confirms that the total intensity of both beams is
constant and equal to the intensity of the incoming beam 10. This beam
is decomposed into two beams of the intensities 10/2, both polarized
in mutually orthogonal planes . If the intensity of one beam after the
166 DEMONSTRATIONAL OPTICS

Ordinary
ray

Natural
light Extraordinary
Ordinary extraord inary
beam beam

a b

Figure 5.5. With normal incidence on a natural face of calcite, the ordinary ray
propagates along the direction of the incident ray, whereas the extraordinary ray is
deflected from this path (a) . Due to the deflection of the extraordinary ray a shift
between the centers of the outgoing rays exists. If the beam diameter is large , the
outgoing beams partly overlap .

analyzer varies as (10/2) cos2 a, then the intensity of the second beam
has to vary according to the expression (10/2) sin 2 a, where a is the
angle between the principle direction of the analyzer and the direction
of polarization of the first beam.
Now we will discuss the mutual spatial arrangement of the natural
faces of calcite in more detail. All its natural faces are rhomboids with

*1I
*
/
,/
/
/

a b

Figure 5.6. The spatial arrangement of the natural faces of calcite. The spatial
diagonal 00' is the optical axis of the cryst al (a) . A beam of natural light has still
no certain polarization state after passing throught the crystal along the optical axis
(b).
OPTICAL ANISOTROPY 167

Symmetry axis
o

o o
f4-- - 2.5.x. - -.-j

a b

Figure 5.7. The negative molecular ion C0 3 is composed of oxygen atoms 0 , at the
corner of an equilateral triangle, and of a carbon atom at the center of the triangle.
The ion C03 has a 3rd order symmetry axis normal to the plane of the equilateral
triangle.

the apex angles of 78° and 102° as shown in Fig.5.6,a. The spatial
diagonal DO' connects two opposite obtuse spatial angles. Cutting the
crystal in a way that the cut surfaces are orthogonal to DO' will result
in a cross section has the shape shown in Fig.5.6,b.
Let us place a crystal cut in such a manner into a narrow beam of
natural light . If this beam passes along the direction 00' (perpendicular
to the cutting faces), there will be no birefringent effect. This special
direction is called the optical axis of the crystal.

2. The structure of calcite


A calcite crystal, which has the chemical formula CaC03, belongs
to the hexagonal system of crystalline lattices. To analyze the optical
properties of the calcite crystal dealing with its optical anisotropy, one
can start with a structural component of the calcite crystal, the neg-
ative molecular ion CO; . This ion has a triangular structure, where
three oxygen atoms, at the corners of an equilateral triangle, surround a
carbon atom placed at the center of this triangle (Fig.5.7,a) . Due to this
structure the ion has a symmetry axis of the 3rd order, which is normal
to the triangle plane and passes through its center (Fig.5.7,b).
An analysis shows that in the visible range of the spectrum, the main
contribution to the polarizability under influence of an electromagnetic
wave is given by the oxygen atoms, whereas the carbon atom of the
ion CO; provides only a small contribution to the total polarizability.
Therefore we can restrict our consideration to the oxygen atoms.
168 DEMONSTRATIONAL OPTICS

E
~->---- k
B

a b

Figure 5.8. Polarization of oxygen atoms by the electric vector of a light wave with
linear polarization. When the light wave propagates perpendicular to the symmetry
axis (a) the resulting dipole moment of the ion is smaller than the dipole moment
indu ced by waves propagating parallel to the symmetry axis (b).

Let the direction of propagation of a linear polarized light wave with


wave vector k be normal to the symmetry axis of the ion (Fig.5.8 ,a), and
let its electric vector E be parallel to the symmetry axis. The induced
elect ric dipoles p of the oxygen atoms have the same direction as the
electric vector of the light wave. The three oxygen atoms mutually
influence each other's dipole moments , leading to a lower p compared
to three single atoms. If the direction k changes to k' , the polarization
has the same value, as long the direction of the electric vector E remains
parallel to the symmetry axis.
We have to compare this case with the situation, where the light wave
propagates along the symmetry axis (perpendicular to the plane of the
tri angle) , and the electric vector is located in the plane of the triangle.
For example, let the electric vector E be parallel to the line joining
two oxygen atoms (A and B) (see Fig .5.8,b) . The electric dipoles of
these atoms are parallel, so they try to reduce each other, whereas they
int ensify the electric dipole of the third oxygen atom C. As a result the
total polarization of the ion is larger with respect to the previous case.
The polarization is nearly independent on the position of the electric
vector with respect to the oxygen atoms. If turning the vector E , the
net polarization of the ion does not change noticeable.
Hence, in the first case the net electric polarization of the C03 ion
is smaller, associated with a smaller value of the refractive index. This
case of propagation perpendicular to the symmetry axis corresponds to
OPTICAL ANISOTROPY 169

Symmetry axis
Upper layer of Ca of 3rd order

Figure 5.9. Three layers of the lattice of calcite .

extraordinary rays. The second case with the wave propagation along
the symmetry axis corresponds to the propagation of ordinary rays.
We should note that the contribution of the Ca atoms to the polariza-
tion is small compared with the contribution of the oxygen atoms. Our
considerations are still valid when Ca atoms and CO;; ions compose a
calcite crystal. Figure 5.9 shows three layers of the lattice. The upper
layer consists of calcium atoms. Every trio of such atoms compose an
equilateral triangle.
The second layer is composed of negative ions of CO;;. The next layer
also consists of calcium atoms, which are placed in another position than
in the upper layer, but each three atoms are also located in the corners of
an equilateral triangle. When we speak about the hexagonal system of
such a crystalline lattice we assume that there is a symmetry axis of the
third order oriented normally to the plane of each equilateral triangle
through its center. The lattice structure stays the same after rotation
by an angle of 1r /3 around the symmetry axis.
We see that the net polarization of atoms due to the action of a
linearly polarized light wave is higher when the electric vector is directed
perpendicular to the symmetry axis, i.e, when the wave travels along
this axis. Contrarily, the net polarization is smaller when the electric
vector is directed along the symmetry axis and the wave travels across
to this axis. This is the case of a wave travelling perpendicular to the
axis with a linear polarization in the plane parallel to the axis. The
170 DEMONSTR ATIONAL OPTICS

symmetry axis is therefore the main optical axis of the calcite crystal. A
smaller polarization is associated with a smaller value of the refractive
index (n e = 1.486) and a higher value of light velocity, i.e , with an
ext raordinary ray. A higher value of polarization connected with a larger
value of th e refractive index (no = 1.658) corresponds to ordinary wave
propagation.
Let us select a C0 3 ion in the crystal (Fig.5.1O,a). Its symmetry axis
DO' is parallel to the optical axis of the crystal. Let a light beam pass
the crystal along the directio n 88' oriented with an arbitrary angle with
respect to DO'. The oscillations of the electric vector of the ordinary
beam occur perpendicular with respect to the plane defined by DO' and
88' . This means that the electric vector E in the ordinary light wave
is paralle l to the molecular plane of C03. Since t he direction 88' was
selected arb itrarily, the fact that vector E is parallel to t he molecular
plane holds for every angle of incidence of the light beam at the crystal
face. Hence t he degree of polarization of the molecular ion C0 3 , t he
refractive index and the phase velocity for an ordinary beam does not
depend on the propagation direction 88' .
For an extraordinary beam the polari zation plane of E always coin-
cides with the plane containing the 88'- and OO'-directions (Fig .5.10,b).
Therefore, the polarization of the C03 ions depends critically on the an-

s' s'

s
~I I
s
~I I
i
I

0 0 1

a b

Figure 5.10. For a given direction of propagation SS' of a natural light beam , os-
cillations of the electric vector E of the ordinary beam always take place in a plane
parallel to that of the molecular ion [a}; oscillations of the electric vector E of the
extraordinary beam always take place in a plane normal to that of the molecular ion
(b) . In the latter case the net induced electric dipole moment p of the ion depends on
the angle between the direction of propagation and the optical axis of t he ion, 00' .
OPTICAL ANISOTROPY 171
gle between 551 and 00 1 and, hence, the refractive index and the phase
velocity of the extraordinary wave are also dependent on the propagation
direction.
Conclusion. For the example of calcite we analyzed the main fea-
tures of the birefringence phenomenon. Spatial anisotropy of the crys-
talline lattice of calcite is the main reason for its optical anisotropy. In
general, the anisotropy leads to two waves which propagate within the
crystal along two different directions. Only two directions within the
crystal exists for which ordinary and extraordinary ray will travel with-
out spatial separation. The first case is propagation along the optical
axis. Here the definition of ordinary and extraordinary rays loses its
sense, both rays propagate with the velocity of the ordinary ray. The
second case is the propagation perpendicular to the optical axis. Here
both rays propagate in the same direction, but with maximally different
velocities given by clno (ordinary ray) and cine. (extraordinary ray).

3. A monochromatic wave in an anisotropic


crystal
3.1 Propogation of energy
Propagation of light in an anisotropic crystal is predicted by MAX-
WELL 'S equations introduced in Chapter 4:

rotH
aD
= 7ft rotE = -aB
- (5.1)
at
The material equations (4.1) and (4.2) written for isotropic media in
Chapter 4, have the same forms here, except that the quantity cleo =
n 2 is dependent on the direction of propagation within an anisotropic
medium. For this reason the direction of propagation of the wave energy
does not always coincide with that of the wave front. As we have seen
in Chapter 4, the direction of energy propagation is determined by the
POYNTING vector S :
S=ExH
We define two unit vectors now: one is t, specifying the direction of
energy propagation, and the other, as before, is s, normal to the wave
front. The vectors t and s differ in space.
According to (4.7) and (4.11) we can write down the relationships
w
Ho = kS x Do and (5.2)
172 DEMONSTRATIONAL OPTICS

D x

s
z

Figure 5.11. In a Cartesian system x,y ,z the vectors E,D,t ,S and s are located in
the Oxz-plane, the magnetic vector H is perpendicular to this plane.

between the complex amplitudes Do, Ho and Eo of a monochromatic


wave. The last expression can be changed to
1
H o = ---Eo x s (5.3)
(/Lv)

Now we substitute the magnitude from (5.3) for the magnetic vector in
(5.2) :
1
- 2- sx (Eo x s) = Do
V/L
The product s x (Eo x s) is then transformed as follows:

sx(Eo x s) = Eo(s· s) - s(Eo . s) = EO.L

where Eo.L is the component of the electri c vector normal to the direction
of propagation, as specified by the unit vector s. Therefore, the two
vectors Do and Eo are not parallel, since, as it follows from the two last
relations,
1 n2
-2-Eo.L
V /L
= -2-EQ.i
C /L
= Do
We see that the vectors E and D are both at right angles to the magnetic
vector H (Fig.5.H). The latter forms a right angle with s as well as
with vector D . Thus Hand D are transversal to the direction of the
propagation s . The electric vector E is also at right angle to H, so
that E , D and s are coplanar. However E includes an angle a with the
direction of propagation, but E is at a right angle to the direction of
energy propagation denoted by the unit vector t. This means that light
energy propagates at the angle a to the normal of wave front s with
a velocity V r , which is not equal to the phase velocity v = cfn . It is
OPTICAL ANISOTROPY 173

optical axis opticat axis


/
./

s
Figure 5.12. Two systems of wavelets propagating away from the face of a negative
uniaxial crystal into the bulk of the crystal. The vector D lies on the wave front
formed by elliptical wavelets, whereas the vector E is outside the wave front . In
cont rast , both vectors D and E are on the wave front formed by spherical wavelets.

clear that the velocity V r , which is called ray-velocity, corresponds to


the energy propagation of an extraordinary ray within an anisotropic
medium. In contrast, for propagation of the ordinary ray the vectors E
and Dare colinear to one another, the phase and ray velocities coincide
in magnitude as well in direction (Fig.5.12) . Two systems of wavelets
propagate away from the face of an uniaxial crystal. The vector D lies
on the wave front formed by elliptical wavelets, whereas the vector E
lies outside of the wave front. In contrast, both vectors D and E are on
the wave front formed by spherical wavelets.
When natural light penetrates through a uniaxial crystal, this differ-
ence between the ray and phase velocities is the reason for two types of
secondary waves, one having elliptical wave fronts and the other spheri-
cal ones (Fig.5.13). If the direction of propagation of the incident wave
is neither along the optical axis nor at a right angle to it, inside the
crystal the elliptical and spherical waves propagate away from the face
of the crystal in different directions and with two mutually orthogonal
polarizations. Even at normal incidence, the ray or energy direction for
the extraordinary wave will deviate from the normal to the boundary.
As we have seen when treating the example of calcite, two principle
indexes of refraction exist : no and n e . The difference ~n = n e - no is
a measure of the birefringence. It has been mentioned that calcite is
a negative uniaxial crystal, since ~n < O. Spherical wave fronts in a
negative crystal, which are associated with ordinary rays, are enclosed
within the elliptical wave fronts of extraordinary rays, because V o < V e .
In contrast, there are other crystals, for example crystalline quartz or
ice, for which ~n > 0 and which are called positive crystals. In the
case of a positive crystal two wavelets emitted by point 0 propagate
within the crystal with different phase velocities, so that the elliptical
174 DEMONSTRATIONAL OPTICS

\ 0 Natural light

0'
Extraordinary Ordinary Direction of
ray ray optical axis

Figure 5.13. If the direction of propagation of natural light includes an arbitrary


angle with the optical axis of the crystal 00', two types of secondary waves are
emitted by the boundary within the crystal: one is the spherical wave, and the second
is the elliptical wave. The spherical waves are associated with ordinary rays, and the
elliptical waves with extraordinary rays . In the direction of the optical axis both wave
fronts coincide, whereas in the direction N N' normal to the optical axis they differ
to the greatest extent .

--
optical axis __~ optical axis __~
~ .

a b

Figure 5.14. In the case of a positive crystal two wavelets emitted by point 0 prop-
agate within the crystal with different phase velocities, so that the elliptical wavelet
is enclosed by the spherical one (a) . For a negative crystal the spherical wavelet has
less velocity than that of the elliptical one, therefore the spherical wavelet is enclosed
by the elliptical one (b) .

wavelet is enclosed by the spherical one (Fig.5.14,a); for a negative crys-


tal the spherical wavelet has less velocity than that of the elliptical one,
therefore the spherical wavelet is enclosed by the elliptical one (5.14,b).
OPTICAL ANISOTROPY 175

Canada
balsam
~/ n=1 .53

90·
Extraordinary ray

Calcite I'lo = 1.66


ne = 1.49

Figure 5.15. The Nicol prism .

3.2 Nicol and Wollaston prisms


There are a large number of optical instruments using the birefrin-
gence phenomenon in calcite crystal. Let us consider two of them: the
NICOL prism and the WOLLASTON prism .
The NICOL prism is fabricated from a calcite crystal with a length 3.5
times larger than its width. Two boundary faces AD and BC are cut at
an angle of 680 with respect to its long sides DC and AB (Fig.5.15) . In
turn AD and AC, as well AC and BC, cross at a right angle. Afterwards,
this crystal is split along the plane perpendicular to the boundary faces
(along AC) . These two parts are glued together by Canada Balsam
without changing their orientation.
This glue has a refractive index of n = 1.549 which is between no
and ne . For an ordinary beam in the crystal, the Canada Balsam is
a medium which is optically less dense and for an extraordinary ray a
medium which is more dense than the crystal. Ordinary beams undergo
total internal reflection at the layer of the Canada Balsam (Fig.5 .15).
This is easy to show by computing the value of the critical angle: <Ptr =
arcsin no/n = arcsin(1.549/1.658) = 69° for the calcite - Canada Balsam
boundary, whereas the incident angle is around 68°. After total reflection
the ordinary beam is absorbed by the cover on the side of the prism.

Figure 5.16. If two Nicol prisms are set into a natural light beam, one before the
other, the first prism will play the role of the polarizer, and the second that of the
analyzer. One may cross the prisms by rotating the analyzer around the optical axis
of the beam .
176 DEMONSTRATIONAL OPTICS

Figure 5.17 In the case


of Wollaston prism , two
orthogonally polarized
beams leave the prism in
two directions.

The extraordinary beam, which can pass the prism, turns out to be
completely polarized. Two NICOL prisms inserted into a beam of natural
light are generally used as crossed polarizers (Fig .5.16). The passing
beam is dislocated by a certain amount from the incident beam.
The WOLLASTON prism allows the creation of two orthogonally po-
larized beams as shown in Fig.5.17. A WOLLASTON prism is made of
two rectangular prisms glued along the hypotenuse in such a way that
its optical axes 01 O~ and 0202 are orthogonal. Ordinary and extra-
ordinary beams propagate in the same direction in the first prism . An
ext raordinary beam leaving the first prism propagates as an ordinary
one in the second prism, and vice versa .
The refractive index for the first beam in the second prism is higher ,
so it is deflected toward the upper face (see Fig.5.17), while the second
beam now enters a medium with a lower refractive index and is refracted
toward the opposite side. A light beam entering perpendicular to the
outer surface of the first prism splits into two perpendicularly polarized
beams deviating symmetrically from the original beam direction.

3.3 A model of a uniaxial crystal


Now we consider a model of a uniaxial crystal. A uniaxial crystal,
having a single axis of symmetry, can be treated as a regular ensemble of
thin parallel equidistant plates located normally to the axis of symmetry
00' (Fig .5.18). In this case the only preferential direction coincides with
00'. A rotation around this axis provides a reproduction of the crystal
structure.
We assume that the separation of adjacent plates, d, is smaller than
the wavelength >. of the incident wave and that all these elementary
plates have the same thicknesses a:

d e:» , a« >. . (5.4)


OPTICAL ANISOTROPY 177

Ez
Dzi====t"J====t-t:===t--
I I
d I
I
I
I

z
z

a b

Figure 5.18. A set of plates modeling an uniaxial crystal. Two cases of propagation
of a linear polarized wave are possible: one is the propagation along the "optical axis"
of the plates and the other is the propagation transversal to this axis . In both cases
the electric vector E is parallel to the faces of the plates (a) . Dependencies of the
projections of the electric displacement vectors D'; and D z on z-coordinate are shown
in (b).

Let s be the dielectric constant of each plate. If the plates are in vacuum,
the dielectric constant of the medium in which they are immersed is equal
to eo.

3.3.1 Vector E parallel to the faces of the plates


Let a linearly polarized plane monochromatic wave pass along direc-
tion 00'. In this case the electric vector E is always parallel to the
elementary plates, modeling the crystalline lattice structure (Fig.5.18).
According to the boundary conditions, the tangential component of
the electrical vector of the light wave varies continuously when the light
beam is propagating across the surface between two media, so that, in
this case, the electric field will have the same value E inside the plates
and in the space between . The electrical displacement in the two regions
are D 1 = eE (inside the plates) and D2 = eoE (inside the spaces) and
have maxima D 1 and minima D2 along the axis 00'. The mean electric
displacement is given by the following relation:
D= D 1a+D2(d-a) =ea+(d-a)eo
E
d d
As the ensemble of plates can be treated as a continuous anisotropic
medium for this case, when the inequalities (5.4) are valid, the factor
[ea + (d - a)eoJld may be treated as the resultant permittivity of the
medium:
ea + (d - a)eo
cli = d
178 DEMONSTRATIONAL OPTICS

o
~:h ~ ~
x a d z
~B
E
a b

Figure 5.1 g. A linearly polarized wave which propagates perpendicularly with re-
spect to the optical axis. The electric vector E is normal to the faces of the plates
(a) . Dependencies of the projections E., D. and E. on z-coordinate are shown in
(b).

The refractive index assignable to cil is given by the expression

(5.5)

where n 2 = c/co is the refractive index of the plates. It is obvious that


the case under discussion is valid only for propagation of a wave along
the optical axis, because there is no dependence on the polarization
direction of the incident wave.

3.3.2 Propagation perpendicular to the optical axis


For a plane wave travelling normally to the optical axis 00' the re-
fractive index changes with the orientation of the electric vector E. If E
is parallel to the surfaces of the plates (Fig.5.18), this case is similar to
those considered above. In other words the refractive index is the same
as before. Therefore the phase velocity is the same as the phase velocity
of the ordinary ray.
But if the electric vector E is directed along the optical axis (normal to
the plates, Fig.5.19), the electric displacement will have the same value
D inside the plates and in the spaces in between. Hence, E 1 = D I co
(inside the spaces) and E2 = D]« (inside the plates). This implies that
the electric vector now undergoes periodic variations inside the medium.
These steps have maxima El and minima E 2 over the period d; hence
the mean electric vector takes the form:
OPTICAL ANISOTROPY 179

y
0'

Figure ' 5.20 The opera-


tion principle of a retarda-
x
tion plate.

The apparent permittivity assignable to this case is given by the expres-


sion
duo
e.L = ---,------:--
de + (e - eo)a
and for the refractive index we obtain, respectively,

2 n 2d
n.L = e.L/eo = n2d + (n2 _ l)a (5.6)

4. Natural optical anisotropy of materials


4.1 Retardation plates
Let us cut a plane out of an uniaxial crystal in a way that the optical
axis is parallel to the faces of the plate, as shown in Fig.5.20. Now let
a linearly polarized monochromatic wave of wavelength >. fall normally
on the plate. If the electric vector E makes an angle with the optical
axis, the orthogonal projections Ex and E y determine the electric field
vectors of the ordinary and extraordinary rays within the plate. Both
rays propagate in the same direction. The orthogonal components are
in phase on the input face of the plate but experience a phase difference
8 due to their different phase velocities in the plate, determined by the
difference of the refractive indexes, n e - no. After a plate of thickness
h, the phase difference is given by

(5.7)
where k = 2rr/ >. is the wave number in vacuum . As we have seen in
Chapter 3 the superposition of two waves polarized in two mutually
orthogonal directions will lead , in general , to an elliptic polarization of
the resulting wave.
It can be seen from (5.7) that in the case of a negative crystal, where
n e - no < 0, the phase velocity of the ordinary ray is smaller than
that of the extraordinary one. Hence, the phase of the ordinary ray is
retarded by 8 with respect to that of the extraordinary ray. Plates of
180 DEMONSTRATIONAL OPTICS

an anisotropic material based on such an operating principle are called


retardation plates.

4.1.1 Experiment with radio waves


It is well known that some sorts of wood consist of stretched layers,
and this striped structure is similar to the pile of plates considered above
(Fig.5.21). A board cut in a special way from a such wood can be used
to a certain degree as a retardation plate for electromagnetic waves of
ultra-high radio frequency (>' = 3 em).
As we already noted, the emitting and detecting horns of the appara-
tu s for demonstration of the ultra-high radio frequency electromagnetic
waves may by turned around their longitudinal axes, so we can fix the
direction of the linear polarization of the emitted wave as we like it .
The layers of this wood structure cause the "optical axis" of such
a st riped structure to be oriented perpendicular to the layers. In our
experiment, let the" optical axis" be vertical and let the polarization of
the electromagnetic wave be at 45° with respect to the "opt ical axis"
(Fig.5.21). This implies that the projections of the electric vector on
the vertical and the horizontal axes have the same magnitudes: Ex =
EoV2j2 and E y = EoV2j2 , where Eo is the amplitude of the emitted
wave. The oscillations of the amplitude of the output wave within the
plane at the back side of the wooden board (having thickness h ) are
given by

E(t ) = Ex(t)
V2 cos(wt) + yEoT
+ Ey(t) = xEoT V2 cos(wt + 0) . (5.8)

In particular, if
2n n
8 = >:(no - ne)h = 2nm + 2"

Detecting hom
45° "" II
O.
Emitting horn Wooden block

Figure 5.21. Anisotropy of wood with respect to radio-waves of 3 em. A wooden


block with a thickness of about 3 em acts like a quater-wave plate, converting the
linearly polarized radio -wave into a wave of circular polarization.
OPTICAL ANISOTROPY 181
for an integer m , or
A
h(no - n e ) = Am + 4" m=0,±1,±2, ... (5.9)

the end of the electric vector E(t) will undergo a circular rotation, be-
cause the relation (5.8) can be reduced to the following form:

J2 J2 .
E(t) = EoxT cos(wt) - EOYT sm(wt)

where x and yare unit vectors of x- and y-axis, respectively.


We see that the retardation plate, which is usually called a quarter-
wave plate (or A/4-plate), converts the linear polarization of an incident
wave into a circularly polarized output wave, when the condition (5.9)
is valid. It should be added that for effective operation asa A/4-plate,
the striped medium must have layers smaller than A. For the case of
the wooden board used in this demonstration, the average separation of
layers was about 1 mm, small compared to A = 3 em.

4.1.2 Quarter-wave plates in optics


For general use, mica quarter-wave plates are perhaps the most com-
mon typ e. They can be made by splitting thick sheets of mica down to
th e appropriate thickness. The effective values of refractive indices for
ordinary and extraordinary rays may vary from one sample to another
one, but the difference between the indices is rather small. For example,
a sample of mica may show the indices no = 1.599 and n e = 1.595 for
yellow light . Since the difference between the velocities of the ordinary
and extraordinary rays is very small, the mica sheets do not need to
be split so thin; typical thicknesses lie between 0.032 to 0.036 mm for
yellow light.
The reason why quarter-wave plates from a mica can be created so
easily, is that the optical axis of mica is practically parallel to the cleav-
age planes (Fig.5.22).
As the optical properties of the quarter-wave plate are dependent on
wavelength, an efficient transformation from linear to circular polariza-
tion takes place within only a certain narrow wavelength range ; for other
wavelengths one generally gets elliptically polarized light. Therefore,
when performing experiments using natural white light, an optical filter
is needed for a clear demonstrational effect (Fig.5.23). It is obvious, that
th e quarter-wave plate can also be used for an inverse transformation of
circularly polarized light to linearly polarized light.
Let us assume a quarter-wave plate to be cut from a positive anisotropic
crystal (n e > no). Its optical axis is vertically oriented (Fig.5.24,a,b).
182 DEMONSTRATIONAL OPTICS

Symmetry plane

x/ /'/: / - - - - -_ 7
Cleavage plane / - 7..:r-
/
/ r>,L 'r-----~--
/
h / /
A / / ,' ./ //
/ t...... /
/ - ,- -- -=- - //
// / / / 0/
/ / , / / //
/ / / Optical axis: " / / /
/
--r---------h-/ /
/ /
/ '/
----------_/
/ I /

Figure 5.22. An elementary cell of mica has a symmetry plane , which contains the
optical axis of the crystal. The direction of the optical axis is deflected from the
cleavage planes of the crystal by a very small amount, which means that a light wave,
falling normal onto the mica crystal, propagates nearly normally to the optical axis .

If the electric vector of a linearly polarized beam falling on the plate


lies at an angle +450 with respect to the positive direction of the x-axis
(Fig .5.24,a), when looking in the direction opposite to the beam propa-
gation, the projections of the electric vector Eo of the incident beam on
the horizontal x-axis and vertical y-axis can be written as follows:

By passing through the plate, these orthogonal components are retarded


by epx = khsi; and epy = khn e , respectively, where h is the thickness of
the plate, and k = 27r/ >. is the magnitude of the wave vector. For the
positive crystal, the difference 8 = epy - epx = 7r/2, hence, after passing

Polarizer Filter Analyzer

~~

Mica plate

Figure 5.23. A setup for observation of circular polarization caused by a mica


quarter-wave plate. Since the retardation of a mica plate is very sensitive to wave-
length, an optical filter must be used .
OPTICAL ANISOTROPY 183
Optical axis , r Optical axis , c
~ ,j' (jj
t' , j"
+
Y, • • y

x
~'~ " , }J4 -plate
X
"
-
, I

, }J4-p late

LJ'
a b
Analyzer
Ana lyze r

O,"~, ~c ~ ~
,
.. ljj
ax"

W~
Opt ical axis .

l~[ " ljj


-:
'l
\ - -

~
.... ",- "

~' \' "'' '


,,, : A/4-plate - ; }J4-plale

c d

Figure 5.24. Test of the polarization state of a light wave by means of a polarizer and
,\/4-plate. There are two ways to transform a linearly polarized beam into a circularly
polarized one by means of a '\/4-plate (a, b). A natural light beam passing through
the '\/4-plate does not change its polarization property which may be examined by
t he anal yzer (c); a circularly polarized beam can be transformed by means of the
,\/4-plate into a linearly polarized beam , shown by the analyzer (d) .

the retardation plate we get

for these components, and left-handed polarization appears (Fig.5.24 ,a).


If the polarization direction is oriented -450 with respect to the positive
direction of the z--axis , then right-handed polarization will take place
(Fig.5.24,b).
A quarter-wave plate can also be used for detecting a circularly po-
larized wave and distinguishing it from unpolarized natural light. The
polarization state of a natural light beam will remain the same after
passing the A/4-plate (Fig.5 .24,d), whereas a circularly polarized beam
becomes linearly polarized, which can easily be detected by an analyzer
(Fig.5.24,d).

4.2 Liquid crystals


Atomic particles in a crystal are firmly located at the nodes of the
three-dimensional crystalline lattice, which is a stable structure. In fact ,
the atomic particles undergo small displacements near the equilibrium
positions due to chaotic thermal motion. Contrary to the microstructure
184 DEMONSTRATIONAL OPTICS

x
a

Figure 5.25. Three domains of a liquid crystal, each grouped around certain di-
rections "1, "2, "1 (a). Prolonged molecules of the liquid crystal grouped around
micro-groves existing on the surface of a glass pla t e (b).

of a crystal, the molecular particles of a liquid do not have fixed positions


and full chaotic motion.
In liquid crystals the molecules do not have so many degrees of freedom
as in an ordinary liquids. The molecules of some kinds of liquid crystals
have a prolonged shape and the axes of adjacent molecules are oriented
in the same direction, but their centres can be positioned arbitrarily
with respect one to another (Fig.5 .25).
The orientation of the molecules is not uniform over a macroscopic
volume of the liquid crystal: only small domains exist with the same
orientation and the local orientations of adjacent domains are different.
There are several ways to artificially create the same macroscopic orien-
tation of all the domains within a liquid crystal. This can be achieved,
for instance, by applying weak magnetic or electric fields. The molecules
in a thin film of a liquid crystal can also be aligned along micro-grooves
existing in the surface of parallel glass plates containing the film. If these
grooves have the same direction, the thin liquid crystal film will show
the properties of an uniaxial crystal with an optical axis oriented along
the grooves.

4.2.1 Optical anisotropy of a thin film of a liquid crystal


For the purpose of demonstration of the optical anisotropy of a liquid
crystal film, the following simple experimental procedure may be carried
out. If a dry clear surface of a glass plate is lightly swept by a soft paper,
micro-groves will be made in its surface. Then a drop of the liquid crystal
is put on this surface and a water drop is placed close to the first drop
to have an object of comparison.
The drops are covered by a glass plate having the same direction of
grooves, made by the same technique. This complex plate is mounted
OPTICAL ANISOTROPY 185

Drop of
liquid crysta l Analyzer
Polar izer
/

Mercury
lam p
Diaph ragm
Heat Drop of
preventing water
filler

Figure 5.26. Setup for the demonstration of the optical anisotropy of a small drop
of liquid crystal. The drop of liquid crystal can lose its anisotropic property under
heating. To avoid this, a transparent heat blocking filter is inserted into the beam
before the glass sample with the drops , if a powerful light source is applied.

between two crossed polarizers and placed into a light .beam as shown
in Fig.5.26. IT the analyzer is absent, the images of two drops will be
identical. In the presence of an analyzer crossed with the polarizer, the
image of the water drop vanishes, whereas the image of the liquid crystal
becomes a color pattern, provided the direction of the grooves is 45° with
respect to the principal direction of the polarizer (Fig.5.26). The thin
film of the liquid crystal behaves similar to a stratified structure, or a
uniaxial anisotropic crystal. In general, this film creates elliptically po-
larized light from linearly polarized incident light. The phase difference
is determined mainly by the local thickness of the film. The outgoing
light rays show the same color after passing the analyzer for areas of the
film having the same thickness.

4.2.2 A liquid crystal prism


The phenomenological fact that two waves polarized in two mutu-
ally orthogonal directions arise after propagation through an anisotropic
crystal may be demonstrated with the help of a liquid crystal prism. This
prism is made from two thin parallel glass plates with micro-grooves on
the inner surfaces oriented identically and parallel to the edge of the
prism, which is mounted in the horizontal direction.
Molecules of the liquid crystal located between the plates are oriented
parallel to the edge of the prism, and the optical axis of this system lies in
the same direction. For observing the spatial separation of the ordinary
and the extraordinary beam , a practically parallel beam of white light
should pass the prism.
A beam from a mercury lamp is focused by means of a lens on a hor-
izontal slit (parallel to the prism edge) (Fig.5.27). An enlarged image
186 DEMONSTRATIONAL OPTICS

y
Analyzer Ordinaryrays
-,
~
----=--=-- - - - ._. -J _.- - _.
Extraordinary rays

x Liquid Screen
Horizontal crystal
slit

Figure 5.27. A setup for the demonstration of refraction of ordinary and exraordinary
rays by a liquid crystal prism . The distance between the prism and the screen is 4 m;
the objective, projecting the image of the vertical slit onto the screen , has the focal
length 50 em. The width of the slit is about 0.3 mm .

of the slit is formed by an objective on a distant screen and the liquid


crystal prism is placed after the objective. If a polarizer is absent, the
beam of natural light is split by the prism into two beams: one is the
ordinary beam, the other the extraordinary one. The refractive indices
corresponding to these beams are different (no> n e) , so the angles of
refraction and the positions of the images of the slit are both different.
Since the edge of the prism is oriented in the horizontal direction, the
spectrum formed by the ordinary rays will be vertically linearly polar-
ized, which is to be verified by an analyzer. Components of the incident
beam having linear polarization in the horizontal direction will give rise
to a spectrum formed by the extraordinary rays, which is also verified
by the analyzer. Due to the difference no - n e and no> n e the ordinary

extraordinary rays ordinary rays

blue - - -
green _ _
yellow ---.

a b

Figure 5.28. Two spectra formed by the liquid crystal prism; one is caused by ex-
traordinary rays (a) , the second by ord inary rays (b) . Deflection of rays of both
spectra .increases from the bottom to top of the pictures. The distance Ay between
two identical spectral lines is approximately equal to 6 mm .
OPTICAL ANISOTROPY 187

F Figure 5.29 A transparent


plate of organic glass sup-
ported at two points expe-
riences the action of a ver-
tical external force F .

rays are deflected by larger angles than the extraordinary rays, and for
this reason , two spatially shifted spectra are observed (Fig.5.28).
It should be noted that the quasi-crystalline state of a liquid crystal is
realized only within a particular temperature range . When heating the
liquid crystal film used in this demonstration by a flow of hot air (about
45°C in our case), the quasi-crystalline state breaks down and the liquid
will show the optical properties of an ordinary isotropic liquid .

5. Artificial birefringence
When a transparent isotropic material is subjected to an external
electric or magnetic field or to mechanical stress, it may become optically
anisotropic.

5.1 Photo-elastic effect


Static mechanical anisotropy of an originally isotropic optically trans-
parent body caused by an imposed extern al force can also lead to an
opti cal anisotropic effect similar to the one discussed above. We shall
illustrate this artificial optical anisotropy due to a mechanical effect . Let
a t ransparent plate of plastic material be supported at two points. An
external force F is exerted at its middle point as shown in Fig.5.29.
An orientation of the non-symmetric molecules of our plastic material
arises under the action of the expanding or compressing external forces
wit hin the originally isotropic material. Such a mechanical anisotropy
is accompanied by an optical anisotropy in such a way that a difference
appears between the refractive indexes of ordinary and extraordinary
rays. This phenomenon is called the photo -elastic effect. The direction
of the optical axis of this induced artificial anisotropy is the same as the
direction of the external mechanical force F (see Fig.5.29).
One can demonstrate this effect easily in the following way. The
surface of the rectangular plastic sample is illuminated by a light beam
and its image is successively projected onto a screen by a lens (Fig .5.30).
An external force is applied to the centre of the sample by a vertically
mounted screw. If the principle axis of the polarizer is at 45° with respect
188 DEMONSTRATIONAL OPTICS

Optical axis

Filter

Screen
Polarizer Analyzer
Transparent
plastic sample

Figure 5.30 . Setup for the demonstration of the photo-elastic effect. A transparent
plasti c sample is placed between two crossed polarizers, one with a horizontal direction
of polarization (x-axis), the other with a vertical dire ction of polarization (y-axis) .
To prov ide a distinct image of the sample, an optical filter has to be inserted into the
beam .

to th e optical axis of the plastic sample and the analyzer is crossed with
the polarizer a dark band will be observed on the screen (Fig.5.31).
This dark band within the image corresponds to a mechanically isotropic
local medium or to a non-stressed region of the sample. Local regions of
the sample which correspond to the parts of the image situated above
this dark band are affected by pressing forces, whereas regions below
thi s band are mechanically stretched.

5.2 The Kerr effect


After inserting into an electric field, some liquids show optical prop-
ert ies which are characteristic for an uniaxial crystal. This phenomenon
is called the KERR effect. A simple demonstration of this effect may be

A L---.....,...-.~

Figure 5.31. The inverse real image of the sample placed between two crossed polar-
izers and stressed in the vert ical direction. The regions AB and B e above the dark
bands are associated with compressed domains within the sample, whereas t he region
be low these bands is associated with mechani cally stetched domains.
OPTICAL ANISOTROPY 189

Kerr ce
Filter
Lamp
A>. -"- I h A n
-v-
I U
'CY
V V
E
U
Polarizer Analyzer Screen

Figure 5.32. Setup for observation of the Kerr effect. A parallel light beam passes
trough the Kerr cell placed between crossed polarizers. The principle direction of the
polarizer is at an angle 45° with the flat plates of the capacitor. Due to the despersion
properties of nitrobenzene, a distinct dark field may be obtained with the use of an
optical filter .

performed as follows. An optical cell filled with a dielectric liquid (ni-


trobenzene is usually used) contains two parallel internal metallic plates
forming an electric capacitor (Fig.5.32). When an electric voltage is ap-
plied to the capacitor plates, the uniform electric field arising between
th e plates will polarize the liquid , orienting the anisotropic polar mole-
cules of nitrobenzene along the electric field lines (Fig.5.33). An optical
anisotropy will appear. The optical axis of this quasi-crystal coincides
wit h the electric field direction.
An empirical law shows that the phase difference 0 resulting from the
difference in the refraction indexes of the ordinary and the extraordi-
nar y ray depends on the square of the electric field strength and on the
thickness of the layer of the liquid as:

(5.10)

where the factor f3 is determined by the nature of the liquid. Nitroben-


zene is characterized by its large value of f3. The glass cell for observation
of this effect is usually called a KERR cell.
If the cell is placed between two crossed polarizers, light will pass
through the cell when a certain electric field is applied to the plates of the
capacitor. This simple experiment is quite demonstrationally significant ,
because , without the field, light will be blocked by the system, providing
the principle direction of the polarizer is at 45° with respect to the optical
axis (the same as the field direction). The effective orientation time of
the nitrobenzene molecules under the influence of the electric field, i.e.
th e characteristic time constant of the effect, is very short and does not
take longer than 10-9 s. The short time scale of the KERR effect enables
its use in many scientific and technical applications such as laser light
modulation, short time shutter, etc.
190 DEMONSTRATIONAL OPTICS

An installation for the demonstration of the KERR effect useful to


estimate the effective time needed for the orientation of the nitrobenzene
molecules in the electric field is shown in Fig.5.34. Light from an electric
spark is guided through the KERR cell by a system of mirrors Ml, M2, M3
and M4. The voltage applied to the electric spark is the same that
charges the capacitor of the KERR cell.
The electric field corresponding to the voltage provides an ordered
orientation of the nitrobenzene molecules. While increasing the voltage,
the capacitor will be charged until a moment to when the spark arises.
Fast variations of the electric field near the spark threshold after to will

+
E E
a b c

Figure 5.99. Formation of optical anisotropy in nitrobenzene. The uniform electric


field, arising inside the capacitor, polarizes molecules of nitrobenzene (a) ; if a molecule
has an arbitrary position with respect to the electric field (b) then it will rotate
around its center and be aligned parallel to the electric field (c) . The movement of
the molecules due to heat acts against a strict alignment.

Crossedpolarizers
Electric
spark Kerr cell

y-~
M1~

Figure 5.94. A setup for the estimation of the relaxation time of nitrobenzene mole-
cules in an electric field.
OPTICAL ANISOTROPY 191

be accompanied by a fast decrease in the electric field strength within


the capacitor. The electric field variation and the orientation processes
are separated in time by a small interval ilt. If the capacitor is at a
distance 1 from the spark, the time ilt is given by

ilt = llc ,
were e is the velocity of light. Now, if L is the geometrical path needed
for light propagation from the spark to the capacitor via M I , M2, M3
and M 4 , the time T corresponding to L is given by

T = Lie .
The geometrical condition of the demonstrational experiment

L» 1

may be easily fulfilled. This condition implies that one can assume the
disorien tation process of the nitrobenzene molecules to start approxi-
mately at the moment of spark ignition, i.e, at the moment to. While
light is passing from the spark over the mirrors, the disorientation of the
molecules can be terminated, if L is large enough. If TO is the character-
istic time constant of the disorientation process , the inequality

T > TO
is valid in this case. By decreasing the geometrical light path L be-
tween th e mirrors, shifting the mirrors MI and M2 towards M3 , M 4 , the
condition
T~TO

can be provided. When this condition has been achieved the light beam
will pass through the KERR cell. Such a clear demonstrational experi-
ment is feasible, because the geometrical length L , corresponding to the
typical orientation time constant of TO ~ 10- 9 S, is equal to 30 em.

5.3 An experiment to determine the velocity of


light
A similar effect, known as POCKELS effect, occurs when applying an
electric field to solid dielectric matter. In some cases, the relaxation
rate of the electric field - indu ced anisotropy in solids is faster than in
nitrobenzene. It is well known that some uniaxial crystals, such as am-
monium dehydrogen phosphate (ADP) and potassium dehydrogen phos-
phate (KDP) , are characterized by a relaxation time of about 10- 11
s.
192 DEMONSTRATIONAL' OPTICS

Crossed polarizers

KDP • crystal
He-Ne laser

Mirror
Cavity

o
Oscilloscope H.F. - Generator (600MHz)

Figure 5.35. Setup for the estimation of the velocity of light . The laser beam is
modulated with help of a KDP crystal.

For a K DP crystal cut along its optical axis and fixed between two
plates of an electric capacitor, the phase difference, arising between the
ordinary and the extraordinary rays, when an electric field is applied,
can be expressed as
C=KU , (5.11)
where U is the voltage applied to the capacitor plates and K is a constant
(d. (5.10)). This phenomenon is known as the POCKELS effect.
A crystal placed between the plates of a capacitor is called a POCKELS
cell. There are two apertures at the centers of the plates to allow light
to pass along the optical axis. Usually the cell is placed between two
crossed polarizers. The POCKELS cell is frequently used as a fast optical
shutter operating under the action of a high-frequency voltage.
We now consider a specific application of this optical shutter for the
creation of a train of short light pulses (the intensity of the light is mod-
ulated with a very high frequency). The arrangement for the demonstra-
tion of this effect is shown in Fig.5.35. The principle of the fast optical
shutter is based on the use of a high-frequency radio-wave cavity which
has the same function as the electric capacitor of the POCKELS cell. This
cavity is made of two coaxial cylinders as shown in Fig.5.36.
The high-frequency electromagnetic energy from a radio frequency
generator excites electromagnetic field vibrations within the cavity by
means of a loop fixed near the side wall of the cavity (Fig.5.36).
The distribution of the electric field is characteristic for standing
waves. The node of the electric field is on the right side, as shown
in the figure, and the antinode of the field is close to the left of the cav-
OPTICAL ANISOTROPY 193
ity, where a K DP crystal, cut along its optical axis, is fixed in a central
hole. The frequency of the field oscillations is 600 MHz.
Light from a He-Ne -laser (>. = 632.8 nm) passes a semi-transparent
plate, a polarizer, the K DP crystal and a second polarizer (crossed with
the first one) and then reaches a mirror, which can be moved along a
slide. The reflected light passes back through all the components and is
reflected to a photomultiplier by the semitransparent plate. The crossed
polarizers form the 'principle part of the optical shutter together with
the cavity and the crystal. We can obtain the relation for the intensity
of the light passing through the POCKELS cell, providing first that the
electric vector of the incident beam is at an angle 45° with the principle
direction of the first polarizer, and second that the absolute value of
phase difference in (5.11) is not larger than unity :

181« 1
This inequality is safely valid, because the value 8 = 7r / 4 is attained
when U ~ 8000 V is applied to the KDP crystal. Under the conditions
of the experiment IUlmax « 8000V.
Let AA' and BB' be the principle directions of the first and second
polarizer, respectively, as shown in Fig.5.37.
If U = a the anisotropy is absent and the oscillations of the electric
vectors for ordinary and extraordinary rays are in the directions x and y ,
respectively, and they have the same amplitude Eo. In the experiment,
the field of the He-Ne laser is polarized along the direction AA' of the
first polarizer. Then the field vector is projected onto the directions x

Electricfield distribution
KDP - crystal

1 cm

Loop

Togenerator

Figure 5.36. The arrangement of the resonance cavity for production of a train of
short light pulses based on the Pockels effect .
194 DEMONSTRATIONAL OPTICS

y y

A Eo Y
\ A B
B

,7~+--- - +-~:>
\ -, / /
\ / -, / /
-,
\
4~~" -,
, -, //
//
/ /

/ -, //

I -, X
,/I~S
1
, -,
-,
\ /
/ / , -,
\ X 8' I /
/
/
-, AI
X B
AI
a b

Figure 5.37. Illustration of t he opera t ional pr inciple of the opti cal shut te r. A K DP
cryst al is placed between two crosse d pol arizers with the principle directions AA'
and BB' . Oscillation s of th e elect ric vector of the light of a He-Ne - laser occur
along dire ct ion AA' (a). With a voltage applied to the crys tal, a small deflection of
th e elect ric vector from direction BB' by an angle {j occ urs, leading to light beam
transmission through th e shutter (b) .

and y . The superposit ion of th ese oscillations is linearly polar ized and
direct ed along AAI , so that the projection of the resulting vect or on BBI
is equal to zero and the light is blocked. With increasing volt age U a
phase difference 8 is introduced, and the projection of the ampli tude of
the resulting field in t he direct ion BB' is given by

Eo cos ~ - Eo cos (~ - 8) = Eo~ sin ( ~ - ~) sin (~) =

= Eo V; [cos (~) sin (~) - sin


2
(~)]
As 8 « 1, the resulting electric vector or t he ampli tude of the wave
passing through the crossed polarizer is

E
.../2.sm8 ~ Eo-.../28
= Eo-8 8
Hence, the intensity of light passing through the crossed polarizers is
given by the expression

(5.12)

On subst it ution of 8 by KU from (5.11), providing that the dependence


U(t ) is a harmonic function of time
U(t) = Uo cos(27r i t)
OPTICAL ANISOTROPY 195

we find
I(U) = Io".2UJ cos2(21r ft) (5.13)
In our case f = 600 MHz. We find that the maxima of the intensity due
to (5.13) is Io",2UJ , when

21rft = 1rm , where m = 0, ±1, ±2 ...

so the light trains have a duration of T = 1/(2f). The propagation


length is l = cr = c/(2f) ~ 25 ern.
Since the optical shutter is periodically closed and opened, the inten-
sity of the reflected light transmitted a second time through the shutter
depends on the position of the mirror M. The interval T is the time
during which the shutter is opened. If L is the distance between the
crystal and the mirror, 2L/ c is the time needed for the light to propa-
gate from the crystal to the mirror and back. Maxima in the intensity
will be observed when
2L
-=nT n=1,2,3, ...
c
By moving the mirror and measuring the positions of two neighboring
maxima L1 and L2, we can estimate the velocity of light . If

L2 - L 1 = (n + 1)cr/2 - ncr/2 = cr/2 ,

and
T = 1/(2f)
we get
c = 2(L2 - L 1) = 4f(L2 - LI) (5.14)
T
Measurements show that £2- L1 = 0.125 m. So for the velocity of light
we find c ~ 4 . 6 . 0.125 . 106 = 3.0· 108 ta]».

6. Optical activity
6.1 Natural quartz
Natural quartz is an anisotropic uniaxial crystal. In order to demon-
strate its optical properties, let a beam from a source of white light pass
through a polarizer and an optical filter, and let it fall normally on a
thin parallel plate of quartz cut across its optical axis (Fig.5.38).
When rotating an analyzer placed after the plate, we find that the
direction of linear polarization of the transmitted beam is at a certain
angle relative to the principle direction of the polarizer, but the light still
remains linearly polarized. This implies that the quartz plate gives rise
196 DEMONSTRATIONAL OPTICS

Polarizer Analyzer
Filter

Lamp

Quartz sample

Figure 5.38. Setup for observation of the optical activity of quartz. A parallel beam
of natural light formed by an objective passes through the polarizer and then through
the quartz sample along its optical axis . The polarization direction of the transmitted
light beam is deflected with respect to its original direction which may be tested by
the analyzer. An optical filter is inserted into the beam because the rotation caused
by the quartz is very sensitive to the wavelength.

to a rotation of the direction of the linear polarization. This phenomenon


is called optical activity.
It is known that natural quartz has two structural modifications: one
called left-rotating quartz and the other called right-rotating quartz.
The direction of rotation is understood to be noticed when viewing into
the output ray. Right -rotating quartz produces a clockwise rotation of
the polarization plane. The absolute values of the rotation angle are the
same for both modifications.
Let two plane parallel quartz plates of equal thickness be glued side
by side, to make a single plate (Fig.5.39). Both of the quartz plates
have been cut across their optical axes, but one of them is made of left-

Figure 5.39. The natural modifications of quartz (one is right- • and another is left-
rotating quartz) cause a rotation of the polarization plane in two opposite directions.
In both cases the light beams propagate along the optical axes of the quartz samples
010~ and 0202, Both samples have identical shapes and sizes.
OPTICAL ANISOTROPY 197

Table 5.1. Rotation angle ¢ in the visible region


for a quartz plate of d = 4 mm, associated with colors and wavelenths A.

Color A (nm) ¢
red 600 63°67'
yellow 570 68°13'
green 540 78°6'
violet 420 130°

rotating quartz and the other from right-rotating quartz. If a linearly


polarized light beam falls on the sample normal to its face, two outgoing
beams are created, both having linear polarization, with one rotated
clockwise, the other rotated counter-clockwise by the same angle ¢. The
value of ¢ is dependent on the wavelength:

d
¢= K.' A2

where the constant K. is approximately equal to 10-8 radian x centimeter


and d is the thickness of the quartz plate. Typical rotation angles for a
4 mm thick plate are given in Table 5.1.
One can observe the dependence of the angular rotation on wave-
length using the experimental arrangement shown in Fig.5.38 without
the optical filter. White light from a lamp passes a polarizer and then
both quartz samples. Rotation of an analyzer placed after the quartz
plate by an angle r.p provides a colored modification of the two halves of
the sample image. When the principle directions of the polarizer and
analyzer are parallel, both parts of the sample appear as a violet field
and a boundary between the two quartz plates is absent. A small rota-
tion of the analyzer yields different colors for both parts of the image,
so the boundary between them is clearly observed.

6.2 The Fresnel experiment


According to a hypothesis by FRESNEL, the optical activity of quartz
is explained by a difference in the velocities of right-handed and left-
handed polarized light passing along the optical axis of the quartz plate.
In other words, the refractive indices of two waves circularly polarized
in two opposite directions are different. Now we consider an experiment
verifying FRESNEL'S hypothesis.
Two 90°-prisms, one made of right-rotating and another ofleft-rotating
quartz, are positioned to form a rectangular solid as shown in Fig. 5.40.
Even though it has parallel entrance and exit surfaces, this rectangular
198 DEMONSTRATIONAL OPTICS

Quartz Analyser
prisms
He-Nelaser

Vertical slit Telescopic system


fJ4 plate Screen

Figure 5.40. Setup for repeating the Fresnel experiment with two quartz prisms, one
is cut from left- , the other from right-rotating quartz.

solid acts as a complex prism. This solid is placed in a linearly polarized


light beam coming from a narrow vertical slit which is illuminated by
the polarized radiation of a He-Ne -laser (>. = 632.8 nm) (Fig.5.40). A
narrow ray of light from the slit falls normally on the face of the first
prism, hence it propagates within the prism without refraction. The
linearly polarized light can be treated as being composed of a left- and
a right- circularly polarized wave, travelling inside the prism with differ-
ent velocities and , therefore, refracted on the inclined exit surface of the
prism at different angles. A spatial separation of these waves will oc-
cur . Then the waves fall on the inclined face of the second prism, which
increases the spatial separation. However, the total angular separation
between the waves after the complex prism is rather small and cannot be
observed directly on a screen. By means of a telescopic system, which
increases the angular separation, two separated images of the slit are
formed on the screen . The operating principle of the telescopic system
is shown in Fig.5.41. In this case the light rays passing the objective
give rise to a real image of the slit , and then the eye-piece produces a
magnified image on the screen.
The separation between the two circular components of a linearl y
polarized wave can be increased using a combination of three prisms as
shown in Fig.5.42 . n r and n/ are the refractive indices for the right-
and the left- circularly polarized wave, respectively. n r < n/ holds for
the first and the third prism, and n r < n/ for the second one. The
angle of refraction of the right circularly polarized wave is smaller than
that of the left circularly polarized wave when the waves leave the first
prism. When passing the second prism , the right- circularly polarized
ray is refracted to a higher degree than the left- circularly polarized ray,
and the angle between the two directions of propagation is effectively
increased. Finally, the spatial separation between the rays increases a
third time after passing the last prism , as shown in Fig.5.42.
An analysis of the polarization of the outgoing beams may be car-
ried out by placing both a quarter-wave plate and an analyzer into the
beams. Circularly polarized light may be decomposed into two linearly
OPTICAL ANISOTROPY 199

Figure 5.41. The operating principle of the telescopic system. A short line p within
the plane of the vertical slit is projected by the objective as a real inverse image p'
near the focal plane of the objective. This image is then transformed by the eye-piece
of the telescope into an enlarged real image p" on the screen. The objective is close
to the slit, whereas the distance between the eye-piece and the screen is about 9 m.

Figure 5.42 Spatial sepa-


ration of two circularly po-
larized beams arises from
Left quartz
a linearly polarized light
beam, characterized by the
E

L
electric vector E . This sep-
aration is caused by two
right- and one left-rotating
quartz prisms. The first
boundary between adja-
cent prisms provides an ini-
Right quartz Right quartz tial spatial separation for
these circularly polarized
beams, whereas the second
boundary enlarges the sep-
aration.

polarized waves having the same amplitudes, mutually orthogonal po-


larizations and a phase difference of 7r/2. As these directions may be
arbitrarily specified without any special conditions or requirements, let
the y-direction coincide with the principle direction of the analyzer, so
that Ex and E y are the amplitudes of the electric vectors of the mutually
orthogonal linearly polarized waves.
The quarter-wave plate placed between the quartz prisms and the
analyzer transforms the circular polarization of the incident beam into a
linearly polarized the output beam, which then propagates through the
analyzer. The polarization transformation introduced by the retardation
plate results in an addition of the phase of T7r/2, depending on left or
right circular polarization. After the analyzer, one observes two mutually
200 DEMONSTRATIONAL OPTICS

Mirror plane y

.-.-.-....--
A~
UW
.- . . '
.- - _.-
a
--- - - - - - - a'

a b

Figure 5.43. A spiral and its mirror image. The point A of the first coil of the spiral
is reflected across the mirror plane to the point A'. Moving a short line a from point
A over the spiral causes th e line a to rotate clockwise, whereas its image a' rotates
ant iclockwise (a) . An electrical current i induced by the electric field E of the incident
wave has two orthogonal components i", and i y (the third i z component is not shown )
(b) .

ort hogonal linear polarizations. Hence, for a position of the analyzer


corresponding to the maximum intensity for one image , the intensity
of other image, having the opposite direction, will be reduced to zero.
Therefore , when rotating the analyzer, one will find that the two images
of the slit will reach their maxima and minima of intensity at alternating
angles.

6.3 Radio - wave rotation


The nature of the optical activity of a substance is due to a specific
asymmetry of microparticles. This fact may be easily demonstrated
using micro-wave radiation. A mechanical model for this type of micro-
asymmetry is a spiral. A spiral and its mirror image are presented in
Fig. 5.43,a. Assume that a right-handed spiral is reflected by a mirror;
its mirror image is a left-handed spiral, providing we look from behind
the spiral.
It is obvious that these two objects cannot coincide, but they can
be considered to be two asymmetric modifications of one object, which
has no mirror symmetry. All the substances showing features of optical
act ivity do not have mirror symmetry, and they are available in nature
with both symmetry modifications - one is left-handed and the other
is right-handed. In addition to the crystal lattice of the two modifica-
OPTICAL ANISOTROPY 201
tions of natural quartz, among others, the molecules of sugar and the
macromolecules of DNA show the same symmetry properties.
We shall now describe an experiment with ultra-high frequency radio
waves of A = 3 em in order to illustrate the principle of optical activity.
This model allows the demonstration of the rotation of a linear polar-
ization of radio waves by wire spirals. Two types of spirals (left- and
right-handed) were produced from a conductive wire coated with dielec-
tric material. The spirals have a length of about 1 em and a diameter of
0.5 em. The spirals were put into two boxes with one type in each box.
First the emitting and detecting horns of the radio wave installa-
tion are fixed vertically and opposite to each other, then the detecting
horn is rotated around its axis until a minimization of the output sig-
nal (Fig .5.44) occures . The emitting horn radiates a linearly polarized
wave, and the detecting horn can only detect a linearly polarized wave.
At mutual displacement of both horns, as considered above, is simi-
lar to the effect of two crossed polarizers. The box with the spirals ,
when placed between the horns , results in an effective rotation of the
linear polarization of the wave while passing through the box. If the box
contains left-rotating spirals, one observes a left-handed rotation of the
polarization plane . To reduce the resulting signal we have to turn the
detecting horn clockwise, providing we are looking in the same direction
as the incident wave. The box containing right-rotating spirals results
in an opposite effect (right-handed circular rotation) that requires us
to rotate the detecting horn counter-clockwise to reduce the resulting
electric signal.

E, E

J. ...
Detecting hom

Emitting hom
Spirals

Figure 5.44. Demonstration of the rotation of the polarization plane of a radio wave
by wire spirals. The distance between the emitting and detecting horns is about 30
cm . A 4 cm wide paper box contains the wire spirals and is placed between these
horns. The rotation of the detecting horn around its axis allows verification of the
polarization state of the transmitted wave.
202 DEMONSTRATIONAL OPTICS

Polarizer Analyzer

Glass samp le

Lamp

Filler

Magnet

Figure 5.45. The setup for observing the Faraday effect with a glass sample. The
sample is placed between the pole shoes of an electromagnet along the lines of the
magnetic field H . Due to dispersion, the angle of rotation depends on wavelength; an
optical filter is therefore inserted into the beam .

A qualitative explanation of this rotation is based on the idea that


electric currents over the spirals are induced by the external electric field
of the incident wave. For a given sort of spiral and direction of prop-
agation of the incident wave, a current over a certain spiral has three
components. Two of them are practically in the plane of the coils, hav-
ing components i x and i y as shown in Fig.5.43,b. These components,
having practically equal magnitudes, radiate a circularly polarized elec-
tromagnetic wave with a certain handedness, depending on the way the
spiral is wound. Therefore, the joint effect caused by a large amount of
identical spirals will be a reduction of the wave velocity for one circular
polarization and an increase in the velocity for the other sort of circular
polarization. The difference in the phase velocities of left-handed and
right-handed circular polarizations leads to the rotation of the polariza-
tion plane .

6.4 The Faraday effect


Some normally amorphous and microscopically homogeneous trans-
parent substances showing no optical activity become optically anisotropic
with respect to propagation of circularly polarized waves when inserted
into a magnetic field.
This phenomenon is called the FARADAY effect. In order to demon-
strate this effect, let a light beam from a source, e.g. a mercury lamp,
fall through a glass bar (length approximately 3 cm) placed between the
two pole shoes of an electromagnet. An optical filter and a polarizer
provide a quai-monochromatic and linearly polarized light beam. After
the light has passed through the glass, one observes a rotation of the
polarization direction similar to optical activity. The value of the angle
OPTICAL ANISOTROPY 203
of the polarization rotation is tested by an analyzer (Fig.5.45). This
angle depends on the magnetic field H, the length of the glass bar l and
the wavelength >.:

<P = f3(>.)IH cos((H) (5.15)

where f3 is a characteristic parameter of the glass material (called VERDET'S


constant) , (H is the angle between the direction of the wave vector k
and the magnetic field H . As we see, for a given field H the sign of
rotation depends on the direction of k. If, for instance, (H = n, the
wave propagates against the direction of the magnetic field vector. In
this case a right-handed (clockwise) rotation takes place, when looking
towards the vector k. If (H = 0, the value of the rotation will be the
same, provided the same conditions are held. In fact, the value of the
rotation angle can be enhanced twice, if the light passing through the
FARADAY glass and a second time back. This effect is used in a lot of
applications, for example in laser physics: A FARADAY rotator inside
the ring cavity of a laser can force additional losses for one direction of
light propagation. A FARADAY rotator, causing a 45° rotation of the
polarization plane , can prevent a laser resonator from being disturbed
by back-reflected light (the rotator is then called an optical isolator).

6.5 Magnetic domains


Thin films of iron or nickel doped with rare earth elements are trans-
parent for visible light but give rise to a high rotation of the polarization
vector. The presence of spontaneous domains of magnetization (or mag-
neti c domains) , which are typical for all ferromagnetic materials, is re-
sponsible for strong and well-expressed magnetic field effects, includ ing
the polarization rotation.
Using a thin film of the ferromagnetic material FeDy03, it is possible
to achieve large magnetic domains which can have the same dimensions
as t he thickness of the film. In t his case two possibilities exist : the ef-
fective magnetic field of the domains is normal to the surface of the film
or it is anti parallel. The general state of such a macroscopic ferromag-
netic domain pattern is governed by the principle of minimum magnetic
energy of a ferromagnetic body (Fig.5.46) .
We turn now to a description of a simple experiment demonstrating
this effect. A linearly polarized light beam falling normally onto the
film surface passes along the magnetic field lines of the micro - domains
(Fig.5.47). Thus a rotation of the polarization in the output beam takes
place. The signs of the rotation angles are contrary for two opposite
directions of the local magnetic fields within the domains. In order to
204 DEMONSTRATIONAL OPTICS

Figure 5.46. Two magnetic domains, - one is marked as the bla ck region , and the
ot her as the bright one - have magnetic moments m directed in opposi t e direction s.
Both vecto rs m are normal to the faces of th e thin film of ferromagnet ic material.
T wo possible directions of linear polarization E 1 and E2 exist in the outgoing beam,
both inclined with resp ect to the electric vector E of the incident wave. The bright
region is associ at ed with E 1 and the da rk one with E2 .

Eyepiece
of microscope
Objective of
Non-polarized microscope
light beam
, - , -, - , - , -~-

Eye
Film of FeDy03

Polarizer Analyzer

Figure 5.47. Setup for observing magnetic domains in a thin film of Fe D y03 by
applying linearly polarized light.

observe the domains, an analyzer should be placed in the beam after t he


film.
Due to the action of the analyzer, th e two kinds of domains produce
different intensities. If the polarization of light passing through domains
with identical directions of the magnetic field corresponds to the prin-
ciple direction of the analyzer, these domains will be represented at the
screen as bright local regions. Adjacent domains having local magnetic
fields in opposite direction will be displayed as dark regions.
A pattern generated by a film of the orthoferromagnetic material
F eDy03 is shown in Fig.5.48 ,a. A film with a thickness of about 0.1
mm was fixed on a microscope table and illuminated by linearl y po-
larized light. The analyzer was positioned after the eye-piece. A bar
magnet located near the film gives rise to re-orientation of the domains.
Applying this external field, the number of domains having local mag-
OPTICAL ANISOTROPY 205

o 0.3 mm
a b

Figure 5.48. The patterns observed by means of a microscope with 70-fold magni-
fication . The pattern of the face of the ferromagnetic film in absence of an external
magnetic field (a) . When a weak external magnetic field is applied, the light and
dark regions of the original move, which may result in two regions, corresponding
to opposite directions of magnetic moment. One is large and dark, and the other is
bright and narrow (b) .

netic fields along the external magnetic field lines increases whereas the
number of the domains having an opposite direction decreases . The vis-
ible effect of such an orientational motion in the external magnetic field
will be an expansion of the boundaries of the local dark areas at the
screen, as shown in Fig.5.48,b .

SUMMARY
As we have seen, optical anisotropy is brought into existence by the
fact that waves of different polarization states propagate inside an opti-
cally anisotropic medium with different phase velocities.This anisotropy
was most thouroughly discussed for the crystalline lattice of calcite,
which possesses a symmetry axis of the 3rd order.
In general, the simplest model of a uniaxial crystal is a layer structure
with a symmetry axis. Such a layer structure model is sufficient for a
qualitative explanation of the phenomena of optical anisotropy of liquid
crystals and of artificial anisotropy of solid and liquid dielectrics.
Microstructures of bodies capable of rotation of the polarization plane
are more complicated. As a rule the positions of atoms and molecules
of such a substance resemble a screw line or a spiral. A number of
optically isotropic media show polarization plane rotation when inserted
into a magnetic field. This phenomenon is known as the FARADAY effect.

PROBLEMS
206 DEMONSTRATIONAL OPTICS

0'
o
-----_ .. _- ---------
Figure 5.49.

Figure 5.50.

5.1 A beam of natural light falls on a plane plate of calcite at the angle
()i= 45°. The plate is cut in such a way that its optical axis is normal
to the plane of incidence (Fig.5.49) . Calculate the angular separation of
the refracted rays, if the indexes for ordinary and extraordinary rays are
no = 1.658 and n e = 1.486, respectively.

5.2 A linearly polarized beam passes through two quarter-wave plates


(Fig.5.50). The first is made of mica and the second of crystalline quartz.
The optical axes of the plates are parallel to each other. The polarization
direction of the incident wave is at an angle 45° with respect to the
optical axis of the first plate. What is the polarization state of the
outgoing beam? How will the polarization change if the second plate is
also made of mica? Take into account that mica is a negative crystal,
n e - no < 0, but crystalline quartz is a positive crystal, n e - no > O.

5.3 It was obtained experimentally that a plate of quartz (4 mm thick)


rotates the polarization plane of an incident wave by ip = 64° for red
light (..\ = 600 nm) and by cp = 78° for green rays (..\ = 540 nm). Esti-
OPTICAL ANISOTROPY 207

Figure 5.51.

~ ._ . _ . _. _._ .- ·-b -· _ .- .- . _. _ ._ . _ ._ .- J~y


Figure 5.52.

mate the difference b.n between the indices for left and right circularly
polarized waves for the two wavelengths mentioned above.

5.4 In the FRESNEL scheme for spatial splitting of a linearly polarized


beam into two beams of circular polarization, two rectangular prisms
of identical sizes are used, one cut from left-rotating quartz and the
other from right-rotating quartz (Fig.5.51). The refractive angle of both
prisms is () = 30°. Derive a formula between the angle of spatial splitting
of the beams 8 and the difference b.n of the refractive indices for left
and right circular polarizations, taking into account a very small value of
8. The light beam propagates along the optical axis within both prisms
(entrance perpendicular to the first surface).

5.5. In the experiment with the liquid crystal prism described above,
two spectra are observed on the screen at a distance of b = 2 m from
the objective (Fig.5.52). The colored spectra formed by ordinary and
extraordinary rays are shifted in the vertical direction by a value of b.y =
0.5 em.
The prism angle is 7°. Estimate the difference b.n oe of the indexes for
ordinary and extraordinary rays in the liquid crystal.

5.6. In the experiment presented in Fig.5.47 concerning the rotation


of the polarization plane by a thin film of FeDy03, one may observe
208 DEMONSTRATIONAL OPTICS

a b

Figure 5.53.

I (U)

I I
E
U

a b

Figure 5.54.

the images shown in Fig.5.53. The first image (Fig.5.53,a) occurs if the
principle plane of the analyzer is parallel to that of light waves passing
through regions associated with a certain direction of magnetic moments .
When rotating the analyzer further , the contrast of the image de-
creases to zero, and then a new distinct image appears at a certain
position of the analyzer. This image is the inverse of the first one
(Fig.5.53,b). Estimate the difference D.n of the indices, corresponding
to left and right circularly polarized waves, at A = 600 nm for a film
thickness of d = 0.1 mm. The analyzer has to be rotated by an angle
D.r.p = 40° to come from one image to the other.
5.7. A linearly polarized laser beam falls through a POCKELS cell as
shown in Fig.5.54.
The direction of polarization of the laser beam is normal to the prin-
ciple direction P of the polarizer. It is well known that the intensity of
OPTICAL ANISOTROPY 209
M
Q
p

Figure 5.55.

the transmitted light beam will vary with the voltage U as follows:

I(U) = IoXU2
provided that U is rather small and I(U) = 0 at U = O. X is a pro-
portional constant. Nevertheless, in practice a linear dependency of the
intensity on the applied voltage is often needed . A simple way to pro-
vide the required linear dependency of the intensity variation is to place
a phase plate in front of the crystal (e.g. a quarter-wave plate). Explain
the action of the quarter-wave plate.

5.8. When producing a quarter-wave plate by cleaving natural mica


plates, a test setup is needed to show that the retardation is close to
>../4. A simple setup which is used to test quarter-wave plates is shown
in Fig.5.55. Light from a source is focused on a pinhole and then is
collimated by an objective into a parallel beam. After passing the optical
filter F , which selects the desired wavelength , the beam is divided by
a beam splitter S. The reflected beam is passes a polarizer P and a
quarter-wave plate Q. A mirror M reflects the beam back to the beam
splitter and then to an observer. Describe the operating principle of this
setup.

5.9. A device generally used to compensate differences in phase retar-


dation is BABINET's compensator. It is usually made of two crystalline
quartz wedges, each with its optical axis in the plane of the face but
with the optical axes rotated by 900 relative to each other (Fig.5.56).
In one wedge the optical axis is parallel to the edge, in the other per-
pendicular. One wedge is movable in the direction perpendicular to the
edges, so that a plane parallel plate of variable thickness is formed. Let
a linearl y polarized beam (polarization plane at angle () with respect to
210 DEMONSTRATIONAL OPTICS

, p

Figure 5.56.

both edges) pass through the compensator and then through an ana-
lyzer. Derive a formula for the phase difference between extraordinary
and ordinary rays for a given wavelength >. and find the angle of the
principle direction of the analyzer and the positions of the dark and
bright bands of the field after the analyzer.

5.10. BABINET'S compensator is adjusted by means of a polarizer and


an analyzer for a certain wavelength >., so that the positions of dark and
bright bands in the light transmitted through the analyzer are known.
Then an elliptically polarized light of the same wavelength is examined
by means of the compensator and analyzer. Derive a formula for the
phase difference 8 between the orthogonal components of the ellipse, and
characterize the ellipse. Use results obtained in the previous problem .

SOLUTIONS
5.1 Since the incident ray is normal to the optical axis, the angles of
refraction Bot, Bet for the ordinary and extraordinary rays are calculated
as follows:
nosin Bot = sin Bi
Substitution of the numerical magnitudes gives Bot ~ 25°B and Bet ~ 280 ,
hence the angular separation between these rays is 8 = 30 •

5.2. Taking the z--axis perpendicular to the optical axis of the quarter-
wave plate of mica, the orthogonal components of the electric vector Eo
of the incident beam may be represented as follows:
OPTICAL ANISOTROPY 211

Figure 5.57.

The ordinary and extraordinary rays are retarded by the mica plate:
2~ 2~
'Px = Tdno, 'Py = Tdne
Since mica is a negative crystal (ne- no < 0), the phase difference 6 =
'Py - 'Px is negative and (per definition for the quarter-wave plate) equal
to - ~ /2. Hence, the components of the electric vector after passing the
quarter-wave plate take the forms:

e; = ~ Eo sin(wt + 'P) ,

where 'P = 'Po + 'Px ·


Crystalline quartz is a positive crystal (ne - no > 0). Since the optical
axes of both plates are parallel one to other, the ordinary beam in mica
will become the extraordinary beam in quartz, and, in turn, the extra-
ordinary beam in mica will be ordinary in quartz. Hence, the quarter-
wave plate of quartz introduces the phase difference of +~ /2 between
the ordinary and extraordinary beams, and for Ex and Ey one gets

Ex = ~Eosin(wt+'P) and

This means that the outgoing beam has a linear polarization along the
same direction as the incident one.
If the quartz plate is replaced by a quarter-wave plate of mica with
its optical axis parallel to that of the first plate, the phase difference
provided by the second plate will be equal to -at /2. In this case the
resulting direction of linear polarization is perpendicular to that of the
incident beam (Fig.5.57) .
212 DEMONSTRATIONAL OPTICS
p
/
/
/

L /

Figure 5.58.

5.3. Let a linear polarized wave fall on the quartz plate at the point
a as shown in Fig.5.58.
The electric vector of the wave P may be represented by two com-
ponents Land R associated with the electric vectors of left and right
circularly polarized waves propagating within the quartz crystal. These
waves get the phase increments <h and <PR while propagating within the
crystal from A to point B (on the exit face of the crystal). If d = AB is
the thickness of the plate, the phase increments may be represented by
the following expressions:
2~ 2~
<PL = nLd-r <PR = nRd-r
where tit. and nR are the refractive indices for left and right circularly
polarized waves of wavelength >.. Due to the difference An = nR - »i:
the resulting electric vector P' composed of two components L' and R'
rotates at some angle <p just before leaving the plate. In the case pre-
sented in Fig.5.58 the component L' gets a smaller phase increment than
R', thus right rotation takes place. Since <p = <PR - <PL, the magnitude
of An can be estimated using
x
An = <P21rd
Substitution of the numerical values for sp, >. and d results in the follow-
ing: Anred = 2.7.10- 5 for red rays and Angreen = 2.9 .10- 5 for green
rays.

5.4. The linearly polarized beam falls normally on the face of the first
prism as shown in Fig.5.59, hence the angle of incidence of the beam on
the second, refracting face is equal to ().
Due to the difference An = nR - nL between the refractive indexes
nR and nL two refracted waves exist outside the prism: one is the right-
OPTICAL ANISOTROPY 213

Figure 5.59.

hand circularly polarized wave refracted with the angle D:R and the other
is the left-hand circularly polarized wave refracted with the angle D:L.
We can write the following expressions for the angles of refraction:

where nR > ni: The refractive face of the second prism is parallel to
that of the first prism , thus the angles of incidence of these circularly
polarized waves on the second prism are equal to D:R and D:L. Since
n.i. > nR is valid for the second prism, the refractive angles f3R and f3L
may be expressed in the following form:

which, with the previous expressions , results in


. f3 R
sm = -nR . ()
sm an d . f3 L
sm , ()
ni. sm
= -nR
nL
Now, the small angular difference 8 = f3R - f3L may be found as follows:
sinf3R - sinf3L = sinf3R - sin(f3R - 8) ~ 8 cosf3R. On the other hand, we
have sin f3R - sin f3L = (nR/nL - nL/nR) sin () from the last expressions.
Hence, for the small angular difference 8, we get the expression 8 =
(sin ()/ cos f3L )(nR/nL - nL/nR) . In fact the angular splitting has a very
small magnitude with respect to the angles of refraction. The angle of
incidence () is also very small, so we can use the approximation f3R ~ ().
Therefore, the expression for 8 takes the form

uJ: -_ - f
sin- [nR
-- - -nL] -_ tan e [nR
- - -nL] -_ tan ()nh - ni
cos () »i. nR nt. nR nRnL
We assume that both indices tit. and n« differ from a mean magni tude of
refractive index no to a very small extent ±6.n = ±(nR-nL). Therefore,
by introducing 6.n and no, the expression for 8 may be transformed:
J:
u = tan
()nh - ni ~ tan
O(nL + nR)6.n
2 ~
026.n
tan - -
nRnL no no
214 DEMONSTRATIONAL OPTICS

Figure 5.60.

In the particular case of e = 30°, we finally get 8 = (2J3/3)(~n/no).

5.5. In order to estimate the difference ~noe, we have to find a relation


between the small angle 8 made up by ordinary and extraordinary rays
and the refractive angle of the liquid prism B (Fig.5.60).
Let us consider a ray falling on the refractive face of the prism at
an angle 'Y. After refraction, the ordinary and extraordinary rays are
deflected at the angles cPo and cPe, respectively, so that the following
expressions are valid:

sin cPo = no sin('Y + B) , sin cPe = n e sin('Y + B)


The refractive angle of the prism B ~ x [h. was given to be 70 = 0.12 rad.
We assume the angle 'Y to be of the same order as B, hence both angles
are sufficiently small to approximate sinb + B) ~ 'Y + B. This is also true
for the angles tPo and cPei hence, sin cPo ~ cPo and sin cPe ~ cPe. With these
assumptions we may get the following relation between 8 = tPo - cPe,
~noe = no - n e and the sum 'Y + B :

The angular separation 8 of the refracted rays may be estimated by


the ratio 8 ~ ~y [b = 2.5 . 10-3 rad . Since 'Y is the angle between
divergent rays leaving the objective, the value of'Y can be approximated
by the ratio 'Y/2 ~ h/b= 0.00125. Therefore, we can neglect 'Y with
respect to B. Finally 8 = ~noeB, which gives the following estimation:
~noe = 8/B ~ 2 .10- 2 .

5.6. This problem is similar to that which has been solved in 5.3. and
we can use the result obtained above:
oX
~n = <fJ 21C'd '

where ip is the angle of rotation of polarization plane (Fig.5.61). If we


assume the film of FeDi03 is to be regarded as a parallel plane plate,
OPTICAL ANISOTROPY 215
p

L
,

Figure 5.61 .

I(U)

612
,0 P
I

6tpl
611
----- I
I
I
E I
I I
U
6U 6U

a b

Figure 5.62.

the two angles shown in Fig.5.61 have the same absolute value, provided
that the initial position of the polarization plane of the incident wave is
that shown by the vertical dashed lines. Hence, for the positive angle ip,
t1tp = 2tp is valid. Therefore, the value of t1n may be estimated from
the expression

Substitution of numerical magnitudes results in t1n = 3.3 . 10-4 .


216 DEMONSTRATIONAL OPTICS

5.7. We assume that the optical axis of the quarter-wave plate 0


placed in front of the crystal has a small angle l:i.cp with respect to the
principle direction of the polarizer P (Fig.5.57,a). The electric vector of
the light beam transmitted through the quarter-wave plate will, there-
fore, have a small component Ell = Eol:i.cp, where Eo is the amplitude
of the incident laser beam and Ell is parallel to the principle direction
of the polarizer. Therefore, even with the applied voltage U = 0, a
transmitted light beam of the intensity III = E1il:i.cp2 is found behind the
polarizer. Since the phase angle l:i.cp is assumed to be small, the action
of the plate is similar to applying a constant voltage Uconst to the faces
of the crystal. Now, increasing the voltage from U = 0 to Ul has the
same effect as an increase from Uconst to Uconst + U, in the case without
a quarter-wave plate. If an alternating voltage , varying between U = 0
and Ul, is applied to the crystal, this voltage will increase the phase l:i.cp
by an additional phase 8. In this case the intensity is dependent on the
phase angle as follows:

where the term 82 is omitted due to 8 « l:i.cp, which also means l:i.U «
Uconst ' It is seen that under these conditions, variations of the intensity
occur practically linearly with that of the alternating voltage . It is clear,
that, for increased ti.cp, a certain variation of U provides a larger varia-
tion of I(U) (see Fig.5.62,b). Note again that with the phase plate for
U = 0, the transmitted intensity is not zero!

5.8. If the test plate has a retardation of nearly A/4 and the principle
direction of the polarizer makes the angle 45° with one of the axes of
the quarter-wave plate, then the transmitted light falling on the mirror
M will be nearly circularly polarized. For example, let this polarization
be right -handed. The light reflected by the mirror has a left circular
polarization state and will be transformed into a linearly polarized one
after passing the quarter wave plate again but with the direction of
polarization normal to the principle direction of the polarizer. Therefore
the observer will see a dark field. If the quarter wave plate is rotated,
there will be four positions per complete rotation where the field will
be dark. If the test plate is slightly thicker than A/4 but still placed at
45° from the principle direction of the polarizer, the light beam will be
slightly elliptically polarized after transmission through the retardation
plate, with the major axis of the ellipse normal to the principle direction
of the polarizer. The field will not become completely dark when turning
the imperfect quarter-wave plate.
OPTICAL ANISOTROPY 217

y y
p

b
a1 x

x
A
C ~h~
h2
c

a .. X

5.9. Let the x-axis be normal to the edges of the quartz wedges
(Fig.5.58,a), and let first a light beam of small diameter (e.g. a laser
beam) pass the arrangement. The amplitudes of the electric vector are
given by

Ex = al cos(wt + cpo) and E y = a2cos(wt + cpo)

The phase retardation of the horizontal beam component after passing


the first wedge, in which it runs as an ordinary ray, is knoh l , where hI
is the thickness of this wedge where the ray passes. This ray is retarded
by kneh2 in the second wedge (where it runs as an extraordinary ray) ,
where h2 is the appropriate thickness . Thus, the total retardation for
the ordinary ray is
Ox = knoh l + kn eh 2
The same magnitude for the vertical beam component is given by

The phase difference 0 between the two orthogonal components is

If 0 = 7rm (m = 0, ±1, ... ) linearly polarized light will result, for which
the directions of polarization follows the relation

which means the principle direction of the analyzer has to be inclined


by the angle 0: = 2B with respect to that of the polarization plane of the
incoming light (Fig.5.63,b).
218 DEMONSTRATIONAL OPTICS

If the compensator is illuminated by an extended plane wave, the


difference h2 - hI changes across the area of the light beam (0 = o(x),
Fig.5.58 ,c). Dark bands appear after the analyzer parallel to the edge
at places for which
.A
h2 - hI = m 2( _ )
ne no
is true. The spatial separation between these bands depends on the edge
angle (3 and on the difference (n e - no).

5.10. By using the results of the previous solution, we can write

Ex = al cos(wt + epx) and E y = a2 cos(wt + epy)

for the orthogonal components of the elliptically polarized light, where


the phase difference 00 = epy - epx. The phase difference between the or-
dinary and the extraordinary ray introduced by BABINET's compensator
is calculated as in Problem 5.9. Hence, one gets a total phase difference
Otot = 0 + 00:
0= 00 + k(n e - n o)(h2 - hI)
If Otot = rrrn, where m = 0, ±1, ±2, ..., the field after the analyzer will
contain a number of dark and bright bands, as considered in the pre-
vious problem . However, the positions of these bands are displaced by
some value Ox (connected with a difference Sli of the thickness of the
compensator plates) with respect to the positions obtained for linearly
polarized light during the adjustment of BABINET 'S compensator. This
value oh is caused by 00 . The magnitude of 00 may be found from the
measurement of Ox. oh is connected to Ox by oh = Ox tan (3, where (3 is
the wedge angle. From the relation

we determine the original phase difference

The orientation of the analyzer parallel to the direction of the linear


polarization of the outgoing beam allows to determine the amplitude
ratio a2/al and therefore to calculate the position of the principle axes
of the polarization ellipse of the incoming light.
Chapter 6

GEOMETRICAL OPTICS

The concept of light beams propagating as straight trajectories within


transparent homogeneous media is widely applied when optical instru-
ments are designed . These light beams can undergo refraction and re-
flection at the boundary of two media with different refraction indexes .
Physical effects devoted to the wave nature of light, like light diffraction,
are neglected in this case. In other words, one treats the linear dimen-
sions of all apertures as very large compared to the wave length (d >> >')
(or one considers the wavelength to be zero). The area of optics dealing
only with phenomena of light beam reflection and refraction is known
as geometrical optics .
Optical instruments constructed by means of the principles of geomet-
rical optics are used for different purposes in science and other applica-
tions, mainly to create optical images of different objects. In principle,
the operation of such optical instruments is reduced to modifications
of the solid angles of propagation of the light beams originating from
an object point. Therefore general properties of optical beams will be
discussed at the beginning of the chapter.

1. Homocentric and astigmatic beams


Let us regard a light beam formed by lines orthogonal to a part of a
spherical surface. For a converging cone (or pencil) of rays , all the ele-
mentary beams intersect at a certain point F (Fig .6.1a) . For a diverging
light pencil (Fig.6 .1b), the extensions of the elementary beams orthogo-
nal to the spherical surface intersect at a point F'. F and F' are called
the focii of the beam. A beam with a focus is called a homocentric beam.
It is obvious that a wave front with a spherical surface corresponds to a
homocentric beam. A special case of a homocentric beam is a parallel

219
O. Marchenko et al., Demonstrational Optics
© Kluwer Academic/Plenum Publishers, New York 2003
220 DEMONSTRATIONAL OPTICS

light beam. In this case the focus of the beam is located at infinity and
the wave front is a plane surface.
A plane mirror is an example of a simple optical device, in which
an image can be constructed with the help of homocentric beams. The
light beams SSI and SS2 originating from a point source S located at a
distance h from the surface of the mirror (Fig.6.2) will propagate along
the linear trajectories SlS~ and S2S;

s S'
2

a b

Figure 6.2. Reflection of rays from a


Figure 6.1. Homocentric beams. point source 8 on a plane mirror. A
Covergent beam (a ) and divergent virtual image 8' is the intersection of
beam (b ). Points of intersection F, F' the extensions of reflected rays 818~
are focii. and 828 2,

As already noted in the introduction, the laws and concepts of geo-


metrical optics are approximate to a certain degree, due to neglection
of all diffraction phenomena. In this respect the concepts of the homo-
centric light beam and the focus are valid only within the framework
of geometrical optics approximations. Nevertheless, even within geo-
metrical optics , deviations from the ideal case of a homocentric light
beam can occur, long before diffraction phenomena cause a violation of
the geometrical optics approach. In the general case, after propagation
through the boundary of a transparent medium, the shape of the wave
front changes. The wave front forms a curved surface, shaped unlike a
plane or spherical surface. Light rays contained within a pencil originat-
ing at the wave front are all normal to the curved surface. It is evident
that when the wave front surface has an arbitrary form, the pencil of
rays will no longer be homocentric. Properties of such pencils can be
found by geometrical characteristics of the curved surface.
Let us regard a small element of a curved wave front ABeD around
point 0 (Fig.6.3). It can be shown by differential geometry that differ-
ent cross sections (e.g. HOF and EOG) of this surface have different
Geometrical Optics 221

Figure 6.3. An astigmatic pencil of rays. A small element of a curved surface ABGD
is limited by circular arcs AB, GD and AD, BG. Cross sections GGE , AH Band
DFG, having the same radius of curvature, are mutually parallel and have centers
of curvature, located on the short line bb'. Another type of arc presented by parallel
cross sections AED, HOF and BGG is specified by a shorter radius of curvature.
The centers of curvature lie on the short line aa'.

curvatures, which are characterized by different radii r and centers of


curvature C located on a line which is orthogonal to the wave front at
point O. The cross sections corresponding to the largest and the small-
est curvature radii are mutually orthogonal. Let the minimal curvature
radius Rrnin correspond to the cross section HOF and the maximal cur-
vature radius R rnax correspond to the cross section EOG. Both cross
sections are arcs of circles. Two of them (AH Band DFC) may be taken
to be vertical , and the other two (AED and BGC) to be horizontal. The
lines normal to the surface ABCD drawn from the points H,O and F
will intersect in point C1 which is the curvature center with the mini-
mal radius, located a distance r1 from the surface. The perpendiculars
to the surface originating from the points E,O and G will intersect, in
turn, at the point C2, located a larger distance r2 from the surface . The
point C2 is the center of the largest curvature of the surface ABCD.
Two vertical cross sections AHBand DFC (parallel to EO G) will have
the same curvature radius r1 . The centers of curvature of these cross
sections are located at points band b'. It can be seen that this line is
orthogonal to OC2. For the horizontal cross sections BGC and ACD
(parallel to HOF) the centers of curvature are located at points a and
r
a.
Therefore all the beams perpendicular to the surface ABCD will have
their conjunction points either on the straight line aa' or on bb'. In other
words, beams originating from an arbitrary curved surface will intersect
at points belonging to two mutually perpendicular lines. These lines are
located at different distances from the radiating surface . A light beam
222 DEMONSTRATIONAL OPTICS

with such properties is called a normal astigmatic beam. Lines aa' and bb'
are called the focal lines and the distance between them, R max - Rmin , is
the astigmatic difference. It is evident that a pencil of astigmatic beams
does not have a single focus point. However, as the astigmatic difference
is reduced, the focal lines will become closer to each other and their
lengths will become smaller . At the limit as r2 - rl --+ 0 the focal lines
will be reduced to a point as the astigmatic beam is transformed into a
homocentric one.
The origin of astigmatism is not linked strictly to light beams origi-
nating from a curved surface. In particular, this phenomenon takes place
under observation of objects through a thick layer of a transparent ma-
terial such as glass or water when the layer is not perpendicular to the
ray SS' (Fig.6.4). Let us demonstrate this optical effect by an example.
A point source S is located under a water layer of thickness h as shown
in Fig 6.4. The homocentric light beam emitted into the surrounding
water from source S is transformed into an astigmatic beam outside the
water due to refraction. Two outgoing light rays SM1 and SM2 , af-
ter refraction propagating outside the water along directions MIPland
M2P2 , will reach an observer positioned at a certain angle with respect
to the water surface. When rotating the picture out of the plane by a
small angle around SO, the rays will still intersect SO at points SI and
S2 , but C 1 moves perpendicular to the plane of the drawing. Therefore,
the ray pencil outside the water has two focal lines, a radial one S1S2 and
a meridian one formed by all points C1 . Now by variation of the angle
of incidence, point Cl will move along the curve SS' between point S
and point S'. This line is called the caustic line. In turn, when rotating
the picture plane around SO , the caustic line SS' will draw a surface ,
which is called the caustic surface.

Figure 6.4 Astigmatism


ansmg by observation of
a point source 8 through
a thick layer of a trans-
»>: s' parent material with a
flat surface. The source ,
at a distance h from the
flat surface, is located on
the perpendicular drawn
through point O. Real rays
51
I 8M1P1 and 8M2P2 seem
I
to originate from point 0 1 .
Geometrical Optics 223
An ideal point light source emits a homocentric light beam . An ideal
optical system creates a point image of the source by transforming this
beam pencil into an outgoing homocentric beam. In principle, all real
elements of every optical system cause astigmatism as described above.
When constructing an optical system, the problem consists mainly in
a reduction of the inherent astigmatism to a reasonable level. Then
the system will create a homocentric outgoing beam with satisfactory
accuracy.
The operation of an imaging system is based on the transformation
of the solid angles confining the incoming and the outgoing light beams.
This transformation usually occurs due to refraction of the incident light
beam at boundaries of optically transparent materials, most often glass
and quartz. These boundaries, called optical surfaces , usually have a
spherical shape. The requirements for these shapes are very strict. The
deviation between the shape of a high precision optical surface and its
ideal geometrical shape should not exceed some fractions of the light
wavelength (e.g., >./10).

2. One refracting surface


Suppose, that two transparent homogeneous substances are separated
by a spherical surface with the curvature radius r, as shown in Fig.6.5.
Let the optical material on the left of the boundary be characterized by
the refractive index ni and on the right by the refractive index n r . A
straight line passing through the point light source 8 located in front of
the spherical surface and its curvature center C is called the main optical
axis. The conjunction of the main optical axis with the surface is called
the apex of the refracting surface,point 0 in this case. Let the origin
of a Cartesian system be placed at the apex with the positive direction
of the z-axis in the direction of light propagation. As usual, one can
regard counterclockwise angles to be positive, and clockwise angles to
be negative .
Let us regard a light beam 8M falling on the spherical surface with
an angle i , which, after refraction, propagates with angle () along the
direction M 8'. Following the refraction law (4.16), we obtain
n i sin i = n r sin () .
ni is the refractive index of the medium containing the incident beam and
n r is the refractive index of the medium containing the refracted beam.
Let us restrict ourselves to so-called paraxial beams which form small
angles u and -u' with the main optical axis (the angles are considered
so small that the sinus and tangents functions can be easily replaced by
the values of the angles and the cosine may be considered equal to one).
224 DEMONSTRATIONAL OPTICS

"I n,
F
---L
-----------o
y ~

___.s -----------l --- S/


x _ c

x f-----<~,---- f/---l. Xl
1------ X,----I+-''''---- X2---~

Figure 6.5. Refraction on a spherical surface . (a) The case of refraction by an op-
tically dense medium (n r > n,) . A light ray 8M is refracted by an optically dense
medium and intersects the main optical axis x at point 8' . (b) Definition of the
focal points F and F' of a spherical refracting surface (b). F and F' are located on
opposite sides of the refracting surface . Focal length IOF'1 is longer than IFOI. (c)
Mutual positions of distances z , Xl, x', X2 and coordinates of focii I, !' with respect
to the apex of the refracting surface are given, as needed for application of Newton's
formula .

Under this approximation, which is called the paraxial approximation,


the refraction law takes the simple form
(6.1)
Expressions for the angles i and () may be obtained from the triangles
SMC and g MC as follows:
i=u -cp () = u' - cp (6.2)
Geometrical Optics 225

where -<p is the angle between the main optical axis and the orthogonal
to the surface, eM. According to (6.2) the following relation between
the angles u, - <p and -u' is valid: nl(u - <p) = -n2(<p - u') , or

(6.3)

Let h be the length of the perpendicular from the point M to the main
optical axis, and Xl and X2 be the coordinates of points Sand S' , re-
spectively. Under the paraxial approximation for the small angle -<p
(please note ip < b) the expression -<p ::::: h/r is valid. In the same way,
for the small angles u one gets u ::::: -h/XI (u> 0 , Xl < 0), and for the
small angle -u' one gets -u' ::::: h/X2 (u' < 0 , X2 > 0). Substituting the
angles obtained by the expressions above into (6.3), one gets
n ni nr - ni
-r - - =
X2 Xl r

or alternatively
n21 1 n21 -1
- -- = (6.4)
X2 Xl r
after introducing the value of the relative refractive index nr/nil desig-
nated as n21. In this way we find (using the paraxial approximation)
that the coordinate X2 of the image position of the point source S' can
be expressed in terms of the quantities Xl, rand n21 . The value X2 de-
pends neither on h nor on the angles of incidence and refraction for light
beams with respect to the refracting surface. As a conclusion, it follows
directly that all paraxial beams emitted by a point source and confined
within the homocentric pencil cross the main optical axis at the same
point S' . In other words, a spherical refracting surface under the parax-
ial approximation transforms a homocentric incident light beam into an
outgoing homocentric beam . The quantity on the right-hand side of the
equat ion (6.4) is named the optical force:

p = n21 -1 (6.5)
r

The coordinate of the second principal, or main, focus F' is determined


by the intersection point of all the refracted light beams when a parallel
pencil of light beams is incident on the refracting surface. The second
principal focal length l = OF' is easily deduced from the expression
(6.4) with Xl tending towards infinity (Xl - -00) :

(6.6)
226 DEMONSTRATIONAL OPTICS

By ,definition, the first principal, or main, locus F is the point on the


main optical axis in which a point light source should be located in
order to obtain a parallel pencil of beams after refraction. We designate
FO = I as the first local length. As X2 tends towards infinity (X2 -+ 00)
in (6.4), for the coordinate I we get:

1= _2- = r (6.7)
P n21-1
The following relations are valid for the principal focal lengths:

(6.8)

Substituting the expressions for the main optical lengths (6.6) and (6.7)
into (6.4) we find a relation between the focal lengths of the refracting
surface and the coordinates of the source and its image on the main
optical axis:
L+£=1
Xl X2
(6.9)

We introduce variables X and x':

Xl = X + I; X2 = X' + I'
Substitution of the above relations into the formula (6.7) leads to:

which after a simple transformation takes the form

xx' = II' (6.10)

known as NEWTON 'S formula.


The symmetry of NEWTON 'S formula results from the principle of
light beam path reversal. This principle means, that when considering
S' as a point source, an image of this source is directly obtained at point
S, just by tracing light beams opposite to the direction shown in Fig.6.6.
I
In this case, Xl, X2, and I'
will satisfy equations (6.9) and (6.10). The
points Sand S' are called conjugated points.

2.1 Positive and negative optical force


In the case of refraction considered above, incident rays propagating
along the main optical axis will be refracted towards the optical axis
and always intersect this axis. Therefore, it is convenient to characterize
Geometrical Optics 227

s~---
_ . __._._._;~._._._ . .- ._.
o~ :;-'.
"j '--~r
a b

c d

Figure 6.6. Refraction of a parallel ray by spherical surfaces with positive or negative
optical force. The ray is refracted by the surface with positive optical force (a.b) .
When nr/ni > 1, the vector n makes a sharp angle with propagation direction x
(a). When ni/nr < 1, the vector n does not make a sharp angle with direction x
(b). Refraction of a parallel ray by a spherical surface with negative optical force
(c,d) . When nr/ni < 1, the vector n makes a sharp angle with direction x (c). When
n r/ni > 1, the vector n does not make a sharp angle with direction x (d) .

a refractive surface which causes a convergent pencil of light rays by a


positive optical force P.
Let us first treat a convex surface of the optically denser medium. The
unit vector n perpendicular to the refracting surface and the center of
the curvature C are located inside the medium. Let n21 be the relative
refractive index of the optically denser medium. The case of refraction
illustrated by Fig.6.6,a is similar to the case considered above, where n r -
ni > 0 (n21 = nr/ni > 1) and the refracted ray propagates within the
optically denser medium. In the case presented in Fig.6.6,b the refracted
ray first passes through the denser medium. Hence, the difference n r -
ni < 0 will be negative and will lead to the negative term 1 - ndn2 =
1-n21' Nevertheless , the optical force of the refracting surface is positive
in both cases. Therefore we have to consider the radius r in eqs. 6.6
and 6.7 to be negative. As a rule, r is positive if the vector n makes a
sharp angle with the x-axis (or with the direction of propagation), but
negative if n is directed against the x-axis (obtuse angle). In both cases
of the convex optically dense medium the optical force is positive. The
focal length belonging to the optically denser medium is always larger
228 DEMONSTRATIONAL OPTICS

than the focal length outside this dense medium. For example, the focal
length of a glass spherical surface (n = 1.5) surrounded by air (n ~ 1)
is r . n2I!(n21 - 1) ~ 3 · r , where r is the radius of the surface and the
outer focal length is r- (n21 - 1) ~ 2· r. The ratio of the first value to
the second one is about the relative refractive index of the glass.
In the case of a concave surface, beams originally propagating along
the main optical axis will be refracted in such a way that they will never
intersect this axis. (Fig .6.6,c,d). One can state that a concave surface of
an optically dense medium always possess a negative optical force. The
unit vector n directed along the internal normal to the surface, as well
as the center of curvature C, will always be located outside the optically
denser medium. If n is directed along the propagation of the refracted
ray (Fig.6.6,c) the value of r should be treated as positive, in the other
case, where n is directed in opposite direction, the value of r should
be considered negative (Fig.6.6,d). The optical force will be negative in
both cases . In the case shown in Fig.6.6,c (r > 0) , the difference n r - ni
is negative, the relat ive refractive index of the optically dense medium
n21 = tuf n ; > 1 and the optical force P = (1 - n21)/r will be negative.
When the refracted ray propagates outside the optically dense medium
(Fig.6.6,d) the value r should be treated as a negative, whereas the
difference n r - ni is positive, which again gives a negative optical force
P = (n 21 -1)/r, where n21 = nr/ni' It should be noted that in the case
of a concave optically dense medium, the focal length belonging to this
medium will always be shorter than the focal length outside the medium.
For a glass spherical surface (n = 1.5) surrounded by air (n ~ 1), the
focal length is r- n2I!(n21 -1) ~ 2· r, where r is the radius of the surface,
and the other focal length (in air) is r . (n21 - 1) ~ 3· r.
Conclusions. The formula (6.5) obtained for the particular case of
a positive optical force is also valid for cases where the optical force is
negati ve or for cases of a concave spherical surface of the optically denser
medium. Formula (6.5) implies that for a given direction of the incident
beam and a given position of the apex, one can obtain the optical force
of the refracting surface and compute t he coordinates of the focii and
of the conjugated points. The sign of the optical force is dependent
on th e shape of the refrac ting surface (convex or concave) and on the
relati ve refractive index, which is specific to this surface. The signs
of the coordinates of the focii and the conjugated points follow from
expressions (6.6-6.7),(6.9-6.10)

2.2 Real and imaginary image


Paraxial approximation enables us to find the image of a spacious
object generated by a spherical surface. The spherical surface surround-
Geometrical Optics 229
ing an optically denser medium (refractive index n r ) can give rise to
an image in two different ways (Fig.6.7,a,b) . In the first, a short verti-
cal arrow 8P representing our object is located at some distance from
the first focus F . The ray PM, propagating from point P parallel to
the main optical axis, is refracted at point M and intersects this axis
at the second focus F', as does any ray passing parallel to the optical
axis. Another ray, originating from point P and passing through the
first focus F, will 'propagate parallel to the optical axis after refraction.
The intersection point of both rays p' will be the image of point P .
The foot of the perpendicular line drawn from p' to the optical axis
(point 8') will be the image of point 8 . This graphical method applied
above to the line 8P is also valid in more general cases. If a short line
is located between the first focus and the apex of the spherical surface
possessing a positive optical force, the image created by the surface will
never be formed by the intersection of two refracted rays (Fig.6. 7,b) . In
this case the position of point p' will be obtained by the intersection of
two rays traced backwards from the emitted rays (the ray F P and the
ray P P') . It can be seen that the image 8' p ' is located on the same
side of the surface as the object. Such an image is called an imaginary
image , contrary to the real image obtained in the previous case by the
intersection of refracted rays.
In cases of refraction by a concave spherical surface which forms the
boundary of the optically denser medium, the image will always be imag-
inary (Fig .6.7,c,d) . The reason is that the second focus F' and the object
are both located on the same side of the refracting surface. Even when
the object is at an arbitrary position Xl with respect to the apex, the
coordinate f of the first focus will always have the opposite sign. On the
other hand, the coordinate f' of the second focus always has opposite
sign of f , which implies that the signs of Xl and f' are the same . Now,
following the expression (6.9) one can state that the sign of the ratio
f IXI has to be negative, whereas the sign of the ratio f'lx2 has to be
positive. This means that the position of an image X2 will always have
the same sign as Xl. Hence, in all cases the object and its image will be
located on the same side with respect to the apex, resulting in imaginary
images.

2.3 Magnification by a spherical refracting


surface
For the examples considered above one can conclude that a spherical
refracting surface transforms the size of an object image, and in particu-
lar cases its direction with respect to the main optical axis as well. This
property of spherical refracting surfaces, valid under the paraxial ap-
230 DEMONSTRATIONAL OPTICS

c d

Figure 6.7. Real and imaginary images caused by a spherical refracting surface. A
surface of positive optical force gives rise to a real image of the object SP placed
behind the first focus of the surface (a) , and an imaginary image if the object is
placed between this focus and the apex (b) . Imaginary images produced by a surface
of negative optical force (c,d) . The object SP is located inside the refracting surface
between the second focus and the apex (c) , or outside the surface behind the second
focus (d).

proximation, permits using a simple graphical method for quantitative


measurement of the linear dimensions of an image.
Transformation of the image dimensions from the initial object di-
mensions is described by the linear magnification (Fig.6.7) :
f3 = S'p' / SP
In the particular case shown in Fig.6.7, the ratio S' P' / SP can be
expressed by means of two distances SF and FO . Point P' belongs
to the refracted ray passing along the main optical axis; hence S' P'
is the distance from the apex to the point on the surface, where this
ray is refracted. Therefore, the ratio of distances SF and FO gives
the linear magnification f3 = S' P' / SP = -FO / SF , where the
Geometrical Optics 231

minus sign indicates that the short lines SP and S' P' have mutually
opposite directions. Please note that this is true only using the paraxial
approximation. Using either the ratio Fa I SF = I Ix or NEWTON'S
formula we get the following expression for the linear magnification :

I x'
{3 = -; = - I" (6.11)

This relation in particular shows that the linear magnification grows


while an object approaches the first focus (x - 0). In the case of an con-
vex dense medium (having positive optical force) (3 is negative (Fig.6.7,a)
if a real image is produced, but {3 is positive for an imaginary image
(Fig.6.7,b). Such an imaginary image points in the same direction as
the object.

2.4 A spherical reflecting surface


A commonly used element in optical instruments is a spherical reflect-
ing mirror. The operating principle of the spherical mirror is similar to
that of a single refracting surface of spherical shape. Let us consider the
case of a concave reflecting surface as shown in Fig.6.8.

o x

Figure 6.8. Reflection by a spherical mirror. The positive direction of the z--axis
makes a sharp angle with the direction of light propagation. The center of the cur-
vature of the reflecting spherical surface C is always located in the region of negative
z . So the values Xl , X2 and / are all negative with respect to the apex.

Evidently, the paraxial approximation is still valid for small angles


of incidence and reflection. Let a ray from a point source S, located a
distance Xl from the apex of the mirror 0, make a small angle u with
the optical axis SO, falling on the reflecting surface at point M. In
the Cartesian system (x, y) placed at the apex of the mirror , the source
is located behind the center of curvature Cj hence, the reflected ray
232 DEMONSTRATIONAL OPTICS

will make a sharp angle u' with the optical axis x and intersect this
axis at point P. Let -i and i be small positive angles of incidence and
reflection , and Xl and X2 be the coordinates of the source S and its image
P , respectively. From the reflection law (4.15) we get

-i = i' . (6.12)
For the angle of refraction i' the following relation is valid:
i' = -cp +u'
where cp is a positive angle between the radius CM and the optical axis.
In turn, for the positive angle -i we can get:
-i = -cp+u .
Using these relations and the expression (6.12) we find:
u+u' = 2cp . (6.13)
The paraxial approximation allows the small angles in the last ex-
pression to be represented by the distance h , by the coordinates Xl , x2
and by the radius of the curvature r (the latter should be taken as a
negative, since the coordinate of the center C is negative):
h h h
'P;:::::: - - , u~--, u'~--
r Xl X2

Substitution of the angles in expression (6.13) by their coordinate rep-


resentations gives t he formula
1 1 2
-+-=-
Xl X2 r
With the notation f = r /2, this formula will take the known form
1 1 1
- +- = - (6.14)
Xl X2 f
The coordinate f is that of the focus of the reflecting surface. The
position of the focus is determined by the requirement Xl -+ -00, which
implies that a parallel beam from a remote point source has to intersect
the optical axis at the focus of the mirror.

3. Centered optical systems


In practice, different types of objectives and eye-pieces are used for
physical experiments and in optical engineering. Lenses are the prin-
ciple elements of these devices. Usually made of optically transparent
Geometrical Optics 233
Surface 00. n-1 n

Figure 6.9. A centered optical system . 00' is the main optical axis of the system .
F l is the first focus of the first spherical surface and F~ is the second focus of the n-th
spherical surface of the system. The coordinates of focal points and that of positions
Xl of the source S and that of images X2n are measured from their respective apexes.

materials, lenses have two refracting surfaces of spherical shape. A typ-


ical optical device, e.g. the objective of a photographic camera, may be
represented as an aggregate of lenses with a large number of refracting
surfaces .
For a description of the operation of such devices, the concept of a
centered optical system is exploited. This system is considered as the sum
of spherical refracting surfaces with their centers of curvature located on
a straight line, the main optical axis of the system. Fig.6.9 shows the
first two and the last two refracting surfaces of a centered system.
Let a point light source 8 be located on the main optical axis and let
it emit a paraxial homocentric pencil of light rays. After refraction by
the first surface this pencil will be transformed into another homocentric
one intersecting the axis at a certain point 8~. This point represents an
image of the light source, resulting from the action of the first surface.
The position X2 of this image on the optical axis (measured from the
apex of the first surface) may be obtained with the help of formula
(6.9) for a given Xl and the focal lengths II,J~ . Source 8~ now acts as
the source imaged by the second surface . Supposing that the distance
between the adjacent refracting surfaces as well as the corresponding
focal lengths 12, f~ are known, it is possible to compute the coordinate
X2 of the image of the source S;. 8 2 acts as a source for the third
surface and so on. The final image is formed by consequent application
of equation (6.9) after treating all refracting surfaces.
The fact that a centered system transforms an incident homocentric
pencil of beams into an outgoing homocentric pencil predetermines the
existence of two principal focii of the system, specified by F s and by
F~ similarly to the principal focii of a single refracting surface . A light
beam originating from a point source located at the first principal focus
F s of the centered system is transformed by this system into an outgoing
parallel light beam. Further, a parallel beam incident on the centered
234 DEMONSTRATIONAL OPTICS

system is transformed by this system into a homocentric pencil of beams


with its focus at the point of the second principal focus F~ (Fig.6.1O).

Figure 6.10. Principal planes of an optical system.

Let a beam S M; propagate parallel to the main optical axis at a


distance h. This beam enters the first refracting surface at point Mi ;
will leave the last surface at point M4 and will then pass through the
second principal focus F~ of the system. Within the homocentric pencil
of beams emitted by a source located at the first principal focus F s ,
one will always find a beam FsM2M3Q leaving the system parallel to
the main optical axis at the same distance h from the axis. Extensions
of these two beams (SMIM4F~ and FsM2M3Q) shown by the dashed
lines will cross near the first refracting surface at point P and near the
second surface at point p'. The two planes H and H' drawn through
the points P and p' perpendicularly to the main optical axis are called
the principal planes of the system. The points P and p' of the principal
planes are the conjugated points, as they result from the conjunction of
the two beams FsM2 and M4F~.

--~- - :~: l-:_"----~p,


N N'

Figure 6.11. Construction of the image of point P produced by an optical system


using the principal planes and focii of the system.

The coordinates of the principal focii Fs,F; and the principal planes
H,H' of a system (Fig .6.11) allow the construction of the image gener-
ated by this centered optical system. Coordinates of the object, of its
image and of the principal focal lengths are taken respectively from the
first and the second principal plane .
Geometrical Optics 235
The position of the point source P in Fig.6.1l is determined by the
length M P = Xl. The light beam PM parallel to the optical axis
will pass the second principal plane H' at point M' and passes then
through the second principal focus F~ after refraction by the system.
Another beam originating from point P passes through the first principal
focus F s , intersects the first principal plane at point N, is refracted and
continous parallel to the optical axis (beam N N'). The intersection
point of the beams M' F~ and N' N gives the location of the image p '
of the source. From the triangles N P M and N FsH we get
HFs NH
MP= NM
Because
MP = -Xl and HFs = -Is ,
where Is is the first principal focal length of the system, we get
Is NH
-=-- (6.15)
Xl NM
In a similar manner, from the triangles M' p' N' and M' H' F~ we obtain
I~ M 'H'
(6.16)
X2 = M'N'

Adding equations (6.15) and (6.16) we can write


Is I~ M'H' NH
(6.17)
Xl + X2 = M' N' + NM
Using
M 'N' =NM and MN=M'H'+NH
the expression (6.17) may be transformed to

Is + I~ = M' H' + N H = 1
Xl X2 MN
Finally, we obtain
Is + I~ = 1 (6.18)
Xl x2
This expression is similar to expression (6.9). In other words, it means
that when the principal focal lengths and the source position are known ,
it is possible to find the location of the image using the paraxial approx-
imat ion. If the positions of the principal planes are known, it is possible
to use the method illustrated by Fig.6.1l for a construction of the image
without analyzing the beam behavior inside the system in detail.
236 DEMONSTRATIONAL OPTICS

4. Lenses
4.1 General relations
A lens is a centered system having two spherical surfaces (as shown
in Fig .6.12). Usually a lens is made of transparent material, for instance
glass. In most cases this material is optically denser than the medium
outside the surfaces. The direction of the main optical axis is specified
by the direction of light propagation. Let us introduce two Cartesian
systems, one (x,y) positioned with its origin at the apex of the first
surface , the other (x, if) with its origin at the apex of the second refract-
ing surface. These two spherical surfaces, each having a positive optical
force, will cause refractions of an incident ray as shown in Fig.6 .12,a.
This ray, propagating parallel to the main optical axis at a distance hI ,
is refracted by the first surface and crosses the optical axis at the sec-
ond focal point of the first surface (at the coordinate ff) . It then leaves
the second surface at a distance ba and intersects the optical axis at
the second focus of the centered system. The coordinate f~s is taken
from the apex of the second surface as a coordinate of the system (x, if).
Extending the incident ray parallel to the optical axis from the point
of incidence to a point of intersection with the refracted ray, one can
find the position of the second principal plane H'. The coordinate XH',
which specifies the position of second principal plane , has to be measured
from the apex of the second refracting surface. For given parameters f~
, f 2 and the distance d between the apexes of the refracting surfaces ,
one can find the coordinates f~s and X H' .
With the distance d between the origins of our coordinate systems,
we can link the coordinates x and x :

x=x+d (6.19)

The light ray intersects the main optical axis at point f~ of the (x, y)
system and, at the same time, at point Xl of the (x,if) system:

(6.20)

The point x~ , the conjugate of point Xl, has the coordinate f~s (the
second focus of the optical system) . Hence, formula (6.18) becomes

h/(f~ - d) + f~/f~s =1
when using Xl from (6.20) and x'l = f~s ' From the last relation, for f~s
we get
I.os, =
f'
2d
d - f~
+h - f~
(6.21)
Geometrical Optics 237

-----~- - - - ---

a
x
fI
1

b
x

Figure 6.12. The principal focal lengths and planes of a positive lens. Cartesian
systems (x , y) and (x,Y) are located at the first and second apex of two refracting
surfaces , respectively. An incident ray parallel to the optical axis intersects the optical
axis at the second focus I~ •. The intersection of the outgoing ray with the extension
of the incoming ray defines the second principal plane H ' (a). Another parallel ray
falling on the second surface intersects the optical axis at point 10. after refraction
on the first surface, and its extension defines the principal plane H (b) .

The location of the principal plane X H , can be found by the relation


Xwli~s = (h2 + hI)lh2 . One can find a similar relation for the coordi-
nates if and d: dl(d- if) = (h2+hI)lh2' Combining these relations to
eliminate hI and h2 we get Xw = i~sdl(d - if). Substituting i~s with
(6.21) one can find
I d
(6.22)
X w = i2 d + h - if
In this way it is also easy to obtain expressions for the coordinates of the
first focus is and the first principal plane X H. Let a ray propagating
parallel to the optical axis fallon the second surface and then intersect
the optical axis at point 12.
Let this point have the coordinate x~ in
the (x,y) system. According to (6.19), one obtains x~ = d+ In turn, 12.
being conjugated points in the (x , y) system, the points x~ and ios have
to satisfy the expression (6.18):

iII ios + i~/(h + d) = 1


238 DEMONSTRATIONAL OPTICS

Transforming the last expression, one may obtain

d+h (6.23)
fos = h d + 12 - f~

The relations XH / fos = (h2 + h 1)/h2 and (h 2 + hI)/h2 = d/x' = d/(d +


h) lead to the coordinate X H :

(6.24)

It should be noted that the coordinates of the first focus fos and of
the second focus f~s are measured from the apexes. It would be useful to
have the coordinates of the focii relative to the positions of the principal
planes : fs = fos - XH, f~ = f~s - XH'· Substituting the magnitudes
fos , f~s ' XH and XH' in the last expressions by (6,21 - 6.24) gives

hh fU2
fs = d+ 12 - f~ and f s' = - d+ 12 - f~
(6.25)

These expressions for the coordinates of the principal focii and the prin-
cipal planes for the general case of an optical system formed by two
centered refracting spherical surfaces are the basis for computations of
simple optical devices. Please note that all equations given in sections
6.3 and 6.4 are valid for general cases. It is not necessary that the re-
fractive indexes of the media have the same values outside the centered
system on the left and on the right side.

4.2 Thick lenses


Let us regard the case of a thick lens, where the distance between the
apexes of the refracting surfaces d is comparable with the focal lengths
h, t; 12, f~ . Suppose, for simplicity, that the lens is in an air envi-
ronment with the refractive index n ~ 1 so that the relative refractive
index n21 for both surfaces is practically equal to the refractive index of
the lens material n. Fig.6.13,a shows a convex-convex lens, for which the
center of curvature of the first refracting surface is on the right side of
the surface and the curvature center of the second refracting surface is
located on its left side. Let us derive an expression for the optical force
of the lens P using the formulae for the second focal length in (6.25):

1
P=,=
fs
Geometrical Optics 239

According to (6.7) and (6.8), we introduce the optical forces of the first
and the second refracting surfaces:

(6.26)

Substituting these relations into the expression for the optical force of
the lens, we find the following formula:

(6.27)

A lens is called positive if its optical force P > 0 and negative in the
opposite case, P < O.
In a similar way, using (6.25) and the relations (6.26, 6.27), one can
find the relation
is = - ~ = -t; (6.28)
for the first focal length is of the lens. Hence, the focal lengths of a thick
lens are expressed by a single optical force. It should be noted that this
result follows from eq. (6.25), where the values is and i; are measured
from their respective principal planes H and H'.
Let us determine the coord inates of the principal planes of the lens in
terms of the optical forces. Following (6.24), (6.26) and (6.27), we get

XH = is.!!.- = ~P2 (6.29)


h n P
for the quantity XH. In the same manner it is possible to obtain the
quantity X H' :
X~ = -~ PI (6.30)
n P
In particular cases, it is convenient to use the expressions (6.21) and
(6.23), where the coordinates ios and i~s are measured from the re-
spective apexes of the refracting surfaces. In the case of the thick lens
presented in Fig.6.13,a, the expressions (6.21) - (6.24) are used with re-
fractive index n = 1.5 and with the ratios rI/r2 = 3/2 and d/rl = 4/7.
Now knowing the positions of the principal planes and focii of the
lens, one can construct an image formed by the lens. The geometrical
paths of some rays emitted by an object towards the thick lens and the
image of the object are shown in Fig.6.13,b.
An important and useful case of a positive lens is a ball lens, where
the refracting surfaces both have the same radius r and the distance
between them is 2r (Fig.6.14). The focii of the ball lens are located
240 DEMONSTRATIONAL OPTICS

H H'

r2 r1
-o- ._ .~ -_ ._ .~ ._ ._ . · ~ - _· _ · ~· _· _ · _ · - o -·

f1 f2 f• f~ f~ f~

Figure 6.13. A thick convex-convex lens. Positions of principal points are calculated
for a relative refractive index n = 1.5, rl/r2 = 3/2 and d/rl = 4/7 (a) . Construction
of an image using the principal planes and the focal lengths of the lens (b) . The dashed
lines indicate ray paths inside the lens obtained by the geometrical construction using
the principal planes and focii. The solid lines inside the lens body show the real light
paths.

I
H,H '

Figure 6.14. Glass ball lens. The given geometrical construction is valid for n = 1.5.
The coordinates of the focii from their respective apexes are both r /2, where r is lens
rad ius. Both principal planes are located at the center of the lens.

symmetrically and are both equal in magnitude to r/2 (for n = 1.5).


The principal planes of the ball lens are both located at the center of
the lens.
An example of a negative thick lens is presented in Fig .6.15. The
positions of the principal planes and focii were obtained for the following
parameters: n = 1.5, rI/r2 = 3/2, d/rl = 4/7.
If the positions of the focii and the principal planes with respect to
the apexes of the refracting surfaces are known, it is also possible to
Geometrical Optics 241

construct the geometrical paths of the refracted rays inside the lens as
shown in Figs.6.13, 6.14, 6.15.

H H'

_ . _ . ~ ._. _ ._ . _ ._ . _._ . _ . -o
._ .- ' - ' - , - , - , 0 · _ · -<)- - · -

f' f'
1 2

Figure 6.15. A thick negative glass lens composed of two concave surfaces . The
geometrical construction of ray paths were obtained for n = 1.5, Tl/T2 = 3/2, d/rl =
1/2 (a) . Construction of an image by means of the known positions of the principal
planes and focii (b). Dashed lines show the ray paths from the inner surfaces to the
principal planes , solid lines show the real light paths inside the lens.

4.3 Thin lenses


For a so-called thin lens the distance d between the apexes of the re-
fracting surfaces is much smaller than the curvature radii of the surfaces.
Therefore, in equations 6.21 - 6.25 the term containing d easily may be
neglected. With this approximation, the coordinates of the principal
planes Hand H' are both equal to zero. The planes Hand H' intersect
the optical axis at the center of the lens (Fig.6 .16,a,b) . In turn, for the
coordinates of the focal lengths we get the following relations:

'
J
f
08 - = - /' = h!I-hif = hiU~
= /=" 8 - if
08 8
(6.31)

In the case of thin lenses, for the optical force P = 1/ i~ of the lens one
gets
1 1
P = -- + - = PI + P2 (6.32)
!I i~
from the definitions of the optical forces of the refracting surfaces (6.27
). Therefore, for the focal length of a thin lens, the following relation is
242 DEMONSTRATIONAL OPTICS

valid:
~ = -!, = (n - 1) (~ - ~) (6.33)
fs fs rl r2
In this case one can regard a thin lens to be characterized by a single
magnitude f, the focal length of the lens. f follows from the radii of
the refracting surfaces and the value of the relative refractive index (eq.
6.31). As in the case of a thick lens there are different kinds of thin lenses,
each being a specific composition of two refracting surfaces. For example,
a thin lens with two convex surfaces will possess a positive optical force.
The focal lengths of the lens are shorter than the shortest focal length
of the refracting surfaces. A thin lens composed of two concave surfaces
has a negative optical force, and the focal lengths will follow the same
rule mentioned above. In the particular case of a thin lens with one
plane surface, the focal length will be determined by the optical force
of the spherical surface. A thin lens composed of two spherical surfaces ,
where one is a concave surface and the other is a convex surface (called
meniscus), can have a positive or negative net optical force, depending
on the ratio of the radii of the surfaces .
For correcting chromatic errors (see section 6.5) it is necessary to com-
bine two thin lenses, having different signs of optical force and consisting
of materials of different relative dispersion to form a lens with a certain
resulting optical force. Even such combinations, known as achromatic
lenses, can be treated in good approximation as one thin lens.
As the thickness of a thin lens may be neglected, one can consider
the ray passing though the center of the lens as not being refracted.
This property of thin lenses is often used for the construction of an
image created by a thin lens. It is seen in Fig.6.16,a,b that the images
are obtained by means of two rays , one passing parallel to the optical
axis, and other one propagating through the center of the lens without
refraction. The intersection of these rays gives the position of the image
point above or below the optical axis.

4.4 Images formed by parallel rays


Sometimes it is necessary to find outgoing rays caused by an extended
source located on a plane normal to the main optical axis drawn though
the first focus of a thin lens (the so called the first focal plane). Within
the paraxial approximation, every emitting point of the source produces
rays which, after refraction by the lens, will propagate as a system of
parallel rays.
Let a point source S located on the first focal plane of a thin lens
emit three rays (Fig.6.17). One ray is parallel to the optical axis, the
second ray travels through the center of the lens and the third ray makes
Geometrical Optics 243

1------- x, --------'~-----

s
r,

t - - - - -- x, -----j

H.H'

Figure 6.16. Image construction for thin lenses having a positive (a) and a negative
optical force (b) . In both cases rays passing through the center of the lens are not
refracted .

a small angle Q with the optical axis. The first ray, falling on the lens
at point M , will int ersect the optical axis at the second focus F' . The
second ray, passing the lens center, is not refracted and forms the angle
a' = IFSI / IFOI with the optical axis. The third ray hits the lens at
point M' and will (after refraction) intersect the second focal plane at
point pl. All outgoing rays are parallel and intersect the second focal
plane (drawn through the second focus normal to the main optical axis)
with the same angle.
In contrast, in most applications using the paraxial approximation one
has the problem of obtaining an image caused by a system of parallel
rays. Following the rays ofFig.6.17 backwards, one sees that all incoming
parallel rays are refracted to point S in the first focal plane. The size of
the image of a far distanced object, such as the Sun , can be determined
if the angles of the incoming rays are known.
244 DEMONSTRATIONAL OPTICS

Figure 6.17. Image construction for rays coming from a point S on the first focal
plane of a thin positive lens. After refraction these rays are parallel (M ' P'IIM F'IIOF')
and form the angle Ct.' = IFSI / IFOI with the optical axis.

4.5 Two thin lenses


One of the simplest optical devices is an aggregate of two thin lenses,
separated by some distance. In fact, the problem of determination of
the focal lengths and the principal planes of such a system can be solved
by applying the general formulae (6.9,6.10) to one lens and then to the
other. This implies that each pair of magnitudes h, f{ and 12, f~ in
expressions (6.21) - (6.25) has to be treated as the focal lengths of the
first lens and the second lens, respectively. Since we have placed the
lenses in air, in expressions (6.26), (6.27) and (6.29) the refractive index
of air (n ~ 1) has to be used. Therefore , we get the following expressions:

(6.34)

P2 P1
XH = d X r = _d (6.35)
P H P
For given focal lengths of the lenses fl, f{ and 12, f~ , the focal lengths
of the optical system can be calculated by means of formulae (6.21) and
(6.23) or by (6.25).
An optical system consisting of two thin lenses is used forf several
simple optical devices such as an eye-piece and different telescopic sys-
tems. An eye-piece is commonly used as an element of a telescope
or microscope to collect light from an extended primary image formed
by the telescope or microscope objective . The advantages of the two-
component eye-piece over a single lens is mainly the fact that it can
provide a higher correction of the imperfections of the lenses forming
the eye-piece (so called aberrations, which will be considered in section
6.5).
As examples let us regard the RAMSDEN eye-piece and the Huv-
GENIAN eye-piece , both built with two plane-convex positive thin lenses.
In the RAMSDEN eye-piece system shown in Fig.6.18,a, the first and
the second lenses have the focal length f and the distance d between
Geometrical Optics 245

the lenses is d = 2f /3. For the optical powers of the lenses we get
PI = P2 = 1/f. Using f{ = -II = f , f 2 = - h = f and d = 2f/3 ,
calculations by means of eqs. (6.21) and (6.23) give
-fos = f~s = f/4 .
In turn, the coordinates of the principal planes obtained for the given
parameters follow from (6.22) and (6.24):
-Xw=XH=f/2.
The positions of the focii and the principal planes with respect to the
components of the RAMSDEN eye-piece are shown in Fig.6.18,a. Let an
object be placed in the first focal plane of the eye-piece. One can regard
the top point of the object as a point source emitting a pair of rays .
One ray propagates parallel to the main optical axis, and the other ray
propagates at a small angle with respect to the first one. It is easy to
see an analogy between this point source and our considerations in the
previous section. Two parallel rays are formed by the optical system
after refraction of the incident rays.
According to the construction rules given in section 6.3, the ray paral-
lel to t he optical axis has to reach the first principal plane H (t he dashed
line, joining the inner surface of the first lens and H), t hen the second
one H' and then it has to intersect the main optical axis at the second
focus of the system f~. The inner part of the ray trace from H' to the
surface of the second lens is shown by a dashed line; the continuation of
this dashed line outside the lens is shown by a solid line. In a similar
way one can find the path of the ray, falling on the first lens at a small
angle, between the lenses. This ray must leave the lens system parallel
to the ray treated before. The real traces of the rays between the lenses
can be found by connecting the entry and exit points of the rays on both
single lenses.
The function of the RAMSDEN eye-piece is similar to that of a single
positive lens. The net optical force of this eye-piece, calculated by the
formula (6.34), is P = 4/(3f) > O. Hence, an equivalent single lens would
have the focal length 3f/4.
In the case of the HUYGENIAN eye-piece the first lens, being positive
and having the focal length 3f, is at a distance d = 2f from the second
positive lens with the focal length f. The positions of the principal planes
are given by

XH = d P2
p = 3f and X H' = -d p
PI
= -f ·
As a result, the principal focii F, F ' are located symmetrically from the
second lens at equal distances ±f/2 (Fig.6 .18,b). The total optical force
246 DEMONSTRATIONAL OPTICS

H' H

---- - -- --

H' H

~ : - ---- --- - - -:::

---------+~-~-- - -~~ ~~~

f~

Figure 6.18. RAMSDEN (a) and HUYGENIA N (b) eye-pieces. The focal length of the
first lens /l , the distance between the lenses d and the focal length of the second lens
h fulfill the ratio /l : d : h = 3 : 2 : 3 for the RAMSDE N eye-piece, and 3 : 2 : 1 for
the HUYGENIAN eye-piece. The object line located in the first focal plane is a primary
image generated by the telescope (or microscope) objective.

of the HUYGENIAN eye-piece calculated using expression (6.34) is P =


2/(3j) , which means that this optical system operates like a positive lens
with the focus length 3f /2. The first focus of the HUYGENIAN eye-piece
is located inside the lens space. Therefore, the first lens is positioned
before the primary image of a real object (this image is shown by the
vertical arrow on the place where it would be located without eye-piece).
We can reconstruct the real paths of the rays passing through the eye-
piece by means of two rays, which would intersect the top point of the
primary image if the eye-piece were absent. This pair of rays , shown by
solid lines in front of the first lens and by dashed lines behind it, will
give rise to two parallel outgoing rays (Fig.6.18,b). It is easy to find
the trace of the ray passing parallel to the optical axis. After refraction
by both lenses this ray will intersect the optical axis at point f~ . The
Geometrical Optics 247
second ray, falling on the first lens at a small angle with respect to the
first ray, will generate the second outgoing ray parallel to the first one.

Conclusions. Using the paraxial approximation, one can describe


different types of lenses in a similar manner, based on the representation
of a lens in terms of a centered optical system. For a lens, this system
consis of two refracting spherical surfaces separated by a certain distance
d. For given parameters d, rl, "z and the relative refractive index n21 ,
the ~oordinates of the principal planes H , H' and of the principal focii
F, F ' can be calculated by means of formulae (6.21) - (6.25). In the
general case, for thick lenses, these coordinates will be different, if fos
and f~s are both measured from the apex of their respective refracting
surfaces (coordinate fos is taken from the first surface, f~s from the
second refracting surface). The latter follows from eqs. (6.21) and (6.23),
where the notations fos, f~s have been used for these coordinates. For the
coordinates of the principal focii measured from their respective principal
planes, the notations fs, f~ were introduced. fs, f~ , have the same
absolute value but differ in sign. In the case of a thin lens the coordinates
of the principal planes are both equal to zero, and the difference between
the magnitudes fos, t, and between f~s' f~ disappears.

5. Errors in optical systems


Until now, our description was limited to paraxial light beams. For
real applicat ions it is of interest to analyze the image formation when
relatively broad light beams are propagating in an·optical system . De-
viation from cases where the paraxial approximation is valid leads to
a violation of the homocentricity of the beams forming an image. The
real optical image, created by an optical system , shows a number of
limitations due to inevitable errors in the imaging system .

Diaphragm

Figure 6.19. Setup for observation of the longitudinal spherical aberration caused
by a lens. A mercury lamp , emitting a set of bright spectral lines, is used . In order
to observe distinct focii of quasi-monocrornatic light , an optical filter is inserted into
the beam.
248 DEMONSTRATIONAL OPTICS

---1----- - -- -- --- - - - - - - -~ Figure 6.20 Longitudinal


spherical aberration is the
I difference of focallenths 8f
from a narrow light beam
(the focus F') to an off-axis
light beam (the focus F ") .

Figure 6.21 A wide light


beam forms an image
within a short line posi-
tioned between F' and F" .
The cross section of the
beam is a circle of minimal
radius on the plane aa'
orthogonal to the optical
axis.

Let us regard the following simple experiments with a positive lens


with two spherical surfaces . Light from a source formed by a condenser
lens propagates as a ray of practically parallel beams (Fig.6 .19). A lens
with a positive optical force is inserted into this beam with a round
diaphragm fixed in front of it to select of a narrower beam. After re-
fraction by the lens the beam is focused at the point F' at a distance
j' from the lens. Let us exchange the diaphragm with a circular ring
diaphragm in order to select the boundary beams of the light falling on
the lens. Boundary beams are more strongly refracted by the lens than
the paraxial beams. For this reason the refracted beams will be focused
at point F" located at a distance l'
smaller than j'. The difference

8f = /' - /
is called the longitudinal spherical aberration (Fig .6.20). The reason for
this behavior is that the lens has two spherical surfaces. Light is focused
into one point only for rays close to the optical axis - thorough Chapter
6 we have used the paraxial approximation up to this point. There are
surfaces which do not show longitudinal aberration. These surfaces have
aspherical shape (called eikonal, see later), but, for technical reasons,
most of the lenses used in applications still have spherical shape.
Longitudinal spherical aberration causes blurring of the image. Ele-
mentary beams falling on the lens at different distances from the optical
axis will be focused within the short length 8f. The cross section of the
Geometrical Optics 249

refracted beams by any plane aa' (Fig.6.21) will have a circular shape. A
cross section with minimal diameter will exist at a certain point between
the focii F' and F". In the case of a lens with a positive optical force
(converging lens) the value of the longitudinal spherical aberration is
negative: 8/ < O. For a lens with a negative optical force this quantity is
positive: 8/ > O. In the latter case the boundary beams are more weakly
refracted than the paraxial ones. This peculiarity is used to design op-
tical systems corrected for the longitudinal spherical aberration.
h1r
1.0

0.75

1_
0.5

0.25

6 r -1

"Figure 6.22. Illustration of the operating principle of an astigmatic lens . The max-
imal value of Of appears at hlr ~ 0.6, where h is the distance from the optical axis
and r is the lens radius. The dashed line shows the variation in Of for a single positive
lens.

Fig.6.22 shows an optical system containing two specially selected


lenses with positive and negative optical forces combined in such a way
that the system possesses a total positive optical force. For the system
shown in Fig.6.22 the longitudinal spherical aberration for the parax-
ial and the boundary beams is practically equal to zero. As it is seen
from the graph, although the quantity 8/ reaches a maximum value at
approximately 65% of the lens radius, it is still much smaller than that
of a single positive lens ofthe same optical"power throughout the whole
region.
Even if the spherical aberration is corrected by means of the above
mentioned technique for beams propagating parallel to the opticalaxis,
it can yet exist for beams travelling under an angle relative to the optical
axis. This effect causes a specific distortion of the point source image.
Let a point source be located at a certain distance from the optical axis
of the lens, as shown in Fig.6.23. An image of this source, formed by a
broad pencil of beams, represents itself as an extended spot. Because it
slightly resembles a comet with a tail , the aberration for the nonaxial
beams is called coma. Coma is the variation of magnification within
the aperture. The amount of coma is the vertical distance between the
250 DEMONSTRATIONAL OPTICS

Figure 6.23 Coma . Rays


, passing from a point P
P :-.'- - - - - i r ' \ . through the outer parts of
,
,, the lens intersect at differ-
, ent points than rays pass-
_ .1_ ._ ._ ._ ing through the lens cen-
ter . Arrows show coma
as the vertical distance be-
tween the central ray and
the intersection point of
the outer refracted rays .

ray, passing through the center of the lens from point source P, and the
intersection of peripheral rays.
In addition to the above described distortions of the pencils of beams
refracted by spherical surfaces, an ordinary astigmatism also takes place.
Astigmatism in optical systems may exist even for narrow pencils of
beams if they are propagating at a considerable inclination to the optical
axis. Astigmatic beams, arising as a result of refraction, are character-
ized by two focus lines, which have properties regarded at the beginning
of this chapter. Astigmatism of an optical system may be corrected by
special selection of the curvature radii and the optical forces of the re-
fracting surfaces. Optical systems corrected for astigmatism are called
anastigmatic.

rr: I 'H '?= -a


l-

I- --I
\-l -r
4.1 1 tJ B
a b c

Figure 6.24. Pincushion (c) and barrel (b) distortion, shown for the example of a
quadratic grid (a) . With pincushion distortion images of points outside the optical
axis are further from the axis of the beam (c), and with barrel distortion they are
closer to the axis of the beam (b). The linear amount of distortion varies with +h2 (c)
and with _h2 (b), the square of the distance h from the axis, thus nonradial straight
lines are imaged as curved lines.
Geometrical Optics 251
White light Red rays
I

---
. _ - _ . _ . _. _ . - . _ . _ . _ . _ . _ . _- -~ ~ ~~
-:;:;. - .- .-
~-,.;.,,=,
.

- -..

Longitudinal
chromatic
aberration

Figure 6.25. Axial chromatic aberration as the longitudinal variation of the focal
position of red rays FR with respect to blue rays FB. The longitudinal chromatic
aberration is the distance FBFR.

Besides spherical aberration, coma, and astigmatism, other errors in


optical systems exist , known as the distortion and curvature of the image
field. Distortion is caused by the fact that the magnification of the op-
tical system is not the same for different distances from the optical axis.
This error violates the geometrical similarity of an object to its image.
Distortion is easily demonstrated by creating an image of a quadratic
net , located in a plane orthogonal to the optical axis (Fig.6.24 ,a) . The
system may lead to t he images, presented in Fig.6.24,b,c. If the distance
of a point from the optical axis is denoted by h, the linear amount of
distortion varies with _h2 , in the case of barrel distortion (b) , thus non-
radial straight lines of the object are imaged as curved lines bent away
from the axes. In the opposite case of pincushion distortion (c) the linear
amount of distortion varies with +h 2 that leads to curved lines of the
image, which are bent toward the axes.
The image of a plane object may be located on a surface which is not
a plane but shows some curvature, even when each point of the object
has been imaged into an ideal point. In some optical instruments, one
has to avoid such a deviation of the image surface from a plane.
The wavelength dependency of the refractive index of the optical ma-
terials of lenses may cause chromatic aberration. Practically all optical
materials used for fabrication of optical lenses show normal dispersion,
where the refractive index grows with a decrease of the wavelength. In
other words the optical force of a single lens increases while the wave-
length decreases from the red to the blue spectral region (the case of
normal dispersion has been considered in Chapter 4).
Chromatic aberration due 'to higher refractivity of the blue beams
with respect to the red ones tends to blur of an image and to color its
252 DEMONSTRATIONAL OPTICS

Crown glass Figure 6.26 Achromat.


Dispersion dn/d>. is com-
pensated using different
Flint glass glass sorts, but a net
refraction is still obtained.

edges. In particular, axial chromatic aberration is the variation of the


focal position with wavelength (Fig.6.25). The longitudinal chromatic
aberration is the distance from the long-wavelength focus (FR for red
rays in Fig.6.25) to the short-wavelength focus (FB for blue rays in
Fig.6.25). At every plane drawn at a right angle to the optical axis
within the distance FRFB , the image of the point source is a colored
circle of finite size.

Image

White light

lateral
chromatic
Aperture aberration

Figure 6.27. Lateral chromatic aberration, or the vertical distance from the off-axis
image of a point obtained by red rays compared to that with blue rays . The aperture
is illuminated by white light .

The effect of chromatic aberration may diminish the quality of an im-


age considerably. Combinations of a positive and a negative lens each
made of glass with a different relative dispersion are designed to com-
pensate for chromatic aberration. For example, so-called crone glass
has a lower relative dispersion than so-called flint glass. For an aggre-
gate of two lenses, one positive lens made of crown glass and another
negative lens made of flint glass (called an achromatic doublet) may prac-
tically compensate the chromatic aberration, while the optical system
as a whole will still possess a positive optical force (Fig.6.26). This type
of complex lens is called an achromat.
Lateral chromatic aberration is the variation of the image size, or mag-
nification, with wavelength. For example, let an aperture illuminated by
Geometrical Optics 253

white rays be imaged by a thin lens as shown in Fig.6.27. The amount


of lateral chromatic aberration is the vertical distance from the off-axis
image of a point in long-wavelength rays (red rays) to the image caused
by short-wave length rays (blue rays).

6. Formation of optical images


6.1 Human eye
Applying the laws of geometrical optics, the eye can be modeled using
only three refracting surfaces: one is the cornea, the boundary of the eye
to air; the other two surfaces form the lens of the eye (Fig.6.28). This
lens has a double convex form. The optical density of the lens matter
increases slightly toward its center, providing a certain correction for
spherical aberration. But the variations of the refractive index of the lens
are small, so a mean value of approximately n = 1.44 can be assumed.
The matter of the vitreous body, which is around both surfaces of the
lens of the eye has a refractive index close to that of water (n = 1.34).

Figure 6.28 Human eye.


H,H' Lens A representation of the eye
in terms of a centered op-
tical system. The princi-
pal planes are both close
f to cornea. The dashed line
_.- _.0- -- -- '-
indicates the position of
the first lens surface for an
unaccommodated eye. In
this case, the focal point is
located before the retina,
providing best vision for
10=25 em. An accommo-
dated eye for looking at in-
finity shifts the focus point
to the retina.

The radius of the curvature of the refracting surfaces of the lens, and
hence its focal length, may be varied slightly by muscular contraction.
Such an adjustment of the focal length of the eye, called accommodation,
provides a sharp image of objects at an infinite distance as well as only
a few centimeters from the eye. Usually the light sensitive layer of the
eye (called the retina) is assumed to be at a distance of 17 mm from
the inner refracting surface of the lens, whereas the focal length for an
unaccommodated eye is about 15 mm. For a normally sighted eye having
such parameters, an object at a distance of about 25 em from the eye is
254 DEMONSTRATIONAL OPTICS

imaged on the retina. This distance is called the distance of best vision
(denoted by 10 in Fig.6.29)

Figure 6.29 An object at


the distance of best vision
10 is observed by an unac-
commodated eye.

For observing objects at infinity, the focal length increases, reaching


17 mm, possibly due to accommodation. In turn, for observation of
objects located closer to the eye than 25 em, the curvature radii of the
lens increase ; hence, the focal length of the lens decreases, which leads
to clear vision of these objects. It should be noted that only about 1/3
of the optical power is caused by the lens. The main part is due to
the curvature of the cornea. Note also that the media on both sides of
the imaging lens system are not the same: outside the eye we have air ,
while the image is formed in the vitreous medium of the inner part of
the eyeball.
An important element of the eye is the contracting iris, which deter-
mines the aperture diameter, also called the entrance pupil of the eye.
The latter may be varied by muscular action, depending on the amount
of light flux passing through the aperture. Under the action of very weak
light fluxes, the iris has a diameter of about 7 mm , whereas for strong
light fluxes, its diameter decreases to 1.5 mm.
The accommodation is accompanied by a variation of t he diameter of
the iris, since the distinctness of vision is also dependent on this diameter
when everything else remains fixed (compare Fig.6.36). For a normal
eye at the smallest diameter (1.5 mm) of the iris, the resolved angular
separation of two point sources which can be observed separately is about
1 arc minute. When the entrance pupil of the eye increases, reaching
5 mm or more, this limiting angular size will increase due to effects of
spherical and chromatic aberration, which implies that the distinction
of vision will decrease, too.
It should be noted that the retina of the eye is one of the most universal
light detectors. An eye adapted to darkness can detect single photons
(with an efficiency of approximately 2 %), but at the other extreme, is
not damaged when looking directly into the Sun (photon flux into the
eye is approximately 1015 Photons/a at an iris diameter of 1.5 mm) . So
the sensitivity covers the enormous range of 1013 !
Geometrical Optics 255

p'

pI
1

Figure 6.30. A magnifier. An object P{' r", placed from the eye at the distance of
best vision 10, is observed at an angle u. A gain in magnification is provided by a
thin lens positioned close to the eye. The object PIP is now located near the first
focal plane and can be observed by the eye at the angle u /. The lens gives rise to an
imaginary, erect image P{PI at the distance of best vision.

6.2 Magnifier
The simplest magnifier consists of one positive lens (Fig.6.30). This
lens L is located in front of the eye so that the observed object PiP is
placed in the vicinity of its focus plane F, closer to the lens. In this case,
an imaginary and magnified image P~ P' appears, located at the distance
of best vision lo, so that the eye may observe it without accommodation.
Considering the lens to be thin, the beams Pi 0 and P~ 0 will pass the
lens without refraction. Therefore, the observation angle u' of the object
is the angle between the beams P~ 0 and P' O. Without the magnifying
lens, the object would be located at Lo and would be observed under the
angle u. The smaller the focal length of the lens is (the closer the object
is located to the lens), the higher the angle u' will be.
For determining the magnification of the magnifier f3, the ratio
-,- --
PiP / PiP
has to be estimated. Assuming that the object is located practically
in the focal plane of the lens, the following approximate value may be
proposed for the magnification:

f3 = p{P' = ~ ~ 25 (6.36)
PiP f f
where f is the focus length of the magnifier, expressed in em.
256 DEMONSTRATIONAL OPTICS

6.3 Telescopic system


6.3.1 Refracting telescope
Optical devices for observation of remote objects are called telescopes.
Often, one distinguishes between astronomical telescopes and telescopic
systems for observation of remote objects on earth (terrestrial telescopes) .
The principle scheme of a refracting telescope, operating on the basis of
light refraction by an objective, is shown in Fig.6.31. In this particular
case, the telescope consists of two collecting lenses: the long focus, usu-
ally achromatic objective, and the short focus eye-piece. This device is
called KEPLER's telescope.

Figure 6.91. Formation of the object image by a system of parallel rays passing
through a telescope. Rays from opposite sides of the object (points A and B) fallon
t he telescope objective at the angle o . Practically parallel rays from point A, inclined
0/2 to the optical axis, form point A', positioned at the first focal plane of the eye-
piece. This point gives rise to parallel rays leaving the eye-piece at an angle f3 /2
> a/2 .

-E)
A
a
o

Figure 6.92. A remote object of ball shape is observed by the eye at point 0 under
an angle 0 « 1.

It is convenient to characterize a remote object by its angular size to


illustrate the operating principle of a telescopic system (Fig.6.32). For
example, for a spherical illuminating source (a model of a planet), the an-
gular size a representing the source, is given by the ratio a = D/ L, where
the distance L from the source to the point of observation 0 is assumed
to be much larger than the diameter of the source D, so that the angle
a « 1. With this approximation, which is valid for all practical cases of
Geometrical Optics 257

astronomical observations, the image of the source will be located in the


focal plane of the telescope objective, since one can regard this image
to be formed by a system of parallel rays. When treating the RAMSDEN
and HUYGENIAN eye-pieces, we saw that the primary image generated
by the telescope objective will give rise to a system of parallel rays leav-
ing the eye-piece, if the focal plane of the eye-piece is located exactly
in the focal plane of the objective. The magnification of a telescopic
system is determined by the ratio of the focal length of the objective
and to that of the eye-piece. Light rays emitted from the border of the
source , entering the telescope objective at the angle of observation 0: ,
will form the primary image A'B' of dimension lA'B'I = o:fob , where
fob is the focal length of the telescope objective. Point A' of this image
gives a system of parallel rays , which will leave the eye-piece at an angle
[3/2. For small angles {3, one gets {3 = A'B' / fep, where fep is the focal
length of the eye-piece. The angles 0: and {3 are measured in opposite
directions; hence, for the angle {3, we write the following expression:

{3 = _o:foo (6.37)
t;
An observer , placing an eye accommodated to infinity behind the eye-
piece, can observe the enlarged image of the object. The larger the ratio
fob/ fep is, the larger the observation angle of the object image in the
focal plane of the objective will be. The construction of a telescopic
system allows a shift of the eye-piece along the optical axis, allowing
th e production of a real image of the object, for example, in the plane
of a photographic plate. In another case, where the eye-piece is shifted
closer towards the objective, th e eye-piece produces an imaginary image
at distance lo. In such a case, the action of the eye-piece is similar to
the magnifier treated before, and the image can be observed with the
eye accommodated to lo.
The image of the object is reversed when observed through such a
telescope. This does not matter when observing the stars of the sky, but
is troublesome when observing terrestrial objects. Therefore, terrestrial
telescopes are practically always used with a combination of prisms or
lenses which turn the image upright.
The objective of the GALLILEO telescopic system consists of a long
focus achromatic lens and a short focus negative lens acting as the eye-
piece (Fig.6.33). The negative lens is mounted in such a way, that its
first focus coincides with the position of the focal plane of the objec-
tive. Advantages provided by an eye-piece with negative focal length
are a shorter total length of the telescopic system and an upright im-
age. A disadvantage is the more difficult correction of imaging errors .
258 DEMONSTRATIONAL OPTICS

Figure 6.33. The GALLILEO telescope. Parallel rays from a remote object falling on
the objective at small angle to the optical axis 00' form a primary image A' B' on
its focal plane . The latter is fitted with the focal plane of a thin negative lens. Point
A' gives rise to two parallel rays leaving the lens.

Such devices are mainly used for small magnifications and lower quality
requirements, e.g. for opera glasses.

6.3.2 Reflecting telescope


The operating principle of a reflecting telescopic system is similar to
that of refracting telescopes. The basic element of such a telescope is a
reflecting concave mirror. Very often, mirrors of a paraboloidal shape
are used to correct for spherical aberration. We have seen in section 6.1
that the reflecting surface of concave spherical shape possesses the ability
to converge light rays. A paraboloid is better suited for large diameters,
since spherical abberation is avoided for rays parallel to the optical axis.
An incident light beam is converted into a converging homocentric pencil
in a similar way as by the objective of a refractive telescope. In an
astronomical telescope system the primary image of a remote object
formed by parallel rays is located in the first focal plane of the mirror.
In fact, the image in the focal plane could be obtained directly by
means of a photographic plate mounted in the focal plane. But in most
cases, the primary image is examined by means of an eye-piece. In or-
der to divert light rays to the eye-piece, a small plane mirror has to be
mounted at an angle 45° with respect to the optical axis of the primary
mirror, as shown in Fig.6.34,a, where a so-called NEWTONIAN telescope
system is presented. The angular magnification of such a system is ex-
pressed by formula (6.37) as the ratio of the focal lengths of the objective
and the eye-piece. In other types of reflecting telescopes, two mirrors
are used to form the primary image: one is the large primary mirror and
the other is a small mirror, usually having a convex shape (Fig.6 .34,b).
The second mirror reflects the rays back and provides the possibility
to mount the eye-piece outside the system of mirrors. In astronomical
Geometrical Optics 259

Figure 6.34 Reflecting


telescope . Geometrical
paths for parallel rays
in case of the Newton
reflecting telescope (a) .
After reflection by the
primary mirror, the rays
are diverted by a plane
mirror into the eye-piece.
Double-mirror system
of a reflecting telescope
(b) . The second convex
mirror reflects rays along
the optical axis of the
telescope . These rays
pass through a hole in
b the primary mirror to the
eye-piece.

reflecting telescopes, different second mirrors are often used for varying
the net focal length of the system of both mirrors.
A principle advantage of the reflecting telescope is the absence of
spherical (in case of paraboloidal shape) and chromatic aberrations in
the mirrors. The primary mirror of a reflecting telescope can be made
larger than the glass lenses of the objective of a refracting telescope ,
since optical inhomogeneities in a mirror block are of no significance .
As the magnification is given by the ratio of the focal lengths of the
objective mirror and the eye-piece lens and has nothing to do with the
mirror diameter, why are astronomers interested in using large-diameter
telescopes? The reason is that the resolving power increases with diam-
eter (see Vol II, Diffraction) and that the point brightness increases with
the square of the diameter of the primary collecting element . In such
a way, a large scale telescope allows the detection of weaker stars with
higher angular resolution.

6.4 Microscope
A microscope is designed for the observation of tiny obje cts when
a significant magnification is needed, that is substantially more than
an ordinary magnifier may achieve. The principle construction of the
microscope comprise a very short-focus objective and an ocular in the
role of the magnifier (Fig .6.35).
An object subjected to observations by a microscope is located close
to the focal plane of the objective, which produces a real magnified image
260 DEMONSTRATIONAL OPTICS

p"

--::~::::>::::::---- Eye

sLiJ~~~~~n:
pi
I-• - - - - - - 10 - - - - - - -

Figure 6.35. Microscope. A short line SP located near the focus FOb of the objective
is imaged on a plane close to the focus F ep of an eye-piece. The latter forms an
enlarged erect image. Point p' of the primary image is imaged to point P" so that
the final image is at the distance of best vision 10 from the eye.

at a considerable distance from the objective. The magnifying power of


the microscope system may be estimated, assuming the objective and
the eye-piece are both thin lenses. Let the object SP at a distance x
(x < 0) from the first focus of the objective and its primary image S' pi
be at a distance x' ( x' > 0) from the second focus. It is clear that x
and x' satisfy the NEWTON formula (6.10), with Io« = -fob' Following
(6.11), the magnification 13ob, created by the objective, is 130b = -fOb/x.
By representing the eye-piece by a thin lens with a focal length fep and
placed in such a way that the primary image is located practically in
the first focus, the lens will give rise to a virtually enlarged image of the
primary image. Hence, this lens operates like a magnifier. The focal
length of the eye-piece is much smaller then the distance of best vision
of the eye lo, and, for the total magnification 13 of the system composing
the microscope, one gets
13 = _~ fOb (6.38)
X fep
It seems that the magnification 13 can achieve extremely large values at
respectively small distances x, when all other parameters remain fixed.
This is true only within the approximations of geometrical optics (,X = 0).
In reality, deviations from straight line light propagation due to its wave
nature will cause restrictions for magnification. These diffraction phe-
nomena limit the appearance of sufficiently small details of the object.
These details are not present in the final image, even if the magnifi-
cation would be appropriate. Due to diffraction, light rays from such
small details will not pass within the entrance aperture of the objective.
(ieo~etricai ()ptics 261

In most practical cases, the magnifying power of a microscope is about


1,61 = 200-300; in some cases, up to 1000. These values are large enough
to examine, for example, biological objects in visible light.

6.5 Limitation of light beams in optical systems


In practical cases, an object usually extends along the optical axis, so
different parts have a different distance from the imaging lens. Strictly
speaking, it is impossible to obtain a plane sharp image of an extensive
object even for an ideally centered optical system, as different parts of
th e object will be transformed into different planes. Let two points of a
spacious object A and B be at different distances from the thin collecting
lens, as shown in Fig .6.36,a. The point A' , as an image of the point A ,
will lie in the plane p' A' , conjugated with the plane of.the object AP.
As it is seen from the figure, the image of B will be focused at the point
B ' lying behind the plane A' p' .
The pencil of beams forming an image of the point B will create a
small circle at this plane. On other words, the image of the point B in
the plane pi A' will be not sharp. The size of this circle depends on the
displacement of the point B relative to the plane AP, as well as on the
angular width of the pencil of beams creating the image. If this pencil
of beams is restrained by a diaphragm D (Fig.6.36,b) , its width will be
smaller and the sharpness of the image of B in the plane A' p' will be
correspondingly higher. Therefore, a limitation of the pencil of beams
by a diaphragm or, naturally, by a lens mount, provides an increase in
the sharpness of the image.
Restriction of the angular width of a beam can be realized by different
means in general cases. A diaphragm, which limits the angle of the pencil
of beams originating from the points of an object, is called the aperture
stop or acting diaphragm (Fig.6.37). It can be a diaphragm placed in
front of a lens or behind it , as well as just a lens mount. The image
of the aperture stop (real or imaginary) constructed by the first part
of the optical system (located close to the aperture stop), is named the
entrance pupil. The image of the aperture stop, created by the part of the
optical system situated after the aperture stop, is called the exit pupil.
In particular, the objective mount of a telescope is the aperture stop and
the entrance pupil at the same time, if there are no optical elements in
front of the objective, whereas its image formed by the eye-piece is the
exit pupil.
Apart from the aperture stop limiting the pencils of beams , there
may be a vision field diaphragm or field stop. This diaphragm limits
the particular part of a spacious object, being examined by an optical
device. Let us regard the effect of the field stop for a refracting telescopic
262 DEMONSTRATIONAL OPTICS

Figure 6.36. Effect of lowering the diameter of a pencil of rays . The image of an
extented object is formed by the lens (a) . Point A located in the plane P is imaged
to point A' on the plane pI, whereas point B, being out of the plane P , is imaged to
point B ' outside of plane r'. The pencil of rays forms a circle in planee: Action of
a diaphragm D (b). A reduction in the diameter of off-axis rays by the diaphragm
D leads to a decrease of the size of the circle in the plane P' .

Entrance

Aperture
slop
I
pupil

Figure 6.37. Aperture stop, entrance and exit pupil.


Geometrical Optics 263

system, which consists of an objective and an eye-piece, both being thin


positive lenses (Fig.6.38). Rays passing inside the field stop all enter
the eye-piece. Without this limiting stop, beams entering the objective
with a higher inclination to the optical axis would be partially shared by
the eye-piece lens. The image of the field stop is located in the object
plane and will limit the part of the object being imaged (in the case
of a telescope at infinity, it will limit the acceptance angle). Without
a correctly placed field stop, a decrease in the illuminance of the image
boundaries would take place due to a partial shearing of non-axial beams.
This effect is know as shadowing or vignetting.
To increase the image quality, it is better to use a double lens eye-
piece instead of a simple lens. Two types of eye-pieces are commonly
used and have been discussed earlier: the RAMSDEN eye-piece and the
HUYGENIAN eye-piece.

Field
stop

- _._
-~

Figure 6.38 The action of


the field stop .

The primary image of a remote object is obtained by practically par-


allel entrance rays in the first focal plane of the objective. In the case of
the RAMSDEN eye-piece, this primary image is located in the first focal
plane of the the eye-piece, close to its first lens. This lens is called the
field lens and collects rays passing from the boundaries of the telescope
objective (Fig. 6.39,a). The aperture stop -and the entrance pupil are
defined by the mounting of the objective. The field stop has to be lo-
cated in the focal plane of the objective. The image of the aperture stop
caused by the eye-piece is the exit pupil. The bundle of outgoing paral-
lel rays is limited by this exit pupil. In a similar way, one can find the
optical paths for rays passing through the HUYGENIAN eye-piece, and
the position of the exit pupil (Fig .6.39,b). Usually it is required that
an exit pupil is situated at a position convenient for the eye. For the
given parameters of the optical system of a double lens eye-piece, the
RAMSDEN eye-piece provides a larger distance between the second lens
and the exit pupil, which is more convenient for placing the eye than in
the case of the HUYGENIAN eye-piece.
264 DEMONSTRATIONAL OPTICS

Field stop

a
Field stop

- -

I Exit pupil
b

Figure 6.39. Field lens in eye-pieces. Geometrical paths of rays through a telescope
with the RAMSDEN eye-piece (a) . The vertical arrow indicates one half of the image
of a remote object. Light rays through the top point of the arrow pass through the
aperture of the field lens, then they form two outgoing parallel rays . Geometrical
paths of rays through a telescope with the HUYGENIAN eye-piece (b). The vertical
arrow , showing one half of the image of a remote object (at its position without the
eye-piece), is located at first focal plane of the eye-piece. Incident rays which pass
through the field lens aperture in a way, that their extensions would intersect at the
top point of the primary image, form outgoing parallel rays .

7. Light propagation inside inhomogeneous media


7.1 Concept of eikonal
So far, all problems of the construction of images generated by optical
systems have been solved on the basis of the laws of light beam refraction
at the boundary of two transparent homogeneous media. Another wide
range of geometrical optics concerns problems of light propagation inside
optically inhomogeneous media. These media are considered to possess
a spatially dependent refractive index:

n = nCr)
In the frame of geometrical optics, we understand a medium to be inho-
mogeneous if noticeable spatial variations of the refractive index n occur
over distances greatly exceeding the light wavelength. Even when treat-
ing the shortest space interval as mathematically infinitesimal, the length
of this interval is much larger than the wave length (and we assumed
,\ = 0). Nevertheless, this interval should be shorter than the character-
istic length of noticeable spatial variations of the refractive index. Then,
Geometrical Optics 265

over very short spatial intervals, we can assume n(r) :::::: const, and within
this characteristic length the behavior of a light wave is similar to that
inside an uniform medium.
Let us regard an infinitely small path element dl between two points
of the light beam trajectory with coordinates ro and r (Fig.6.4D):

I dr I = I r - ro I = dl
The unit vector dr / dl will be the tangent to the trajectory at the point
roo For this infinitesimal part of the beam trajectory, the light wave can
be regarded as a plane wave with the wave vector k oriented tangentially
to the trajectory, fulfilling

where s is the beam vector and ko = 27T/ "\0, where ..\0 is the wavelength
in vacuum. Using the beam vector, it is possible to represent the electric
field of this plane wave at point ro in the following form:

E = Eo exp{i( cp + kro - wt)} = Eo exp{ -wt} exp{ikonsro + icp} ,


(6.39)
where n = n(ro) is the refractive index and ip is the initial phase.

s Figure 6.40 Trajectory of


dr a light ray in an inhomo-
r, geneous medium . The vec-
=
tor dr rl - ro is parallel
y to the beam vector s drawn
from point ro o
x

The coordinate dependency of the refractive index leads to a com-


plicated coordinate dependency of the wave phase. It is convenient to
represent the behavior of the phase in terms of a spatial function S(r)
called the eikonal:
exp{ikoS(r)} (6.40)
The physical dimension of the eikonal is that of an optical path, which is
represented by the product nlr l. Let us examine relationships between
an infinitesimal geometrical path change and the change in the eikonal,
dS. After introducing the Cartesian system (x, y, z) for the infinitesimal
change in the scalar function dS(r), corresponding to the change of the
266 DEMONSTRATIONAL OPTICS

coordinate dr , it is convenient to use the following geometrical relations


(F ig.6.41,a) . By definition

dr = idx+jdy+kdz ,

where i, j , k are the unit vectors associated with the Cartesian system
(x , y , z). The same is valid for the scalar function

as as as
dS(r) = ax dx+ aydy+ azdz ,

which can be written in another form:

dS(r) = (~~ i) . (idx) + (:j) .(jdy)+


+ ( ~~ k) . (kdx) = dr-grad S , (6.41)

where
as. as. as as
gradS = -l+-J+
ax ay
-k=-
az ar
is the gradient of S at point r. The expression (6.41) shows that dS
is represented by a scalar product of two vectors: gradS and dr. Now
consider P and Q to be two points on a surface S(r) = C , where C is
a const ant. These points are chosen so that Q is at distance dr from P.
Moving from P to Q, the change dS(r) in S(r) = C will be still zero,
since C is a constant. Hence,

dS(r) = ds-qrad S = 0 ,

which means, the vector grad S is perpendicular to dr. dr may have


any direction from P . As long as Q stays on the surface with arbitrary
position with respect to P , the vector grad S will be normal to the
surface S(r) = C.
Now let point P belong to surface S = C1 and point Q to an adjacent
surface S = C2, and dr go from P to Q (Fig.6.41,b). According to (6.41)
for an increment dS one gets :

dS= C2 - C1 = dr-qrad S . (6.42)

For a given dS, the distance Idrl is a minimum when dr is chosen parallel
to grad S , or parallel to the normal vector of the first surface, drawn at
P. In another words , for a given distance Idrl, the change in the scalar
function S is maximized by choosing dr parallel to grad S. Thus grad S
Geometrical Optics 267

grad S(r)

a b

Figure 6.41. Geometrical representation of the gradient of scalar function S(r). The
vector r is drawn to point P on the surface S(r) = C with fixed Cj the vector dr
drawn from P to Q can take an arbitrary position on the surface , so that dS(r) = OJ
grad S(r) is parallel to the unit normal vector n drawn at point P (a). Two surfaces
S(r) = C1 and S(r) = C2 separated by an infinitesimal distance IP Rlj the vector dr
drawn from P to Q can be chosen arbitrarily while Q belongs to the surface S(r) = C2j
the minimum of Idrl is obtained for dr II gradS(r). The vector gradS(r) is normal
to S(r) = C 1 at point P (b).

becomes the vector pointing in the direction of the fastest space rate of
change of the function S.
When applying the differential forms considered above to the problem
of light propagation within an inhomogeneous medium, it is assumed
that variations in the refractive index occur smoothly. Particularly, it
means that any small element of the surface S(ro) = C around a fixed
point ro must be regarded as a plane. Using the given ko, according
to the definition of the eikonal (6.40), the product koS(r) represents a
phase within an element of a plane wave front; then the change in phase
over the interval dr is represented by
kodS(r) = kodr·gradS
Now, one can regard the phase of the light wave to be changing over the
same infinitesimal interval dr, similar to that in case of a plane wave in
(6.39):
t::.cp = konsdr
The last expressions enables one to get the following:
kodr ·grad S = konsdr
Hence, for the gradient of eikonal, the final equation can be written as
ns = grad S (6.43)
268 DEMONSTRATIONAL OPTICS

which is known as the eikonal equation.

Figure 6.42 To the eikonal


equation. Two surfaces
5
grad S(r) S(ro) = C 1 and S(rd =
Light C2 separated by an infini-
path tesimal interval dl joining
two points ro and rl of
y light trajectory. The beam
vector s drawn from point
x
ro is parallel to the vector
grad S at the same point.

The eikonal equation represents the fact that the light beam vector s is
directed along the shortest path Idrl joining two points of wave surfaces
S = C1 and S = C2. In other words, the light trajectory specified by the
vector s, penetrating surface S = C1 at point rl, is directed along the
normal to the surface S = C 1 . Therefore, the vector grad S(r) and the
vector s both have the same direction.
When considering light beam vectors at two adjacent points, one can
now take into account changes in the direction of these vectors caused by
variations of the refractive index n = n(r). For an infinitesimal element
dl between two points r1 and r2, let the vectors 81 and 82 make a small
angle a (Fig.6.42). As we assume that the refractive index changes
smoothly, it is enough to take into account the first order derivative
dnfdl. From the eikonal equation we get
d(ns) d as a dS (6.44)
"dl = dl Or = or df
Now,
dS as or
dl or Ol
Since ar/Ol = s and as/or = ns, the derivative dS/dl can be expressed
as dS/dl = n. Substituting this result into (6.44) we get
d(ns) an
~ = or = grad n . (6.45)

As lsi = 1 it is evident, that the following relation is valid for the small
angle a:
a = Idsl = Idsl (6.46)
lsi
The non-zero value of Idsl indicates thatthe light path between rl
and r2 has to be characterized by a curvature radius R, originating from
Geometrical Optics 269

grad n
--=----..

Figure 6.43. TI ~ relationship 1/R = (ngradn)/n. The vertical vector gradn makes
a sharp angle with the light trajectory; the latter is represented by the arc of a circle
with the curvature radius R and center C. Two adjacent beam vectors Sl and Sl
drawn from the ends of an infinitesimal light path dl make a small angle 0 . The inner
unit normal vector n of the light trajectory is parallel to the vector ds .

the curvature center C. By treating dl as an infinitesimal arc element ,


the angle a may be represented by the radius R as follows: a = dljR.
So, taking (6.46) into account, we have

Idsl = dljR
As the normal unit vector n at the point ro has the same direction as
ds, the last equation can be written in the following vector form:
ds n
= (6.47)
dl R
Differentiation of the left-hand side of expression (6.46) gives
d(ns) dn ds
----;{l = diS + n dl (6.48)

It is simplified via a scalar multiplication of both sides by the unit vector


n. The beam vector s is tangential and the n vector is orthogonal to the
trajectory at point ro; therefore, ns = 0, and we get
d(ns) ds
n ----;{l = nn dl

Substituting nj R for dsjdl from (6.47) gives


d(ns) 1
n----;{l =R
On the other hand, substitution for d(ns)jdl by gradn from (6.45) gives
the following equation for the curvature radius of the trajectory in terms
270 DEMONSTRATIONAL OPTICS

Heigth
"0

S~dS

S4'--~ ds
•orad "

Figure 6.44- Geometrical representation of light passing in an inhomogeneous


medium . gradn indicates the vertical increase of n from no to n1j horizontal lines
are drawn for the same interval An . gradn makes a sharp angle with an initial beam
vector 81 that causes an inclination in the following vector 82 with respect to 81 , and
so on.

of gradient of the refractive index:

~ =n gradn (6.49)
R n
The left-hand side of the last equation is never negative ; therefore, vec-
tors nand gradn form a sharp angle. In turn, according to (6.49), the
unit vector n is always directed along a positive increment of the beam
vector ds. Therefore , this positive increment ds always makes a sharp
angle with the vector grad n, and the latter always points towards the
direction of increasing refractive index. Hence, the light beams prop-
agating in the optically inhomogeneous medium are bent towards the
region of increasing refractive index.
For example, let an inhomogeneous medium be characterized by a
vertically decreasing refractive index (Fig.6.44). For a given no < nl,
there is a set of horizontal lines, each drawn over a fixed increment
~n < 0, so that the gradient of the refractive index points downwards
in the picture. At first let the beam vector 81 make a sharp angle with the
gradient vector. An infinitesimal increment ds of 81 gives a new vector
82. This vector, making a sharp angle with gradn, is directed towards
the region of increasing refractive index and results to the following
vector 83, and so on. Hence, a light trajectory composed by a sequence
of vectors 81, 82, 83, 84, will be bent towards the region with a larger
refractive index.

7.2 Light beam bending


Light beam bending can be demonstrated by the following simple ex-
periment, shown in Fig.6.45,a. A long glass cell is partly filled with
Geometrical Optics 271

glycerine (nl = 1.5) and carefully covered by water (no = 1.3). The
relatively sharp boundary between water and glycerine soon disappears
because of diffusion and a liquid with a continous variation of the re-
fractive index appears: at the bottom of the cell n ~ nl and at the top
n ~ no. The refractive index gradient in this optically inhomogeneous
liquid is a vertical vector directed downwards. While passing through
this liquid, a thin horizontal pencil of beams bends in the direction of
the larger refractive index, as seen in the Fig.6.45,b.

! Optical
density

! grad n

Figure 6.45. A medium with spatially varying optical density (a); bending of a laser
beam propagating through an optically inhomogeneous liquid (b) .

7.2.1 Mirages
The natural phenomenon of a mirage is explained by the bending of
light beams in inhomogeneously heated air (Fig.6.46,a). The layers of
the hotter and less dense air are located near the surface of the hot
desert sand . The temperature drops with altitude, whereas the density
increases. The refractive index of air, proportional to the local density,
varies smoothly from a certain minimal value at the sand surface up to
larger values within higher layers of air. In this case, the refractive index
gradient is directed vertically upwards ; therefore, a horizontal light beam
is bent upwards from the heated surface (Fig.6.46,a,b).
Light beams originating from a bright region of the sky near the hori-
zon pass to an observer along a curved trajectory with its lowest part
passing in the vicinity of the surface of the earth. Therefore, an apparent
image of this region of the sky is located close to the earth's surface, at
the (dashed) extension of the line of sight. The illusion of a shining water
surface arises. The same illusion occurs for highway asphalt heated by
272 DEMONSTRATIONAL OPTICS

grad n
Light ray
_II __

Heated surface

a b

Figure 6.46. Mirage caused by a heated surface (a) ; mutual positions of the light
path and the gradient of refractive index (b) .

sunlight . An observer sees "pools", which disappear when approaching


to them .

~ Ray of sight
_ Igrad n
- - - - - - - - -- - - . - Li~
1
Cold water surface
a b

Figure 6.47. Mirage caused by a cold surface (a); mutual positions of the light path
and the gradient of refractive index (b)

Opposite bending of rays can occur when observing a remote object


along a cold surface, for example, a cold water surface (Fig.6.47,a). The
gradient vector of the refractive index of air will be directed towards the
surface. The light rays are bent towards the surface, so that the object
appears slightly higher than that it would appear with rays propagating
along straight lines (Fig.6.47,b).
Another example of light beam bending is provided by the astronom-
ical refraction phenomenon. This effect is caused by variation of the
refractive index of air with altitude. The density and the refraction in-
dex of air decrease with altitude, so the gradient of the refractive index
is oriented along the Earth's radius towards the surface of the Earth
(Fig.6.48). The apparent altitude of remote sources (outside the Earth's
atmosphere, such as stars) over the horizon seems to be higher than its
real altitude.

7.3 Fermat '8 principle


In order to substantiate the laws of light propagation within the frame
of geometrical optics, the principle of the minimal propagation time was
put forward by FERMAT (in 1657), and is known as FERMAT'S principle.
Geometrical Optics 273

~~
AtrnOSPh ere
Figure 6.48 Astronomical
Earth
refraction.

According to this principle, light propagates so that the time required


to cover a certain path is always minimal.
Let a pencil of light beams pass through a certain region of an optically
inhomogeneous medium in such a way that these beams do not cross in
the medium (Fig.6.49,a). Using the eikonal equation (6.43), an integral
along a short part of the trajectory ACB of one of the beams, between
the points A and B, will take the form

J J~~ = J
B B B

nsdr = dr dS = S(B) - S(A) (6.50)


A A A

As the integration path was arbitrarily selected, the value of the integral
on the left hand side of the last equation is equal to the eikonal difference
of the points A and B, and does not depend on the integration path.
Introducing an infinitesimal element of the optical path ndl let us
compare an integral over the optical path for an arbitrary curved tra-
jectory

J
B

ndl
A
with the integral (6.50). The scalar product nsdr may be expressed in
terms of the infinitesimal element of the integration path in the following
manner:
nsdr = ndl cos(;,di-) ,
where (S4r) is the angle between the vectors s and dr . It is evident,
that
nsdr = ndl cos(;,di-) ~ ndl .
Therefore, the optical path along an arbitrary curve is always larger
than the optical path corresponding to the light beam trajectory:

J J
B B

nsdr:5 ndl
A A
274 DEMONSTRATIONAL OPTICS

a b

Figure 6.4g. FERMAT 'S principle . Curved lines with arrows show light trajectories
(a) . Two optical paths between fixed points A and B (b) . The optical path along the
shortest path ACB is associated with a light trajectory.

Equality between these optical paths is attained only in the case when
the integration path coincides with the light beam path. Let us compare,
for example, the curve ACB associated with the optical path of light ,
and another arbitrary path ADB, which does not coincide with the
first one (Fig.6.49,b) . It is clear that the net path-length over ADB is
longer than that over ACB. Therefore, light beam propagation always
occurs along the shortest optical path. An element of the optical path
is connect ed to the time interval dt via the velocity of light in a medium
v:
dl = vdt .
Therefore, light propagation between two points takes place in the min-
imal amount of time .

SUMMARY
Geometrical optics uses the concept of light rays as mathematical lines
along which light propagates. Nothing is presupposed about the nature
of light with the exception that no deviations from light paths due to
effects like diffraction takes place. Light pencils are considered here as
a set of independently propagating rays obeying the well known laws
of straight line propagation within a homogeneous medium, as well as,
refraction and reflection laws at the boundary between two media.
Basic laws of geometrical optics may be derived from the wave concept
of light for the limiting case .A -- 0, when the wave nature of light
becomes non-essential. For most practical problems , the approximation
in terms of the geometrical optics gives excellent results. This fact is
due to the very small wavelength of light compared to the dimensions
of the optical elements usually applied (mirrors, lenses, diaphragms and
so on) .
Geometrical Optics 275

Nevertheless, it is necessary to exceed the framework of the geomet-


rical optics approximation to consider problems connected with light
distribution around the focal planes of optical instruments, such as the
resolving power problem for telescopes and microscopes. The corre-
sponding chapters can be found in volume II of this textbook.

PROBLEMS
6.1. In order to get a rough estimate of the dispersive properties of
glass used for the production of lenses, the quantity
1:::.= nF-nC
iio -1
is generally applied, where nc, no and nF are the refractive indexes
for the FRAUNHOFER C, D and F lines (>" = 659 nm, >.. = 589 nm and
>.. = 486 nm). Derive a formula for the change of in the focal length of
a thin lens caused by a change in the refractive index on of the glass.
The quantity on should be expressed by 1:::..

6.2 Derive the formula for the focal length of two thin lenses separated
by d. (Fig.6.50). The focal lengths of the lenses are fa and fb, respec-
tively. Consider the case of two positive (a) and positive and negative

_._.--- _._.-p-'~
lenses (b).
;4 d ..;

-L_' ~i .=:1._

Figure 6.50.

6.3. Two thin lenses of the focal lengths !l and 12, respectively, are
used for designing a simple achromatic system to roughly reduce the
longitudinal chromatic aberration. Derive a formula for the distance d
between these lenses, which provides conditions of achromaticity of the
system for wavelengths in the region between blue and red FRAUNHOFER
lines (see. Problem 6.1). Discuss the particular case both lenses are made
of an identical material.

6.4. An achromatic doublet is composed by two thin lenses: one is


positive and made from crown glass and the other is negative, made
276 DEMONSTRATIONAL OPTICS

from flint glass. Both lenses are in contact. For the given focal length
of the achromatic doublet, f = 50 em, calculate the focal lengths of
both lenses, using the following values of the refractive indices of crown
n Ce) and flint n(f) glasses: : n~) = 1.530, n~) = 1.527, n~) = 1.525,
n~) = 1.575, n~) = 1.565, n~) = 1.555 (the letters F,D,C refers to the
corresponding FRAUNHOFER lines mentioned in Problem 6.1).

t ~
_ .- ._._._ .-F,F' ~ \~
o-.- ._._._. _._._._._._ .- d-. -

-l+-----f,J!
Figure 6.51.

6.5 A well defined time delay T of a short-time laser pulse can be


obt ained by means of two concave mirrors arranged as shown in Fig .6.51.
The spherical mirrors M, and M2 have radii rl and r2, respectively
(r l > r2). The second mirror has a round hole of diameter d in its
center. The focii of both mirrors coincide at the same point on the
optical axis. Let a narrow laser beam, propagating parallel to the optical
axis at distance h, reflect by the first mirror. Estimate the time delay T
and calculate its magnitude for rl = 1 m, r2 = 0.5 m, d = 0.5 em, h = 3
em.

6.6. Fig.6.52 shows the experimental scheme for observation of the


FRESNEL's experiment, discussed in Chapter 5. The telescopic system
provides an enlarged real image (within the plane p") of the verticalslit
positioned on the plane p. The objective (focal length fo = 20 em) of
the telescopic system forms a real image of the slit on plane p' close to
th e focal plane of the eye-piece (focal length fep = 0.75 em).
The plane p" is at a distance L = 9 m from the vertical slit. The
distance between the slit and the objective is equal to a = 1.2 m. Esti-
mat e t he difference of refractive indices Lln for left and right circularly
polarized waves of A = 632.8 nm, which occur after the double prism
Geometrical Optics 277

Figure 6.52.

composed by left- and right-rotating quartz prisms. The separation of


the images of the slit on plane p" is equal to s = 2.5 em.

6.7. Discuss the operating principle of an autocollimator used for


measurements of the focal length of a centered optical system, which
is shown schematically by two positive lenses placed in front of a plane
mirror M (Fig.6.53). The optical system is illuminated by a small source
of monochromatic light.

Figure 6.53.

-~. rl
. .

Figure 6.54.

6.8. The convergent meniscus shown in Fig.6.54 has two refracting


surfaces ofradii r1.and T2 = 4rl, respectively. The distance between th e
278 DEMONSTRATIONAL OPTICS

--~"
-~- - _
--=:----------------
r2

Figure 6.55.

apexes of the refracting surfaces is equal to rI/2, the refractive index of


glass n = 1.5. Calculate the focal lengths of the meniscus, as well as the
positions of its focal planes F and F' and its principal planes Hand H' .
Draw an image of an object placed on the first focal plane of the lens.

6.9 The divergent meniscus shown in Fig.6.55 has two spherical re-
fracting surfaces; r2 = rI/4 and the index of the glass is n = 1.5. The
separation between the surfaces is d = r2/2. Calculate all magnitudes
as in problem 6.8, and construct an image of an object placed on the
second focal plane of the lens.

SOLUTIONS

6.1 From the expression for the focal length f of a thin lens (6.33),
one can obtain
1 = (..!..+..!..) ,
fen - 1) rl r2
where rl, r2 are the radii of the spherical surfaces of the lens. One can
see that the quantity fen - 1) does not dependent on the wavelength;
therefore , the variation o(J{n - 1)) = 0, or (n - l)of + fon = 0 may be
written as follows:
of on
f + (n-l) = 0 .
For estimating the longitudinal chromatic aberration of the thin lens,
the quantity on/en -1) may, roughly, be replaced by the quantity t1, so
that the following approximation for the distance of is valid:

of=ft1 ,
where 8f is the distance between the images obtained by parallel red
and blue rays.

6.2 Focal lengths of thin lenses are usually measured from the apexes
and , for this reason, we use the expressions (6.21) and (6.23) for the
Geometrical Optics 279
coordinates of focii f~s and fos of the system consisting of two thin
lenses:

f f d+h /,' f' d - f{


J as = 1d +h - f{ as = 2d +h - f{

where II = -fa, f{ = fa, h = - fb and f~ = fb. The signs of the


coordinates are assigned by means of Fig.6.12,a,b. With substitution for
the coordinates by the appropriate magnitudes represented in terms of
the focal lengths fa and fb, one can rewrite the expressions for fos and
f~s in the following form:

d- fb , d- fa
fos = -fa d - fb - fa fos = fb f f
d - [b - [b
or
fos = fafb + fad
fb + fa - d fb + fa - d
/,' = fafb fbd
as fb + fa - d fb + fa - d
Further, XH = fad/Ub + fa - d) and XH, = -fbd/Ub + fa - d) are the
coordinates of the principal planes Hand H' measured from the apexes
of the first and second lenses, respectively (see (6.22,6.24)). In turn,
the coordinates of the focii measured from the points XH and Xw are
fs = fos - XH and f~ = f~s - XiI. Using such a representation of the
coordinates of the optical system, the focii fs and f~ may be expressed
in the form:
1 lId 1 lId
------+- -=-+---
[» - fb fb fafb ' f~ fb fb fafb

Figure 6.56.

Since both representations are equivalent, we will use last form, as-
suming that the appropriate coordinates of the principal planes H, H'
are known. The focal length f may either be positive, if fa + fb - d > 0,
or negative, if fa + fb - d < O. In a particular case where d = fa + fb,
280 DEMONSTRATIONAL OPTICS

a parallel beam passing through the system will still stand parallel, as
shown in Fig.6 .56. In the case of a combination of a positive lens ( fa)
and a negative one (- fb), several special cases may also appear. For
example, if d = 0 and fa = - fb' the action of both lenses cancels out
and no refraction occurs. The same is true for fa - d = - fb .

6.3 Let us represent the focal length f of the system of two thin lenses
in the following form (see. Problem 6.2):
1 fI+h-d 1 1 d
7= fIh = fI + 12 - fIh .
The change of due to changes of refractive indices will be equal to zero
if of/ f = O. The variation of the right-hand side of the expression for
1/f gives the following:
of = ofI + 012 _ dofI _ doh _ 0
f fI 12 If 12 fdi - ,
or
of = ofI + 012 _ .s: (OfI + 012) = 0
f fI 12 fIh fI 12
By introducing the quantities of the longitudinal chromatic aberration
for both lenses, 6. 1 and 6.2, respectively, we can substitute 6fdh by
6. 1 and 612/12 by 6. 2, which gives

d = 126.1 + fI6.2
6.1 + ~2
for the appropriate distance d. In the particular case where both lenses
are made of the same material with ~1 = 6.2, the separation d = (fI +
12)/2 is roughly independent of wavelengths, at least between the blue
and red Fraunhofer lines.

6.4. Since the achromatic doublet is assumed to be composed of two


thin lenses in contact, d is equal to zero in the expression for the focal
length f (Problem 6.2) and in the expression obtained in Problem 6.3.
Therefore, we have to fulfill two conditions for the focal lengths of the
positive and negative lens:
1 1 1
-=-+-
f fI 12 '
where
Geometrical Optics 281

Here .6. 1 belongs to the positive crown lens and .6.2 to the negative flint
lens. For the focal lengths, one then can get

Substitution of the numerical values gives !I = 30.84 em and h =


-80.51 cm.

6.5 Let the laser beam fall on the first mirror M 1 at point A, as shown
in Fig.6.57. Since the incident beam is parallel to the optical axis of the
system, the reflected beam will intersect this axis at the focal point F . In
turn, the beam propagating from point F to the second mirror M2 will
pass parallel to the optical axis after reflection at a distance hI from this
axis. Because the focal length of a spherical mirror is equal to r /2, the
ratio h/h 1 is equal to that of the radii: h/h1 = rdr2' or hI = h(r2/rl);
hence, hI < h. As a result, the beam, passing between the mirrors, will
approach the optical axis.

~-----------------"">iC

Figure 6.57.

After the second pair of reflections, the distance between the passing
beam and the optical axis will be h 2 = h(r2/rl)2, after the third reflec-
tion h 3 = h(r2/rl)3 is observed, and so on. The beam can leave the
system after n pairs of reflections if the following inequality is true:

d > 2h n = 2h(r2/r;)n .
Since the approximation of paraxial rays is valid, one may regard all
angles between the traveling beams and the optical axis as being small, so
that the net path length l between two adjacent reflections , for example
the path ABC, is roughly equal to l = 2(rd2 + r2/2) = rl + r2. Hence,
if the beam leaves the system after n pairs of reflections , the net path
282 DEMONSTRATIONAL OPTICS

between the mirrors will be equal to in = n(rl + r2), and the delay may
be est imated by the following expression:

where c is light velocity.


For the given numerical magnitudes of h, d, rl and r2 we can calculate
t he value of n from the inequality 2n > 120. It is easy to find that n = 8
in order to satisfy the last condit ion. By using the expression for T , one
can estimate the time delay to be ~ 4 . 10-8 S = 40 ns.

6.6 We use the solution of Problem 5.4. where a relation between the
angular separation 8 and the difference b.n was obtained (for the case
where the refractive angles of both prisms are 300 ) :

We may assume no = 1.55, corresponding to red radiation. In turn, the


angular separation 8 may be expressed by the angle 'Y between the centers
of the two images of the vertical slit and by the angular magnification /3
of t he telescopic system:
'Y = 8/3 .
For the given distance L between the slit and the plane p" and the
linear separation s between the centers of the images, we estimate 'Y =
s] L = 2.5/900 ~ 2.8 . 10-3 rad. Let us calculate the distance b between
t he plane p' and the objective, using the relation: l/b + Y]« = 1/10 .
Substi tution of the numeric al values for a (120 em) and 10 (20 cm)
gives b = 24 em; hence, the magnification of the objective is /b/al = 0.2.
Nevertheless, the eye-piece provides a huge magnification, which may be
estimated by the value I(L-a)/ lepl ~ 1000, so that the net magnification
of the telescopic system is /3 ~ 200. Substitution of the numerical values
gives 8 = 1.4 . 10-5 rad. Finally, the quantity b.n may be estimated as
b.n = (V3/2)n o8 = 1.8 . 10-5 .

6.7. With the fixed position of the pinhole 8, so that its center lies
a little above the optical axis, the optical system is adjusted by a dis-
placement of the lens system along the optical axis to provide a distinct
image 8' of the pinhole at a position immediately adjacent to the pin-
hole. The pinhole and its image lie on the same plane if and only if this
plane is the focal plane of the optical system. That means, if the rays
fall parallel on the mirror M , they are then reflected back , as shown in
Fig.6.58. If this is the case, the distance from the apex of the first lens
to t he plane of the pinhole and image is equal to the magnitude los.
Geometrical Optics 283

M
f05 - - - - -

Figure 6.58.

6.8. According to the geometrical construction presented in Fig.6.12,


distances are measured from the apexes of the refracting surfaces and
will be regarded as positive if the appropriate distance is directed in the
propagation direction of the light rays, and as negative if directed in the
opposite direction.

H H'

--- -- Ft
Fs 81 82

Figure 6.59.

For the coordinates !l and fi, connected with the first surface which
has a positive optical force, and measured from the apex aI, we can
therefore write

!l=-~ and
n-1
The second surface of radius r2 has a negative optical force; hence, the
appropriate coordinates measured from the apex a2 are expressed as
follows:
f2 = -~
n -1
II _
J2 - -
nr2
n -1
In order to calculate the coordinates of the focii of the meniscus we find
the expressions for the focal lengths of the optical system, f~s and fos ,
284 DEMONSTRATIONAL OPTICS

using the formulae (6.21, 6.23):

f.' = nr2. nrl - (n - l)d and


os n - 1 (n - l)d + r2 - nrl

J. - _ nrl. r2 + (n - l)d
os - n - 1 (n - l)d + r2 - nrl
Both focal lengths are measured from the apexes of the refracting sur-
faces. In a similar way, the coordinates of the principal planes are cal-
culated by means of the expressions (6.22, 6.24):
d
Xw = -nr2 and
(n - l)d + r2 - nrl
d
XH = -rl-;----:--,-------
(n - l)d + r2 - nrl
We substitute r2 = 4rl for r2 and d = rI/2 for d in all expressions and
obtain

fos = - 3r l Xw = -rl and Xw = -O. 17r l


The meniscus and the positions of its focal planes and H -planes are
shown proportionally in Fig.6.59. It can be seen that this lens has a
positive optical force, so that rays drawn from an object placed in the
first focal plane will leave the lens as parallel rays.

6.9. We may use the results of the previous problem, considering the
convergent meniscus, but we have to take into account that rl is now
larger than r2 , opposite to the previous case.

-- - _. _ . _ . _ . _ . -0 - -_ . -
F.'
s Fs

H HI

Figure 6.60.

Taking into account this fact , we rewrite the magnitudes obtained for
the convergent meniscus in the following form:

f.' = _ nr2 . nrl - (n - 1)d


os n - 1 nrl - (n - l)d - r2
REFERENCES 285

fos = nrl. r2 + (n - 1)d


n - 1 nrl - (n - l)d - r2
d
X H , = nr2-----,...-----,.---
nrl - (n - l)d - r2
d
XH = rl--........,.-----..,.-,---
nrl-(n-1)d-r2
If d is rather small with respect to rl and, therefore, nrl - (n -l)d - r2 >
0, then f~s will be negative and fos positive. This means that the first
focal plane lies on the right-hand side of the lens and the second focus
on the left-hand side. However, planes H and H' are both located on
the right-hand side of the appropriate apexes (Fig .60). Substitution of
the relations rl = 4r2 and d = r2/2 , and using n = 1.5 gives

fos = 2.7r2 , f~s = -3. 6r 2


We see that this meniscus has a negative optical force. If an object is
placed on the second focal plane F'; the top point of the short line will
be imaged by the meniscus through a system of parallel rays . A pair of
these rays is shown in Fig.6.60. The first ray, travelling parallel to the
optical axis, is deflected away from the optical axis, so that the reverse
continuation of this ray intersects the axis at the second focus. The
other ray, leaving the top of the short line at some angle, will also be
deflected away from the optical axis and propagate outside the meniscus
parallel to the first ray.

References

[1] Ch. Huygens, Traite de la lumiere. Leyden 1670. Ostwald's Klassiker No 20

[2] T .Young, Phil. Trans Roy. Soc. (London) 1802, v.12, p. 387

[3] A.Fresnel, Ann. Chern. et Phys . 1816, (2), 1, Olwres, vol.I

[4] M.Faraday. Experimental Researchs in Electricity, London 1839

[5] J.K.Maxwell, A Treatise of Electricity and Magnetism. 2 Vols, Oxford 18.

[6] M.Planck , Vert. d. Deutsch Phys. ber .2, 1900,202, 237; Ann d.Phys. (4)

[7] A.Einstein , Ann d. Phys . (4), 17, 1905, 132; 29, 1906, 199.

[8] N.Bohr, Phil. Mag. (6) 26 1913; 1,476,857

[9] A.A.Michelson, E.W.Morley, Amer. J . Sci. (3), 34, 833, 1887; Phil. Mag.12, 449,
1887
286 DEMONSTRATIONAL OPTICS

[10] A.A.Michelson, Phil. Mag. 34, 291, 1892

[11] A.A.Michelson, Astrophys. J . 51, 257, 1920

[12] P.H.Van Cittert, Physica I, 201, 1934; 6, 1129, 1939

[13] F .Zernike, Physica 5, 785, 1938

[14] R.Hanbury-Brown, R.Q.Twiss, Phil. Mag. 45, 603, 1954

[15] R.Hanbury-Brown, R.Q .Twiss, Nature 7, 1956

[16] R.Hanbury-Brown , R.Q.Twiss , Nature 45, 1956

[17] E .Wolf., Proc . Roy. Soc. A, 230, 1955; Nuovo Cimento 12, 1954

[18] A.Forrester, R.Gudmundsen, P. Johnson , Phys . Rev. 99, No 6, 1955

[19] J .R.Klauder, E.C.G.Sudarshan, Fundamentals of Quantum Optics, W .A. Ben-


jamin NY, 1968

[20] E.L.O'Neill, Introduction to Statistical Optics. NY Addison , Reading 1963

[21] J .W.Goodman, Statistical Optics, Whiley Int .Publ. NY 1985

[22] M.Born , E.Wolf, Principles of Optics , 2nd edition , Pergamon Press NY 1964

[23] H.F.Meiners, Physics demonstration Experiments, Roland Press Corp . NY 1970

[241 M.Francon , N.Krauzman, J .P. Mathien, M.May, Experiments in Physical Optics,


Gordon and Breach, NY, London, Paris 1970

[25] Pohl, R.W. Einfuhrung in die Optik. Vierte und furfte verbesserte anflage. Berlin ,
Springer - Verlag, 1943 (Einfuhrung in die Physik ).

[26] Cherenkov P.A. Soviet Academy of Sciences Doklady 1934, v.2, p.451;

[27] Vavilov S.N. Soviet Academy of Sciences Doklady 1934, v.2, p. 457.

[28] Tamm I.E ., Franck I.M. Soviet Academy of Sciences Doklady 1937, v.14, p. 107.

[29] Ginzburg V.L., Franck I.M. JETP 1946, v.16, p.15.

[30] Bracewell R.H. The Fourier Tansformation and its Applications. 2-nd ed.- N.Y.:
Mc.Graw-Hili Book Company 1979.

[31] Papolis A. Systems and Transformations with Applications in Optics.- N.Y.:


Mc.Graw-Hili Book Company 1968.

[32] Zernike F. , Zs.Techn.Phys . 16,454 (1935).

[33] Porter A.B., Phil.Mag. (6), 11, 154 (1906)

[34] Elias P., Gray D.S., Robinson D.Z., Journ, Opt . Soc. Amer. 42, 127, (1952)

[35] Elias R., Journ. Opt. Soc. Amer. 42, 229, (1953)
REFERENCES 287
[36] R. Hanbury Brown, R.C.Jennison, M.K. Das Gupta. Nature 170, 1061 (1952)

[37] M.Born , E.Wolf. Principles of Optics. 557 - 560. Cambridge University


Press ,.1999.

[38] ibid, 619 - 621.

[391 ibid, 25 - 28.

[40] ibid, 790 - 796.

[41] E .Kreyszig. Advanced Engineering Mathematics. 482 - 488. N.Y. Jonh Wiley &
Sons, Inc, 1993.

[421 ibid, 522 - 529.

[43] ibid, 544,556.


Index

Aberration, 247 Cauchy formula, 131


for nonaxial beams, 249 Caustic
longitudinal spherical, 248 line, 222
chromatic, 251-252, 259 surface, 222
spherical, 249, 251, 254-255 , 258, 260 Chromatic aberration
Accomodation, 253 axial ,251
Accommodation, 254, 257-258 lateral, 252
Achromat, 252-253 Coherence
Acting diaphragm, 261 theory, 14
Aether, 3, 6, 13, 72 length, 13, 49
Analytical signal , 41 Coma, 249-251
Analyz er, 79-80 Cornea, 254
Angle Correlation
of incidence , 107-108 measurements, 16
of reflection, 108 Concave surface, 228-229 , 231
Aperture Convex surface, 227-228
ent rance, 260 Conjugated points, 226, 228
stop, 261 Crystal
Apex of reflecting surface, 223
ADP , 191
Ast igmatic difference , 222
calcite, 164, 167, 169-170, 175
Ast igmatism, 222-223 , 251
optical axis, 166-167
Bartholinus, 2, 163
KDP, 191-194
Beam
astigmatic, 219, 222 liquid, 184-187
normal astigmatic, 222 mica, 181-182
focus, 220 negative, 173-174
homocentric, 219, 222, 225, 227 positive, 173-174
paraxial, 223, 225, 227 quartz natural, 195
pencil, 223, 225, 227, 253 quartz left - rotating, 196-197
Biot - Savart - Laplace law, 18, 23 quartz right - rotating, 196-197
Birefringence, 2, 163 uniaxial, 173, 176, 179, 195
artificial, 187 Crystalline quartz, 173
Black body De Broglie, 12
radiation , 10 Diffraction, 2, 220
Bohr, 11 Diffraction grating, 7
Boltzmann, 9 Dipole moment, 55
Brewster Dipole radiation, 23, 57, 90-92
angle , 90, 123 Dispersion, 126
window, 124-125 anomalous, 130
Carrier frequency, 37-38, 44-45, 50-52, 137 curve, 129-130

289
290 DEMONSTRATIONAL OPTICS

norm al, 130 Huygens-Fresnel principle, 4, 146


Displacement current , 6, 19 Image
Distortion, 249-251 imag inary, 229-231
Doppler, 46 real, 229-231
effect, 46, 48-49 Inhomogeneous media, 264-265, 267, 270,
Eikon al , 265, 267-268 273-274
equation, 268 Intensity, 44-45
Einst ein , 9-10, 46, 71-72 Interference
Electrodynamic const ant , 18 pattern, 4, 13-14
Electromagnetic contrast, 14
en ergy, 41 Interferometer, 8
Elliptically polarized wave, 76-78 Fabry-Perot, 8
Energy Jamin, 8
den sity, 41-42 Mach-Zender, 8
quantization, 10 Michelson stellar, 13
Entrance pupil , 254, 262 Twyman-Green, 8
Entropy, 9 Brown -Twiss stellar, 15
Exit pupil, 263-264 stellar, 13
Eye-piece Jeans, 9
double lens, 263 Johnson , 16
Huygenian, 244-246, 257, 263-264 J ones vectors, 82
Ramsden, 244, 246, 257, 263-264 Kerr
Faraday, 19,23 cell, 190-191
concept, 6 effect, 188-190
effect , 202, 205 Kirchhoff, 9
rot ator, 203 Laplace op erator, 24
Fermat's prin ciple, 272, 274 Lens
Field ball, 239-240
lens , 263 convex-convex, 238
stop, 263-264 negative, 239
Fizeau, 29 positive, 239
Focal thick , 238-240
line, 222 thin, 241-242
Focal length, 225-226 Light, 1
optical syst em, 235 beam, 2
Focal plan e beam bending, 270
first, 242 diffra ction , 2, 4
second ,243 pressure, 7, 54, 59
Focus, 225 velocity, 6, 28-29
first , 229 wave momentum, 58, 60, 62
second principal, 225 wave angular momentum , 62-63
Forr ester, 16 Lorentz, 7, 72
Fresnel, 4, 71 curve , 51, 130
diffra ction, 4 force, 56-57
exp erim ent, 4, 197 electron theory, 7
formulae, 6, 119-121 Mach-Zender, 8
Fraunhofer, 7 Magn etic
Galil eo domains, 203-205
teles copic system, 257 induction, 17-18, 23
Gauss, 17 Magnification
Gudmundsen, 16 angular, 258
Heisenberg, 11 linear, 230, 260
Human eye, 253 Magnifier, 255, 259
Hanberry-Brown, 15-16 Material equations, 102
Huygens, 2-3 Maxwell, 6, 17, 72
geometrical repr esentation, 2 equat ions, 17, 19, 24, 104
principle, 2, 4 Michelson, 8, 12-13, 71
INDEX 291
interferometer, 14 Pockels
Microscope, 259 cell, 192-1 93
Mirages, 271 effect, 191
Monoc hromatic wave Polar ization
plane monochrom at ic wave, 28 circular, 35 -3 6, 73, 75
spherical monochromat ic wave, 33 degree , 81-82
Monochromaticity, 49, 51 ellipt ical, 74, 76
degree, 51, 59 linear, 35-36
Morl ey, 12 partial , 81
Natu ral light, 5, 52-5 3, 87-88 plan e, 35-3 6
Newton formula, 226, 231 rotation , 196
Newtonian telescop e system, 258 st ate , 35, 73-74
Norr enb erg polariscop e, 123 Polariz er, 78-81
Optical Poynting vect or , 42, 58, 103, 171
axis main, 225, 229 Planes principal , 234-2 36
beats , 16 Pr ism
acti vity, 196-197,200-202 liquid cryst al, 185-186
fiber, 113-114 Nicol, 175-176
for ce, 225, 227-228 refraction, 108
force negative, 226, 228 Wollaston, 175-176
force positive, 227-228 Propagation
shutter, 192-193 numb er , 28
sur face, 223 Quarter-wave plat e, 181-182
Opt ical syst em Quantum
centere d, 233-234 th eory, 12
plan es, 234 mechanics, 12
Oscillations Quasi-mon ochr omat ic wave, 37, 44, 46,
quasi -monochromatic, 37-3 8 49- 52
Par ax ial Rad io
ap proximation , 224 inte rferomete r, 15
Par t icle concept of light , 10 wave rota tion , 200
Permeabili ty Ra msden eye-piece, 244-246
magnetic, 6, 18, 102-103 Ra y
Permi ttivity extraordinary, 2, 163, 165
electric, 6 ordi na ry, 2, 163, 165
dielectr ic, 102 Rayleigh,9
free spa ce, 17 formula, 140
Ph ase Reflect ion
of plan e mono chrom atic wave, 29, 32 law, 107
plan e, 27 tot al, 111, 110-11 4
initial ph ase, 28 Reflect ivit y
Ph ot ocurrent , 15, 43-45 power , 119
Phot odet ector, 43-45 Refra ction, 2
Ph ot oeffect, 10 astronomical, 272-273
Ph ot o- elastic effect, 187 law, 108
Ph ot oelectron, 42-44 Refractive
Ph ot omultiplier, 16 index, 108
Ph ot ons , 10, 12, 60-61 index gradient , 272
energy, 60-61 index relative, 225, 227-228
mom entum , 61 Re lat ivity
angular momentum, 61- 62 t heory, 12
Physical field , 39 Ret ardation plate, 179- 180
Planck, 8- 10 Rowla nd ,8
constant, 12, 61 Rutherford, 10
Plane mon ochromatic wave, 28 plan etary model of atom , 10
Plan e of incidence, 107 SchrOdinger , 12
Pl an e wave, 27 Smith -Par sell effect, 150
292 DEMONSTRATIONAL OPTICS

Snell's law, 107 Velocity


Sommerfeld, 11 of electromagnetic waves, 6
Spectrum group , 136-141, 143
line, 110 phase, 172-174
continuous, 110, 133-134 ray, 173
Spherical Verdet constant, 203
rel1ecting mirror, 231 Visibility function, 14
rel1ecting surface, 231 Wave, 2
refracting surface, 229-230 concept of light, 1, 10
Stefan, 9 evanescent, 110
Stokes front, 2, 32
parameters, 82, 84-86, 92-98 secondary, 2
Superposition spherical, 33-34
principle, 2 superposition, 3
surface , 110-111
Telescope
train , 47, 140-141
rel1ecting, 258-259
transversal, 6
refracting, 256
vector, 32
terrestrial, 256-257
wavelength, 28
Transparency, 119 Wien, 9
Twiss , 15 Wolf,16
Twym an -Green, 8 Wood experiment, 133
Van-Zittert, 14 Young, 3, 71
Vavilov-Cherenkov radiation, 144-147 experiment, 3
Zeeman, 7

You might also like